1: % ****** Start of file apssamp.tex ******
2: % See the REVTeX 4 README file for restrictions and more information.
3: %
4: % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
5: % as well as the rest of the prerequisites for REVTeX 4.0
6: %
7: % See the REVTeX 4 README file
8:
9: \documentclass[prb,twocolumn,showpacs,amsmath,amssymb,eqsecnum,superscriptaddress]{revtex4}
10:
11: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
12:
13: % Some other (several out of many) possibilities
14: %\documentclass[preprint,aps]{revtex4}
15: %\documentclass[preprint,aps,draft]{revtex4}
16: %\documentclass[prb]{revtex4}% Physical Review B
17:
18: \usepackage{graphicx}% Include figure files
19: \usepackage{dcolumn}% Align table columns on decimal point
20: \usepackage{bm}% bold math
21: \setlength{\parskip}{0pt}
22: %\nofiles
23:
24: \begin{document}
25:
26: \newcommand{\lin}{l_{\textrm{in}}}
27:
28: \newcommand{\me}{m_{\textrm{e}}}
29:
30: \newcommand{\tperp}{t_{\perp}}
31:
32: \newcommand{\Bperp}{B_{\perp}}
33:
34: \newcommand{\dagg}{^{\dagger}}
35:
36: \newcommand{\phdagg}{^{\phantom{\dagger}}}
37:
38: \newcommand{\lB}{l_{\textrm{B}}}
39:
40: \newcommand{\lel}{l_{\textrm{el}}}
41:
42: \newcommand{\LT}{L_{\textrm{T}}}
43:
44: \newcommand{\vF}{v_{\textrm{F}}}
45:
46: \newcommand{\kB}{k_{\textrm{B}}}
47:
48: \newcommand{\U}{\mathcal{U}}
49:
50: \newcommand{\Uzero}{\mathcal{U}_{\,0}}
51:
52: \newcommand{\Ulin}{\mathcal{U}_{\,\textrm{lin}}}
53:
54: \newcommand{\V}{\mathcal{V}}
55:
56: \newcommand{\Vzero}{\mathcal{V}_{\,0}}
57:
58: \newcommand{\Vlin}{\mathcal{V}_{\,\textrm{lin}}}
59:
60: \newcommand{\modsq}[1]{\vert #1 \vert^2}
61:
62: \newcommand{\Tr}{\textrm{Tr}}
63:
64: \newcommand{\hberg}{^{(H)}}
65:
66: \newcommand{\Ham}{\mathcal{H}}
67:
68: \newcommand{\order}[1]{\mathcal{O}(#1)}
69:
70: \newcommand{\overlap}[2]{\langle #1\vert #2\rangle}
71:
72: \newcommand{\ket}[1]{\vert #1\rangle}
73:
74: \newcommand{\bra}[1]{\langle #1\vert}
75:
76: \newcommand{\pages}[1]{ \emph{(#1 pages)}}
77:
78: \newcommand{\mb}[1]{\mathbf{#1}}
79:
80: \newcommand{\subtxt}[1]{_{\textrm{#1}}}
81:
82: \newcommand{\bmat}{\begin{displaymath}}
83:
84: \newcommand{\emat}{\end{displaymath}}
85:
86: \newcommand{\bit}{\begin{itemize}}
87:
88: \newcommand{\eit}{\end{itemize}}
89:
90: \newcommand{\beq}{\begin{equation}}
91:
92: \newcommand{\eeq}{\end{equation}}
93:
94: \newcommand{\bspl}{\begin{split}}
95:
96: \newcommand{\espl}{\end{split}}
97:
98: \newcommand{\twomat}[4]{\ensuremath{\left( \begin{array}{rr} #1 &
99: #2\\#3 & #4\end{array}\right)}}
100:
101: \newcommand{\threemat}[9]{\ensuremath{\left( \begin{array}{rrr} #1 &
102: #2 & #3\\#4 & #5 & #6\\#7 & #8 & #9 \end{array} \right)}}
103:
104: \newcommand{\twomatc}[4]{\ensuremath{\left( \begin{array}{cc} #1 &
105: #2\\#3 & #4\end{array}\right)}}
106:
107: \newcommand{\threematc}[9]{\ensuremath{\left( \begin{array}{ccc} #1 &
108: #2 & #3\\#4 & #5 & #6\\#7 & #8 & #9 \end{array} \right)}}
109:
110: \newcommand{\twovec}[2]{\ensuremath{\left( \begin{array}{r} #1 \\
111: #2\end{array}\right)}}
112:
113: \newcommand{\threevec}[3]{\ensuremath{\left(\begin{array}{r}
114: #1\\#2\\#3\end{array}\right)}}
115:
116: \newcommand{\augmat}[8]{\ensuremath{\left( \begin{array}{rr|rr}
117: #1\\#5 \end{array}\right)}}
118:
119:
120: \preprint{APS/123-QED}
121:
122: \title{Transport between edge states in multilayer integer
123: quantum Hall systems:\\
124: exact treatment of Coulomb interactions and disorder}
125: %The Rudolph Peierls Centre for
126: \author{J. W. Tomlinson} %\email{tomljw@thphys.ox.ac.uk}
127: \affiliation{Theoretical Physics, University of Oxford, 1 Keble
128: Road, OX1 3NP, United Kingdom.}
129: \author{J.-S. Caux} %\email{jcaux@science.uva.nl }
130: \affiliation{Institute for Theoretical Physics, University of
131: Amsterdam, Valckenierstraat 65, 1018 XE Amsterdam, The
132: Netherlands.}
133: \author{J. T. Chalker} %\email{chalker@thphys.ox.ac.uk}
134: \affiliation{Theoretical Physics, University of Oxford, 1 Keble
135: Road, OX1 3NP, United Kingdom.}
136:
137: \date{\today}
138: % It is always \today today, but any date may be explicitly specified
139:
140: \begin{abstract}
141:
142: A set of stacked two-dimensional electron systems in a perpendicular magnetic field
143: exhibits a three-dimensional version of the quantum Hall effect if interlayer
144: tunneling is not too strong. When such a sample is in a quantum Hall
145: plateau, the edge states of each layer combine to form a chiral
146: metal at the sample surface. We study the interplay of interactions
147: and disorder in transport properties of the chiral metal,
148: in the regime of weak interlayer tunneling. Our starting point is
149: a system without interlayer tunneling, in which the only excitations
150: are harmonic collective modes: surface magnetoplasmons.
151: Using bosonization and working perturbatively in the interlayer tunneling amplitude,
152: we express transport properties in terms of the spectrum for
153: these collective modes, treating electron-electron interactions
154: and impurity scattering exactly.
155: We calculte the conductivity as a function of temperature,
156: finding that it increases with increasing temperature
157: as observed in recent experiments. We also calculate the
158: autocorrelation function of mesoscopic conductance fluctuations
159: induced by changes in a magnetic field component perpendicular
160: to the sample surface, and its dependence
161: on temperature. We show that conductance fluctuations are characterised by
162: a dephasing length that varies inversely with temperature.
163:
164:
165: \end{abstract}
166:
167: \pacs{73.20.-r, 73.23.-b, 72.20.-i, 73.21.Ac}
168:
169: %73.20.-r Electron states at surfaces and interfaces
170: %73.23.-b Electronic transport in mesoscopic systems
171: %72.20.-i Conductivity phenomena in semiconductors and insulators
172: %73.21.Ac Electron states and collective excitations in multi-layers
173:
174: %\keywords{Suggested keywords}
175: %Use showkeys class option if keyword display desired
176:
177: \maketitle
178:
179:
180:
181:
182: \section{Introduction}\label{sec:intro}
183:
184: Multilayer quantum Hall systems offer a setting in which
185: to study the influence of electron-electron interactions
186: and impurity scattering on tunneling between
187: quantum Hall edge states. Specifically, consider
188: a layered conductor in a magnetic field that is perpendicular
189: to the layers, with the field strength chosen so that a
190: single layer in isolation would have quantised Hall conductance.
191: Then, if interlayer tunneling is not too strong, the multilayer
192: system exhibits a three-dimensional version of the
193: quantum Hall effect and the bulk is insulating at low temperatures.
194: Under these conditions, edge states are present in each layer at the
195: sample surface and are coupled by interlayer tunneling to form a surface phase,
196: which is a chiral, two-dimensional metal.\cite{Dohmen,Balents/Fisher}
197: The contribution of this surface phase to the interlayer electron
198: transport properties of such systems has been isolated in experiments
199: on semiconductor multilayers,\cite{UCSB1}
200: and is dominant if samples are sufficiently small and cold.
201:
202: The consequences of impurity scattering for transport in the chiral metal
203: have been discussed extensively from a theoretical viewpoint
204: \cite{Dohmen,Balents/Fisher,Mathur,Gruzberg,Cho,Plerou,Sondhi,Betouras}
205: and have been probed experimentally in several
206: ways.\cite{UCSB1,UCSB0,MULTILAYER,UCSB9,Kuraguchi/Osada,UCSB5,UCSB7,UCSB2,UCSB11,UCSB10,UCSB12}
207: Crucially, the chiral motion of electrons along the layer edges
208: means that localisation is suppressed.\cite{Dohmen,Balents/Fisher}
209: As a result, the surface conductivity in the interlayer direction
210: has a low-temperature limit that is non-zero, even though its
211: measured value may be much smaller than $e^2/h$.\cite{UCSB1,MULTILAYER,Kuraguchi/Osada}
212: Separately, theoretical discussions
213: of conductance fluctations\cite{Mathur,Gruzberg,Cho,Plerou,Betouras}
214: have examined both their dependence on geometry in fully
215: phase-coherent samples, and their dependence on the inelastic scattering length
216: when this is smaller than sample size.
217: Observations of reproducible mesoscopic conductance
218: fluctuations,\cite{UCSB9,UCSB12}
219: induced by small changes of magnetic field
220: within a quantum Hall plateau, demonstrate that interlayer
221: hopping is quantum-mechanically coherent and also provide a
222: way to determine the inelastic scattering length.
223: In addition, magnetoresistance in response to a field component
224: perpendicular to the sample surface has been proposed\cite{Sondhi}
225: and used\cite{UCSB5,UCSB7,UCSB11} as a method for measuring the
226: elastic scattering length.
227:
228: In contrast to these studies of disorder effects,
229: past theoretical work on effects due to electron-electron interactions
230: in the chiral metal has been limited. There have
231: been discussions, first, of the
232: temperature dependence of the inelastic
233: scattering length\cite{Balents/Fisher,Betouras}
234: and, second, of the fact that there is no zero-bias
235: anomaly in the tunneling density
236: of states (or any related contribution to the conductivity), because
237: of ballistic motion of charge in the in-layer
238: direction.\cite{Balents/Fisher,Betouras}
239:
240: Against this background, recent experiments finding a significant temperature
241: dependence to the surface conductivity\cite{UCSB2,UCSB10} are striking as
242: likely indications of interaction effects, and provide one of the motivations
243: for the work we present here. In particular, the fact that conductivity
244: is observed to {\it increase} with increasing temperature presents a puzzle for theory.
245: Some straightforward potential explanations are specifically excluded
246: by the experimental design: large ratios of sample perimeter to cross-sectional area
247: ensure that surface states make the dominant contribution to the measured
248: conductance; and sample perimeters much longer than the inelastic scattering length
249: ensure that weak localisation effects are absent. For samples
250: studied in Ref.~\onlinecite{UCSB10}, the measured conductivity $\sigma(T)$
251: increases by about $7\%$ in the temperature range from $50$mK to $300$mK,
252: implying a temperature scale of
253: $\sigma(T)\cdot [d\sigma(T)/dT]^{-1} \sim 4$K,
254: which is similar to that for other
255: interaction effects in quantum Hall systems
256:
257: In this paper we study interactions and disorder in the
258: chiral metal, working in the experimentally-relevant limit
259: of weak interlayer tunneling. Treating tunneling perturbatively,
260: Coulomb interactions and impurity scattering can be handled
261: exactly by means of a straighforward application of bosonization.
262: We calculate the full temperature dependence of the conductivity.
263: We also study conductance fluctuations induced by magnetic field changes,
264: obtaining their autocorrelation function and its dependence on
265: temperature. Making appropriate parameter choices, our results
266: for both quantities are consistent with
267: experimental findings. A short account of this work
268: has been presented previously, in Ref.~\onlinecite{PRL}.
269:
270: Our work differs from most of the extensive literature on
271: tunneling between quantum Hall edges states in two important ways.
272: First, while much previous work has been concerned with edge
273: states of fractional quantum Hall systems,\cite{Wen1,Wen2,Wen3,EdgeReview}
274: including multilayer
275: samples,\cite{Naud1,Naud2}
276: our focus is on the integer quantum Hall effect.
277: Second, whereas most past work (with some exceptions: see
278: Refs.~\onlinecite{Moon,Zulicke,Oreg,Pryadko}) has been restricted to systems
279: with only short-range interactions, we find that the long-range
280: nature of Coulomb interactions, which we treat in full, is
281: central for the results we obtain.
282:
283:
284: The remainder of this paper is organised as follows. We develop a
285: model for the chiral metal in Sec.~\ref{sec:model} and show how
286: bosonization can be used to give an exact description of the
287: collective excitations. Sec.~\ref{sec:cond} contains calculations
288: of the temperature dependence of the conductivity. We study
289: conductance fluctuations in Sec.~\ref{sec:condflucs}, and discuss
290: our results in Sec.~\ref{sec:discussion}.
291:
292:
293:
294: \section{Modelling the chiral metal}\label{sec:model}
295:
296: In this section we summarise the physical ingredients
297: that are important for modelling transport between edge
298: states in multilayer conductors and set out the lengthscales
299: that characterise the system. We introduce a Hamiltionian
300: in terms of fermionic operators for edge electrons. We
301: bosonize this Hamiltonian, obtaining a result which
302: is quadratic in boson operators if interlayer tunneling
303: is omitted. Finally, we express the two-electron correlation function
304: that is central to transport calculations in terms of
305: boson correlators.
306:
307: \subsection{Ingredients, lengthscales, and parameters}\label{ssec:model:ingreds}
308:
309: A multilayer conductor is illustrated in Fig~\ref{fig:stack}.
310: We use coordinates with the $x$-axis parallel to the
311: layer edges, and treat a sample of $N$ layers with layer index $n$
312: and layer spacing $a$. Consider the system in the presence of a
313: perpendicular magnetic field of strength $B$, with the chemical
314: potential lying between the lowest and first excited Landau levels.
315: In the bulk of the sample single particle states at energies
316: close to the chemical potential are localised by disorder.
317: At the sample surface in this energy range, edge states
318: propagate in the confining potential $V_{\rm edge}(y)$ at a velocity $v$.
319: Interactions modify the confining potential and the edge velocity:
320: we denote by $v_{\rm F}$ the velocity allowing for Hartree contributions.
321: Edge states have a width $w$ in the $y$-direction, which is set by the magnetic length
322: $l_{\rm B}$ in a clean sample, and by the bulk localisation length $\xi$
323: in the presence of impurities. We use a one-dimensional decription
324: of the edge state in each layer, projected onto the $x$-coordinate in the standard
325: way.
326:
327: Out theoretical treatment takes account only of one edge state
328: in each layer and is therefore appropriate for a
329: system in which electrons are spin polarised.
330: %or alternatively one in which electrons with opposite spin directions
331: %contribute additively and incoherently to the
332: %conductivity.
333: In fact, some of the experiments we refer to,
334: including those on the temperature-dependence of conductivity,\cite{UCSB10}
335: are for systems with Landau level filling factor per layer of $\nu=2$.
336: It is appropriate to apply our theory
337: to these systems provided electrons with opposite spin directions
338: contribute additively and incoherently to the
339: conductivity.
340:
341: The system of edge states can be characterised using three lengthscales.
342: First, impurities, which generate only forward scattering
343: with a phase shift, result in
344: an elastic mean free path $l_{\rm el}$, the distance over which a phase shift
345: of order $2\pi$ is accumulated.
346: Second, temperature $T$ in combination
347: with the velocity $v_{\rm F}$ can be expressed in terms of the thermal
348: length $L_{\rm T} = \hbar v_{\rm F}/k_{\rm B}T$. Third, interlayer tunneling
349: with amplitude $t_\perp$ can be parameterised
350: by the characteristic distance $l_\perp$ through which electrons move
351: in the chiral direction between tunneling events. The value of
352: $l_\perp$ can be expressed in terms of the interlayer diffusion constant
353: $D$: since, for small $t_\perp$, interlayer hops are of length $a$ and occur at
354: a rate $v_{\rm F}/l_\perp$, one has $l_\perp = a^2 v_{\rm F} / D$.
355: In turn, this can be expressed in terms of the conductivity,
356: using the Einstein relation and the fact that the
357: density of states is $n=1/2\pi a\hbar v_{\rm F}$, giving
358: $l_\perp = a (e^2/2\pi \hbar\sigma)$.\cite{Betouras}
359:
360: Parameter values for the experiments of Refs.~\onlinecite{UCSB1}, \onlinecite{UCSB11}
361: and \onlinecite{UCSB10} are as follows. Samples consist of $N \sim 50$ -- $100$
362: layers with spacing $a=30$nm. The mean free path is estimated \cite{UCSB11} to be
363: $l_{\rm el} \sim 30$nm. An upper bound on $v_{\rm F}$, reached in samples
364: with a steep confining potential is $v_{\rm F}\sim\omega\subtxt{C}l_{\rm B}$,
365: where $\omega\subtxt{C}$ is the cyclotron frequency.
366: It has the value
367: $\omega\subtxt{C}\lB=1.7\times10^5\textrm{ms}^{-1}$
368: in GaAs at 6.75 T.
369: With this value, $L_{\rm T}\sim 10\mu$m at $T=100$mK.
370: Finally, for a surface conductivity of $\sigma = 1.3 \times 10^{-3}e^2/2\pi \hbar$
371: (which lies within
372: the observed range at $\nu=2$), $l_\perp = 40\mu$m. We are therefore concerned with the
373: regime $\lel\ll\LT\ll l_\perp$, and this motivates our approach,
374: based on a perturbative treatment of tunneling.
375: \begin{figure}
376: \begin{center}
377: \includegraphics[width=0.40\textwidth]{1.eps}
378: \caption{\label{fig:stack} A multilayer conductor, showing the orientation of
379: axes in our coordinate system, with edge states propagating in the $x$-direction.
380: The form of the confining potential $V_{\rm edge}(y)$ is illustrated top left.
381: Interlayer tunneling amplitude and spacing are denoted by $t_\perp$ and $a$,
382: respectively. }
383: \end{center}
384: \end{figure}
385:
386:
387:
388: \subsection{Fermionic Hamiltonian}\label{ssec:model:fermi}
389:
390: Our model Hamiltonian,
391: $\Ham=\Ham_0+\Ham_{\textrm{dis}}+\Ham_{\textrm{hop}}+\Ham_{\textrm{int}}$,
392: has single-particle terms $\Ham_0$, $\Ham_{\textrm{dis}}$
393: and $\Ham_{\textrm{hop}}$, representing, respectively, free motion along
394: each edge, impurity scattering and interlayer hopping,
395: and a contribution $\Ham_{\textrm{int}}$ from Coulomb interactions.
396: We write it in terms of the electron creation operator $c\dagg_{qn}$
397: for an edge state with wavevector $q$ in layer $n$, taking sample
398: perimeter $L$ so that $q=2\pi n_q/L$, where $n_q$ is integer.
399: The creation operator at a point is
400: \begin{equation}
401: \psi_n^{\dagger}(x)=\frac{1}{\sqrt{L}}\sum_{q=-\infty}^{\infty}
402: e^{-i q x }c_{qn}^{\dagger}\label{eq:ctopsi}\,.
403: \end{equation}
404: We normal order the Hamiltonian with respect to a vacuum in
405: which states are occupied for $q\leq 0$ and empty otherwise.
406: Then
407: \begin{eqnarray}
408: \Ham_0&=&
409: %\hbar v\sum_{q,n}q :\!c^{\dagger}_{q,n}c_{q,n}\!:
410: %\label{eq:H0c}\\
411: %&\equiv&
412: -i\hbar v\sum_n\int dx
413: :\!\psi^{\dagger}_n(x)\partial_x\psi_n(x)\!:\label{eq:H0psi}\,,
414: \end{eqnarray}
415: and
416: \begin{align}
417: \Ham_{\textrm{hop}}&=
418: %\tperp\sum_n\sum_q(c\dagg_{q+1,n}c\phdagg_{qn}+\textrm{h.~c.})\\
419: %&\equiv
420: \sum_n\int dx [t_{\perp}\psi^{\dagger}_{n+1}(x)
421: \psi_n(x)+\textrm{H.~c.}]\,.
422: \end{align}
423: The interaction contribution, written in terms of the projected density
424: $\rho(x)=\psi^{\dagger}_n(x) \psi_n(x)$ with a two-particle potential
425: $U_{n-m}(x-x')$, is
426: \begin{equation}
427: %\begin{split}
428: \Ham_{\textrm{int}}
429: %&=\frac{1}{2L}\sum_{nm}\sum_{kpg}J(q,p,m)\\
430: %&\quad\times:c\dagg_{k+q,n}c\dagg_{k+p,m+n}
431: %c\phdagg_{k+p+q,m+n}c\phdagg_{k,n}:\label{eq:Hintkspace},
432: %\end{split}
433: =\frac{1}{2}\sum_{nm}\int dx \int dx' :\rho_n(x) U_{n-m}(x-x')\rho_m(x'):\,.
434: \end{equation}
435: %where the exact expression for the vertex $J$ is complicated by
436: %the non-trivial $y$ dependence of the wavefunctions. It
437: %simplifies though, given that only a small band of states are
438: %active, to
439: %\begin{equation}
440: %J(p,q,m)=\eta(p)\int dx\,e^{iqx}U_m(x,w)\label{eq:defJ}.
441: %\end{equation}
442: %The function $\eta(p)$ is unity for small $\vert p\vert$ and
443: %decays on a scale far larger than any momentum separating states
444: %within the active band. It is required due to complications
445: %arising from the infinite Fermi-sea. The depth of the chiral
446: %surface $w$ comes from assuming that this is the average
447: %separation in $y$ of any pair of electrons.
448:
449: Finally, writing the impurity potential projected
450: onto the edge coordinate in the $n$th layer as $V_n(x)$,
451: we have
452: \begin{equation}
453: \Ham_{\textrm{dis}}=\sum_n\int dx
454: V_n(x):\psi^{\dagger}_n(x)\psi_n(x):
455: \, .
456: \end{equation}
457: We take $V_n(x)$ to be Gaussian distributed with
458: zero-range correlations and strength $\Delta$:
459: $[V_n(x)]_{\textrm{av}}=0$ and
460: $[V_n(x)V_{n'}(x')]_{\textrm{av}}=\Delta\delta_{n,n'}\delta(x-x')$.
461: This disorder term can be removed by
462: means of a gauge transformation on the fermionic field
463: operators, under which
464: \begin{equation}
465: \psi_n^{\dagger}(x)\to
466: e^{i\theta_n(x)}\psi_n^{\dagger}(x)\label{eq:gaugexform},
467: \end{equation}
468: where
469: \begin{equation}
470: \theta_n(x)=\frac{1}{\hbar v}\int_0^xdx'V_n(x')
471: \end{equation}
472: is the phase shift acquired under forward scattering
473: from the impurities. The elastic scattering length is
474: related to the disorder strength $\Delta$ by $\lel=\hbar^2
475: v^2/\Delta$. Under this gauge transformation,
476: $\Ham_0+\Ham\subtxt{dis}\to\Ham_0$. The hopping term, however, picks
477: up a dependence on the disorder, and after the transformation
478: is
479: \begin{equation}
480: \Ham_{\textrm{hop}}=\sum_n\int dx
481: [\tperp(n,x)\psi^{\dagger}_{n+1}(x)
482: \psi_n(x)+\textrm{H.~c.}]\label{eq:Hhop},
483: \end{equation}
484: where
485: \begin{equation}
486: \tperp(n,x)=\tperp
487: e^{i(\theta_{n+1}(x)-\theta_n(x))}\label{eq:hopphase}.
488: \end{equation}
489: We ignore the effects of this gauge transformation
490: on the boundary conditions applying to $\psi_n(x)$,
491: which is justified at temperatures large compared to the
492: single-particle level spacing.
493: With this, $\Ham_{\textrm{0}}+\Ham_{\textrm{int}}$ is unaffected
494: by the gauge transformation, and gauge transformed operators
495: $c_{qn}^{\dagger}$ can be defined by inverting Eq.~(\ref{eq:ctopsi}).
496: All further references in this
497: paper to fermionic operators are to the gauge-transformed ones.
498:
499: %
500: % --------------------
501: %
502: \subsection{Bosonised Hamiltonian}\label{ssec:model:bose}
503:
504: We bosonize the Hamiltonian in the standard way, expressing
505: $\Ham_{\textrm{0}}+\Ham_{\textrm{int}}$ in terms of non-interacting
506: collective modes. Since $\Ham_{\textrm{hop}}$ transforms
507: into a cosine function of the boson creation and annihilation operators,
508: we treat it perturbatively. To justify this, we require that $t_\perp$
509: should be small. Since $t_\perp$ is a relevant perturbation,\cite{Naud1}
510: we also require that temperature should not be too small:
511: $L_{\rm T} \ll l_\perp$.
512:
513: Boson creation operators are defined in the
514: usual way (see, for example, Ref.~\onlinecite{vonDelft}) as
515: \begin{equation}
516: b_{qm}^{\dagger}=\frac{i}{(n_q)^{1/2}}\sum_{r=-\infty}^{\infty}
517: c^{\dagger}_{r+q,m} c\phdagg_{r,m}
518: \end{equation}
519: for $q>0$.
520: Fourier transforming the interaction potential and expressing the
521: result as a velocity, we introduce
522: \begin{equation}
523: u_{n-m}(q) = (2\pi \hbar)^{-1}\int dx e^{iqx}U_{n-m}(x)\,.
524: \end{equation}
525: The Fermi velocity renormalised by Hartree interactions is
526: $v_{\rm F} = v - \sum_n u_n(0)$, where the divergence
527: which arises in the sum in the case of Coulomb interactions
528: is cancelled by contributions to $v$ from a neutralising background.
529: The Hamiltonian in the absence of hopping (and omitting
530: fermion number terms which appear at electron densities different from
531: that of our vacuum) is
532: \begin{equation}
533: \Ham_0+\Ham_{\textrm{int}}=\sum_{mn}\sum_{q>0}\hbar[v_{\rm F} +
534: u_{n-m}(q)]q b\dagg_{qn}b\phdagg_{qm}\,.
535: \end{equation}
536:
537: The combination $\Ham_0+\Ham\subtxt{int}$ is diagonalised by Fourier
538: transform in the layer index $n$. We impose periodic boundary conditions
539: on $n$, define the wavevector $k=2n_k\pi/Na$, with $n_k$ integer and
540: $-\pi/a \leq k < \pi/a$, and set
541: \begin{equation}
542: b^{\dagger}_{qk}=\frac{1}{\sqrt{N}}\sum_{n=1}^N
543: e^{inka}b^{\dagger}_{qn}\,,
544: \end{equation}
545: and
546: \begin{equation}
547: u(q,k) = \sum_n e^{inka} u_n(q)\,.
548: \end{equation}
549: Then
550: \begin{equation}
551: \Ham_0+\Ham_{\textrm{int}}=\sum_k
552: \sum_{q>0}\hbar\omega(q,k)b^{\dagger}_{qk}b\phdagg_{qk}\label{eq:bosonH}
553: \end{equation}
554: where the excitation frequencies are
555: \begin{equation}
556: \omega(q,k)=[\vF+u(q,k)]q\label{eq:defomega}.
557: \end{equation}
558:
559: The Coulomb interaction, regularised at short distances by a finite width
560: $w$ for edge states, has the form
561: \begin{equation}
562: U_n(x)=\frac{e^2}{4\pi\epsilon_0\epsilon_r}\frac{1}{\sqrt{x^2 +n^2a^2 + w^2}}\label{eq:rsCoulomb}\,.
563: \end{equation}
564: The edge state width $w$ is set by the localisation length $\xi$
565: of localised states in the bulk of the sample at the Fermi energy.
566: In a clean sample with well-separated Landau levels, $\xi \sim l_{\rm B}$,
567: but in a highly disordered sample with Landau levels that are broad in energy
568: one may have $\xi \gg l_{\rm B}$. The value of $w$ proves important
569: in matching our results to experiment, as we discuss in Sec.~\ref{ssec:cond:results}.
570:
571: We write the Fourier transform, using the Poisson summation formula, as
572: \begin{equation}
573: u(q,k)=v_F\frac{\kappa }{2\pi} \sum_p \iint
574: dxdz\frac{e^{-i(qx+kz+2\pi p z/a)}}{\sqrt{x^2+z^2+w^2}}\,.
575: \end{equation}
576: and find
577: \begin{equation}
578: \omega(q,k)=\vF
579: q\left(1+\kappa\sum_{p\,\in\mathbb{Z}}Q_p^{-1}e^{-wQ_p}\right)\label{eq:CoulombFT}
580: \end{equation}
581: with $Q_p^2=q^2+(k+2\pi p/a)^2$ and $p$ integer. Here, the inverse screening length
582: $\kappa\equiv e^2/4\pi\epsilon\subtxt{r}\epsilon_0 \hbar \vF a$ characterises
583: the interaction strength.
584:
585: For isolated layers, taking the limit of large $a$, the sum on $p$
586: may be replaced with an integral and one recovers the
587: dispersion relation of edge magnetoplasmons in
588: a single layer system, known from previous work.\cite{Volkov1,Volkov2}
589:
590: For the
591: multilayer system the
592: expression for the dispersion relation may be simplified in two stages. First,
593: if the layer spacing is small ($a\ll w$) the sum on $p$ may be
594: omitted, so that
595: \begin{equation}
596: \omega(q,k)=\vF q\left(1+\frac{\kappa
597: e^{-w\sqrt{q^2+k^2}}}{\sqrt{q^2+k^2}}\right)\label{eq:dispwide}\,.
598: \end{equation}
599: If, in addition, interactions are weak ($w \ll \kappa^{-1}$)
600: \begin{equation}
601: \omega(q,k)=\vF
602: q\left(1+\frac{\kappa}{\sqrt{q^2+k^2}}\right)\label{eq:dispnarrow}.
603: \end{equation}
604: In the following we obtain detailed results for systems
605: with wide edges using the dispersion relation of Eq.~(\ref{eq:dispwide}),
606: and for systems with narrow edges using the dispersion
607: relation of Eq.~(\ref{eq:dispnarrow}).
608:
609:
610: %
611: % --------------------
612: %
613: \subsection{Two-particle correlation function}\label{ssec:model:G}
614:
615: A central quantity in our calculations of transport properties is the
616: two-fermion correlation function
617: \begin{equation}
618: G(x,t)\equiv\langle\psi_n^{\dagger}(x,t)\psi\phdagg_{n+1}(x,t)
619: \psi^{\dagger}_{n+1}(0,0)\psi\phdagg_n(0,0)\rangle\label{eq:defG}\,,
620: \end{equation}
621: where $\langle\ldots\rangle\equiv\textrm{Tr}(e^{-\beta
622: \Ham}\ldots)/\textrm{Tr}(e^{-\beta \Ham})$ and
623: operators are written in the Heisenberg representation, with $\mathcal{O}(t)=e^{i\Ham
624: t/\hbar}\mathcal{O}e^{-i\Ham t/\hbar}$.
625: We evaluate this in the absence of tunneling, so that
626: $\Ham=\Ham_{0}+\Ham_{\textrm{int}}$.
627:
628: As a first step, define the boson field operator\cite{footnote0}
629: \begin{equation}
630: \phi_n(x)=-\sum_{q>0}n_q^{-1/2}\:
631: \left(e^{-iqx}b_{qn}^{\dagger}+ e^{iqx}b_{qn}\right)e^{-\epsilon
632: q/2}\label{eq:defphiboson}
633: \end{equation}
634: where $\epsilon$ is a short-distance cut-off.
635: Omitting Klein factors (which cancel from
636: $G(x,t)$), the fermion and boson field operators are related by
637: \begin{equation}
638: \psi_n(x)=(2\pi\epsilon)^{-1/2}\exp{(-i\phi_n(x))}\,.
639: \end{equation}
640: The correlation function is
641: \begin{equation}
642: G(x,t)\!\!=\!\frac{1}{(2\pi\epsilon)^2}\langle
643: e^{i\phi_n(x,t)}e^{-i\phi_{n+1}(x,t)}e^{i\phi_{n+1}(0,0)}e^{-i\phi_n(0,0)}\rangle.
644: \end{equation}
645: We define its logarithm $S$ via
646: \begin{equation}
647: G(x,t)\equiv\frac{1}{(2\pi)^2}e^S\label{eq:GtoS}\,.
648: \end{equation}
649: Because $\Ham$ is harmonic, $S$ can be expressed as
650: \begin{equation}
651: \begin{split}
652: S=&-\frac{1}{2}\left<(\phi_n(x,t)\!-\!\phi_{n+1}(x,t)\!+\!\phi_{n+1}(0,0)\!-\!\phi_n(0,0))^2\right>\\
653: &+\!\frac{1}{2}\left[\phi_n(x,t)\!-\!\phi_{n+1}(x,t),\phi_n(0,0)\!-\!\phi_{n+1}(0,0)\right]\\
654: &-2\log{\epsilon}.
655: \end{split}
656: \end{equation}
657: The thermal average and the commutator appearing in this expression can be
658: evaluated in the standard way via a mode expansion, by expressing
659: $\phi_n(x,t)$ in terms of boson creation and annihilation
660: operators using Eq.~(\ref{eq:defphiboson}).
661: Taking the thermodynamic limit and so
662: replacing wavevector sums with integrals, with $\beta = 1/k_{\rm B} T$,
663: we arrive at
664: \begin{align}
665: &S(x,t,T)=-2\log{\epsilon}-\frac{a}{\pi}\!\int^{\pi/a}_{-\pi/a}\!\!\!\!\!\!dk(1-\cos{ak})
666: \!\!\int_0^{\infty}\!\!\frac{dq}{q}e^{-\epsilon q}\nonumber\\
667: &\!\times\!\!\bigg(\!\!\coth{\!\left(\,\beta\hbar\omega(q,k)/\,2\,\right)}\!
668: \left[1\!-\!\cos{(qx\!-\omega(q,k)t)}\right]\label{eq:fullS}\\
669: &\qquad\qquad\qquad\qquad\qquad\qquad+i\sin{(qx\!-\omega(q,k)t)}
670: \!\!\bigg)\nonumber.
671: \end{align}
672: It is useful to note that
673: \begin{equation}
674: G(-x,-t)=G(x,t)^{\ast}\label{eq:Gstar},
675: \end{equation}
676: and also to define a frequency-dependent correlator,
677: \begin{equation}
678: \tilde{G}(x,\Omega)=\int dt e^{i\Omega t}G(x,t).
679: \end{equation}
680:
681:
682:
683:
684: \section{Conductivity}\label{sec:cond}
685:
686: In this section we express the conductivity $\sigma(T)$
687: obtained from a Kubo formula in terms of the two-fermion
688: correlation function calculated in Sec.~\ref{ssec:model:G}.
689: We also set out the steps required for a numerical evaluation
690: of $\sigma(T)$, present our results, and compare them with the experimental
691: data of Ref.~\onlinecite{UCSB10}.
692:
693: \subsection{Kubo formula for conductivity}\label{ssec:cond:kubo}
694:
695: The operator for the interlayer current density between layers $n$
696: and $n+1$ is
697: \begin{equation}
698: j_n(x)=\frac{ie}{\hbar}
699: \left(\tperp(n,x)\psi_{n+1}^{\dagger}(x)\psi_n(x)-\textrm{H.~c.}\right).\label{eq:defj}
700: \end{equation}
701: %Applying a spatially constant electric field $E(t)$ to the chiral
702: %metal perpendicular to the layers gives rise to a time-dependent
703: %perturbation to the Hamiltonian: $\Ham\to \Ham+\Ham_1(t)$ with
704: %\begin{equation}
705: %\Ham_1(t)=-E(t)\sum_n naQ_n\label{eq:defHpert}.
706: %\end{equation}
707: The real part of the conductivity at frequency $\Omega$
708: is given by the Kubo formula\cite{footnote}
709: \begin{align}
710: \sigma(\Omega,T)\!=\!\frac{ia}{\hbar\Omega
711: L}\sum_m\!\int_{-\infty}^{\infty} \!\!&dt\sin{\Omega t}\!\!\int
712: \!dx\!\int\! dx'\nonumber\\
713: &\times\left\langle\!
714: j_n(x,t)j_m(x'\!,0)\!\right\rangle\label{eq:sigmajj}\, .
715: \end{align}
716: To leading order, the interlayer hopping appears only in the current
717: operators, and we evaluate the thermal average
718: using a Hamiltonian from which interlayer hopping is omitted.
719:
720: Substituting for $j_n(x,t)$ using
721: Eq.~(\ref{eq:defj}) gives an expression for the conductivity of
722: the chiral metal with a given configuration of disorder: to
723: leading order in $t_{\perp}(n,x)$,
724: \begin{align}
725: \sigma(\Omega,T)&=\frac{2iaL}{\hbar\Omega}\left(\frac{e}{\hbar
726: L}\right)^2\!\!\int\! dx\!\int\!
727: dx'\!\int_{-\infty}^{\infty}\!\!\!\!
728: dt\sin{\Omega t}\label{eq:sigmadis}\nonumber\\
729: &\times\tperp(n,x)\tperp^\ast(n,x')\nonumber\\
730: &\times\langle\psi_n^{\dagger}(x,t)\psi_{n+1}(x,t)\psi^{\dagger}_{n+1}(x',0)\psi_n(x',0)\rangle\,.
731: \end{align}
732: Averaging over disorder configurations yields
733: \begin{equation}
734: [\tperp(n,x)\tperp^\ast(n,x')]_{\rm av} = \tperp^2 e^{-|x|/l_{\rm el}}
735: \end{equation}
736: and hence
737: \begin{align}
738: \sigma(\Omega,T)&=\frac{e^2}{h}\frac{8\pi i a \lel
739: t_{\perp}^2}{\Omega\hbar^2}\int \frac{dx}{2\lel}\, e^{-\vert
740: x\vert/\lel}\!\int_{-\infty}^{\infty}\!dt \sin{\Omega t}\nonumber\\
741: &\times\langle\psi_n^{\dagger}(x,t)\psi_{n+1}(x,t)\psi^{\dagger}_{n+1}(0,0)\psi_n(0,0)\rangle\label{eq:sigmaGfull}.
742: \end{align}
743: This result can be expressed in terms of the time or frequency
744: dependent two-particle correlation functions defined in
745: Sec.~\ref{ssec:model:G}. Setting $\Omega=0$ we find
746: \begin{align}\label{sigma}
747: \sigma(T)&\!=-\frac{e^2}{h}\frac{8\pi a \lel
748: t_{\perp}^2}{\hbar^2}\!\!\int\!\!\frac{dx}{2\lel}\, e^{-\vert
749: x\vert/\lel}\!\!\!\int_{-\infty}^{\infty}\!\!\!\!\!\!dt\,t\,\textrm{Im}G(x,t)\\
750: &\equiv\frac{e^2}{h}\frac{8\pi a \lel
751: t_{\perp}^2}{\hbar^2}\!\!\int\!\!\frac{dx}{2\lel}\, e^{-\vert
752: x\vert/\lel}\textrm{Re}\left[\partial_{\,\Omega}
753: \tilde{G}(x,\Omega)\big\vert_{\Omega=0}\right]\nonumber.
754: \end{align}
755:
756: For a boson dispersion relation $\omega(q,k)= v_{\rm F}q$, as results
757: from the Hartree approximation, the fermion correlation function factorises
758: into independent contributions from each layer. These
759: have the form
760: \begin{equation}
761: \langle\psi_n^{\dagger}(x,t)\psi_{n}(0,0)\rangle=\frac{1}{2\pi}\int_{-\infty}^{\infty}dk
762: \frac{e^{ik(\vF t-x)}}{1+e^{\,\beta\hbar\vF k}}
763: \end{equation}
764: and we find a temperature-independent conductivity
765: \begin{equation}
766: \sigma(\Omega,T)=\frac{e^2}{h}\frac{2t_{\perp}^2\lel
767: a}{\hbar^2\vF^2}\frac{1}{1+\Omega^2\lel^2/\vF^2}\,,
768: \end{equation}
769: which in the zero-frequency limit has the value
770: \begin{equation}
771: \sigma_0 =\frac{e^2}{h}\frac{2t_{\perp}^2l_{el}a}{\hbar^2\vF
772: ^2}\label{eq:sigma0}\,.
773: \end{equation}
774:
775: More generally, with an arbitrary boson dispersion relation a simplification
776: of Eq.~(\ref{sigma}) is possible for $\lel\ll\LT$, since $G(x,t)$
777: varies with $x$ only on the scale $\LT$ while the
778: correlator $[\tperp(n,x)\tperp^\ast(n,x')]_{\rm av}$ has range $\lel$.
779: We get
780: \begin{align}
781: \sigma(T)&=-4\pi\sigma_0\vF^2\int_{-\infty}^{\infty}\!\!\!dt\,t\,
782: \textrm{Im}\,G(0,t)\label{eq:sigmaGIm}\nonumber\\
783: &\equiv\phantom{-}4\pi\sigma_0
784: \vF^2\textrm{Re}\left[\partial_{\,\Omega}\tilde{G}(0,\Omega)
785: \big\vert_{\Omega=0}\right].
786: %\label{eq:sigmaGomega}
787: \end{align}
788: %
789: % --------------------
790: %
791: \subsection{Evaluation of $\sigma(T)$}\label{ssec:cond:evalsigma}
792:
793: To find the temperature dependence of the conductivity we must
794: combine Eqs. (\ref{eq:GtoS}), (\ref{eq:fullS}), and
795: (\ref{eq:sigmaGIm}). A first step before numerical evaluation
796: is to isolate the dependence on the cut-off
797: $\epsilon$ and take the limit $\epsilon \rightarrow 0$,
798: as we describe in this subsection.
799:
800: We start from the expression given in Eq.~(\ref{eq:fullS})
801: for the
802: logarithm of the two-particle correlation function,
803: which we evaluate at $x=0$. It is convenient to separate
804: out a zero-temperature contribution by writing
805: \begin{equation}
806: S(t,T)\equiv S(t,0)+\Delta S(t,T)
807: \end{equation}
808: and also to split $S(t,0)$ into real and imaginary parts, with
809: \begin{equation}
810: S(t,0)\equiv \mathcal{U}(t)-i\mathcal{V}(t)\,,
811: \end{equation}
812: where $\mathcal{U}(t)$ and $\mathcal{V}(t)$ are real for $t$ real.
813: Then,
814: writing
815: \begin{equation}
816: \sigma(T)=\sigma(0)+\Delta\sigma(T),
817: \end{equation}
818: we obtain from
819: Eq.~(\ref{eq:sigmaGIm})
820: \begin{align}
821: \sigma(0)=&\frac{2\sigma_0
822: \vF^2}{\pi}\int_0^{\infty}\!\!\!dt\,t
823: \,e^{\mathcal{U}(t)}\sin{\mathcal{V}(t)}\label{eq:sigma0UV}
824: \end{align}
825: and
826: \begin{align}
827: \Delta\sigma(T)\!\!=&\frac{2\sigma_0\vF^2}{\pi}\!\!
828: \int_0^{\infty}\!\!\!\!\!dt\,t\,e^{\mathcal{U}(t)}\sin{\mathcal{V}(t)}\!
829: \left[e^{\Delta S(t,T)}\!-1\right]\label{eq:DeltasigmaUV}.
830: \end{align}
831: In the case of a linear boson dispersion relation,
832: $\omega(q,k)= v_{\rm F}q$, the functions $\U(t)$ and $\V(t)$
833: have the forms
834: \begin{align}
835: \mathcal{U}_{\,\textrm{lin}}(t)&=-\log{(\epsilon^2+\vF^2t^2)}\label{eq:defUlin}\\
836: \mathcal{V}_{\,\textrm{lin}}(t)&=\pi-2\tan^{-1}(\epsilon/\vF
837: t)\label{eq:defVlin}.
838: \end{align}
839: Adding and subtracting these expressions from the ones for $\U(t)$
840: and $\V(t)$ with a general dispersion relation, we find
841: \begin{align}
842: &\mathcal{U}(t)=\mathcal{U}_{\,\textrm{lin}}(t)+\frac{a}{\pi}\int_{-\pi/a}^{\pi/a}\!\!\!dk(1-\cos{ak})\nonumber\\
843: &\times\int_0^{\infty}\frac{dq}{q}\,e^{-\epsilon
844: q}[\cos{(\omega(q,k)t})-\cos{(\vF qt)}]\label{eq:defU}
845: \end{align}
846: and
847: \begin{align}
848: &\mathcal{V}(t)=\mathcal{V}_{\,\textrm{lin}}(t)+\frac{a}{\pi}\int_{-\pi/a}^{\pi/a}\!\!\!dk(1-\cos{ak})\nonumber\\
849: &\times\int_0^{\infty}\frac{dq}{q}\,e^{-\epsilon q}
850: \left[\sin{(\omega(q,k)t)}-\sin{(\vF qt)}\right]\label{eq:defV}.
851: \end{align}
852: Finally, we have
853: \begin{align}
854: \Delta S(t,T)&=-\frac{a}{\pi}\int_{-\pi/a}^{\pi/a}\!\!\!dk(1-\cos{ak})\label{eq:defDeltaS}\\
855: \times\int_0^{\infty}\frac{dq}{q}&e^{-\epsilon
856: q}(1-\cos{\omega(q,k)t})
857: \left[\coth{\left(\frac{\beta\hbar\omega(q,k)}{2}\right)}-1\right]\nonumber.
858: \end{align}
859:
860: The advantage of casting the equations for the conductivity in
861: this form is that the momentum integrals in Eqs.~(\ref{eq:defU}),
862: (\ref{eq:defV}) and (\ref{eq:defDeltaS}) can be performed at $\epsilon=0$, since the
863: integrands decay fast enough at large $q$ for convergence.
864: Dependence on $\epsilon$ is confined for small $\epsilon$
865: to the functions $\Ulin(t)$ and $\Vlin(t)$, and from
866: Eqs.~(\ref{eq:defUlin}) and (\ref{eq:defVlin}) one sees
867: that it is important only for
868: $t\sim\order{\epsilon}$.
869: It is therefore convenient to separate
870: the integration range in
871: Eq.~(\ref{eq:sigma0UV}) into two parts, $0\leq t<R$ and
872: $R\leq t <\infty$, with $\epsilon\ll R\ll 1$.
873: In the first interval $\mathcal{U}(t)=\Ulin(t)$ and $\mathcal{V}=\Vlin(t)$;
874: in the second interval one can set $\epsilon=0$.
875:
876: Let the contributions to $\sigma(0)$ from the two intervals be
877: $\sigma^{(1)}$ and
878: $\sigma^{(2)}$. Writing
879: $t'=\vF t/\epsilon$ we have
880: \begin{equation}
881: \sigma^{(1)}=\frac{2\epsilon^2\sigma_0}{\pi}\int_0^{R\epsilon^{-1}}\!\!\!\!\!dt'\,t'\,
882: e^{\mathcal{U}(t')}\sin{\mathcal{V}(t')}
883: \end{equation}
884: %and, as $\epsilon\to0$, $\mathcal{U}$ and $\mathcal{V}$ are given
885: %by
886: %\begin{align}
887: %&\mathcal{U}(t')\sim \mathcal{U}_{\,\textrm{lin}}(t')\sim-2\log{\epsilon}-\log{(1+t'^2)}\\
888: %&\mathcal{V}(t')\sim\mathcal{V}_{\,\textrm{lin}}(t')\sim2\tan^{-1}(t'^{-1}).
889: %\end{align}
890: %This gives
891: which gives
892: \begin{equation}
893: \sigma^{(1)}=\frac{2\sigma_0}{\pi}\int_{0}^{\infty}dt'\,t'\,
894: \frac{1}{1+t'^2}\frac{2t'}{1+t'^2}=\sigma_0\,.
895: \end{equation}
896: Evaluation of $\sigma^{(2)}$ requires a numerical calculation,
897: and we present results in Sec.~\ref{ssec:cond:results}.
898:
899: Finally, turning to the conductivity at non-zero temperature,
900: we note that there are no extra difficulties
901: %the limit $\epsilon \rightarrow 0$
902: %introduces no extra subtlties
903: in the evaluation of $\Delta\sigma$
904: using Eq.~(\ref{eq:DeltasigmaUV}).
905: The function $\Delta S(t,T)$,
906: %$\mathcal{U}(t)$, and $\mathcal{V}(t)$
907: can be computed
908: numerically with $\epsilon=0$, and $\Delta S(t,T)\rightarrow 0$
909: as $t \rightarrow 0$, so that $\Delta\sigma(T)$ has no contribution
910: from the integration interval $0\leq t<R$ in the limit
911: $\epsilon \rightarrow 0$.
912:
913: In summary, when evaluating $\sigma(0)$ or $\Delta\sigma(T)$ using
914: Eqs.~(\ref{eq:sigma0UV}) and (\ref{eq:DeltasigmaUV}), the functions
915: $\mathcal{U}(t)$, $\mathcal{V}(t)$, and $\Delta S(t,T)$ may be
916: evaluated numerically by setting $\epsilon=0$ in
917: Eqs.~(\ref{eq:defU}), (\ref{eq:defV}), and (\ref{eq:defDeltaS}),
918: and the results used in Eq.~(\ref{eq:sigma0UV})
919: to find $\sigma^{(2)}$. To this one must add
920: $\sigma^{(1)}=\sigma_0$ in order to obtain the zero temperature
921: conductivity $\sigma(0)$. These equations combine
922: with Eq.~(\ref{eq:DeltasigmaUV}) for $\Delta\sigma(T)$ to give a
923: computationally tractable, though non-trivial, expression for $\sigma(T)$.
924: %
925: % --------------------
926: %
927: \subsection{Conductivity at zero temperature}\label{ssec:cond:sigma0T}
928:
929: The conductivity at zero temperature and zero frequency is determined
930: solely by the low energy limit of the group velocity for excitations,
931: since no other modes are excited as $T,\Omega \rightarrow 0$.
932: This zero frequency limit is reached as $q$, the wavevector component
933: in the chiral direction, approaches zero. The
934: group velocity, $\partial \omega(q,k) / \partial q\vert_{q=0} \equiv \vF \alpha(k)$,
935: is in general a function of $k$, the wavevector component in
936: the interlayer direction.
937:
938: To determine $\sigma(0)$, a useful procedure is to consider a model dispersion
939: relation which is exactly linear
940: in $q$: $\omega(q,k)=\vF
941: q\alpha(k)$. A linear dispersion relation is also of interest in its own right.
942: It arises from an interaction that in real space is short range
943: in the chiral direction, $x$: $U_n(x)=g_n\delta(x)$, giving
944: $\alpha(k)=1+(2\pi\hbar\vF)^{-1}\sum_{n}e^{ikna}g_n$.
945: With a linear dispersion relation, $q$-integrals in
946: the expressions leading to $G(x,t)$ can be evaluated analytically,
947: greatly simplifying the calculation of conductivity.
948: As we show in the following, for the limit $l_{\rm el} \ll L_{\rm T}$ that we consider,
949: a dispersion relation linear in $q$ yields a temperature-independent
950: value of conductivity.
951: For interactions, such as the Coulomb potential, that are not short
952: range in $x$, linearisation of the dispersion relation gives only an approximation
953: to $G(x,t)$. The value of $\sigma(0)$ that results
954: from integrating this approximate form for $G(x,t)$
955: is nevertheless exact (at the leading order in $t_{\perp}$
956: considered throughout this paper).
957: This fact is clear on physical grounds, since
958: we have correctly accounted for the dispersion relation at low energy.
959: It may also be derived formally, as follows.
960:
961: Starting from Eq.~(\ref{eq:sigmaGIm}),
962: we deform the contour for the time integral
963: into the semicircle at infinity in the lower half of the complex plane,
964: writing $t=t_R + {\rm i} t_I$ with $t_R$ and $t_I$ real.
965: Then in Eq.~(\ref{eq:fullS}) we have the factor
966: \begin{equation}
967: \int_0^{\infty} dq \frac{1}{q} \exp(-\epsilon q -iqx -it_R\omega(q,k) + t_I \omega(q,k))\,.
968: \end{equation}
969: This must be evaluated for all values of $t$ lying on the deformed time integration contour.
970: When $|t_R|$ is large, $\exp(-it_R\omega(q,k))$ is a rapidly oscillating function of $q$,
971: and the $q$-integral can be computed using the method of stationary phase: since
972: $\omega(q,k)$ is a monotonically increasing function of $q$, the dominant contribution
973: comes from the vicinity of the end-point at $q=0$. Similarly, when $t_I$ is large and negative,
974: $\exp(t_I\omega(q,k))$ is small for most values of $q$, and
975: the $q$-integral can be computed using steepest descents: again, the dominant contribution
976: comes from the vicinity of $q=0$. In both instances we may approximate
977: $\omega(q,k)$ by its form linearised about $q=0$; after linearisation
978: the $q$-integral can be evaluated analytically.
979:
980: This calculation yields
981: % for $G(x,t)$
982: %the expression
983: %is described in Appendix \ref{ap:GlowT}. We find
984: \begin{equation}
985: \begin{split}
986: G(0,t)=\frac{1}{(2\pi)^2}\left(\frac{\pi t/\beta\hbar}{\sinh{(\pi
987: t/\beta\hbar)}}\right)^2\frac{1}{\vF^2}\frac{1}{(\epsilon+it)^2}\\
988: \times\exp{\left(-\frac{2a}{\pi}\int_0^{\pi/a}\!\!\!dk\,(1-\cos{ak})\log{\alpha(k)}\right)}.
989: \end{split}
990: \end{equation}
991: Substituting this into Eq.~(\ref{eq:sigmaGIm}) we obtain
992: \begin{align}
993: \sigma(T)=\,&\frac{2\sigma_0}{\pi}\exp{\left(-\frac{2a}{\pi}
994: \int_0^{\pi/a}\!\!\!dk\,(1-\cos{ak})\log{\alpha(k)}\right)}\nonumber\\
995: &\times\int\frac{dt\,\epsilon
996: t^2}{(\epsilon^2+t^2)^2}\left(\frac{\pi t/\beta\hbar}{\sinh{(\pi
997: t/\beta\hbar)}}\right)^2.
998: \end{align}
999: In the limit $\epsilon\to 0$, the $t$ integral gives $\pi/2$
1000: regardless of temperature, demonstrating that, for systems
1001: with a linear dispersion relation, in the regime $l_{\rm el} \ll L_{\rm T}$,
1002: $\sigma(T)$ is independent of $T$. We find
1003: \begin{equation}
1004: \sigma(T)=\sigma_0\exp{\left(-\frac{2a}{\pi}\int_0^{\pi/a}
1005: \!\!\!dk(1-\cos{ak})\log{\alpha(k)}\right)}\label{eq:sigmaT=0}\,.
1006: \end{equation}
1007: This is our final result for the dependence of $\sigma(0)$
1008: on the dispersion relation as parameterised by $\alpha(k)$.
1009: %It is easy to evaluate this $k$ integral
1010: %numerically.
1011: %
1012: % --------------------
1013: %
1014: \subsection{Results}\label{ssec:cond:results}
1015:
1016: We are now in a position to calculate the conductivity
1017: for a system with Coulomb interactions by
1018: evaluating numerically the formulae we have derived: first, the
1019: zero-temperature value using the results from Sec.~\ref{ssec:cond:sigma0T},
1020: and then the full temperature-dependent conductivity using the
1021: results from Sec.~\ref{ssec:cond:evalsigma}. We investigate variation
1022: of the conductivity with two parameters, the Fermi velocity $\vF$ and
1023: the edge state depth $w$, and seek values of these parameters for
1024: which our results match the experimental data of Ref.~\onlinecite{UCSB10}.
1025: The parameters enter
1026: the dispersion relation $\omega(q,k)$ directly, and $\vF$ also
1027: appears in the inverse screening length $\kappa$.
1028: The interaction strength is set by the combination $\kappa a$
1029: (recall that $a$ is the layer spacing).
1030: %$\kappa=e^2/4\pi\epsilon_0\epsilon\subtxt{r}\vF\hbar a$.
1031: A scale for temperature is set by $\vF$ and $a$, via
1032: $T_0\equiv \hbar\vF/a\kB$, so that $T/T_0=a/\LT$.
1033: A scale for conductivity
1034: is given by $\sigma_0$, its value in the Hartree approximation.
1035:
1036: At a qualitative level, the effect of interactions on the conductivity
1037: can be anticipated by starting from the expression
1038: given in Eq.~(\ref{eq:sigma0}) for this quantity
1039: within the Hartree approximation. In turn, that expression
1040: can be understood in terms of a calculation
1041: of the interlayer tunneling rate, based on the Fermi golden rule:
1042: the rate involves the square of a matrix element between initial and
1043: final states on adjacent layers, and a power of the density of states
1044: for both the initial and the final states. The squared matrix element, allowing for
1045: disorder which affects phases of initial and final states separately,
1046: contributes a factor of $t_\perp^2 \lel$ to $\sigma_0$. The
1047: form of the density of states on a single edge, $1/2\pi \hbar \vF$,
1048: implies that $\sigma_0 \propto \vF^{-2}$. Returning to
1049: a full treatment of the interacting system, we note that the
1050: effect of interactions is to generate an energy-dependent group
1051: velocity in place of a constant value, $\vF$.
1052: %Moreover, Coulomb
1053: %interactions generate a group velocity
1054: %that is larger than $\vF$
1055: %at low energy and that approaches $\vF$ at high energy.
1056: In effect, the value of $\sigma(T)$ at a particular temperature
1057: involves a thermal average of the inverse square of the group velocity.
1058: Because Coulomb interactions increase the group velocity at low
1059: energy, they decrease conductivity at low temperature;
1060: equally, because the group velocity approaches $\vF$ at high energy,
1061: the conductivity approaches $\sigma_0$ at high temperature.
1062:
1063: Turning to detailed results,
1064: the dependence of $\sigma(0)$ on
1065: $w/a$ and $\kappa a$ is shown in Fig.~\ref{fig:sigmaT0},
1066: as obtained from Eq.~(\ref{eq:sigmaT=0}) using $\alpha(k) = 1 +\kappa e^{-w|k|}/|k|$.
1067: Interactions reduce the value of the conductivity, by a factor
1068: which is large if $\kappa a$ is large.
1069: \begin{figure}
1070: \begin{center}
1071: \includegraphics[width=0.40\textwidth]{2.eps}
1072: \caption{\label{fig:sigmaT0} Conductivity at zero temperature, as a function
1073: of interaction strength, parameterised by
1074: inverse screening length $\kappa$, for various edge state widths $w$.}
1075: \end{center}
1076: \end{figure}
1077: The variation of $\sigma(T)$ with $T$ is illustrated in
1078: Fig.~\ref{fig:sigmaTw=0kappa=1,5} for a system with
1079: the dispersion relation appropriate for
1080: narrow edge states,
1081: Eq.~(\ref{eq:dispnarrow}). In this case
1082: the $k$ integrals
1083: in Eqs.~(\ref{eq:defU}) and (\ref{eq:defV}) can be done
1084: analytically, leaving only the $q$ and $t$ integrals to
1085: be evaluated numerically.
1086: \begin{figure}
1087: \begin{center}
1088: \includegraphics[width=0.40\textwidth]{3.eps}
1089: \caption{\label{fig:sigmaTw=0kappa=1,5} Dependence of conductivity on
1090: temperature for narrow edge states, with interaction strengths $\kappa a=1$ and
1091: $\kappa a=5$. }
1092: \end{center}
1093: \end{figure}
1094: Finally, the behaviour of $\sigma(T)$ for a system with
1095: wide edge states ($w\geq a$) is presented in Fig.~\ref{fig:sigmaTw=4kappa=1,50}.
1096: In this case the dispersion relation is as given in Eq.~(\ref{eq:dispwide}),
1097: analytical progress does not seem possible,
1098: and integrals on $k$, $q$ and $t$ must be evaluated numerically to obtain
1099: $\sigma(T)$.
1100: \begin{figure}
1101: \begin{center}
1102: \includegraphics[width=0.40\textwidth]{4.eps}
1103: \caption{
1104: \label{fig:sigmaTw=4kappa=1,50}
1105: Dependence of conductivity on
1106: temperature for wide edge states
1107: with $w=4a$ and interaction strengths
1108: $\kappa
1109: a=1$
1110: and
1111: $\kappa
1112: a=50$.}
1113: \end{center}
1114: \end{figure}
1115: %Fig.~\ref{fig:sigmaTw=4kappa=1,50} shows the temperature
1116: %dependence of the conductivity calculated using this dispersion
1117: %relation for wide edges with $w=4a$ and $\kappa a=1$ and $\kappa a=50$.
1118: We note in passing that we checked that there are only
1119: small changes to the results presented when using the more complete form
1120: of the interaction given in Eq.~(\ref{eq:CoulombFT}), including the sum on $p$.
1121:
1122: Examining these results, it is evident that the general shape of $\sigma(T)$
1123: does not vary greatly with parameters: the temperature
1124: dependence is quadratic at low temperatures, has a roughly
1125: linear region at intermediate temperatures, and approaches $\sigma_0$
1126: in the high temperature limit.
1127: The quadratic dependence at low temperature is universal,
1128: but the extent of the roughly linear region at intermediate
1129: temperature is model-dependent. Moreover,
1130: scales in this temperature dependence change dramatically
1131: with parameter values. The value of the dimensionless temperature
1132: $T/T_0$ at the crossover between the low and intermediate
1133: temperature regimes is dependent on $\kappa$ (see
1134: Fig.~\ref{fig:sigmaTw=0kappa=1,5}) and varies even more strongly
1135: with $w$ (compare Figs. \ref{fig:sigmaTw=0kappa=1,5} and
1136: \ref{fig:sigmaTw=4kappa=1,50}). In addition, the magnitude of
1137: the variation in $\sigma(T)$ between low and high $T$ depends
1138: very much on the values of $w$ and $\kappa a$.
1139: In order to reproduce the experimental observation
1140: of a nearly linear increase in $\sigma(T)$, by about 7\% between the
1141: temperatures of 50mK and 300mK,\cite{UCSB10} we require parameters which
1142: place the experimental temperature window in the intermediate regime
1143: for behaviour, so that quadratic variation of $\sigma(T)$ with $T$
1144: occurs only in a temperature range below 50 mK, and saturation of $\sigma(T)$
1145: occurs only above 300mK.
1146: %This may be achieved by making the
1147: %dimensionless temperature at upper limit of the quadratic region
1148: %sufficiently small (by
1149: %increasing $w$), or by decreasing the temperature scale $T_0$ (by
1150: %reducing $\vF $), or by a combination of both these
1151: %procedures.
1152: Since the available data is not sufficiently detailed
1153: to justify a formal fitting procedure, we instead survey
1154: the consequences of a range of parameter choices in our results
1155: and examine the match to experimental observations.
1156:
1157: We begin by considering
1158: narrow edges states, using the results shown in
1159: Fig.~\ref{fig:sigmaTw=0kappa=1,5}. Supposing $\vF \sim \omega_{\rm C} l_{\rm B}$,
1160: which represents an upper bound on $\vF$,
1161: %appropriate for clean samples with a sharp confining
1162: %potential for edge states,
1163: we have $\vF = 1.7 \times 10^5{\rm ms}^{-1}$.
1164: With $a=30$nm, we find $\kappa a \sim 1$
1165: and $T_0 \sim 40$K. Taking these values, the variation in $\sigma(T)$
1166: over the experimental temperature range is
1167: very small and quadratic, in disagreement with observations.
1168: A reduction in the value of $\vF$ serves to decrease the temperature scale $T_0$,
1169: and also increases $\kappa$.
1170: It is possible to generate approximately linear variation of $\sigma(T)$
1171: with $T$ in the experimental temperature range by using a sufficiently small
1172: value of $\vF$ (reduced from the upper bound by $\sim\order{10^3}$),
1173: but we know of no reason for $\vF$ to be so small.
1174:
1175: We therefore turn to theoretical results for wide edge states,
1176: as illustrated in Fig.~\ref{fig:sigmaTw=4kappa=1,50}.
1177: In this case, we find that large values of $w$
1178: greatly reduce the temperature range over which $\sigma(T)$ varies
1179: quadratically with $T$, and can
1180: lead to approximately linear variation in the experimental
1181: temperature range. A second consequence of large $w$ is that
1182: the conductivity change $\sigma(\infty) - \sigma(0)$ is reduced.
1183: This tendency can be counteracted by increasing the
1184: interaction strength $\kappa a$.
1185: We find that observed behaviour can be reproduced
1186: by taking $w=4a=120{\rm nm}$ and
1187: $\vF=3\times10^3\textrm{ms}^{-1}$ (giving $\kappa a=50$). The temperature dependence
1188: of $\sigma(T)$ obtained using these parameter values is
1189: shown in Fig.~\ref{fig:sigma(T)w=4kappa=50} for temperatures
1190: below 400mK.
1191: \begin{figure}
1192: \begin{center}
1193: \includegraphics[width=0.40\textwidth]{5.eps}
1194: \caption{\label{fig:sigma(T)w=4kappa=50}Dependence of conductivity
1195: on temperature for
1196: $w=4a$, with $a=30\textrm{nm}$,
1197: $\vF =3\times10^{3}\textrm{ms}^{-1}$ and
1198: $\sigma_0=1.893 \times 10^{-3}e^2/2\pi \hbar$ (full line),
1199: compared with experimental data (points) taken from Fig 2
1200: of Ref.~\onlinecite{UCSB10} (data set for Fractal 2).
1201: }
1202: \end{center}
1203: \end{figure}
1204:
1205: This choice of parameters, and its implications, merit further discussion.
1206: First, we note that there are two separate experimental indications
1207: that edge states have a width closer to the value we have adopted, of
1208: $120 {\rm nm}$, than to the conventionally expected value of
1209: $l_{\rm B} \simeq 10{\rm nm}$.
1210: One comes from measurements of bulk hopping transport
1211: in multilayer samples\cite{UCSB13},
1212: which give a localisation length of
1213: $\xi = 120{\rm nm}$: one expects $w\simeq \xi$. The other comes
1214: from studies of
1215: conductance fluctuations,\cite{UCSB12}
1216: discussed in Sec.~\ref{sec:condflucs}.
1217: These yield a value for the inelastic scattering length,
1218: from the amplitude of fluctuations, and a value for
1219: the area of a phase-coherent region perpendicular to the
1220: applied field, from the correlation field for
1221: fluctuations. The ratio of this phase-coherent area to the inelastic scattering length
1222: implies an edge state width which is also much larger than
1223: $l_{\rm B}$: $w\simeq 70{\rm nm}$.
1224: Next, turning to the value of $\vF$, which we have taken $50$
1225: times smaller than for edge states in a steep confining potential,
1226: we note that large edge state width favours a small value for
1227: $\vF$, because wide edge states penetrate into the bulk of the sample
1228: where both the confining potential gradient and the drift velocity
1229: of electrons moving in this potential are small.
1230: Finally, we comment on the fact that accepting a small value for
1231: $\vF$ implies a large value for $\sigma_0$,
1232: if other parameters are unchanged. In fact, large $w$ acts in
1233: the opposite direction, to reduce the effective tunneling amplitude
1234: $t_\perp$ between edge states, since different portions of the edge
1235: contribute to the amplitude with different phases, so that there are
1236: partial cancellations. To account for the magnitude of the measured\cite{UCSB10}
1237: conductivity, $1.5\times 10^{-3} e^2/2\pi \hbar$, using
1238: the value for the mean free path $\lel = 30 {\rm nm}$ derived from
1239: magnetoresistance measurements\cite{UCSB11} requires an effective value of $t_\perp$
1240: about 50 times smaller than bare estimate\cite{UCSB1} of $0.12$ meV.
1241: This is a surprisingly strong supression of tunneling, though possible if edge states
1242: in successive layers have different displacements from the surface, as suggested in
1243: Ref.~\onlinecite{UCSB10}.
1244:
1245:
1246: %\bibitem{footnote0}
1247: %Eqns.~(\ref{eq:defphiboson}) and (\ref{eq:fullS}) correct sign errors
1248: %in Ref.~\onlinecite{PRL} Eqns.~(8) and (13).
1249: %None of the results of Ref.~\onlinecite{PRL} are
1250: %changed by these corrections.
1251: %\bibitem{footnote}
1252: %This version of the Kubo formula does not distinguish
1253: %between the external, applied field and an internal, screened
1254: %field. The distinction is unimportant in our context because
1255: %screening in the chiral metal of an electric
1256: %field in the interlayer direction vanishes in the
1257: %long wavelength limit.\cite{Betouras} See Ref.~\onlinecite{Oreg}
1258: %for a related discussion. One of use (J. T. C.) is grateful to
1259: %D. G. Polyakov for a discussion of this point.
1260:
1261:
1262: \section{Conductance fluctuations}\label{sec:condflucs}
1263:
1264: It is found experimentally that mesoscopic fluctuations
1265: in the conductance of the chiral metal are induced
1266: by small changes of magnetic field within a quantum Hall
1267: plateau.\cite{UCSB9,UCSB12} These conductance fluctuations are
1268: observed in samples with a perimeter that is several times larger
1269: than the estimated inelastic scattering length.
1270: Under such conditions, it is not initially clear why the
1271: magnetic field component perpendicular to layers in the sample
1272: should influence conductance in this way, since
1273: in the simplest picture electron trajectories
1274: enclose flux only by encircling the sample. More realistically,
1275: a number of possibilities are evident:\cite{UCSB12}
1276: the sample walls may lie at an angle to the layer normal,
1277: either on average or because of surface roughness,
1278: or finite edge state width may be important.
1279: In our theoretical treatment of conductance fluctuations we
1280: avoid specific assumptions about this aspect of the system
1281: by considering fluctuations that result from
1282: variations in a magnetic field component $B_\perp$ perpendicular to
1283: the sample surface. The amplitude of fluctuations
1284: is not affected by this choice. By contrast, the scale for the
1285: correlation field of fluctuations is
1286: dependent on the model chosen for flux linkage.
1287:
1288: In a general setting, there are two possible
1289: reasons for the amplitude of conductance fluctuations
1290: to decrease with inceasing temperature. One is because of a decrease in
1291: the inelastic scattering length; the other is
1292: because of thermal smearing. In the case of a chiral metal
1293: only the first mechanism operates, because states at
1294: different energies are perfectly correlated.\cite{Betouras}
1295: In this sense, conductance fluctuations offer a
1296: rather direct probe of interaction effects.
1297:
1298: In this section, in place of conductivity $\sigma$,
1299: we are concerned with the conductance
1300: $g=\sigma L/Na$ of a finite sample and fluctuations
1301: $\delta g = g - [g]_{\rm av}$ about its average value.
1302: We denote the average within the Hartree approximation
1303: by $g_0\equiv \sigma_0 L/Na$.
1304: We derive an analytic expression for the autocorrelation
1305: function of conductance fluctuations induced by $B_\perp$.
1306: We focus on its temperature dependence at low temperatures,
1307: obtaining a scaling form for the
1308: regime in which $\sigma(T) \approx \sigma(0)$.
1309: We compute the scaling function, evaluate our expressions numerically,
1310: and compare our results with the observations of Ref.~\onlinecite{UCSB12}.
1311:
1312:
1313: %
1314: % --------------------
1315: %
1316: \subsection{Correlation function}\label{ssec:condflucs:corfn}
1317:
1318: The conductance autocorrelation function
1319: \begin{equation}
1320: F(\delta B)=[\delta g(B_{\perp}) \delta g(B_{\perp}+\delta
1321: B)]_{\textrm{av}}
1322: \end{equation}
1323: is characterised by the amplitude $F(0)$
1324: and by the correlation field. An obvious field scale is set by a
1325: flux density of one flux quantum $\Phi_0$ through a rectangle with sides
1326: proportional to the layer spacing and the thermal length, and we define
1327: $B_0=\Phi_0/2\pi a\LT =\hbar/eaL_{\rm T}$. We also introduce a dimensionless
1328: field variation $b=\delta B/B_0$, which depends on temperature through
1329: $L_{\rm T}$, and a temperature-independent reduced field $h$
1330: which has dimensions of wavevector:
1331: $h=b/L_{\rm T} \equiv e \delta B/a\hbar$.
1332:
1333: With a suitable choice of gauge, the transverse field enters
1334: the Hamiltonian only as a phase for interlayer hopping.
1335: Taking for convenience $B_\perp =0$, in the
1336: presence of non-zero $\delta B$
1337: Eq.~(\ref{eq:hopphase}) is modified to
1338: \begin{align}
1339: \tperp(n,x)=\tperp e^{i(\theta_{n+1}(x)-\theta_n(x)+hx)}.
1340: \end{align}
1341: This additional, field-dependent phase alters $\Ham\subtxt{hop}$
1342: and consequently the current operator.
1343:
1344: An expression for the conductance of a sample with a specific
1345: disorder configuration is obtained by scaling
1346: Eq.~(\ref{eq:sigmadis}) with the sample dimensions.
1347: Taking account of the field-dependent phases in the current operator and
1348: substituting into the definition of $F(\delta B)$, after some
1349: manipulation we arrive at
1350: \begin{align}
1351: F(\delta B)&=\frac{g_0^2\pi^2\vF^4}{L^2\lel^2N^2}\sum_{n,m}\int
1352: dx\int dx'\int dy\int dy'\label{eq:flucsfull}\\
1353: \times&\int dt\,it\,G(x-x',t)\int
1354: dt'\,it'\,G(y-y',t')\nonumber\\
1355: \times&\left(e^{ih(x-x')}+e^{-ih(x-x')}\right)e^{C(x,x')}e^{C(y,y')}\nonumber\\
1356: \times&
1357: \left(e^{D_{nm}(x,x';\,y,y')}+e^{-D_{nm}(x,x';\,y,y')}-2\right)\nonumber\,.
1358: \end{align}
1359: Two contributions to this expression arise from the disorder
1360: average:
1361: \begin{equation}
1362: C(x,x')\!=-\frac{1}{2}[(\theta_{n+1}(x)\!-\!\theta_n(x)
1363: \!-\!\theta_{n+1}(x')\!+\!\theta_n(x'))^2]_{\textrm{av}}
1364: \end{equation}
1365: and
1366: \begin{align}
1367: D_{nm}&(x,x';y,y')\!=\!
1368: \big[(\theta_{n+1}(x)\!-\!\theta_n(x)\!-\!\theta_{n+1}(x')\!+\!\theta_n(x'))\nonumber\\
1369: \times&(\theta_{m+1}(y)\!-\!\theta_m(y)\!-\!\theta_{m+1}(y')\!+\!\theta_m(y'))\big]_{\textrm{av}}.
1370: \end{align}
1371: Both may be evaluated using the result (for $x,\,y>0$)
1372: \begin{equation}
1373: [\theta_n(x) \theta_m(y)]_{\textrm{av}}=\frac{\delta_{nm}}{\lel}\min\{x,y\}\,.
1374: \end{equation}
1375: The equation for $C$ gives
1376: \begin{equation}
1377: e^{C(x,x')}=e^{-|x-x'|/\lel},
1378: \end{equation}
1379: which in the limit of small $\lel$ can be written $2\lel\delta(x-x')$.
1380: The expression for $D$ is more complicated: one finds
1381: \begin{align}
1382: D_{nm}(x,x';y,y')\!=\!\frac{R(x,x';y,y')}{\lel}
1383: \left(2\delta_{nm}\!\!-\!\delta_{n+1,m}\!\!-\!\delta_{n-1,m}\right).
1384: \end{align}
1385: The function $R(x,x';y,y')$ gives the overlap between the two
1386: directed intervals on the real line $x\to x'$ and $y\to y'$:
1387: for example, $R(1,5;4,9)=-R(5,1;4,9)=1$. On substituting these
1388: expressions for $C$ and $D$ into Eq.(\ref{eq:flucsfull}), we
1389: obtain
1390: \begin{align}
1391: &F(b)=\frac{g_0^2\pi^2\vF^4}{L^2\lel^2N}\!\!\int\!\!
1392: dx\!\!\int\!\! dx'\!\!\int\!\! dy\!\!\int\!\! dy'\big(e^{ih(x-x')}+e^{-ih(x-x')}\big)\nonumber\\
1393: &\times\int dt\,it\,G(x-x',t)\int
1394: dt'\,it'\,G(y-y',t')e^{-|x-x'|/\lel}\nonumber\\
1395: &\times
1396: e^{-|y-y'|/\lel}\bigg\{e^{2R(x,x';\,y,y')/\lel}+e^{-2R(x,x';\,y,y')/\lel}-2\nonumber\\
1397: &+\,2e^{R(x,x';\,y,y')/\lel}+2e^{-R(x,x';\,y,y
1398: ')/\lel}-4\bigg\}\label{eq:huge}.
1399: \end{align}
1400:
1401: Examining where the weight of the integrand lies with respect to
1402: the spatial integrals in Eq.~(\ref{eq:huge}), one sees that the
1403: term in braces vanishes except in places where $R\ne0$. We
1404: consider different types of contributions from these regions, and
1405: keep only those which are leading order for $\LT \gg\lel$. First,
1406: consider regions in which $\vert x-y\vert\sim \lel$ but $\vert
1407: x-x'\vert\gg \lel$. The small factor
1408: $e^{-|x-x'|/\lel}$ is compensated by
1409: the first term in the braces if $\vert x'-y'\vert\sim\lel$. Then
1410: \begin{align}
1411: e^{-|x-x'|/\lel}e^{-|y-y'|/\lel}e&^{2R(x,x';\,y,y')/\lel}
1412: =\\&\qquad e^{(-|x-y|-|x'-y'|)/\lel}\nonumber.
1413: \end{align}
1414: Since $G(x,t)$ has a range in $x$ of order $\LT$, the
1415: resulting contribution to $F(\delta B)$ is $\order{\LT/L}$. Another
1416: contribution of the same order arises from regions where $\vert
1417: x-y'\vert\sim\lel$ and $\vert x'-y\vert\sim\lel$. Subleading
1418: contributions come from regions where all four spatial variables
1419: are within an elastic length of one another. These contributions
1420: are $\order{\lel/L}$.
1421:
1422: Keeping only the leading order terms, the
1423: expression for the correlation function has the much simplified form
1424: \begin{align}
1425: F(\delta B)&=\frac{4g_0^2\pi^2\vF^4}{NL}\int dx (e^{ihx}+e^{-ihx})\int
1426: \!\!dt\,it\,\int\!\! dt'\,it'\nonumber\\
1427: &\times\left(G(x,t)G(x,t')+G(x,t)G(-x,t')\right).
1428: \end{align}
1429: Using the symmetry of $G(x,t)$ (see Eq.~(\ref{eq:Gstar})) one finds
1430: \begin{equation}
1431: F(\delta B)=\frac{g_0^2}{NL}\,\int_{-\infty}^{\infty}\!\!dx\,
1432: e^{ihx}[f(x)]^2\label{eq:cfcorr}\,,
1433: \end{equation}
1434: where
1435: \begin{equation}
1436: f(x)\equiv
1437: -4\pi\vF^2\int_{-\infty}^{\infty}\!\!dt\,t\,\textrm{Im}G(x,t)\label{eq:deff}\,.
1438: \end{equation}
1439: %
1440: % --------------------
1441: %
1442: \subsection{Computing the correlation function}\label{ssec:condflucs:computing}
1443:
1444: In order to compare our theory for conductance fluctuations with
1445: experiment, we need to be able to calculate $F(\delta B)$ for
1446: various values of the temperature and parameters $\vF$ and $w$.
1447: Although it is possible to use a computer to evaluate the form of
1448: $F(B_{\perp})$ given in Eq.~(\ref{eq:cfcorr}) without further
1449: approximation, it is far easier to make progress by calculating
1450: $G(x,t)$ for a linearised
1451: dispersion relation.
1452: %(as described in Appendix \ref{ap:GlowT}).
1453: This
1454: approach is exact in the low-temperature regime defined by the condition
1455: $\sigma(T)\approx \sigma(0)$, and we proceed to use it
1456: in our calculations.
1457:
1458: In the low temperature regime where the linearised dispersion
1459: relation may be used, $F(B_{\perp})$ has a scaling form.
1460: To make this apparent, it is helpful to recast equations in terms of dimensionless
1461: variables,
1462: characterising $\delta B$ by $b$ in place of $h$,
1463: and introducing $\hat{x}=x/L_T$ and $\hat{t}= \vF t/\LT$.
1464: %\begin{equation}
1465: %\hat{x}=x/L_T,\quad \hat{t}= \vF t/\LT,\,.\label{eq:dimlessvars}
1466: %\end{equation}
1467: Writing $G(x,t)=(2\pi \LT)^{-2} \hat{G}(\hat{x},\hat{t})$ and
1468: $f(\LT \hat{x})=\hat{f}(\hat{x})$, for
1469: a linear dispersion
1470: relation, $\omega(q,k)=q\vF\alpha(k)$, we have
1471: \begin{align}
1472: &\hat{G}(\hat{x},\hat{t})=\exp\bigg\{\frac{-2a}{\pi}\int_0^{\pi/a}
1473: \!\!\!\!\!dk(1-\cos{ak})\nonumber\\
1474: &\times\bigg[\log|\hat{x}-\alpha(k)
1475: \hat{t}|\nonumber-\log\bigg(\frac{\pi[\alpha(k)
1476: \hat{t}-\hat{x}]/\alpha(k)}{\sinh(\pi[\alpha(k)
1477: \hat{t}-\hat{x}]/\alpha(k))}\bigg)\bigg]\bigg\}\\
1478: &\times\exp\bigg\{\!\!-ia\!\!\int_0^{\pi/a}
1479: \!\!\!\!\!dk(1-\cos{ak})\,\textrm{sgn}\,(\hat{x}-\alpha(k)
1480: \hat{t})\bigg\}\label{G-hat}
1481: \end{align}
1482: and
1483: \begin{equation}
1484: \hat{f}(\hat{x})=-\frac{1}{\pi}\int_{-\infty}^{\infty}d\hat{t}\,\hat{t}\,
1485: \textrm{Im}\{\hat{G}(\hat{x},\hat{t})\}\label{eq:fredefine}\,.
1486: \end{equation}
1487: Then the conductance autocorrelation function has the form
1488: \begin{equation}
1489: F(\delta B)=\frac{g_0^2\LT}{NL}\,C\left(\delta
1490: B/{B_0}\right)\label{eq:corrfinal}
1491: \end{equation}
1492: with scaling function
1493: \begin{equation}\label{C(b)}
1494: C(b)=\int_{-\infty}^{\infty}
1495: d\hat{x}\,e^{ib\hat{x}}[\hat{f}(\hat{x})]^2.
1496: \end{equation}
1497: In this form $F(\delta B)$ depends on temperature $T$ and magnetic field difference $\delta B$
1498: only through the scaling variables $\LT/L$ and $\delta B/B_0$.
1499: The thermal length $\LT$ plays
1500: the role of an inelastic scattering length, in the sense that it determines both the amplitude
1501: of conductance fluctuations and (through $B_0$) their correlation field.
1502: Such behaviour is initally surprising, since $\LT$ is independent of interaction strength.
1503: In fact, of course, the form of the scaling function $C(b)$
1504: depends parametrically on interaction strength.
1505:
1506: For weak interactions this dependence of $C(b)$ on $\kappa$ can be extracted analytically,
1507: as follows. First, note from Eq.~(\ref{eq:dispnarrow}) that $\alpha(k) = 1+\kappa/|k|$.
1508: Also, in Eqs.~(\ref{G-hat}), (\ref{eq:fredefine}) and (\ref{C(b)}), change variables from
1509: $\hat{x},\hat{t}$ to $y,p$ with $\hat{x}=y/\kappa$ and $\hat{t}=yp+y/\kappa$.
1510: Then
1511: \begin{align}
1512: \lim_{\kappa\rightarrow 0}\,&\hat{G}(y/\kappa,p+y/\kappa)\equiv g(y,p)\nonumber\\
1513: =\,&\exp\bigg\{\frac{-2a}{\pi}\int_0^{\pi/a}
1514: \!\!\!\!\!dk(1-\cos{ak})\nonumber\\
1515: &\times\bigg[\log|y(p+1/k)|
1516: -\log\bigg(\frac{\pi y[p+1/k]}
1517: {\sinh(\pi y[p+1/k])}\bigg)\bigg]\bigg\}\nonumber\\
1518: &\times\exp\bigg\{ia\!\!\int_0^{\pi/a}
1519: \!\!\!\!\!dk(1-\cos{ak})\,\textrm{sgn}\,(y[p+1/k])\bigg\}\nonumber
1520: \end{align}
1521: and
1522: \begin{equation}
1523: \lim_{\kappa\rightarrow 0} \,\hat{f}(y/\kappa ) \equiv \tilde{f}(y) =
1524: -\frac{y^2}{\pi}\int_{-\infty}^{\infty}dp\,p\
1525: \textrm{Im}\{{g}(y,p)\}\,.\nonumber
1526: \end{equation}
1527: The $\kappa$-dependence of the scaling function is hence isolated for small $\kappa$
1528: as
1529: \begin{equation}
1530: C(b) = \frac{1}{\kappa}\int_{-\infty}^{\infty} dy \exp(iyb/\kappa) [\tilde{f}(y)]^2\,,
1531: \end{equation}
1532: demonstrating that the amplitude of conductance fluctuations
1533: grows and that the correlation field shrinks as interactions are made weaker.
1534: In both cases, the variation implies an inelastic scattering length that diverges
1535: as $\kappa^{-1}$ for weak interactions.
1536: Such a dependence of the inelastic scattering length on interaction strength
1537: is long-established in non-chiral, one-dimensional conductors.\cite{Apel}
1538:
1539: In order to find the form of the scaling function and to study its
1540: $\kappa$-dependence at general $\kappa$,
1541: a three-dimensional numerical integration is necessary.
1542: We compute $\hat{G}(\hat{x},\hat{t})$, then
1543: $\hat{f}(\hat{x})$, and then the scaling function $C(b)$ itself.
1544: % Although the
1545: %computation is not trivial, we are able to obtain plots of $C$
1546: %against applied field for a range of the parameters $\vF$ and $w$.
1547: %
1548: % --------------------
1549: %
1550: \subsection{Results}\label{ssec:condflucs:results}
1551:
1552: We illustrate the form of the
1553: scaling function $C(\delta B/ B_0)$ for a range of parameter
1554: values in a sequence of three figures. Its dependence on interaction strength
1555: $\kappa a$ is shown for narrow edge states in Fig.~\ref{fig:flucsw=0}
1556: and for $w=a$ in Fig.~\ref{fig:flucsw=1}. In both cases, smaller
1557: interaction strength leads to a larger amplitude for conductance fluctuations
1558: and a smaller correlation field, as may be anticipated on the grounds that
1559: weaker interactions lead to a longer inelastic scattering length.
1560: In Fig.~\ref{fig:Fscaling} $C(\delta B/B_0)$ is shown for
1561: $\kappa=50$ and $w=4a$, the
1562: parameter values suggested by the comparison of our conductivity
1563: calculations with experiment. We discuss experimental data on
1564: conductance fluctuations in Sec.~\ref{ssec:condflucs:expt}.
1565: \begin{figure}
1566: \begin{center}
1567: \includegraphics[width=0.40\textwidth]{6.eps}
1568: \caption{\label{fig:flucsw=0} $C(\delta B/B_0)$ for narrow edge states and
1569: $\kappa a=0.6$, $0.8$, and $1$. }
1570: \end{center}
1571: \end{figure}
1572: \begin{figure}
1573: \begin{center}
1574: \includegraphics[width=0.40\textwidth]{7.eps}
1575: \caption{\label{fig:flucsw=1} $C(\delta B/B_0)$ for $w=a$ and
1576: $\kappa a=0.6$, $0.8$, and $1$. }
1577: \end{center}
1578: \end{figure}
1579: \begin{figure}
1580: \begin{center}
1581: \includegraphics[width=0.385\textwidth]{8.eps}
1582: \caption{\label{fig:Fscaling} $C(\delta B/B_0)$ at $w=4a$ and
1583: $\kappa a=50$.}
1584: \end{center}
1585: \end{figure}
1586: Finally, the increase in the amplitude of conductance fluctuations
1587: with dereasing $\kappa$ is illustrated in
1588: Fig.~\ref{fig:flucskappa}.
1589: \begin{figure}
1590: \begin{center}
1591: \includegraphics[width=0.40\textwidth]{9.eps}
1592: \caption{\label{fig:flucskappa} Conductance fluctuation amplitude
1593: as a function of interaction strength $\kappa a$ at $w=0$ (full line),
1594: and asymptotic behaviour calculated analytically for small $\kappa a$
1595: (dashed line)}
1596: \end{center}
1597: \end{figure}
1598: %
1599: % --------------------
1600: %
1601: \subsection{Comparison with experiment and previous
1602: theory}\label{ssec:condflucs:expt}
1603:
1604: The exact treatment of disorder and interactions
1605: provided by the calculations we have decribed presents an opportunity
1606: to test the standard theoretical treatment of conductance fluctuations,
1607: in which a single inelastic scattering length $l_{\rm in}$,
1608: or equivalently a scattering rate $\vF/l_{\rm in}$ is used as a cut-off
1609: in perturbation theory. For the chiral metal, such calculations have been
1610: described in Ref.~\onlinecite{Betouras}. They
1611: yield a Lorentzian scaling function
1612: \begin{equation}
1613: F(\delta
1614: B)=\frac{2g_0^2}{NL}\frac{l\subtxt{in}}{1+z^2}\label{eq:CBcorrelator}
1615: \end{equation}
1616: with $z=2\pi \delta Bl\subtxt{in}a/\Phi_0$.
1617: A comparison between the functional form we obtain for
1618: $F(\delta B)$ and a Lorenztian is given in
1619: Fig.~\ref{fig:Fscaling}: while the two functions are similar,
1620: the discrepancies are worth attention because they
1621: indicate behaviour which cannot be characterised by a single
1622: relaxation time. A similar comparison can be made in
1623: the Fourier transformed domain, in terms of the function $f(x)$.
1624: To reproduce Eq.~(\ref{eq:CBcorrelator}) from our Eq.~(\ref{eq:cfcorr}),
1625: we would require $l\subtxt{in}=\LT$ and
1626: \begin{equation}
1627: \hat{f}(\hat{x})= e^{-\vert \hat{x}\vert /2}\,,
1628: \end{equation}
1629: where exponential decay is indicative of a single lifetime
1630: $l\subtxt{in}/\vF$ for excitations. The form we obtain
1631: for $\hat{f}(\hat{x})$ is shown in Fig.~\ref{fig:f}.
1632: The absence of a cusp at $x=0$ indicates that there is
1633: of a range of relaxation times in the system.
1634: In addition, the fact that $f(0)\not= 1$ is an interaction
1635: effect (from Eq.~(\ref{eq:sigmaGIm}) one sees that $f(0)=\sigma(0)/\sigma_0$)
1636: not allowed for in the standard perturbative treatment.
1637: \begin{figure}
1638: \begin{center}
1639: \includegraphics[width=0.40\textwidth]{10.eps}
1640: \caption{\label{fig:f} $\hat{f}(\hat{x})$ calculated at $\kappa a=50$
1641: and $w=4a$ (solid line) compared with the best fitting exponential
1642: decay (dashed line).}
1643: \end{center}
1644: \end{figure}
1645:
1646: We close this section with a comparison between the experiments
1647: of Ref.~\onlinecite{UCSB12} and our results, using the same parameters,
1648: $\kappa a=50$ and $w=4a$, that provided a match for the behaviour of $\sigma(T)$.
1649: For the experimental base temperature of
1650: $T=70\textrm{mK}$, we use our approach to determine the amplitude
1651: of conductance fluctuations. As a way to present the result,
1652: we then follow the experimental analysis\cite{UCSB12} in using
1653: Eq.~(\ref{eq:CBcorrelator}) to obtain a value for $\lin$ of
1654: $0.3\mu\textrm{m}$. The experimental value,
1655: extracted in the same way, is
1656: $\lin\sim 1\mu\textrm{m}$.
1657: Since the calculated amplitude of conductance fluctuations
1658: varies by several orders of magnitude over the range of
1659: parameter values we have investigated, and since no
1660: new adjustment of parameters was involved in our
1661: discussion of conductance fluctuations,
1662: we find the rough agreement between these two values of $\lin$ very encouraging.
1663:
1664:
1665:
1666: \section{Conclusions}\label{sec:discussion}
1667:
1668: In summary, for the system of weakly coupled
1669: quantum Hall edge states that we have studied, bosonisation
1670: provides a very complete treatment of the interplay between
1671: electron-electron interactions and
1672: disorder. We have shown that interaction effects
1673: can account for the observed temperature dependence
1674: of interlayer conductivity, provided we allow for
1675: finite edge state width and adopt a value for the edge state velocity
1676: that is rather smaller than previously supposed.
1677: We have investigated conductance fluctuations within
1678: the same theoretical approach, showing how they are
1679: suppressed with increasing temperature, with a characteristic lengthscale
1680: $\LT \propto T^{-1}$. Encouragingly,
1681: the same parameter values used to match
1682: the measured behaviour of conductivity
1683: reproduce approximately the observed
1684: fluctuation amplitude.
1685: From a theoretical
1686: viewpoint, it is interesting that such dephasing effects
1687: can be generated from a description based on harmonic
1688: collective modes, simply via the nonlinear
1689: relation between boson and fermion operators.
1690:
1691:
1692:
1693:
1694:
1695: \begin{acknowledgments}
1696:
1697: We thank E. G. Gwinn for very helpful discussions and J. J.
1698: Betouras for previous collaborations. The work was supported by
1699: EPSRC under Grant GR/R83712/01 and by the Dutch FOM foundation.
1700:
1701: \end{acknowledgments}
1702:
1703: %\input{appendices}
1704: \bibliographystyle{h-physrev}
1705: \bibliography{bibliography}
1706:
1707: \end{document}
1708: