1: \documentclass[twocolumn,floatfix,eqsecnum,rmp]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{amsmath}
4: \usepackage{bm}
5: % Revtex fails if there are more than 255 references;
6: % only the rmp class works; the \bibpunct command then
7: % restores the superscript references (as in the prb class);
8: % unfortunately, I could not get \onlinecite to work
9: \bibpunct{}{}{,}{s}{}{\textsuperscript{,}}
10: %\citestyle{nature}
11: \hyphenation{Been-ak-ker Molen-kamp Sta-ring nano-struc-tures nano-struc-ture meso-scopic}
12: \newcommand{\be}{\begin{equation}}
13: \newcommand{\ee}{\end{equation}}
14: \begin{document}
15: \title{Quantum Transport in Semiconductor Nanostructures}
16: \author{C. W. J. Beenakker and H. van Houten}
17: \affiliation{Philips Research Laboratories, Eindhoven, The Netherlands}
18: \begin{abstract}
19: {\tt Published in Solid State Physics, {\bfseries 44}, 1-228 (1991)}
20: \end{abstract}
21: \maketitle
22:
23: \tableofcontents
24:
25: \section{\label{secI} Introduction}
26:
27: \subsection{\label{sec1} Preface}
28:
29: In recent years semiconductor nanostructures have become the model systems of choice for investigations of electrical conduction on short length scales. This development was made possible by the availability of semiconducting materials of unprecedented purity and crystalline perfection. Such materials can be structured to contain a thin layer of highly mobile electrons. Motion perpendicular to the layer is quantized, so that the electrons are constrained to move in a plane. As a model system, this {\em two-dimensional electron gas (2DEG)} combines a number of desirable properties, not shared by thin metal films. It has a low electron density, which may be readily varied by means of an electric field (because of the large screening length). The low density implies a large Fermi wavelength (typically 40 nm), comparable to the dimensions of the smallest structures (nanostructures) that can be fabricated today. The electron mean free path can be quite large (exceeding 10 $\mu{\rm m}$). Finally, the reduced dimensionality of the motion and the circular Fermi surface form simplifying factors.
30:
31: Quantum transport is conveniently studied in a 2DEG because of the combination of a large Fermi wavelength and large mean free path. The quantum mechanical phase coherence characteristic of a microscopic object can be maintained at low temperatures (below 1 K) over distances of several microns, which one would otherwise have classified as macroscopic. The physics of these systems has been referred to as {\em mesoscopic},\cite{ref1} a word borrowed from statistical mechanics.\cite{ref2} Elastic impurity scattering does not destroy phase coherence, which is why the effects of quantum interference can modify the conductivity of a disordered conductor. This is the regime of diffusive transport, characteristic for disordered metals. Quantum interference becomes more important as the dimensionality of the conductor is reduced. Quasi-one dimensionality can readily be achieved in a 2DEG by lateral confinement.
32:
33: Semiconductor nanostructures are unique in offering the possibility of studying quantum transport in an artificial potential landscape. This is the regime of {\em ballistic transport}, in which scattering with impurities can be neglected. The transport properties can then be tailored by varying the geometry of the conductor, in much the same way as one would tailor the transmission properties of a waveguide. The physics of this transport regime could be called {\em electron optics\/} in the solid state.\cite{ref3} The formal relation between conduction and transmission, known as the Landauer formula,\cite{ref1,ref4,ref5} has demonstrated its real power in this context. For example, the quantization of the conductance of a quantum point contact\cite{ref6,ref7} (a short and narrow constriction in the 2DEG) can be understood using the Landauer formula as resulting from the discreteness of the number of propagating modes in a waveguide.
34:
35: Two-dimensional systems in a perpendicular magnetic field have the remarkable property of a quantized Hall resistance,\cite{ref8} which results from the quantization of the energy in a series of Landau levels. The magnetic length $(\hbar/eB)^{1/2}$ ($\approx$ 10 nm at $B = 5\,{\rm T}$) assumes the role of the wavelength in the quantum Hall effect. The potential landscape in a 2DEG can be adjusted to be smooth on the scale of the magnetic length, so that inter-Landau level scattering is suppressed. One then enters the regime of adiabatic transport. In this regime truly macroscopic behavior may not be found even in samples as large as 0.25 mm.
36:
37: In this review we present a self-contained account of these three novel transport regimes in semiconductor nanostructures. The experimental and theoretical developments in this field have developed hand in hand, a fruitful balance that we have tried to maintain here as well. We have opted for the simplest possible theoretical explanations, avoiding the powerful --- but more formal --- Green's function techniques. If in some instances this choice has not enabled us to do full justice to a subject, then we hope that this disadvantage is compensated by a gain in accessibility. Lack of space and time has caused us to limit the scope of this review to metallic transport in the plane of a 2DEG at small currents and voltages. Transport in the regime of strong localization is
38: excluded, as well as that in the regime of a nonlinear current-voltage dependence. Overviews of these, and other, topics not covered here may be found in Refs. \onlinecite{ref9,ref10,ref11}, as well as in recent conference proceedings.\cite{ref12,ref13,ref14,ref15,ref16,ref17}
39:
40: We have attempted to give a comprehensive list of references to theoretical and experimental work on the subjects of this review. We apologize to those whose contributions we have overlooked. Certain experiments are discussed in some detail. In selecting these experiments, our aim has been to choose those that illustrate a particular phenomenon in the clearest fashion, not to establish priorities. We thank the authors and publishers for their kind permission to reproduce figures from the original publications. Much of the work reviewed here was a joint effort with colleagues at the Delft University of Technology and at the Philips Research Laboratories, and we are grateful for the stimulating collaboration.
41:
42: The study of quantum transport in semiconductor nanostructures is motivated by more than scientific interest. The fabrication of nanostructures relies on sophisticated crystal growth and lithographic techniques that exist because of the industrial effort toward the miniaturization of transistors. Conventional transistors operate in the regime of classical diffusive transport, which breaks down on short length scales. The discovery of novel transport regimes in semiconductor nanostructures provides options for the development of innovative future devices. At this point, most of the proposals in the literature for a quantum interference device have been presented primarily as interesting possibilities, and they have not yet been critically analyzed. A quantitative comparison with conventional transistors will be needed, taking circuit design and technological considerations into account.\cite{ref18} Some proposals are very ambitious, in that they do not only consider a different principle of operation for a single transistor, but envision entire computer architectures in which arrays of quantum devices operate phase coherently.\cite{ref19}
43:
44: We hope that the present review will convey some of the excitement that the workers in this rewarding field of research have experienced in its exploration. May the description of the variety of phenomena known at present, and of the simplest way in which they can be understood, form an inspiration for future investigations.
45:
46: \subsection{\label{sec2} Nanostructures in Si inversion layers}
47:
48: Electronic properties of the two-dimensional electron gas in Si MOSFET's (metal-oxide-semiconductor field-effect transistors) have been reviewed by Ando, Fowler, and Stern,\cite{ref20} while general technological and device aspects are covered in detail in the books by Sze\cite{ref21} and by Nicollian and Brew.\cite{ref22} In this section we only summarize those properties that are needed in the following. A typical device consists of a $p$-type Si substrate, covered by a ${\rm Si0}_{2}$ layer that serves as an insulator between the (100) Si surface and a metallic gate electrode. By application of a sufficiently strong positive voltage $V_{\rm g}$ on the gate, a 2DEG is induced electrostatically in the $p$-type Si under the gate. The band bending leading to the formation of this inversion layer is schematically indicated in Fig.\ \ref{fig1}. The areal electron concentration (or sheet density) $n_{\rm s}$ follows from $en_{\rm s}=C_{\rm ox}(V_{\rm g} - V_{\rm t})$, where $V_{\rm t}$ is the threshold voltage beyond which the inversion layer is created, and $C_{\rm ox}$ is the capacitance per unit area of the gate electrode with respect to the electron gas. Approximately, one has $C_{\rm ox} = \varepsilon_{\rm ox}/d_{\rm ox}$ (with $\varepsilon_{\rm ox} = 3.9\,\varepsilon_{0}$ the dielectric constant of the ${\rm Si0}_{2}$ layer),\cite{ref21} so
49: \begin{equation}
50: n_{\rm s}=\frac{\varepsilon_{\rm ox}}{ed_{\rm ox}}(V_{\rm g}-V_{\rm t}).\label{eq2.1}
51: \end{equation}
52: The linear dependence of the sheet density on the applied gate voltage is one of the most useful properties of Si inversion layers.
53:
54: \begin{figure}
55: \centerline{\includegraphics[width=8cm]{figures/fig1}}
56: \caption{
57: Band-bending diagram (showing conduction band $E_{\rm c}$, valence band $E_{\rm v}$, and Fermi level $E_{\rm F}$) of a metal-oxide-semiconductor (MOS) structure. A 2DEG is formed at the interface between the oxide and the $p$-type silicon substrate, as a consequence of the positive voltage $V$ on the metal gate electrode.
58: \label{fig1}
59: }
60: \end{figure}
61:
62: The electric field across the oxide layer resulting from the applied gate voltage can be quite strong. Typically, $V_{\rm g} - V_{\rm t} = 5\,{\rm V}$ and $d_{\rm ox} = 50\,{\rm nm}$, so the field strength is of order 1 MV/cm, at best a factor of 10 lower than typical fields for the dielectric breakdown of ${\rm Si0}_{2}$. It is possible to change the electric field at the interface, without altering $n_{\rm s}$, by applying an additional voltage across the $p$-$n$ junction that isolates the inversion layer from the $p$-type substrate (such a voltage is referred to as a substrate bias). At the Si-${\rm Si0}_{2}$ interface the electric field is continuous, but there is an electrostatic potential step of about 3 eV. An approximately triangular potential well is thus formed
63: at the interface (see Fig.\ \ref{fig1}). The actual shape of the potential deviates somewhat from the triangular one due to the electronic charge in the inversion layer, and has to be calculated self-consistently.\cite{ref20} Due to the confinement in one direction in this potential well, the three-dimensional conduction band splits into a series of two-dimensional subbands. Under typical conditions (for a sheet electron density $n_{\rm s} = 10^{11}-10^{12}\,{\rm cm}^{-2}$) only a single two-dimensional subband is occupied. Bulk Si has an indirect band gap, with six equivalent conduction band valleys in the (100) direction in reciprocal space. In inversion layers on the (100) Si surface, the degeneracy between these valleys is partially lifted. A twofold valley degeneracy remains. In the following, we treat these two valleys as completely independent, ignoring complications due to intervalley scattering. For each valley, the (one-dimensional) Fermi surface is simply a circle, corresponding to free motion in a plane with effective electron mass\cite{ref20} $m = 0.19\,m_{\rm e}$. For easy reference, this and other relevant numbers are listed in Table \ref{table1}.
64:
65: The electronic properties of the Si inversion layer can be studied by capacitive or spectroscopic techniques (which are outside the scope of this review), as well as by transport measurements in the plane of the 2DEG. To determine the intrinsic transport properties of the 2DEG (e.g., the electron mobility), one defines a wide channel by fabricating a gate electrode with the appropriate shape. Ohmic contacts to the channel are then made by ion implantation, followed by a lateral diffusion and annealing process. The two current-carrying contacts are referred to as the source and the drain. One of these also serves as zero reference for the gate voltage. Additional side contacts to the channel are often fabricated as well (for example, in the Hall bar geometry), to serve as voltage probes for measurements of the longitudinal and Hall resistance. Insulation is automatically provided by the $p$-$n$ junctions surrounding the inversion layer. (Moreover, at the low temperatures of interest here, the substrate conduction vanishes anyway due to carrier freeze-out.) The electron mobility $\mu_{\rm e}$ is an important figure of merit for the quality of the device. At low temperatures the mobility in a given sample varies nonmonotonically\cite{ref20} with increasing electron density $n_{\rm s}$ (or increasing gate voltage), due to the opposite effects of enhanced screening (which reduces ionized impurity scattering) and enhanced confinement (which leads to an increase in surface roughness scattering at the Si-${\rm Si0}_{2}$ interface). The maximum low-temperature mobility of electrons in high-quality samples is around $10^4\,{\rm cm}^2 /{\rm Vs}$. This review deals with the modifications of the transport properties of the 2DEG in narrow geometries. Several lateral confinement schemes have been tried in order to achieve narrow inversion layer channels (see Fig. \ref{fig2}). Many more have been proposed, but here we discuss only those realized experimentally.
66:
67: \begin{table*}
68: \caption{
69: Electronic properties of the 2DEG in GaAs-AlGaAs heterostructures and Si inversion layers.
70: \label{table1}
71: }
72: \begin{ruledtabular}
73: \begin{tabular}{lllll}
74: &&GaAs(100)&Si(100)&Units\\
75: \hline
76: Effective Mass&$m$&0.067&0.19&$m_{\rm e}=9.1\times10^{-28}\,{\rm g}$\\
77: Spin Degeneracy&$g_{\rm s}$&2&2&\\
78: Valley Degeneracy&$g_{\rm v}$&1&2&\\
79: Dielectric Constant&$\varepsilon$&13.1&11.9&$\varepsilon_{0}=8.9\times 10^{-12}\,{\rm Fm}^{-1}$\\
80: Density of States&$\rho(E)=g_{\rm s} g_{\rm v} (m/2\pi\hbar^{2})$&0.28&1.59&$10^{11}\,{\rm cm}^{-2}\,{\rm meV}^{-1}$\\
81: Electronic Sheet Density\footnotemark[1]
82: &$n_{\rm s}$&4&1--10&$10^{11}\,{\rm cm}^{-2}$\\
83: Fermi Wave Vector&$k_{\rm F} =(4\pi n_{\rm s} /g_{\rm s} g_{\rm v})^{1/2}$&1.58&0.56--1.77&$10^{6}\,{\rm cm}^{-1}$\\
84: Fermi Velocity&$v_{\rm F} =\hbar k_{\rm F} /m$&2.7&0.34--1.1&$10^7 \,{\rm cm/s}$\\
85: Fermi Energy&$E_{\rm F} =(\hbar k_{\rm F})^2/2m$&14&0.63--6.3&meV\\
86: Electron Mobility\footnotemark[1]&$\mu_{\rm e}$&$10^4-10^6$&$10^4$&${\rm cm}^2/{\rm Vs}$\\
87: Scattering Time&$\tau=m\mu_{\rm e}/e$&0.38--38&1.1&ps\\
88: Diffusion Constant&$D=v_{\rm F}^2\tau/2$&140--14000&6.4--64&${\rm cm}^2/{\rm s}$\\
89: Resistivity&$\rho=(n_{\rm s} e\mu_{\rm e})^{-1}$&1.6--0.016&6.3--0.63&${\rm k}\Omega$\\
90: Fermi Wavelength&$\lambda_{\rm F}=2\pi/k_{\rm F}$&40&112--35&nm\\
91: Mean Free Path&$l=v_{\rm F}\tau$&$10^2-10^4$&37--118&nm\\
92: Phase Coherence Length\footnotemark[2]&$l_{\phi}=(D\tau_{\phi})^{1/2}$&200--...&40--400&nm$(T/{\rm K})^{-1/2}$\\
93: Thermal Length&$l_{\rm T}=(\hbar D/k_{\rm B} T)^{1/2}$&330--3300&70--220&nm$(T/{\rm K})^{-1/2}$\\
94: Cyclotron Radius&$l_{\rm cycl}=\hbar k_{\rm F}/eB$&100&37--116&nm$(B/{\rm T})^{-1}$\\
95: Magnetic Length&$l_{\rm m}=(\hbar/eB)^{1/2}$&26&26&nm$(B/{\rm T})^{-1/2}$\\
96: &$k_{\rm F}l$&15.8--1580&2.1--21&\\
97: &$\omega_{\rm c}\tau$&1--100&1&$(B/{\rm T})$\\
98: &$E_{\rm F}/\hbar\omega_{\rm c}$&7.9&1--10&$(B/{\rm T})^{-1}$
99: \end{tabular}
100: \end{ruledtabular}
101: \footnotetext[1]{
102: A typical (fixed) density value is taken for GaAs-AlGaAs heterostructures, and a typical range of values in the metallic conduction regime for Si MOSFET's. For the mobility, a range of representative values is listed for GaAs-AlGaAs heterostructures, and a typical ``good'' value for Si MOSFET's. The variation in the other quantities reflects that in $n_{\rm s}$ and $\mu_{\rm e}$.}
103: \footnotetext[2]{
104: Rough estimate of the phase coherence length, based on weak localization experiments in laterally confined heterostructures\protect\cite{ref23,ref24,ref25,ref26,ref27} and Si MOSFET's.\protect\cite{ref28,ref29} The stated $T^{-1/2}$ temperature dependence should be regarded as an indication only, since a simple power law dependence is not always found (see, for example, Refs.\ \protect\onlinecite{ref30} and \protect\onlinecite{ref25}). For high-mobility GaAs-AlGaAs heterostructures the phase coherence length is not known, but is presumably\protect\cite{ref31} comparable to the (elastic) mean free path $l$.}
105: \end{table*}
106:
107: Technically simplest, because it does not require electron beam lithography, is an approach first used by Fowler et al., following a suggestion by Pepper\cite{ref32,ref33,ref34} (Fig.\ \ref{fig2}a). By adjusting the negative voltage over $p$-$n$ junctions on either side of a relatively wide gate, they were able to vary the electron channel width as well as its electron density. This technique has been used to define narrow accumulation layers on $n$-type Si substrates, rather than inversion layers. Specifically, it has been used for the exploration of quantum transport in the strongly localized regime\cite{ref32,ref35,ref36,ref37} (which is not discussed in this review). Perhaps the technique is particularly suited to this highly resistive regime, since a tail of the diffusion profile inevitably extends into the channel, providing additional scattering centers.\cite{ref34} Some studies in the weak localization regime have also been reported.\cite{ref33}
108:
109: \begin{figure}
110: \centerline{\includegraphics[width=8cm]{figures/fig2}}
111: \caption{
112: Schematic cross-sectional views of the lateral pinch-off technique used to define a narrow electron accumulation layer (a), and of three different methods to define a narrow inversion layer in Si MOSFET's (b-d). Positive ($+$) and negative ($-$) charges on the gate electrodes are indicated. The location of the 2DEG is shown in black.
113: \label{fig2}
114: }
115: \end{figure}
116:
117: The conceptually simplest approach (Fig.\ \ref{fig2}b) to define a narrow channel is to scale down the width of the gate by means of electron beam lithography\cite{ref38} or other advanced techniques.\cite{ref39,ref40,ref41} A difficulty for the characterization of the device is that fringing fields beyond the gate induce a considerable uncertainty in the channel width, as well as its density. Such a problem is shared to some degree by all approaches, however, and this technique has been quite successful (as we will discuss in Section \ref{secII}). For a theoretical study of the electrostatic confining potential induced by the narrow gate, we refer to the work by Laux and Stern.\cite{ref42} This is a complicated problem, which requires a self-consistent solution of the Poisson and Schr\"{o}dinger equations, and must be solved numerically.
118:
119: The narrow gate technique has been modified by Warren et al.\cite{ref43,ref44} (Fig.\ \ref{fig2}c), who covered a multiple narrow-gate structure with a second dielectric followed by a second gate covering the entire device. (This structure was specifically intended to study one-dimensional superlattice effects, which is why multiple narrow gates were used.) By separately varying the voltages on the two gates, one achieves an increased control over channel width and density. The electrostatics of this particular structure has been studied in Ref.\ \onlinecite{ref43} in a semiclassical approximation.
120:
121: Skocpol et al.\cite{ref29,ref45} have combined a narrow gate with a deep self-aligned mesa structure (Fig.\ \ref{fig2}d), fabricated using dry-etching techniques. One advantage of their method is that at least an upper bound on the channel width is known unequivocally. A disadvantage is that the deep etch exposes the sidewalls of the electron gas, so that it is likely that some mobility reduction occurs due to sidewall scattering. In addition, the deep etch may damage the 2DEG itself. This approach has been used successfully in the exploration of nonlocal quantum transport in multiprobe channels, which in addition to being narrow have a very small separation of the voltage probes.\cite{ref45,ref46} In another investigation these narrow channels have been used as instruments sensitive to the charging and discharging of a single electron trap, allowing a detailed study of the statistics of trap kinetics.\cite{ref46,ref47,ref48}
122:
123: \subsection{\label{sec3} Nanostructures in GaAs-AlGaAs heterostructures}
124:
125: \begin{figure}
126: \centerline{\includegraphics[width=8cm]{figures/fig3}}
127: \caption{
128: Band-bending diagram of a modulation doped GaAs-${\rm Al}_x{\rm Ga}_{1-x}{\rm As}$ heterostructure. A 2DEG is formed in the undoped GaAs at the interface with the $p$-type doped AlGaAs. Note the Schottky barrier between the semiconductor and a metal electrode.
129: \label{fig3}
130: }
131: \end{figure}
132:
133: In a modulation-doped\cite{ref49} GaAs-AlGaAs heterostructure, the 2DEG is present at the interface between GaAs and ${\rm Al}_x{\rm Ga}_{1-x}{\rm As}$ layers (for a recent review, see Ref.\ \onlinecite{ref50}). Typically, the Al mole fraction $x = 0.3$. As shown in the band-bending diagram of Fig.\ \ref{fig3}, the electrons are confined to the GaAs-AlGaAs interface by a potential well, formed by the repulsive barrier due to the conduction band offset of about 0.3 V between the two semiconductors, and by the attractive electrostatic potential due to the positively charged ionized donors in the $n$-doped AlGaAs layer. To reduce scattering from these donors, the doped layer is separated from the interface by an undoped AlGaAs spacer layer. Two-dimensional sub bands are formed as a result of confinement perpendicular to the interface and free motion along the interface. An important advantage over a MOSFET is that the present interface does not interrupt the crystalline periodicity. This is possible because GaAs and AlGaAs have almost the same lattice spacing. Because of the absence of boundary scattering at the interface, the electron mobility can be higher by many orders of magnitude (see Table \ref{table1}). The mobility is also high because of the low effective mass $m = 0.067\,m_{\rm e}$ in GaAs (for a review of GaAs material properties, see Ref.\ \onlinecite{ref51}). As in a Si inversion layer, only a single two-dimensional subband (associated with the lowest discrete confinement level in the well) is usually populated. Since GaAs has a direct band gap, with a single conduction band minimum, complications due to intervalley scattering (as in Si) are absent. The one-dimensional Fermi surface is a circle, for the commonly used (100) substrate orientation.
134:
135: Since the 2DEG is present ``naturally'' due to the modulation doping (i.e., even in the absence of a gate), the creation of a narrow channel now requires the selective depletion of the electron gas in spatially separated regions. In principle, one could imagine using a combination of an undoped heterostructure and a narrow gate (similarly to a MOSFET), but in practice this does not work very well due to the lack of a natural oxide to serve as an insulator on top of the AlGaAs. The Schottky barrier between a metal and (Al)GaAs (see Fig.\ \ref{fig3}) is too low (only 0.9 V) to sustain a large positive voltage on the gate. For depletion-type devices, where a negative voltage is applied on the gate, the Schottky barrier is quite sufficient as a gate insulator (see, e.g., Ref.\ \onlinecite{ref52}).
136:
137: \begin{figure}
138: \centerline{\includegraphics[width=8cm]{figures/fig4}}
139: \caption{
140: Schematic cross-sectional views of four different ways to define narrow 2DEG
141: channels in a GaAs-AlGaAs heterostructure. Positive ionized donors and negative charges on a
142: Schottky gate electrode are indicated. The hatched squares in d represent unremoved resist used as a gate dielectric.
143: \label{fig4}
144: }
145: \end{figure}
146:
147: The simplest lateral confinement technique is illustrated in Fig.\ \ref{fig4}a. The appropriate device geometry (such as a Hall bar) is realized by defining a deep mesa, by means of wet chemical etching. Wide Hall bars are usually fabricated in this way. This approach has also been used to fabricate the first micron-scale devices, such as the constrictions used in the study of the breakdown of the quantum Hall effect by Kirtley et al.\cite{ref53} and Bliek et al.,\cite{ref54} and the narrow channels used in the first study of quasi-one-dimensional quantum transport in heterostructures by Choi et al.\cite{ref55} The deep-mesa confinement technique using wet\cite{ref25,ref56} or dry\cite{ref57} etching is still of use for some experimental studies, but it is generally felt to be unreliable for channels less than 1 $\mu{\rm m}$ wide (in particular because of the exposed sidewalls of the structure).
148:
149: The first working alternative confinement scheme was developed by Thornton et al.\cite{ref58} and Zheng et al.,\cite{ref24} who introduced the split-gate lateral confinement technique (Fig.\ \ref{fig4}b). On application of a negative voltage to a split Schottky gate, wide 2DEG regions under the gate are depleted, leaving a narrow channel undepleted. The most appealing feature of this confinement scheme is that the channel width and electron density can be varied continuously (but not independently) by increasing the negative gate voltage beyond the depletion threshold in the wide regions (typically about $-0.6$ V). The split-gate technique has become very popular, especially after it was used to fabricate the short and narrow constrictions known as quantum point contacts\cite{ref6,ref7,ref59} (see Section \ref{secIII}). The electrostatic confinement problem for the split-gate geometry has been studied numerically in Refs.\ \onlinecite{ref60} and \onlinecite{ref61}. A simple analytical treatment is given in in Ref. \protect\onlinecite{ref62}. A modification of the split-gate technique is the grating-gate technique, which may be used to define a 2DEG with a periodic density modulation.\cite{ref62}
150:
151: The second widely used approach is the shallow-mesa depletion technique (Fig.\ \ref{fig4}c), introduced in Ref.\ \onlinecite{ref63}. This technique relies on the fact that a 2DEG can be depleted by removal of only a thin layer of the AlGaAs, the required thickness being a sensitive function of the parameters of the heterostructure material, and of details of the lithographic process (which usually involves electron beam lithography followed by dry etching). The shallow-mesa etch technique has been perfected by two groups,\cite{ref64,ref65,ref66} for the fabrication of multi probe electron waveguides and rings.\cite{ref67,ref68,ref69,ref70} Submicron trenches\cite{ref71} are still another way to define the channel. For simple analytical estimates of lateral depletion widths in the shallow-mesa geometry, see Ref.\ \onlinecite{ref72}.
152:
153: A clever variant of the split-gate technique was introduced by Ford et al.\cite{ref73,ref74} A patterned layer of electron beam resist (an organic insulator) is used as a gate dielectric, in such a way that the separation between the gate and the 2DEG is largest in those regions where a narrow conducting channel has to remain after application of a negative gate voltage. As illustrated by the cross-sectional view in Fig.\ \ref{fig4}d, in this way one can define a ring structure, for example, for use in an Aharonov-Bohm experiment. A similar approach was developed by Smith et al.\cite{ref75} Instead of an organic resist they use a shallow-mesa pattern in the heterostructure as a gate dielectric of variable thickness. Initially, the latter technique was used for capacitive studies of one- and zero-dimensional confinement.\cite{ref75,ref76} More recently it was adopted for transport measurements as well.\cite{ref77} Still another variation of this approach was developed by Hansen et al.,\cite{ref78,ref79} primarily for the study of one-dimensional subband structure using infrared spectroscopy. Instead of electron beam lithography, they employ a photolithographic technique to define a pattern in the insulator. An array with a very large number of narrow lines is obtained by projecting the interference pattern of two laser beams onto light-sensitive resist. This technique is known as {\em holographic illumination\/} (see Section \ref{sec11b}).
154:
155: As two representative examples of state-of-the-art nanostructures, we show in Fig.\ \ref{fig5}a a miniaturized Hall bar,\cite{ref67} fabricated by a shallow-mesa etch, and in Fig.\ \ref{fig5}b a double-quantum-point contact device,\cite{ref80} fabricated by means of the split-gate technique.
156:
157: Other techniques have been used as well to fabricate narrow electron gas channels. We mention selective-area ion implantation using focused ion beams,\cite{ref81} masked ion beam exposure,\cite{ref82} strain-induced confinement,\cite{ref83} lateral $p$-$n$ junctions,\cite{ref84,ref85} gates in the plane of the 2DEG,\cite{ref86} and selective epitaxial growth.\cite{ref87,ref88,ref89,ref90,ref91,ref92} For more detailed and complete accounts of nanostructure fabrication techniques, we refer to Refs.\ \onlinecite{ref9} and \onlinecite{ref13,ref14,ref15}.
158:
159: \begin{figure}
160: \centerline{\includegraphics[width=8cm]{figures/fig5a}}\medskip
161:
162: \centerline{\includegraphics[width=8cm]{figures/fig5b}}
163: \caption{
164: Scanning electron micrographs of nanostructures in GaAs-AlGaAs heterostructures. (a, top) Narrow channel (width 75 nm), fabricated by means of the confinement scheme of Fig.\ \protect\ref{fig4}c. The channel has side branches (at a 2-$\mu{\rm m}$ separation) that serve as voltage probes. Taken from M. L. Roukes et al., Phys.\ Rev.\ Lett.\ {\bf 59}, 3011 (1987). (b, bottom) Double quantum point contact device, based on the confinement scheme of Fig. \protect\ref{fig4}b. The bar denotes a length of $1\,\mu{\rm m}$. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
165: \label{fig5}
166: }
167: \end{figure}
168:
169: \subsection{\label{sec4} Basic properties}
170:
171: \subsubsection{\label{sec4a} Density of states in two, one, and zero dimensions}
172:
173: The energy of conduction electrons in a single subband of an unbounded 2DEG, relative to the bottom of that subband, is given by
174: \begin{equation}
175: E(k)=\hbar^{2}k^{2}/2m,\label{eq4.1}
176: \end{equation}
177: as a function of momentum $\hbar k$. The effective mass $m$ is considerably smaller than the free electron mass $m_{e}$ (see Table \ref{table1}), as a result of interactions with the lattice potential. (The incorporation of this potential into an effective mass is an approximation\cite{ref20} that is completely justified for the present purposes.) The density of states $\rho(E)\equiv dn(E)/dE$ is the derivative of the number of electronic states $n(E)$ (per unit surface area) with energy smaller than $E$. In $k$-space, these states are contained within a circle of area $A=2\pi mE/\hbar^2$ [according to Eq.\ (\ref{eq4.1})], which contains a number $g_{\rm s} g_{\rm v} A/(2\pi)^2$ of distinct states. The factors $g_{\rm s}$ and $g_{\rm v}$ account for the spin degeneracy and valley degeneracy, respectively (Table \ref{table1}). One thus finds that $n(E) = g_{\rm s} g_{\rm v} mE/2\pi \hbar^2$, so the density of states corresponding to a single subband in a 2DEG,
178: \begin{equation}
179: \rho(E) = g_{\rm s} g_{\rm v} mE/2\pi \hbar^2,\label{eq4.2}
180: \end{equation}
181: is {\em independent\/} of the energy. As illustrated in Fig. \ref{fig6}a, a sequence of subbands is associated with the set of discrete levels in the potential well that confines the 2DEG to the interface. At zero temperature, all states are filled up to the Fermi energy $E_{\rm F}$ (this remains a good approximation at finite temperature if the thermal energy $k_{\rm B} T\ll E_{\rm F}$). Because of the constant density of states, the electron (sheet) density $n_{\rm s}$ is linearly related to $E_{\rm F}$ by $n_{\rm s} = E_{\rm F} g_{\rm s} g_{\rm v} m/2\pi \hbar^2$. The Fermi wave number $k_{\rm F} =(2mE_{\rm F} /\hbar^2)^{1/2}$ is thus related to the density by $k_{\rm F} = (4\pi n_{\rm s} /g_{\rm s} g_{\rm v})^{1/2}$. The second subband starts to be populated when $E_{\rm F}$ exceeds the energy of the second band bottom. The stepwise increasing density of states shown in Fig. \ref{fig6}a is referred to as {\em quasi}-two-dimensional. As the number of occupied subbands increases, the density of states eventually approaches the $\sqrt{E}$ dependence characteristic for a three-dimensional system. Note, however, that usually only a single subband is occupied.
182:
183: \begin{figure}
184: \centerline{\includegraphics[width=8cm]{figures/fig6a}}\medskip
185:
186: \centerline{\includegraphics[width=8cm]{figures/fig6b}}\medskip
187:
188: \centerline{\includegraphics[width=8cm]{figures/fig6c}}
189: \caption{
190: Density of states $\rho(E)$ as a function of energy. (a) Quasi-2D density of states, with only the lowest subband occupied (hatched). Inset: Confinement potential perpendicular to the plane of the 2DEG. The discrete energy levels correspond to the bottoms of the first and second 2D subbands. (b) Quasi-1D density of states, with four 1D subbands occupied. Inset: Square-well lateral confinement potential with discrete energy levels indicating the 1D subband bottoms. (c) Density of states for a 2DEG in a perpendicular magnetic field. The occupied 0D subbands or Landau levels are shown in black. Impurity scattering may broaden the Landau levels, leading to a nonzero density of states between the peaks.
191: \label{fig6}
192: }
193: \end{figure}
194:
195: If the 2DEG is confined laterally to a narrow channel, then Eq.\ (\ref{eq4.1}) only represents the kinetic energy from the free motion (with momentum $\hbar k$) {\em parallel\/} to the channel axis. Because of the lateral confinement, a single two-dimensional (2D) subband is split itself into a series of one-dimensional (1D) subbands, with band bottoms at $E_n$, $n = 1,2,\ldots$ The total energy $E_n (k)$ of an
196: electron in the $n$th 1D subband (relative to the bottom of the 2D subband) is given by
197: \begin{equation}
198: E_n (k)=E_n + \hbar^2 k^2 /2m.\label{eq4.3}
199: \end{equation}
200: Two frequently used potentials to model analytically the lateral confinement are the square-well potential (of width $W$, illustrated in Fig.\ \ref{fig6}b) and the parabolic potential well (described by $V(x)=\frac{1}{2}m\omega_{0}^2 x^2$). The confinement levels are then given either by $E_n = (n\pi \hbar)^2 /2mW^2$ for the square well or by $E_n = (n - \frac{1}{2})\hbar\omega_0$ for the parabolic well. When one considers electron transport through a narrow channel, it is useful to distinguish between states with positive and negative $k$, since these states move in opposite directions along the channel. We denote by $\rho_n^{+}(E)$ the density of states with $k > 0$ per unit channel length in the $n$th 1D subband. This quantity is given by
201: \begin{eqnarray}
202: \rho_{n}^{+}(E)&=&g_s g_v \left(2\pi\frac{dE_{n}(k)}{dk}\right)^{-1}\nonumber\\
203: &=&g_s g_v \frac{m}{2\pi\hbar^{2}}\left(\frac{\hbar^{2}}{2m(E-E_n)}\right)^{1/2}.\label{eq4.4}
204: \end{eqnarray}
205: The density of states $\rho_{n}^{-}$ with $k < 0$ is identical to $\rho_{n}^{+}$. (This identity holds because of time-reversal symmetry; In a magnetic field, $\rho_{n}^{+}\neq \rho_{n}^{-}$, in general.) The total density of states $\rho(E)$, drawn in Fig.\ \ref{fig6}b, is twice the result (\ref{eq4.4}) summed over all $n$ for which $E_n < E$. The density of states of a quasi-one-dimensional electron gas with many occupied 1D subbands may be approximated by the 2D result (\ref{eq4.2}).
206:
207: If a magnetic field $B$ is applied perpendicular to an unbounded 2DEG, the energy spectrum of the electrons becomes fully discrete, since free translational motion in the plane of the 2DEG is impeded by the Lorentz force. Quantization of the circular cyclotron motion leads to energy levels at\cite{ref93}
208: \begin{equation}
209: E_n =(n-{\textstyle\frac{1}{2}})\hbar\omega_{\rm c},\label{eq4.5}
210: \end{equation}
211: with $\omega_{\rm c} = eB/m$ the cyclotron frequency. The quantum number $n = 1, 2,\ldots$ labels the Landau levels. The number of states is the same in each Landau level and equal to one state (for each spin and valley) per flux quantum $h/e$ through the sample. To the extent that broadening of the Landau levels by disorder can be neglected, the density of states (per unit area) can be approximated by
212: \begin{equation}
213: \rho(E)=g_s g_v \frac{eB}{h}\sum_{n=1}^{\infty}\delta(E-E_{n}),\label{eq4.6}
214: \end{equation}
215: as illustrated in Fig.\ \ref{fig6}c. The spin degeneracy contained in Eq.\ (\ref{eq4.6}) is resolved in strong magnetic fields as a result of the Zeeman splitting $g\mu_{\rm B}B$ of the Landau levels ($\mu_{\rm B}\equiv e\hbar/2m_{\rm e}$ denotes the Bohr magneton; the Land\'{e} $g$-factor is a complicated function of the magnetic field in these systems).\cite{ref20} Again, if a large number of Landau levels is occupied (i.e., at weak magnetic fields), one recovers approximately the 2D result (\ref{eq4.2}). The foregoing considerations are for an unbounded 2DEG. A magnetic field perpendicular to a narrow 2DEG channel causes the density of states to evolve gradually from the 1D form of Fig.\ \ref{fig6}b to the effectively 0D form of Fig.\ \ref{fig6}c. This transition is discussed in Section \ref{sec10}.
216:
217: \subsubsection{\label{sec4b} Drude conductivity, Einstein relation, and Landauer formula}
218:
219: In the presence of an electric field $\mathbf{E}$ in the plane of the 2DEG, an electron
220: acquires a drift velocity $\mathbf{v}=-e\mathbf{E}\Delta t/m$ in the time $\Delta t$ since the last impurity collision. The average of $\Delta t$ is the scattering time $\tau$, so the average drift
221: velocity $\mathbf{v}_{\mathrm{drift}}$ is given by
222: \begin{equation}
223: v_{\mathrm{drift}}=-\mu_{\mathrm{e}}\mathrm{E},\;\;\mu_{\mathrm{e}}=e\tau/m.\label{eq4.7}
224: \end{equation}
225: The electron mobility $\mu_{\mathrm{e}}$ together with the sheet density $n_{\mathrm{s}}$ determine the conductivity $\sigma$ in the relation $-en_{\mathrm{s}}\mathbf{v}_{\mathrm{drift}}=\sigma \mathbf{E}$. The result is the familiar Drude conductivity,\cite{ref94} which can be written in several equivalent forms:
226: \begin{equation} \sigma=en_{\mathrm{s}}\mu_{\mathrm{e}}=\frac{e^{2}n_{\mathrm{s}}\tau}{m}=g_{\mathrm{s}}
227: g_{\mathrm{v}}\frac{e^{2}}{h}\frac{k_{\mathrm{F}}l}{2}.\label{eq4.8}
228: \end{equation}
229: In the last equality we have used the identity $n_{\mathrm{s}}=g_{\mathrm{s}}g_{\mathrm{v}}k_{\mathrm{
230: F}}^{2}/4\pi$ (see Section \ref{sec4a}) and have defined the mean free path $l=v_{\mathrm{F}}\tau$. The dimensionless quantity $k_{\mathrm{F}}l$ is much greater than unity in metallic systems (see Table \ref{table1} for typical values in a 2DEG), so the conductivity is large compared with the quantum unit $e^{2}/h\approx(26\,\mathrm{k}\Omega)^{-1}$.
231:
232: From the preceding discussion it is obvious that the current induced by the applied electric field is carried by {\it all\/} conduction electrons, since each electron acquires the same average drift velocity. Nonetheless, to determine the conductivity it is sufficient to consider the response of electrons {\it near the Fermi level\/} to the electric field. The reason is that the states that are more than a few times the thermal energy $k_{\mathrm{B}}T$ below $E_{\mathrm{F}}$ are all filled so that in response to a weak electric field only the distribution of electrons among
233: states at energies close to $E_{\mathrm{F}}$ is changed from the equilibrium Fermi-Dirac
234: distribution
235: \begin{equation}
236: f(E-E_{\mathrm{F}})=\left(1+ \exp\frac{E-E_{\mathrm{F}}}{k_{\mathrm{B}}T}\right)^{-1}.\label{eq4.9}
237: \end{equation}
238: The Einstein relation\cite{ref94}
239: \begin{equation}
240: \sigma=e^{2}\rho(E_{\mathrm{F}})D\label{eq4.10}
241: \end{equation}
242: is one relation between the conductivity and Fermi level properties (in this case the density of states $\rho(E)$ and the diffusion constant $D$, both evaluated at $E_{\mathrm{F}})$. The Landauer formula\cite{ref4} [Eq.\ (\ref{eq4.21})] is another such relation (in terms of the transmission probability at the Fermi level rather than in terms of the diffusion constant).
243:
244: The Einstein relation (\ref{eq4.10}) for an electron gas at zero temperature follows
245: on requiring that the sum of the drift current density $-\sigma \mathbf{E}/e$ and the diffusion
246: current density $-D\nabla n_{\mathrm{s}}$ vanishes in thermodynamic equilibrium, characterized by a spatially constant electrochemical potential $\mu$:
247: \begin{equation}
248: -\sigma \mathbf{E}/e-D\nabla n_{\mathrm{s}}=0,\;\; \mathrm{when}\;\; \nabla\mu=0.\label{eq4.11}
249: \end{equation}
250: The electrochemical potential is the sum of the electrostatic potential energy $-eV$ (which determines the energy of the bottom of the conduction band) and the chemical potential $E_{\mathrm{F}}$ (being the Fermi energy relative to the conduction band bottom). Since (at zero temperature) $dE_{\mathrm{F}}/dn_{\mathrm{s}}=1/\rho(E_{\mathrm{F}})$, one has
251: \begin{equation}
252: \nabla\mu=e\mathbf{E}+\rho(E_{\mathrm{F}})^{-1}\nabla n_{\mathrm{s}}.\label{eq4.12}
253: \end{equation}
254: The combination of Eqs.\ (\ref{eq4.11}) and (\ref{eq4.12}) yields the Einstein relation (\ref{eq4.10}) between $\sigma$ and $D$. To verify that Eq.\ (\ref{eq4.10}) is consistent with the earlier expression (\ref{eq4.8}) for the Drude conductivity, one can use the result (see below)
255: for the 2D diffusion constant:
256: \begin{equation}
257: D={\textstyle\frac{1}{2}}v_{\mathrm{F}}^{2}\tau={\textstyle\frac{1}{2}}v_\mathrm{F}l,\label{eq4.13}
258: \end{equation}
259: in combination with Eq.\ (\ref{eq4.2}) for the 2D density of states.
260:
261: At a finite temperature $T$, a chemical potential (or Fermi energy) gradient
262: $\nabla E_{\mathrm{F}}$ induces a diffusion current that is smeared out over an energy range of
263: order $k_{\mathrm{B}}T$ around $E_{\mathrm{F}}$. The energy interval between $E$ and $E+dE$ contributes to the diffusion current density $\mathbf{j}$ an amount $d\mathbf{j}$ given by
264: \begin{eqnarray}
265: d\mathbf{j}_{\mathrm{diff}}&=&-D\nabla\{ \rho(E)f(E-E_{\mathrm{F}})dE\}\nonumber\\
266: &=&-dED \rho(E)\frac{df}{dE_{\mathrm{F}}}\nabla E_{\mathrm{F}},\label{eq4.14}
267: \end{eqnarray}
268: where the diffusion constant $D$ is to be evaluated at energy $E$. The total
269: diffusion current density follows on integration over $E$:
270: \begin{equation}
271: \mathbf{j}=-\nabla E_{\mathrm{F}}e^{-2}\int_{0}^{\infty} dE\, \sigma(E,0) \frac{df}{dE_{\mathrm{F}}},\label{eq4.15}
272: \end{equation}
273: with $\sigma(E,0)$ the conductivity (\ref{eq4.10}) at temperature zero for a Fermi energy equal to $E$. The requirement of vanishing current for a spatially constant
274: electrochemical potential implies that the conductivity $\sigma(E_{\mathrm{F}},T)$ at temperature $T$ and Fermi energy $E_{\mathrm{F}}$ satisfies
275: \[
276: \sigma(E_{\mathrm{F}},T)e^{-2}\nabla E_{\mathrm{F}}+\mathbf{j}=0.
277: \]
278: Therefore, the finite-temperature conductivity is given simply by the energy average of the zero-temperature result
279: \begin{equation}
280: \sigma(E_{\mathrm{F}},T)=\int_{0}^{\infty} dE\, \sigma(E,0)\frac{df}{dE_{F}}.\label{eq4.16}
281: \end{equation}
282: As $T\rightarrow 0,$ $df/dE_{\mathrm{F}}\rightarrow\delta(E-E_{\mathrm{F}})$, so indeed only $E=E_{\mathrm{F}}$ contributes to the
283: energy average. Result (\ref{eq4.16}) contains exclusively the effects of a finite
284: temperature that are due to the thermal smearing of the Fermi-Dirac
285: distribution. A possible temperature dependence of the scattering processes is
286: not taken into account.
287:
288: We now want to discuss one convenient way to calculate the diffusion
289: constant (and hence obtain the conductivity). Consider the diffusion current
290: density $j_{x}$ due to a small constant density gradient, $n(x)=n_{0}+cx$. We write
291: \begin{eqnarray}
292: j_{x}&=&\lim_{\Delta t\rightarrow\infty}\langle v_{x}(t=0)n(x(t=-\Delta t))\rangle\nonumber\\
293: &=&\lim_{\Delta t\rightarrow\infty}c\langle v_{x}(0)x(-\Delta t)\rangle\nonumber\\
294: &=& \lim_{\Delta t\rightarrow\infty}-c\int_{0}^{\Delta t}dt\langle v_{x}(0)v_{x}(-t)\rangle,\label{eq4.17}
295: \end{eqnarray}
296: where $t$ is time and the brackets $\langle\cdots\rangle$ denote an isotropic angular average over the Fermi surface. The time interval $\Delta t\rightarrow\infty$, so the velocity of the electron at time $0$ is uncorrelated with its velocity at the earlier time $-\Delta t$. This allows us to neglect at $x(-\Delta t)$ the small deviations from an isotropic velocity distribution induced by the density gradient [which could not have
297: been neglected at $x(0)]$. Since only the time difference matters in the velocity
298: correlation function, one has $\langle v_{x}(0)v_{x}(-t)\rangle=\langle v_{x}(t)v_{x}(0)\rangle$. We thus obtain for the diffusion constant $D=-j_{x}/c$ the familiar linear response formula\cite{ref95}
299: \begin{equation}
300: D= \int_{0}^{\infty}dt\langle v_{x}(t)v_{x}(0)\rangle.\label{eq4.18}
301: \end{equation}
302: Since, in the semiclassical relaxation time approximation, each scattering
303: event is assumed to destroy all correlations in the velocity, and since a
304: fraction $\exp(-t/\tau$) of the electrons has not been scattered in a time $t$, one has
305: (in 2D)
306: \begin{equation}
307: \langle v_{x}(t)v_{x}(0)\rangle=\langle v_{x}(0)^{2}\rangle \mathrm{e}^{-t/\tau}={\textstyle\frac{1}{2}}v^{2}_\mathrm{F}\mathrm{e}^{-t/\tau}. \label{eq4.19}
308: \end{equation}
309: Substituting this correlation function for the integrand in Eq.\ (\ref{eq4.18}), one
310: recovers on integration the diffusion constant (\ref{eq4.13}).
311:
312: The Drude conductivity (4.8) is a {\it semiclassical\/} result, in the sense that
313: while the quantum mechanical Fermi-Dirac statistic is taken into account,
314: the dynamics of the electrons at the Fermi level is assumed to be classical. In
315: Section \ref{secII} we will discuss corrections to this result that follow from
316: correlations in the diffusion process due to quantum interference. Whereas
317: for classical diffusion correlations disappear on the time scale of the
318: scattering time $\tau$ [as expressed by the correlation function (\ref{eq4.19})], in quantum
319: diffusion correlations persist up to times of the order of the phase coherence
320: time. The latter time $\tau_{\phi}$ is associated with inelastic scattering and at low
321: temperatures can become much greater than the time $\tau$ associated with elastic
322: scattering.
323:
324: In an experiment one measures a {\it conductance\/} rather than a conductivity.
325: The conductivity $\sigma$ relates the local current density to the electric field,
326: $j=\sigma E$, while the conductance $G$ relates the total current to the voltage drop,
327: $I=GV$. For a large homogeneous conductor the difference between the two
328: is not essential, since Ohm's law tells us that
329: \begin{equation}
330: G=(W/L)\sigma\label{eq4.20}
331: \end{equation}
332: for a 2DEG of width $W$ and length $L$ in the current direction. (Note that $G$
333: and $\sigma$ have the same units in two dimensions.) If for the moment we disregard
334: the effects of phase coherence, then the simple scaling (\ref{eq4.20}) holds provided
335: both $W$ and $L$ are much larger than the mean free path $l$. This is the diffusive
336: transport regime, illustrated in Fig.\ \ref{fig7}a. When the dimensions of the sample
337: are reduced below the mean free path, one enters the {\it ballistic\/} transport
338: regime, shown in Fig.\ \ref{fig7}c. One can further distinguish an intermediate {\it quasi-ballistic\/} regime, characterized by $W<l<L$ (see Fig.\ \ref{fig7}b). In ballistic
339: transport only the conductance plays a role, not the conductivity. The
340: Landauer formula
341: \begin{equation}
342: G=(e^{2}/h)T\label{eq4.21}
343: \end{equation}
344: plays a central role in the study of ballistic transport because it expresses the
345: conductance in terms of a Fermi level property of the sample (the transmission probability $T$, see Section \ref{sec12}). Equation (\ref{eq4.21}) can therefore be
346: applied to situations where the conductivity does not exist as a local quantity,
347: as we will discuss in Sections \ref{secIII} and \ref{secIV}.
348:
349: \begin{figure}
350: \centerline{\includegraphics[width=8cm]{figures/fig7}}
351: \caption{
352: Electron trajectories characteristic for the diffusive ($l < W, L$), quasi-ballistic ($W < l < L$), and ballistic ($W, L < l$) transport regimes, for the case of specular boundary scattering. Boundary scattering and internal impurity scattering (asterisks) are of equal importance in the quasi-ballistic regime. A nonzero resistance in the ballistic regime results from back scattering at the connection between the narrow channel and the wide 2DEG regions. Taken from H. van Houten et al., in {\em Physics and Technology of Submicron Structures\/} (H. Heinrich, G. Bauer, and F. Kuchar, eds.). Springer, Berlin, 1988.
353: \label{fig7}
354: }
355: \end{figure}
356:
357: If phase coherence is taken into account, then the minimal length scale
358: required to characterize the conductivity becomes larger. Instead of the
359: (elastic) mean free path $l\equiv v_{\mathrm{F}}\tau$, the phase coherence length $l_{\phi}\equiv(D\tau_{\phi})^{1/2}$
360: becomes this characteristic length scale (up to a numerical coefficient
361: $l_{\phi}$ equals the average distance that an electron diffuses in the time $\tau_{\phi}$). Ohm's
362: law can now only be applied to add the conductances of parts of the sample
363: with dimensions greater than $l_{\phi}$. Since at low temperatures $l_{\phi}$ can become
364: quite large (cf.\ Table \ref{table1}), it becomes possible that (for a small conductor) phase coherence extends over a large part of the sample. Then only the conductance
365: (not the conductivity) plays a role, even if the transport is fully in the diffusive
366: regime. We will encounter such situations repeatedly in Section \ref{secII}.
367:
368: \subsubsection{\label{sec4c} Magnetotransport}
369:
370: In a magnetic field $B$ perpendicular to the 2DEG, the current is no longer
371: in the direction of the electric field due to the Lorentz force. Consequently,
372: the conductivity is no longer a scalar but a tensor $\bm{\sigma}$, related via the Einstein
373: relation $\bm{\sigma}=e^{2}\rho(E_{\mathrm{F}})\mathbf{D}$ to the diffusion tensor
374: \be
375: \mathbf{D}=\int_{0}^{\infty}dt\langle v(t)v(0)\rangle.\label{eq4.22}
376: \ee
377: Equation (\ref{eq4.22}) follows from a straightforward generalization of the argument leading to the scalar relation (\ref{eq4.18}) [but now the ordering of $v(t)$ and $v(0)$
378: matters]. Between scattering events the electrons at the Fermi level execute
379: circular orbits, with cyclotron frequency $\omega_{\mathrm{c}}=eB/m$ and cyclotron radius
380: $l_{\mathrm{cycl}}=mv_{\mathrm{F}}/eB$. Taking the 2DEG in the $x-y$ plane, and the magnetic field in
381: the positive $z$-direction, one can write in complex number notation
382: \be
383: \tilde{v}(t)\equiv v_{x}(t)+iv_{y}(t)=v_{\mathrm{F}}\exp(i\phi+i\omega_{\mathrm{c}}t).\label{eq4.23}
384: \ee
385: The diffusion tensor is obtained from
386: \begin{eqnarray}
387: D_{xx}+iD_{yx}&=& \int_{0}^{2\pi}\frac{d\phi}{2\pi}\int_{0}^{\infty}dt\,\tilde{v}(t)v_{\mathrm{F}}\cos\phi \mathrm{e}^{-t/\tau}\nonumber\\
388: &=&\frac{D}{1+(\omega_{\mathrm{c}}\tau)^{2}}(1+i\omega_{\mathrm{c}}\tau),\label{eq4.24}
389: \end{eqnarray}
390: where $D$ is the zero-field diffusion constant (\ref{eq4.13}). One easily verifies that
391: $D_{yy}=D_{xx}$ and $D_{xy}=-D_{yx}$. From the Einstein relation one then obtains the
392: conductivity tensor
393: \be
394: {\bm\sigma}=\frac{\sigma}{1+(\omega_{\mathrm{c}}\tau)^{2}}\left(\begin{array}{ll}
395: 1 & -\omega_{\mathrm{c}}\tau\\
396: \omega_{\mathrm{c}}\tau & 1
397: \end{array}\right),\label{eq4.25}
398: \ee
399: with $\sigma$ the zero-field conductivity (\ref{eq4.8}). The resistivity tensor ${\bm\rho}\equiv{\bm\sigma}^{-1}$ has the
400: form
401: \be
402: {\bm\rho}=\rho\left(\begin{array}{ll}
403: 1 & \omega_{\mathrm{c}}\tau\\
404: -\omega_{\mathrm{c}}\tau & 1
405: \end{array}\right), \label{eq4.26}
406: \ee
407: with $\rho=\sigma^{-1}=m/n_{\mathrm{s}}e^{2}\tau$ the zero-field resistivity.
408:
409: The off-diagonal element $\rho_{xy}\equiv R_{\mathrm{H}}$ is the classical {\it Hall\/} resistance of a
410: 2DEG:
411: \be
412: R_{\mathrm{H}}= \frac{B}{n_{\mathrm{s}}e}=\frac{1}{g_{\mathrm{s}}g_{\mathrm{v}}}\frac{h}{e^{2}}\frac{\hbar\omega_{\mathrm{c}}}{E_{\mathrm{F}}}. \label{eq4.27}
413: \ee
414: Note that in a 2D channel geometry there is no distinction between the Hall
415: {\it resistivity\/} and the Hall {\it resistance}, since the ratio of the Hall voltage
416: $V_{\mathrm{H}}=WE_{x}$ across the channel to the current $I=Wj_{y}$ along the channel does
417: not depend on its length and width (provided transport remains in the
418: diffusive regime). The diagonal element $\rho_{xx}$ is referred to as the {\it longitudinal\/}
419: resistivity. Equation (\ref{eq4.26}) tells us that classically the magnetoresistivity is
420: zero (i.e., $\rho_{xx}(B)-\rho_{xx}(0)=0)$. This counterintuitive result can be understood
421: by considering that the force from the Hall voltage cancels the average
422: Lorentz force on the electrons. A general conclusion that one can draw from
423: Eqs.\ (\ref{eq4.25}) and (\ref{eq4.26}) is that the classical effects of a magnetic field are
424: important only if $\omega_{\mathrm{c}}\tau\gtrsim 1$. In such fields an electron can complete several
425: cyclotron orbits before being scattered out of orbit. In a high-mobility 2DEG
426: this criterion is met at rather weak magnetic fields (note that $\omega_{\mathrm{c}}\tau=\mu_{\mathrm{e}}B$, and
427: see Table \ref{table1}).
428:
429: \begin{figure}
430: \centerline{\includegraphics[width=8cm]{figures/fig8}}
431: \caption{
432: Schematic dependence on the reciprocal filling factor $\nu^{-1}\equiv 2eB/hn_{\rm s}$ of the longitudinal resistivity $\rho_{xx}$ (normalized to the zero-field resistivity $\rho$) and of the Hall resistance $R_{\rm H}\equiv \rho_{xy}$ (normalized to $h/2e^{2}$). The plot is for the case of a single valley with twofold spin degeneracy. Deviations from the semiclassical result (\ref{eq4.26}) occur in strong magnetic fields, in the form of Shubnikov-De Haas oscillations in $\rho{xx}$ and quantized plateaus [Eq.\ (\ref{eq4.30})] in $\rho_{xy}$.
433: \label{fig8}
434: }
435: \end{figure}
436:
437: In the foregoing application of the Einstein relation we have used the zero-field density of states. Moreover, we have assumed that the scattering time is
438: $B$-independent. Both assumptions are justified in weak magnetic fields, for
439: which $E_{\mathrm{F}}/h\omega_{\mathrm{c}}\gg 1$, but not in stronger fields (cf.\ Table \ref{table1}). As illustrated in Fig.\
440: \ref{fig8}, deviations from the semiclassical result (\ref{eq4.26}) appear as the magnetic field is
441: increased. These deviations take the form of an oscillatory magnetoresistivity
442: (the {\it Shubnikov-De Haas effect}) and plateaux in the Hall resistance (the
443: {\it quantum Hall effect}). The origin of these two phenomena is the formation of
444: Landau levels by a magnetic field, discussed in Section \ref{sec4a}, that leads to the $B$-dependent density of states (\ref{eq4.6}). The main effect is on the scattering rate $\tau^{-1}$,
445: which in a simple (Born) approximation\cite{ref96} is proportional to $\rho(E_{\mathrm{F}})$:
446: \be
447: \tau^{-1}=(\pi/\hbar)\rho(E_{\mathrm{F}})c_{\mathrm{i}}u^{2}. \label{eq4.28}
448: \ee
449: Here $c_{\mathrm{i}}$ is the areal density of impurities, and the impurity potential is
450: modeled by a 2D delta function of strength $u$. The diagonal element of the
451: resistivity tensor (\ref{eq4.26}) is $\rho_{xx}=(m/e^{2}n_{\mathrm{s}})\tau^{-1}\propto\rho(E_{\mathrm{F}})$. Oscillations in the
452: density of states at the Fermi level due to the Landau level quantization are
453: therefore observable as an oscillatory magnetoresistivity. One expects the
454: resistivity to be minimal when the Fermi level lies between two Landau levels,
455: where the density of states is smallest. In view of Eq.\ (\ref{eq4.6}), this occurs when
456: the Landau level {\it filling factor\/} $\nu\equiv(n_{\mathrm{s}}/g_{\mathrm{s}}g_{\mathrm{v}})(h/eB)$ equals an integer $N=1,2$,
457: $\ldots$ (assuming spin-degenerate Landau levels). The resulting Shubnikov-De
458: Haas oscillations are periodic in $1/B$, with spacing $\Delta(1/B)$ given by
459: \be
460: \Delta(\frac{1}{B})=\frac{e}{h}\frac{g_{\mathrm{s}}g_{\mathrm{v}}}{n_{\mathrm{s}}}, \label{eq4.29}
461: \ee
462: providing a means to determine the electron density from a magnetoresistance measurement. This brief explanation of the Shubnikov-De Haas effect
463: needs refinement,\cite{ref20} but is basically correct. The quantum Hall effect,\cite{ref8} being
464: the occurrence of plateaux in $R_{\mathrm{H}}$ versus $B$ at precisely
465: \be
466: R_{\mathrm{H}}= \frac{1}{g_{\mathrm{s}}g_{\mathrm{v}}}\frac{h}{e^{2}}\frac{1}{N}, \;\;N=1,2,\ldots, \label{eq4.30}
467: \ee
468: is a more subtle effect\cite{ref97} to which we cannot do justice in a few lines (see
469: Section \ref{sec18}). The quantization of the Hall resistance is related on a fundamental level to the quantization in zero magnetic field of the resistance of a
470: ballistic point contact.\cite{ref6,ref7} We will present a unified description of both these
471: effects in Sections \ref{sec12} and \ref{sec13}.
472:
473: \section{\label{secII} Diffusive and quasi-ballistic transport}
474:
475: \subsection{\label{sec5} Classical size effects}
476:
477: In metals, the dependence of the resistivity on the size of the sample has been the subject of study for almost a century.\cite{ref98} Because of the small Fermi wave length in a metal, these are {\em classical\/} size effects. Comprehensive reviews of this field have been given by Chambers,\cite{ref99} Br\"{a}ndli and Olsen,\cite{ref100} Sondheimer,\cite{ref101} and, recently, Pippard.\cite{ref102} In semiconductor nanostructures both classical and quantum size effects appear, and an understanding of the former is necessary to distinguish them from the latter. Classical size effects in a 2DEG are of intrinsic interest as well. First of all, a 2DEG is an ideal model system to study known size effects without the complications of nonspherical Fermi surfaces and polycrystallinity, characteristic for metals. Furthermore, it is possible in a 2DEG to study the case of nearly complete specular boundary scattering, whereas in a metal diffuse scattering dominates. The much smaller cyclotron radius in a 2DEG, compared with a metal at the same magnetic field value, allows one to enter the regime where the cyclotron radius is comparable to the range of the scattering potential. The resulting modifications of known effects in the quasi-ballistic transport regime are the subject of this section. A variety of new classical size effects, not known from metals, appear in the ballistic regime, when the resistance is measured on a length scale below the mean free path. These are discussed in Section \ref{sec16}, and require a reconsideration of what is meant by a resistance on such a short length scale.
478:
479: In the present section we assume that the channel length $L$ (or, more generally, the separation between the voltage probes) is much larger than the mean free path $l$ for impurity scattering so that the motion remains diffusive along the channel. Size effects in the resistivity occur when the motion across the channel becomes ballistic (i.e., when the channel width $W < l$). Diffuse boundary scattering leads to an increase in the resistivity in a zero magnetic field and to a nonmonotonic magnetoresistivity in a perpendicular magnetic field, as discussed in the following two subsections. The 2D channel geometry is essentially equivalent to the 3D geometry of a thin metal plate in a parallel magnetic field, with the current flowing perpendicular to the field. Size effects in this geometry were originally studied by Fuchs\cite{ref103} in a zero magnetic field and by MacDonald\cite{ref104} for a nonzero field. The alternative configuration in which the magnetic field is perpendicular to the thin plate, studied by Sondheimer\cite{ref105} does not have a 2D analog. We discuss in this section only the classical size effects, and thus the discreteness of the 1D subbands and of the Landau levels is ignored. Quantum size effects in the quasi-ballistic transport regime are treated in Section \ref{sec10}.
480:
481: \subsubsection{\label{sec5a} Boundary scattering}
482:
483: In a zero magnetic field, scattering at the channel boundaries increases the resistivity, unless the scattering is specular. {\em Specular scattering\/} occurs if the confining potential $V(x, y)$ does not depend on the coordinate $y$ along the channel axis. In that case the electron motion along the channel is not influenced at all by the lateral confinement, so the resistivity $\rho$ retains its 2D bulk value $\rho_0 =m/e^2 n_{\rm s}\tau$. More generally, specular scattering requires any roughness of the boundaries to be on a length scale smaller than the Fermi wavelength $\lambda_{\rm F}$. The confining potential created electrostatically by means of a gate electrode is known to cause predominantly specular scattering (as has been demonstrated by the electron focusing experiments\cite{ref59} discussed in Section \ref{sec14}). This is a unique situation, not previously encountered in metals, where as a result of the small $\lambda_{\rm F}$ (on the order of the interatomic separation) diffuse boundary scattering dominates.\cite{ref102}
484:
485: {\em Diffuse\/} scattering means that the velocity distribution at the boundary is isotropic for velocity directions that point away from the boundary. Note that this implies that an incident electron is reflected with a (normalized) angular distribution $P(\alpha) = \frac{1}{2} \cos\alpha$, since the reflection probability is proportional to the flux normal to the boundary. Diffuse scattering increases the resistivity above $\rho_0$ by providing an upper bound $W$ to the effective mean free path. In order of magnitude, $\rho\sim(l/W)\rho_{0}$ if $l\gtrsim W$ (a more precise expression is derived later). In general, boundary scattering is neither fully specular nor fully diffuse and, moreover, depends on the angle of incidence (grazing incidence favors specular scattering since the momentum along the channel is large and not easily reversed). The angular dependence is often ignored for simplicity, and the boundary scattering is described, following Fuchs,\cite{ref103} by a single parameter $p$, such that an electron colliding with the boundary is reflected specularly with probability $p$ and diffusely with probability $1-p$.
486: This specularity parameter is then used as a fit parameter in comparison with experiments. Soffer\cite{ref106} has developed a more accurate, and more complicated, modeling in terms of an angle of incidence dependent specularity parameter.
487:
488: In the extreme case of fully diffuse boundary scattering ($p = 0$), one is justified in neglecting the dependence of the scattering probability on the angle of incidence. We treat this case here in some detail to contrast it with fully specular scattering, and because diffuse scattering can be of importance in 2DEG channels defined by ion beam exposure rather than by gates.\cite{ref107,ref108} We calculate the resistivity from the diffusion constant by means of the Einstein relation. Fuchs takes the alternative (but equivalent) approach of calculating the resistivity from the linear response to an applied electric field.\cite{ref103} Impurity scattering is taken as isotropic and elastic and is described by a scattering time $\tau$ such that an electron is scattered in a time interval $dt$ with probability $dt/\tau$, regardless of its position and velocity, This is the commonly employed ``scattering time'' (or ``relaxation time'') approximation.
489:
490: The channel geometry is defined by hard walls at $x=\pm W/2$ at which the
491: electrons are scattered diffusely. The stationary electron distribution function
492: at the Fermi energy $F(\mathbf{r}, \alpha)$ satisfies the Boltzmann equation
493: \be
494: \mathbf{v}\cdot\frac{\partial}{\partial\mathbf{r}}F=-\frac{1}{\tau}F+\frac{1}{\tau}\int_{0}^{2\pi}\frac{d\alpha}{2\pi}F, \label{eq5.1}
495: \ee
496: where $\mathbf{r}\equiv(x, y)$ is the position and $\alpha$ is the angle that the velocity $ \mathbf{v}\equiv v_{\mathrm{F}}(\cos \alpha, \sin \alpha)$ makes with the $x$-axis. The boundary condition corresponding to
497: diffuse scattering is that $F$ is independent of the velocity direction for
498: velocities pointing away from the boundary. In view of current conservation
499: this boundary condition can be written as
500: \begin{eqnarray}
501: F(\mathbf{r}, \alpha)&=& \frac{1}{2}\int_{-\pi/2}^{\pi/2}d\alpha^{\prime}\,F(\mathbf{r}, \alpha^{\prime})\cos \alpha^{\prime},\nonumber\\
502: &&{\rm for}\;\; x= \frac{W}{2},\; \frac{\pi}{2}<\alpha<\frac{3\pi}{2},\nonumber\\
503: &=& \frac{1}{2}\int_{\pi/2}^{3\pi/2}d\alpha^{\prime}\,F(\mathbf{r}, \alpha^{\prime})\cos \alpha^{\prime},\nonumber\\
504: &&{\rm for}\;\; x=- \frac{W}{2},\; - \frac{\pi}{2}<\alpha<\frac{\pi}{2}. \label{eq5.2}
505: \end{eqnarray}
506: To determine the diffusion constant, we look for a solution of Eqs.\ (\ref{eq5.1}) and
507: (\ref{eq5.2}) corresponding to a constant density gradient along the channel,
508: $F(\mathbf{r}, \alpha)=-cy+f(x, \alpha)$. Since there is no magnetic field, we anticipate that the
509: density will be uniform across the channel width so that
510: $\int_{0}^{2\pi}fd\alpha=0$. The
511: Boltzmann equation (\ref{eq5.1}) then simplifies to an ordinary differential equation
512: for $f$, which can be solved straightforwardly. The solution that satisfies the
513: boundary conditions (\ref{eq5.2}) is
514: \be
515: F(\mathbf{r}, \alpha)=-cy+cl\sin\alpha \left[1- \exp\left(-\frac{W}{2l|\cos \alpha|}-\frac{x}{l\cos \alpha}\right)\right], \label{eq5.3}
516: \ee
517: where we have written $l\equiv v_{\mathrm{F}}\tau$. One easily verifies that $F$ has indeed a uniform
518: density along $x$. The diffusion current
519: \be
520: I_{y}=v_{\mathrm{F}} \int_{-W/2}^{W/2}dx\int_{0}^{2\pi}d\alpha\,F\sin\alpha \label{eq5.4}
521: \ee
522: along the channel in response to the density gradient $\partial n/\partial y=-2\pi c$
523: determines the diffusion constant $D=-(I_{y}/W)(\partial n/\partial y)^{-1}$. The resistivity
524: $\rho=E_{\mathrm{F}}/n_{\mathrm{s}}e^{2}D$ then follows from the Einstein relation (\ref{eq4.10}), with the 2D
525: density of states $n_{\mathrm{s}}/E_{\mathrm{F}}$. The resulting expression is
526: \be
527: \rho=\rho_{0}\left[1-\frac{4l}{\pi W}\int_{0}^{1}d\xi\,\xi(1-\xi^{2})^{1/2}(1-\mathrm{e}^{-W/l\xi})\right]^{-1},\label{eq5.5}
528: \ee
529: which can be easily evaluated numerically. It is worth noting that the
530: above result\cite{ref109} for $\rho/\rho_{0}$ in a 2D channel geometry does not differ much
531: (less than 20\%) from the corresponding result\cite{ref103} in a 3D thin film.
532:
533: For $l/W\ll 1$ one has
534: \be
535: \rho=\rho_0 \left( 1+\frac{4}{3\pi}\frac{l}{W}\right),\label{eq5.6}
536: \ee
537: which differs from Eq.\ (\ref{eq5.5}) by less than 10\% in the range $l/W\lesssim 10$. For $l/W\gg 1$ one has asymptotically
538: \begin{eqnarray}
539: \rho&=&\frac{\pi}{2}\rho_{0}\frac{l}{W}\frac{1}{\ln(l/W)}\nonumber\\
540: &=&\frac{\pi}{2}\frac{mv_{\rm F}}{n_{\rm s}e^2 W}\frac{1}{\ln(l/W)}.\label{eq5.7}
541: \end{eqnarray}
542: In the absence of impurity scattering (i.e., in the limit $l\rightarrow\infty$), Eq.\ (\ref{eq5.7}) predicts a vanishing resistivity. Diffuse boundary scattering is ineffective in establishing a finite resistivity in this limit, because electrons with velocities nearly parallel to the channel walls can propagate over large distances without collisions and thereby short out the current. As shown by Tesanovic et al.,\cite{ref110} a small but nonzero resistivity in the absence of impurity scattering is recovered if one goes beyond the semiclassical approximation and includes the effect of the quantum mechanical uncertainty in the transverse component of the electron velocity.
543:
544: \subsubsection{\label{sec5b} Magneto size effects}
545:
546: In an unbounded 2DEG, the longitudinal resistivity is magnetic-field independent in the semiclassical approximation (see Section \ref{sec4c}). We will discuss how a nonzero magnetoresistivity can arise classically as a result of boundary scattering. We consider the two extreme cases of specular and diffuse boundary scattering, and describe the impurity scattering in the scattering time approximation. Shortcomings of this approximation are discussed toward the end of this subsection.
547:
548: We consider first the case of specular boundary scattering. In a zero magnetic field it is obvious that specular scattering cannot affect the resistivity, since the projection of the electron motion on the channel axis is not changed by the presence of the channel boundaries. If a magnetic field is applied perpendicular to the 2DEG, the electron trajectories in a channel cannot be mapped in this way on the trajectories in an unbounded system. In fact, in an unbounded 2DEG in equilibrium the electrons perform closed cyclotron orbits between scattering events, whereas a channel geometry supports open orbits that skip along the boundaries. One might suppose that the presence of these {\em skipping orbits\/} propagating along the channel would increase the diffusion constant and hence reduce the (longitudinal) resistivity below the value $\rho_0$ of a bulk 2DEG. That is not correct, at least in the scattering time approximation, as we now demonstrate.
549:
550: The stationary Boltzmann equation in a magnetic field $\bf B$ in the $z$-direction (perpendicular to the 2DEG) is
551: \be
552: \mathbf{v}\cdot\frac{\partial}{\partial\mathbf{r}}F+\omega_{\rm c}\frac{\partial}{\partial\alpha}F=
553: -\frac{1}{\tau}F+\frac{1}{\tau}\int_{0}^{2\pi}\frac{d\alpha}{2\pi}F.\label{eq5.8}
554: \ee
555: Here, we have used the identity $-em^{-1}(\mathbf{v}\times\mathbf{B})\cdot\partial/\partial\mathbf{v}\equiv\omega_{\rm c}\partial/\partial\alpha$ (with $\omega_{\rm c}\equiv eB/m$ the cyclotron frequency) to rewrite the term that accounts for the Lorentz force. The distribution function $F(\mathbf{r},\alpha)$ must satisfy the boundary conditions for specular scattering,
556: \be
557: F(\mathbf{r},\alpha)=F(\mathbf{r},\pi-\alpha),\;\;{\rm for}\;\;x=\pm W/2.\label{eq5.9}
558: \ee
559: One readily verifies that
560: \be
561: F(\mathbf{r},\alpha)=-c(y+\omega_{\rm c}\tau x)+cl\sin\alpha\label{eq5.10}
562: \ee
563: is a solution of Eqs.\ (\ref{eq5.8}) and (\ref{eq5.9}). The corresponding diffusion current $I_y = \pi cWv_{\rm F}l$ and density gradient along the channel $\partial n/\partial y= -2\pi c$ are both the same as in a zero magnetic field. It follows that the diffusion constant $D = I_y /2\pi c W$ and, hence, the longitudinal resistivity $\rho = E_{\rm F} /n_{\rm s} e^2 D$ are $B$-independent; that is, $\rho=\rho_0 \equiv m/n_{\rm s} e^2 \tau$, as in an unbounded 2DEG. More generally, one can show that in the scattering time approximation the longitudinal resistivity is $B$-independent for {\em any\/} confining potential $V(x, y)$ that does not vary with the coordinate $y$ along the channel axis. (This statement is proven by applying the result of Ref.\ \onlinecite{ref111}, of a $B$-independent $\rho_{yy}$ for periodic $V(x)$, to a set of disjunct parallel channels (see Section \ref{sec11b}); the case of a single channel then follows from Ohm's law.)
564:
565: \begin{figure}
566: \centerline{\includegraphics[width=8cm]{figures/fig9}}
567: \caption{
568: Illustration of the effect of a magnetic field on motion through a channel with diffuse boundary scattering. (a) Electrons which in a zero field move nearly parallel to the boundary can reverse their motion in weak magnetic fields. This increases the resistivity. (b) Suppression of back scattering at the boundaries in strong magnetic fields reduces the resistivity.
569: \label{fig9}
570: }
571: \end{figure}
572:
573: In the case of diffuse boundary scattering, the zero-field resistivity is enhanced by approximately a factor $1 + l/2W$ [see Eq.\ (\ref{eq5.6})]. A sufficiently strong magnetic field suppresses this enhancement, and reduces the resistivity to its bulk value $\rho_0$. The mechanism for this negative magnetoresistance is illustrated in Fig.\ \ref{fig9}b. If the cyclotron diameter $2l_{\rm cycl}$ is smaller than the channel width $W$, diffuse boundary scattering cannot reverse the direction of motion along the channel, as it could for smaller magnetic fields. The diffusion current is therefore approximately the same as in the case of specular scattering, in which case we have seen that the diffusion constant and, hence, resistivity have their bulk values. Figure \ref{fig9} represents an example of {\em magnetic reduction of backscattering}. Recently, this phenomenon has been understood to occur in an extreme form in the quantum Hall effect\cite{ref112} and in ballistic transport through quantum point contacts.\cite{ref113} The effect was essentially known and understood by MacDonald\cite{ref104} in 1949 in the course of his magnetoresistivity experiments on sodium wires. The ultimate reduction of the resistivity is preceded by an initial increase in weak magnetic fields, due
574: to the deflection toward the boundary of electrons with a velocity nearly parallel to the channel axis (Fig.\ \ref{fig9}a). The resulting nonmonotonic $B$-dependence of the resistivity is shown in Fig.\ \ref{fig10}. The plot for diffuse scattering is based on a calculation by Ditlefsen and Lothe\cite{ref114} for a 3D thinfilm geometry. The case of a 2D channel has been studied by Pippard\cite{ref102} in the limit $l/W\rightarrow\infty$, and he finds that the 2D and 3D geometries give very similar results.
575:
576: \begin{figure}
577: \centerline{\includegraphics[width=8cm]{figures/fig10}}
578: \caption{
579: Magnetic field dependence of the longitudinal resistivity of a channel for the two cases of diffuse and specular boundary scattering, obtained from the Boltzmann equation in the scattering time approximation. The plot for diffuse scattering is the result of Ref.\ \onlinecite{ref114} for a 3D thin film geometry with $l=10W$. (A 2D channel geometry is expected to give very similar results.\cite{ref102})
580: \label{fig10}
581: }
582: \end{figure}
583:
584: \begin{figure}
585: \centerline{\includegraphics[width=8cm]{figures/fig11}}
586: \caption{
587: Experimental magnetic field dependence of the resistance of channels of different widths, defined by ion beam exposure in the 2DEG of a GaAs-AlGaAs heterostructure ($L= 12\,\mu{\rm m}$, $T = 4.2\,{\rm K}$). The nonmonotonic magnetic field dependence below $1\,{\rm T}$ is a classical size effect due to diffuse boundary scattering, as illustrated in Fig.\ \ref{fig9}. The magnetoresistance oscillations at higher fields result from the quantum mechanical Shubnikov-De Haas effect. Taken from T. J. Thornton et al., Phys.\ Rev.\ Lett.\ {\bf 63}, 2128 (1989).
588: \label{fig11}
589: }
590: \end{figure}
591:
592: An experimental study of this effect in a 2DEG has been performed by Thornton et al.\cite{ref107} In Fig.\ \ref{fig11} their magnetoresistance data are reproduced for channels of different widths $W$, defined by low-energy ion beam exposure. It was found that the resistance reaches a maximum when $W\approx 0.5\,l_{\rm cycl}$, in excellent agreement with the theoretical predictions.\cite{ref114,ref102} Thornton et al.\ also investigated channels defined electrostatically by a split gate, for which one expects predominantly specular boundary scattering.\cite{ref59} The foregoing analysis would then predict an approximately $B$-independent resistance (Fig.\ \ref{fig10}), and indeed only a small resistance maximum was observed in weak magnetic fields. At stronger fields, however, the resistance was found to decrease substantially. Such a monotonically decreasing resistance in channels with predominantly specular boundary scattering was first reported by Choi et al.,\cite{ref55} and studied for a narrower channel in Ref.\ \onlinecite{ref27} (see Section \ref{sec9b} for some of these experimental results). We surmise that a classical negative magnetoresistance in the case of specular boundary scattering can result if the cyclotron radius becomes smaller than some characteristic correlation length in the distribution of impurities (and in the resulting potential landscape). Correlations between the positions of impurities and the channel boundaries, which are neglected in the scattering time approximation, will then play a role. For an example, see Fig.\ \ref{fig12}, which shows how an isolated impurity near the boundary can reverse the direction of electron motion in a zero magnetic field but not in a sufficiently strong magnetic field. In metals, where the cyclotron radius is much larger than in a 2DEG, an electron will effectively experience a random impurity potential between subsequent boundary collisions, so the scattering can well be described in terms of an average relaxation time. The experiments in a 2DEG suggest that this approximation breaks down at relatively weak magnetic fields.
593:
594: \begin{figure}
595: \centerline{\includegraphics[width=8cm]{figures/fig12}}
596: \caption{
597: Electron trajectories in a channel with specular boundary scattering, to illustrate how a magnetic field can suppress the back scattering by an isolated impurity close to a boundary. This effect would lead to a negative magneto resistivity if one would go beyond the scattering time approximation.
598: \label{fig12}
599: }
600: \end{figure}
601:
602: \subsection{\label{sec6} Weak localization}
603:
604: The temperature dependence of the Drude resistivity $\rho = m/n_{\rm s}e^2 \tau$ is contained in that of the scattering time $\tau$, since the electron density is constant in a degenerate electron gas. As one lowers the temperature, inelastic scattering processes (such as electron-phonon scattering) are suppressed, leading to a decrease in the resistivity. The residual resistivity is due entirely to elastic scattering (with stationary impurities or other crystalline defects) and is temperature-independent in the semiclassical theory. Experimentally, however, one finds that below a certain temperature the resistivity of the 2DEG starts to rise again. The increase is very small in broad samples, but becomes quite pronounced in narrow channels. This is illustrated in Fig.\ \ref{fig13}, where the temperature dependencies of the resistivities of wide and narrow GaAs-AlGaAs heterostructures are compared.\cite{ref63}
605:
606: The anomalous resistivity increase is due to long-range correlations in the diffusive motion of an electron that are purely quantum mechanical. In the semiclassical theory it is assumed that a few scattering events randomize the electron velocity, so the velocity correlation function decays exponentially in time with decay time $\tau$ [see Eq.\ (\ref{eq4.19})]. As discussed in Section \ref{sec4c}, this assumption leads to the Drude formula for the resistivity. It is only in recent years that one has come to appreciate that purely elastic scattering is not effective in destroying correlations in the phase of the electron wave function. Such correlations lead to quantum interference corrections to the Drude result, which can explain the anomalous increase in the resistivity at low temperatures.
607:
608: \begin{figure}
609: \centerline{\includegraphics[width=8cm]{figures/fig13}}
610: \caption{
611: Temperature dependence of the resistivity of a wide 2DEG in a GaAs-AlGaAs heterostructure (circles) and of two narrow channels of lithographic width $W_{\rm lith}= 1.5\,\mu{\rm m}$ (squares) and $W_{\rm lith}= 0.5\,\mu{\rm m}$ (triangles). The channel length $L = 10\,\mu{\rm m}$. The resistivity is estimated from the measured resistance $R$ by multiplying by $W_{\rm lith}/L$, disregarding the difference between the conducting and lithographic width in the narrow channels. Taken from H. van Houten et al., Appl.\ Phys.\ Lett.\ {\bf 49}, 1781 (1986).
612: \label{fig13}
613: }
614: \end{figure}
615:
616: A striking effect of quantum interference is to enhance the probability for backscattering in a disordered system in the metallic regime. This effect has been interpreted as a precursor of localization in strongly disordered systems and has thus become known as {\em weak localization}.\cite{ref115,ref116,ref117} In Section \ref{sec6a} we describe the theory for weak localization in a zero magnetic field. The application of a magnetic field perpendicular to the 2DEG suppresses weak localization,\cite{ref118} as discussed in Section \ref{sec6b}. The resulting negative magnetoresistivity is the most convenient way to resolve experimentally the weak localization correction.\cite{ref119} The theory for a narrow channel in the quasiballistic transport regime\cite{ref109,ref120} differs in an interesting way from the theory for the diffusive regime,\cite{ref121} as a consequence of the flux cancellation effect.\cite{ref122} The diffusive and quasi-ballistic regimes are the subjects of Sections \ref{sec6b} and \ref{sec6c}, respectively.
617:
618: \subsubsection{\label{sec6a} Coherent backscattering}
619:
620: The theory of weak localization was developed by Anderson et al.\cite{ref116} and Gorkov et al.\cite{ref117} This is a diagrammatic perturbation theory that does not lend itself easily to a physical interpretation. The interpretation of weak localization as {\em coherent backscattering\/} was put forward by Bergmann\cite{ref123} and by Khmel'nitskii and Larkin,\cite{ref124,ref125} and formed the basis of the path integral theory of Chakravarty and Schmid.\cite{ref126} In this description, weak localization is understood by considering the interference of the probability amplitudes for the classical trajectories (or ``Feynman paths'') from one point to another, as discussed later. For reviews of the alternative diagrammatic approach, we refer to Refs.\ \onlinecite{ref127} and \onlinecite{ref128}.
621:
622: \begin{figure}
623: \centerline{\includegraphics[width=8cm]{figures/fig14}}
624: \caption{
625: Mechanism of coherent back scattering. The probability amplitudes $A_i$ and $A_j$ of two trajectories from $\mathbf{r}$ to $\mathbf{r}'$ have uncorrelated phases in general (a), but the amplitudes $A^+$ and $A^-$ of two time-reversed returning trajectories are equal (b). The constructive interference of $A^+$ and $A^-$ increases the probability for return to the point of departure, which is the origin of the weak localization effect. The volume indicated in black is the area $\lambda_{\rm F}v_{\rm F}dt$ covered by a flux tube in a time interval $dt$, which enters in Eq.\ (\ref{eq6.2}) for the conductivity correction.
626: \label{fig14}
627: }
628: \end{figure}
629:
630: In a Feynman path description\cite{ref129} of diffusion, the probability $P(\mathbf{r},\mathbf{r}', t)$ for motion from point $\mathbf{r}$ to point $\mathbf{r}'$ in a time $t$ consists of the absolute value squared of the sum of probability amplitudes $A_i$, one for each trajectory from $\mathbf{r}$ to $\mathbf{r}'$ of duration $t$:
631: \be
632: P(\mathbf{r},\mathbf{r}',t)=\left| \sum_{i}A_i \right|^2 =\sum_i |A_i|^2 +\sum_{i\neq j}A_i A_j^{\ast}.\label{eq6.1}
633: \ee
634: The restriction to {\em classical\/} trajectories in the sum over Feynman paths is allowed if the separation between scattering events is much larger than the wavelength (i.e., if $k_{\rm F}l\gg 1$). The classical diffusion probability corresponds to the first term on the right-hand side of Eq.\ (\ref{eq6.1}), while the second term accounts for quantum interference. In the diffusive transport regime there is a very large number of different trajectories that contribute to the sum. One might suppose that for this reason the interference term averages out, because different trajectories have uncorrelated phases. This is correct if the beginning and end points $\mathbf{r}$ and $\mathbf{r}'$ are different (Fig.\ \ref{fig14}a), but not if the two coincide (Fig.\ \ref{fig14}b). In the latter case of ``backscattered'' trajectories, one can group the contributions to the sum (\ref{eq6.1}) in time-reversed pairs. Time-reversal invariance guarantees that the probability amplitudes $A^+$ and $A^-$ for clockwise and counterclockwise propagation around the closed loop are identical: $A^+ = A^- \equiv A$. The coherent backscattering probability $|A^+ + A^-|^2 =4|A|^2$ is then twice the classical result. The enhanced probability for return to the point of departure reduces the diffusion constant and, hence, the conductivity. This is the essence of weak localization. As phrased by Chakravarty and Schmid,\cite{ref126} ``it is one of those unique cases where the superposition principle of quantum mechanics leads to observable consequences at the macroscopic level.''
635:
636: The magnitude of the weak localization correction $\delta\sigma_{\mathrm{loc}}$ to the Drude
637: conductivity $\sigma$ is proportional to the probability for return to the point of
638: departure.\cite{ref126} Since $\delta\sigma_{\mathrm{loc}}$ is assumed to be a small correction, one can estimate this probability from classical diffusion. Let $C(t)d\mathbf{r}$ denote the classical
639: probability that an electron returns after a time $t$ to within $d\mathbf{r}$ of its point of
640: departure. The weak localization correction is given by the time integral of
641: the return probability:
642: \be
643: \frac{\delta\sigma_{\mathrm{loc}}}{\sigma}=-\frac{2\hbar}{m}\int_{0}^{\infty}dt\,C(t)\mathrm{e}^{-t/\tau_{\phi}}. \label{eq6.2}
644: \ee
645: The correction is negative because the conductivity is reduced by coherent
646: backscattering. The factor $\hbar/m\propto\lambda_{\mathrm{F}}v_{\mathrm{F}}$ follows in the path integral formalism
647: from the area covered by a flux tube of width $\lambda_{\mathrm{F}}$ and length $v_{\mathrm{F}}dt$ (see Fig.\ \ref{fig14}b). The factor $\exp(- t/\tau_{\phi})$ is inserted ``by hand'' to account for the loss of phase
648: coherence after a time $\tau_{\phi}$ (as a result of inelastic scattering). The return
649: probability $C(t)$ in a 2D channel of width $W$ is given for times $t\gg \tau$ in the
650: diffusive regime by
651: \begin{subequations}
652: \label{eq6.3}
653: \begin{eqnarray}
654: C(t)&=&(4\pi Dt)^{-1},\;\;{\rm if}\;\;t\ll W^{2}/D, \label{eq6.3a}\\
655: C(t)&=&W^{-1}(4\pi Dt)^{-1/2},\;\;{\rm if}\;\;t\gg W^{2}/D. \label{eq6.3b}
656: \end{eqnarray}
657: \end{subequations}
658: The $1/t$ decay of the return probability (\ref{eq6.3a}) assumes unbounded diffusion in
659: two dimensions. A crossover to a lower $1/\sqrt{t}$ decay (\ref{eq6.3b}) occurs when the
660: root-mean-square displacement $(2Dt)^{1/2}$ exceeds the channel width, so
661: diffusion occurs effectively in one dimension only. Because the time integral of
662: $C(t)$ itself diverges, the weak localization correction (\ref{eq6.2}) is determined by the
663: behavior of the return probability on the phase coherence time $\tau_{\phi}$, which
664: provides a long-time cutoff. One speaks of 2D or 1D weak localization,
665: depending on whether the return probability $C(\tau_{\phi})$ on the time scale of $\tau_{\phi}$ is
666: determined by 2D diffusion (\ref{eq6.3a}) or by 1D diffusion (\ref{eq6.3b}). In terms of the
667: phase coherence length $l_{\phi}\equiv(D\tau_{\phi})^{1/2}$, the criterion for the dimensionality is
668: that 2D weak localization occurs for $l_{\phi}\ll W$ and 1D weak localization for
669: $l_{\phi}\gg W$. On short time scales $t\lesssim\tau$, the motion is ballistic rather than diffusive,
670: and Eq.\ (\ref{eq6.3}) does not apply. One expects the return probability to go to zero
671: smoothly as one enters the ballistic regime. This short-time cutoff can be
672: accounted for heuristically by the factor $1- \exp(- t/\tau)$, to exclude those
673: electrons that at time $t$ have not been scattered.\cite{ref109} The form of the short-time
674: cutoff becomes irrelevant for $\tau_{\phi}\gg \tau$. (See Ref.\ \onlinecite{ref130} for a theoretical study of weak localization in the regime of comparable $\tau_{\phi}$ and $\tau$.)
675:
676: The foregoing analysis gives the following expressions for the 2D and 1D
677: weak localization corrections:
678: \begin{widetext}
679: \begin{subequations}
680: \label{eq6.4}
681: \begin{eqnarray}
682: \delta\sigma_{\mathrm{loc}}&=&-\frac{2\hbar}{m}\sigma\int_{0}^{\infty}dt\,(4\pi Dt)^{-1}(1-e^{-t/\tau})e^{-t/\tau_{\phi}}=-g_{\mathrm{s}}g_{\mathrm{v}} \frac{e^{2}}{4\pi^{2}\hbar}\ln\left(1+\frac{\tau_{\phi}}{\tau}\right),\;\;{\rm if}\;\; l_{\phi}\ll W, \label{eq6.4a}\\
683: \delta\sigma_{\mathrm{loc}}&=&-\frac{2\hbar}{m}\sigma\int_{0}^{\infty}dt\,W^{-1}(4\pi Dt)^{-1/2}(1-e^{-t/\tau)}e^{-t/\tau_{\phi}}=-g_{\mathrm{s}}g_{ \mathrm{v}} \frac{e^{2}}{2\pi \hbar}\frac{l_{\phi}}{W}\left(1-\left(1+\frac{\tau_{\phi}}{\tau}\right)^{-1/2}\right),\;\;{\rm if}\;\; l_{\phi}\gg W,\nonumber\\&& \label{eq6.4b}
684: \end{eqnarray}
685: \end{subequations}
686: \end{widetext}
687: where we have used the expression for the Drude conductivity $\sigma=e^{2}\rho(E_{\mathrm{F}})D$
688: with the 2D density of states (\ref{eq4.2}). The ratio of the weak localization
689: correction to the Drude conductivity $\delta\sigma_{\mathrm{loc}}/\sigma$ is of order $1/k_{\mathrm{F}}l$ for 2D weak
690: localization and of order $(l_{\phi}/W)(1/k_{\mathrm{F}}l)$ for 1D weak localization. In the 2D
691: case, the correction is small (cf.\ the values of $k_{\mathrm{F}}l$ given in Table \ref{table1}), but still
692: much larger than in a typical metal. The correction is greatly enhanced in the
693: 1D case $l_{\phi}\gg W$. This is evident in the experimental curves in Fig.\ \ref{fig13}, in which
694: the resistivity increase at low temperatures is clearly visible only in the
695: narrowest channel.
696:
697: The weak localization correction to the conductance $\delta G_{\mathrm{loc}}\equiv(W/L)\delta\sigma_{\mathrm{loc}}$ is
698: of order $(e^{2}/h)(W/L)$ in the 2D case and of order $(e^{2}/h)(l_{\phi}/L)$ in the 1D case. In
699: the latter case, the conductance correction does not scale with the channel
700: width $W$, contrary to what one would have classically. The conductance does
701: scale with the reciprocal of the channel length $L$, at least for $L\gg l_{\phi}$. The factor
702: $l_{\phi}/L$ in $\delta G_{\mathrm{loc}}$ in the 1D case can be viewed as a consequence of the classical
703: series addition of $L/l_{\phi}$ channel sections. It will then be clear that the scaling
704: with $L$ has to break down when $L\lesssim l_{\phi}$, in which case the weak localization
705: correction saturates at its value for $L\approx l_{\phi}$. The maximum conductance
706: correction in a narrow channel is thus of order $e^{2}/h$, independent of the
707: properties of the sample. This ``universality'' is at the origin of the phenomenon of the universal conductance fluctuations discussed in Section \ref{sec7}.
708:
709: \subsubsection{\label{sec6b} Suppression of weak localization by a magnetic field}
710:
711: {\bf (a) Theory.} The resistance enhancement due to weak localization can be
712: suppressed by the application of a weak magnetic field oriented perpendicular to the 2DEG. The suppression results from the fact that a
713: magnetic field breaks time-reversal invariance. We recall that in a zero
714: magnetic field, time-reversal invariance guarantees that trajectories that form
715: a closed loop have equal probability amplitudes $A^{+}$ and $A^{-}$ for clockwise
716: and counterclockwise propagation around the loop. The resulting constructive interference enhances the backscattering probability, thereby leading to the weak localization effect. In a weak magnetic field, however, a phase
717: difference $\phi$ develops between $A^{+}$ and $A^{-}$, even if the curvature of the
718: trajectories by the Lorentz force can be totally negected. This Aharonov-Bohm phase results from the fact that the canonical momentum $\mathbf{p}=m \mathbf{v}-e\mathbf{A}$
719: of an electron in a magnetic field contains the vector potential $\mathbf{A}$. On
720: clockwise $(+)$ and counterclockwise $(-)$ propagation around a closed loop,
721: one thus acquires a phase difference
722: \begin{eqnarray}
723: \phi&=&\hbar^{-1}\oint_{+}\mathbf{p}^{+}\cdot d{\mathbf l}-\hbar^{-1}\oint_{-}\mathbf{p}^{-}\cdot d\mathbf{l}\nonumber\\
724: &=& \frac{2e}{\hbar}\int(\nabla\times \mathbf{A})\cdot d\mathbf{S}=\frac{2eBS}{\hbar}\equiv\frac{2S}{l_{\mathrm{m}}^{2}}\equiv 4\pi\frac{\Phi}{\Phi_{0}}.\nonumber\\
725: && \label{eq6.5}
726: \end{eqnarray}
727: The phase difference is twice the enclosed area $S$ divided by the square of the
728: magnetic length $l_{\mathrm{m}}\equiv(\hbar/eB)^{1/2}$, or, alternatively, it is $4\pi$ times the enclosed
729: flux $\Phi$ in units of the elementary flux quantum $\Phi_{0}\equiv h/e$.
730:
731: Many trajectories, with a wide distribution of loop areas, contribute to the
732: weak localization effect. In a magnetic field the loops with a large area $S\gtrsim l_{\mathrm{m}}^{2}$
733: no longer contribute, since on average the counterpropagating trajectories no
734: longer interfere constructively. Since trajectories enclosing a large area
735: necessarily take a long time to complete, the effect of a magnetic field is
736: essentially to introduce a long-time cutoff in the integrals of Eqs.\ (\ref{eq6.2}) and
737: (\ref{eq6.4}), which is the magnetic relaxation time $\tau_{B}$. Recall that the long-time cutoff
738: in the absence of a magnetic field is the phase coherence time $\tau_{\phi}$. The
739: magnetic field thus begins to have a significant effect on weak localization if
740: $\tau_{B}$ and $\tau_{\phi}$ are comparable, which occurs at a characteristic field $B_{\mathrm{c}}$. The weak
741: localization effect can be studied experimentally by measuring the negative
742: magnetoresistance peak associated with its suppression by a magnetic field.
743: The significance of such experiments relies on the possibility of directly
744: determining the phase coherence time $\tau_{\phi}$. The experimental data are most
745: naturally analyzed in terms of the conductance. The magnitude of the zero-field conductance correction $\delta G_{\mathrm{loc}}(B=0)$ follows directly from the saturation
746: value of the magnetoconductance, according to
747: \be
748: G(B\gg B_{\mathrm{c}})-G(B=0)=-\delta G_{\mathrm{loc}}(B=0). \label{eq6.6}
749: \ee
750: Once $\delta G_{\mathrm{loc}}(B=0)$ is known, one can deduce the phase coherence length $l_{\phi}$
751: from Eq.\ (\ref{eq6.4}), since $D$ and $\tau$ are easily estimated from the classical part of the
752: conductance (which dominates at slightly elevated temperatures). The magnetoconductance contains, in addition, information on the channel width $W$,
753: which is a parameter difficult to determine otherwise, as will become clear in
754: the discussion of the experimental situation in subsection (b).
755:
756: \begin{table*}
757: \caption{Magnetic relaxation time $\tau_{B}$ and characteristic field $B_{\rm c}$ for the suppression of 2D and 1D weak localization.
758: \label{table2}
759: }
760: \begin{ruledtabular}
761: \begin{tabular}{ccccc}
762: &\multicolumn{2}{c}{Dirty Metal\footnotemark[1]\footnotemark[2] $(l\ll W)$}&\multicolumn{2}{c}{Pure Metal\footnotemark[1]\footnotemark[3] $(W\ll l)$}\\
763: &2D ($l_{\phi}\ll W)$&1D ($W\ll l_{\phi}$)&1D weak field ($l_{\rm m}^{2}\gg Wl$)& 1D strong field ($Wl\gg l_{\rm m}^2 \gg W^2$)\\
764: \hline
765: &&&&\\
766: $\tau_{B}$&$\displaystyle\frac{l_{\rm m}^2}{2D}$&$\displaystyle\frac{3l_{\rm m}^4}{W^2 D}$&$\displaystyle\frac{C_1 l_{\rm m}^4}{W^3 v_{\rm F}}$&$\displaystyle\frac{C_2 l_{\rm m}^2 l}{W^2 v_{\rm F}}$\\
767: &&&&\\
768: $B_{\rm c}$&$\displaystyle\frac{\hbar}{e}\frac{1}{2 l_{\phi}^2}$&$\displaystyle\frac{\hbar}{e}\frac{3^{1/2}}{W l_{\phi}}$&$\displaystyle\frac{\hbar}{e}\frac{1}{W}\left(\frac{C_{1}}{Wv_{\rm F}\tau_{\phi}}\right)^{1/2}$&$\displaystyle\frac{\hbar}{e}\frac{C_{2}l}{W^2 v_{\rm F}\tau_{\phi}}$
769: \end{tabular}
770: \end{ruledtabular}
771: \footnotetext[1]{
772: All results assume a channel length $L\gg l_{\phi}$, a channel width $W\gg\lambda_{\rm F}$, as well as $\tau_{\phi}\gg\tau$.}
773: \footnotetext[2]{
774: From Refs.\ \onlinecite{ref118,ref131}, and \onlinecite{ref121}. The diffusion constant $D=\frac{1}{2}v_{\rm F}l$. If $W\ll l_{\phi}$, a transition to 2D weak localization occurs when $l_{\rm m}\lesssim W$.}
775: \footnotetext[3]{
776: From Ref.\ \onlinecite{ref109}. The constants are given by $C_1 =9.5$ and $C_2 = 24/5$ for specular boundary scattering ($C_1 =4\pi$ and $C_2 =3$ for a channel with diffuse boundary scattering). For pure metals, the case $l_{\rm m}<W$ is outside the diffusive transport regime for weak localization.}
777: \end{table*}
778:
779: The effectiveness of a magnetic field in suppressing weak localization (as
780: contained in the functional dependence of $\tau_{B}$ on $B$, or in the expression for $B_{\mathrm{c}}$)
781: is determined by the average flux enclosed by backscattered trajectories of a
782: given duration. One can distinguish different regimes, depending on the
783: relative magnitude of the channel width $W$, the mean free path $l\equiv v_{\mathrm{F}}\tau$, the
784: magnetic length $l_{\mathrm{m}}$, and the phase coherence length $l_{\phi}\equiv(D\tau_{\phi})^{1/2}$. In Table \ref{table2}
785: the expressions for $\tau_{B}$ and $B_{\mathrm{c}}$ are summarized, as obtained by various
786: authors.\cite{ref109,ref118,ref121,ref131} In the following, we present a simple physical interpretation that explains these results, except for the numerical prefactors. We will
787: not discuss the effects of spin-orbit scattering\cite{ref131} or of superconducting
788: fluctuations,\cite{ref132} since these may be neglected in the systems considered in this
789: review. In this subsection we only discuss the dirty metal regime $l\ll W$. The
790: pure metal regime $l\gg W$, in which boundary scattering plays an important
791: role, will be discussed in Section \ref{sec6c}.
792:
793: If $l_{\phi}\ll W$ the {\it two--dimensional\/} weak localization correction to the conductivity applies, given by Eq.\ (\ref{eq6.4a}) for a zero magnetic field. The typical area
794: $S$ enclosed by a backscattered trajectory on a time scale $\tau_{B}$ is then of the order
795: $S\sim D\tau_{B}$ (assuming diffusive motion on this time scale). The corresponding
796: phase shift is $\phi\sim D\tau_{B}/l_{\mathrm{m}}^{2}$, in view of Eq.\ (\ref{eq6.5}). The criteria $\phi\sim 1$ and $\tau_{B}\sim\tau_{\phi}$
797: thus imply
798: \be
799: \tau_{B}\sim l_{\mathrm{m}}^{2}/D;\;\;B_{\mathrm{c}}\sim h/eD\tau_{\phi}\equiv h/el_{\phi}^{2}. \label{eq6.7}
800: \ee
801: The full expression for the magnetoconductance due to weak localization
802: is\cite{ref118,ref131}
803: \begin{widetext}
804: \be
805: \delta G_{\mathrm{loc}}^{2\mathrm{D}}(B)-\delta G_{\mathrm{loc}}^{2\mathrm{D}}(0)=\frac{W}{L}g_{\mathrm{s}}g_{ \mathrm{v}}\frac{e^{2}}{4\pi^{2}\hbar}\left[\Psi\left(\frac{1}{2}+\frac{\tau_{B}}{2\tau_{\phi}}\right)-\Psi\left(\frac{1}{2}+\frac{\tau_{B}}{2\tau}\right)+\ln\left(\frac{\tau_{\phi}}{\tau}\right)\right],
806: \label{eq6.8}
807: \ee
808: \end{widetext}
809: where $\Psi(x)$ is the digamma function and $\tau_{B}=l_{\mathrm{m}}^{2}/2D$. The digamma function
810: has the asymptotic approximation $\Psi(x)\approx\ln(x)-1/x$ for large $x$; thus, in a
811: zero magnetic field result (\ref{eq6.4a}) is recovered (assuming also $\tau_{\phi}\gg \tau$). In the
812: case of 2D weak localization the characteristic field $B_{\mathrm{c}}$ is usually very weak.
813: For example, if $l_{\phi}=1\,\mu \mathrm{m}$, then $B_{\mathrm{c}}\approx 1\,\mathrm{mT}$. The suppression of the weak
814: localization effect is complete when $\tau_{B}\lesssim\tau$, which occurs for
815: $B\gtrsim \hbar/eD\tau\sim \hbar/el^{2}$. These fields are still much weaker than classically strong
816: fields for which $\omega_{\mathrm{c}}\tau\gtrsim 1$ (as can be verified by noting that when $B=\hbar/el^{2}$, one
817: has $\omega_{\mathrm{c}}\tau=1/k_{\mathrm{F}}l\ll 1$). The neglect of the curvature of electron trajectories in
818: the theory of weak localization is thus entirely justified in the 2D case. The
819: safety margin is narrower in the 1D case, however, since the characteristic
820: fields can become significantly enhanced.
821:
822: The {\it one-dimensional\/} case $W\ll l_{\phi}$ in a magnetic field has first been treated
823: by Al'tshuler and Aronov\cite{ref121} in the dirty metal regime. This refers to a narrow
824: channel with $l\ll W$ so that the wall-to-wall motion is diffusive. Since the
825: phase coherence length exceeds the channel width, the backscattered trajectories on a time scale $\tau_{B}$ have a typical enclosed area $S\sim W(D\tau_{B})^{1/2}$ (see Fig.
826: \ref{fig15}). Consequently, the condition $S\sim l_{\mathrm{m}}^{2}$ for a unit phase shift implies
827: \be
828: \tau_{B}\sim l_{\mathrm{m}}^{4}/DW^{2};\;\; B_{\mathrm{c}}\sim h/eWl_{\phi}. \label{eq6.9}
829: \ee
830: The difference with the 2D case is that the enclosed flux on a given time scale
831: is reduced, due to the lateral compression of the backscattered trajectories.
832: This leads to an enhancement by a factor $l_{\phi}/W$ of the characteristic field scale $B_{\mathrm{c}}$, compared with Eq.\ (\ref{eq6.7}). The full expression for the weak localization
833: correction if $l_{\phi}, l_{\mathrm{m}}\gg W\gg l$ is\cite{ref121}
834: \be
835: \delta G_{\mathrm{loc}}^{1\mathrm{D}}(B)=-g_{\mathrm{s}}g_{ \mathrm{v}}\frac{e^{2}}{h}\frac{1}{L}\left(\frac{1}{D\tau_{\phi}}+\frac{1}{D\tau_{B}}\right)^{-1/2}, \label{eq6.10}
836: \ee
837: with $\tau_{B}=3l_{\mathrm{m}}^{4}/W^{2}D$. For an elementary derivation of this result, see Ref.\ \onlinecite{ref109}.
838: At $l_{\mathrm{m}}\sim W$ a crossover from 1D to 2D weak localization occurs [i.e., from Eq.\
839: (\ref{eq6.10}) to Eq.\ (\ref{eq6.8})]. The reason for this crossover is that the lateral
840: confinement becomes irrelevant for the weak localization when $l_{\mathrm{m}}\lesssim W$,
841: because the trajectories of duration $\tau_{B}$ then have a typical extension
842: $(D\tau_{B})^{1/2}\lesssim W$, according to Eq.\ (\ref{eq6.9}). This crossover from 1D to 2D restricts
843: the available field range that can be used to study the magnetoconductance
844: associated with 1D weak localization.
845:
846: \begin{figure}
847: \centerline{\includegraphics[width=8cm]{figures/fig15}}
848: \caption{
849: Typical closed electron trajectory contributing to 1D weak localization ($l_{\phi}\gg W$) in the dirty metal regime ($l\ll W$). The asterisks denote elastic scattering events. Taken from H. van Houten et al., Acta Electronica {\bf 28}, 27 (1988).
850: \label{fig15}
851: }
852: \end{figure}
853:
854: The magnetic relaxation time $\tau_{B}$ in the dirty metal regime is found to be
855: inversely proportional to the diffusion constant $D$, in 2D as well as in 1D. The
856: reason for this dependence is clear: faster diffusion implies that less time is
857: needed to complete a loop of area $l_{\mathrm{m}}^{2}$. It is remarkable that in the pure metal
858: regime such a proportionality no longer holds. This is a consequence of the
859: flux cancellation effect discussed in Section \ref{sec6c}.
860:
861: {\bf (b) Experiments in the dirty metal regime.} Magnetoresistance experiments
862: have been widely used to study the weak localization correction to the
863: conductivity of wide 2D electron gases in Si\cite{ref28,ref30,ref133,ref134,ref135} and GaAs.\cite{ref23,ref136,ref137}
864: Here we will discuss the experimental magnetoresistance studies of weak
865: localization in narrow channels in Si MOSFETs\cite{ref34,ref38,ref40,ref138} and GaAs-AlGaAs heterostructures.\cite{ref24,ref25,ref58} As an illustrative example, we reproduce in
866: Fig.\ \ref{fig16} a set of experimental results for $\delta R/R\equiv[R(0)-R(B)]/R(0)$ obtained
867: by Choi et al.\cite{ref25} in a wide and in a narrow GaAs-AlGaAs heterostructure.
868: The quantity $\delta R$ is positive, so the resistance decreases on applying a
869: magnetic field. The 2D results are similar to those obtained earlier by
870: Paalanen et al.\cite{ref137} The qualitative difference in field scale for the suppression
871: of 2D (top) and 1D (bottom) weak localization is nicely illustrated by the data
872: in Fig.\ \ref{fig16}. The magnetoresistance peak is narrower in the 2D case, consistent
873: with the enhancement in 1D of the characteristic field $B_{\mathrm{c}}$ for the suppression
874: of weak localization, which we discussed in Section \ref{sec6b}(a). The solid curves in
875: Fig.\ \ref{fig16} were obtained from the 2D theoretical expression (\ref{eq6.8}) and the 1D
876: dirty metal result (\ref{eq6.10}), treating $W$ and $l_{\phi}$ as adjustable parameters. A
877: noteworthy finding of Choi et al.\cite{ref25} is that the effective channel width $W$ is
878: considerably reduced below the lithographic width $W_{\mathrm{lith}}$ in narrow channels
879: defined by a deep-etched mesa (as in Fig.\ \ref{fig4}a). Differences $W-W_{\mathrm{lith}}$ of about
880: $0.8\,\mu \mathrm{m}$ were found.\cite{ref25} Significantly smaller differences are obtained\cite{ref27,ref63} if a
881: shallow-etched mesa is used for the lateral confinement, as in Fig.\ \ref{fig4}c. A split-gate device (as in Fig.\ \ref{fig4}b) of variable width has been used by Zheng et al.\cite{ref24} to study weak localization in GaAs-AlGaAs heterostructure channels. Magnetoresistance experiments in a very narrow split-gate device (fabricated using
882: electron beam lithography) were reported by Thornton et al.\cite{ref58} and analyzed
883: in terms of the dirty metal theory. Unfortunately, in their experiment the
884: mean free path of $450\,\mathrm{nm}$ exceeded the width inferred from a fit to Eq.\ (\ref{eq6.10})
885: by an order of magnitude, so an analysis in terms of the pure metal theory
886: would have been required.
887:
888: \begin{figure}
889: \centerline{\includegraphics[width=8cm]{figures/fig16}}
890: \caption{
891: A comparison between the magnetoresistance $\Delta R/R\equiv[R(0)-R(B)]/R(0)$ due to 2D weak localization in a wide channel (upper panel) and due to 1D weak localization in a narrow channel (lower panel), at various temperatures. The solid curves are fits based on Eqs.\ (\ref{eq6.8}) and (\ref{eq6.10}). Taken from K. K. Choi et al., Phys.\ Rev.\ B {\bf 36}, 7751 (1987).
892: \label{fig16}
893: }
894: \end{figure}
895:
896: Early magnetoresistance experiments on narrow Si accumulation layers
897: were performed by Dean and Pepper,\cite{ref34} in which they observed evidence for a
898: crossover from the 2D to the 1D weak localization regime. A comparison of
899: weak localization in wide and narrow Si inversion layers was reported by
900: Wheeler et al.\cite{ref38} The conducting width of the narrow channel was taken to be
901: equal to the lithographic width of the gate (about $400\,\mathrm{nm}$), while the mean
902: free path was estimated to be about $100\,\mathrm{nm}$. This experiment on a low-mobility Si channel thus meets the requirement $l\ll W$ for the dirty metal
903: regime. The 1D weak localization condition $l_{\phi}\gg W$ was only marginally
904: satisfied, however. Licini et al.\cite{ref40} reported a negative magnetoresistance peak
905: in 270-nm-wide Si inversion layers, which was well described by the 2D
906: theory at a temperature of $2.2\,\mathrm{K}$, where $l_{\phi}=120\,\mathrm{nm}$. Deviations from the 2D
907: form were found at lower temperatures, but the 1D regime was never fully
908: entered. A more recent study of 1D weak localization in a narrow Si
909: accumulation layer has been performed by Pooke et al.\cite{ref138} at low temperatures, and the margins are somewhat larger in their case.
910:
911: We note a difficulty inherent to experiments on 1D weak localization in
912: semiconductor channels in the dirty metal regime. For 1D weak localization
913: it is required that the phase coherence length $l_{\phi}$ is much larger than the
914: channel width. If the mean free path is short, then the experiment is in the
915: dirty metal regime $l\ll W$, but the localization will be only marginally one-dimensional since the phase coherence length $l_{\phi}\equiv(D\tau_{\phi})^{1/2}=(v_{F}l\tau_{\phi}/2)^{1/2}$ will
916: be short as well (except for the lowest experimental temperatures). If the mean
917: free path is long, then the 1D criterion $l_{\phi}\gg W$ is easily satisfied, but the
918: requirement $l\ll W$ will now be hard to meet so that the experiment will tend
919: to be in the pure metal regime. A quantitative comparison with the theory
920: (which would allow a reliable determination of $l_{\phi}$) is hampered because the
921: asymptotic regimes studied theoretically are not accessible experimentally
922: and because the channel width is not known a priori. Nanostructures are thus
923: not the best candidates for a quantitative study of the phase coherence length,
924: which is better studied in 2D systems. An altogether different complication is
925: that quantum corrections to the conductivity in semiconductor nanostructures can be remarkably large (up to 100\% at sufficiently low temperatures\cite{ref27,ref34}), which puts them beyond the range of validity of the perturbation theory.
926:
927: \subsubsection{\label{sec6c} Boundary scattering and flux cancellation}
928:
929: {\bf (a) Theory.} In the previous subsection we noticed that the pure metal
930: regime, where $l\gg W$, is characteristic for 1D weak localization in semiconductor nanostructures. This regime was first theoretically considered by
931: Dugaev and Khmel'nitskii,\cite{ref120} for the geometry of a thin metal film in a
932: parallel magnetic field and for diffuse boundary scattering. The geometry of a
933: narrow 2DEG channel in a perpendicular magnetic field, with either diffuse
934: or specular boundary scattering, was treated by the present authors.\cite{ref109} Note
935: that the nature of the boundary scattering did not play a role in the dirty
936: metal regime of Section \ref{sec6b}, since there the channel walls only serve to impose
937: a geometrical restriction on the lateral diffusion.\cite{ref121} The {\it flux cancellation effect\/} is characteristic of the pure metal regime, where the electrons move
938: ballistically from one wall to the other. This effect (which also plays a role in
939: the superconductivity of thin films in a parallel magnetic field\cite{ref122}) leads to a
940: further enhancement of the characteristic field scale $B_{\mathrm{c}}$. Flux cancellation
941: results from the fact that typically backscattered trajectories for $l\gg W$ self-intersect (cf.\ Fig.\ \ref{fig17}) and are thus composed of smaller loops that are
942: traversed in opposite directions. Zero net flux is enclosed by closed trajectories involving only wall collisions (as indicated by the shaded areas in Fig.\
943: \ref{fig17}, which are equal but of opposite orientation), so impurity collisions are
944: required for phase relaxation in a magnetic field. This is in contrast to the
945: dirty metal regime considered before, where impurity scattering hinders
946: phase relaxation by reducing the diffusion constant. The resulting nonmonotonous dependence of phase relaxation on impurity scattering in the dirty
947: and pure metal regimes is illustrated in Fig.\ \ref{fig18}, where the calculated\cite{ref109}
948: magnetic relaxation time $\tau_{B}$ is plotted as a function of $l/W$ for a fixed ratio
949: $l_{\mathrm{m}}/W$.
950:
951: \begin{figure}
952: \centerline{\includegraphics[width=8cm]{figures/fig17}}
953: \caption{
954: Illustration of the flux cancellation effect for a closed trajectory of one electron in a narrow channel with diffuse boundary scattering. The trajectory is composed of two loops of equal area but opposite orientation, so it encloses zero flux. Taken from C. W. J. Beenakker and H. van Houten, Phys.\ Rev.\ B. {\bf 38}, 3232 (1988).
955: \label{fig17}
956: }
957: \end{figure}
958:
959: Before continuing our discussion of the flux cancellation effect, we give a
960: more precise definition of the phase relaxation time $\tau_{B}$. The effect of a
961: magnetic field on weak localization is accounted for formally by inserting the
962: term
963: \be
964: \langle {\mathrm e}^{\mathrm{i}\phi(t)}|\mathbf{r}(t)=\mathbf{r}(0)\rangle=e^{-t/\tau_{B}},\;\; W\ll l_{\mathrm{m}}, l_{\phi}, \label{eq6.11}
965: \ee
966: in the integrand of Eq.\ (\ref{eq6.2}). The term (\ref{eq6.11}) is the conditional average over all
967: closed trajectories having duration $t$ of the phase factor $\mathrm{e}^{\mathrm{i}\phi(t)}$, with $\phi$ the phase
968: difference defined in Eq.\ (\ref{eq6.5}). It can be shown\cite{ref109} that in the case of 1D weak
969: localization (and for $l_{\mathrm{m}}\gg W$), this term is given by an exponential decay
970: factor $\exp(- t/\tau_{B})$, which defines the magnetic relaxation time $\tau_{B}$. In this
971: regime the weak localization correction to the conductivity in the presence of
972: a magnetic field is then simply given by Eq.\ (\ref{eq6.4b}), after the substitution
973: \be
974: \tau_{\phi}^{-1}\rightarrow\tau_{\phi}^{-1}+\tau_{B}^{-1}. \label{eq6.12}
975: \ee
976: Explicitly, one obtains
977: \begin{widetext}
978: \be
979: \delta G_{\mathrm{loc}}(B)=-g_{\mathrm{s}}g_{ \mathrm{v}}\frac{e^{2}}{h}\frac{1}{L}\left(\left[\frac{1}{D\tau_{\phi}}+\frac{1}{D\tau_{B}}\right]^{-1/2}-\left[\frac{1}{D\tau_{\phi}}+\frac{1}{D\tau_{B}}+\frac{1}{D\tau}\right]^{-1/2}\right).
980: \label{eq6.13}
981: \ee
982: \end{widetext}
983:
984: One can see from Fig.\ \ref{fig18} and Table \ref{table2} that in the pure metal regime $l\gg W$,
985: a weak and strong field regime can be distinguished, depending on the ratio
986: $Wl/l_{\mathrm{m}}^{2}$. This ratio corresponds to the maximum phase change on a closed
987: trajectory of linear extension $l$ (measured along the channel). In the {\it weak\/} field
988: regime $(Wl/l_{\mathrm{m}}^{2}\ll 1)$ many impurity collisions are required before a closed
989: electron loop encloses sufficient flux for complete phase relaxation. In this
990: regime a further increase of the mean free path does not decrease the phase
991: relaxation time (in contrast to the dirty metal regime), because as a
992: consequence of the flux cancellation effect, faster diffusion along the channel
993: does not lead to a larger enclosed flux. On comparing the result in Table \ref{table2}
994: for $B_{\mathrm{c}}$ in the weak field regime with that for the dirty metal regime, one sees an
995: enhancement of the characteristic field by a factor $(l/W)^{1/2}$. The {\it strong\/} field
996: regime is reached if $Wl/l_{\mathrm{m}}^{2}\gg 1$, while still $l_{\mathrm{m}}\gg W$. Under these conditions, a
997: single impurity collision can lead to a closed trajectory that encloses sufficient
998: flux for phase relaxation. The phase relaxation rate $1/\tau_{B}$ is now proportional
999: to the impurity scattering rate $1/\tau$ and, thus, to $1/l$. The relaxation time $\tau_{B}$
1000: accordingly {\it increases\/} linearly with $l$ in this regime (see Fig.\ \ref{fig18}). For
1001: comparison with experiments in the pure metal regime, an analytic formula
1002: that interpolates between the weak and strong field regimes is useful. The
1003: following formula agrees well with numerical calculations:\cite{ref109}
1004: \be
1005: \tau_{B}=\tau_{B}^{\mathrm{weak}}+\tau_{B}^{\mathrm{strong}}. \label{eq6.14}
1006: \ee
1007: Here $\tau_{B}^{\mathrm{weak}}$ and $\tau_{B}^{\mathrm{strong}}$ are the expressions for $\tau_{B}$ in the asymptotic weak and strong field regimes, as given in Table \ref{table2}.
1008:
1009: \begin{figure}
1010: \centerline{\includegraphics[width=8cm]{figures/fig18}}
1011: \caption{
1012: Phase relaxation time $\tau_{B}$ in a channel with specular boundary scattering, as a function of the elastic mean free path $l$. The plot has been obtained by a numerical simulation of the phase relaxation process for a magnetic field such that $l_{\rm m}=10\,W$. The dashed lines are analytic formulas valid in the three asymptotic regimes (see Table \ref{table2}). Taken from C. W. J. Beenakker and H. van Houten, Phys.\ Rev.\ B {\bf 38}, 3232 (1988).
1013: \label{fig18}
1014: }
1015: \end{figure}
1016:
1017: So far, we have assumed that the transport is diffusive on time scales
1018: corresponding to $\tau_{\phi}$. This will be a good approximation only if $\tau_{\phi}\gg \tau$.
1019: Coherent diffusion breaks down if $\tau_{\phi}$ and $\tau$ are of comparable magnitude (as
1020: may be the case in high-mobility channels). The modification of weak
1021: localization as one enters the ballistic transport regime has been investigated
1022: by Wittmann and Schmid.\cite{ref130} It would be of interest to see to what extent the
1023: ad hoc short-time cutoff introduced in our Eq.\ (\ref{eq6.4}), which is responsible for
1024: the second bracketed term in Eq.\ (\ref{eq6.13}), is satisfactory.
1025:
1026: {\bf (b) Experiments in the pure metal regime.} Because of the high mobility
1027: required, the pure metal regime has been explored using GaAs-AlGaAs
1028: heterostructures only. The first experiments on weak localization in the pure
1029: metal regime were done by Thornton et al.,\cite{ref58} in a narrow split-gate device,
1030: although the data were analyzed in terms of the theory for the dirty metal
1031: regime. An experimental study specifically aimed at weak localization in the
1032: pure metal regime was reported in Refs.\ \onlinecite{ref26} and \onlinecite{ref27}. In a narrow channel
1033: defined by the shallow-mesa etch technique of Fig.\ \ref{fig4}c (with a conducting
1034: width estimated at $0.12\,\mu \mathrm{m}$), a pronounced negative magnetoresistance effect
1035: was found, similar to that observed by Thornton et al.\cite{ref58} A good agreement of
1036: the experimental results with the theory\cite{ref109} for weak localization in the pure
1037: metal regime was obtained (see Fig.\ \ref{fig19}), assuming specular boundary
1038: scattering (diffuse boundary scattering could not describe the data). The
1039: width deduced from the analysis was consistent with independent estimates
1040: from other magnetoresistance effects. Further measurements in this regime
1041: were reported by Chang et al.\cite{ref70,ref139} and, more recently, by Hiramoto et al.\cite{ref81}
1042: These experiments were also well described by the theory of Ref.\ \onlinecite{ref109}.
1043:
1044: \begin{figure}
1045: \centerline{\includegraphics[width=8cm]{figures/fig19}}
1046: \caption{
1047: Magnetoconductance due to 1D weak localization in the pure metal regime ($W = 120\,{\rm nm}$, $L = 350\,{\rm nm}$). The solid curves are one-parameter fits to Eq.\ (\ref{eq6.13}). Only the field range $l_{\rm m} > W$ is shown in accordance with the condition of coherent diffusion imposed by the theory. The phase coherence length $l_{\phi}$ obtained from the data at various temperatures is tabulated in the inset. Taken from H. van Houten et al., Surf.\ Sci.\ {\bf 196}, 144 (1988).
1048: \label{fig19}
1049: }
1050: \end{figure}
1051:
1052: \subsection{\label{sec7} Conductance fluctuations}
1053:
1054: Classically, sample-to-sample fluctuations in the conductance are negligible in the diffusive (or quasi-ballistic) transport regime. In a narrow-channel geometry, for example, the root-mean-square $\delta G_{\mathrm{class}}$ of the classical
1055: fluctuations in the conductance is smaller than the average conductance $\langle G\rangle$
1056: by a factor $(l/L)^{1/2}$, under the assumption that the channel can be subdivided
1057: into $L/l\gg 1$ {\it independently\/} fluctuating segments. As we have discussed in the
1058: previous section, however, quantum mechanical correlations persist over a
1059: phase coherence length $l_{\phi}$ that can be much larger than the elastic mean free
1060: path $l$. Quantum interference effects lead to significant sample-to-sample
1061: fluctuations in the conductance if the size of the sample is not very much
1062: larger than $l_{\phi}$. The Al'tshuler-Lee-Stone theory of {\it Universal Conductance Fluctuations}\cite{ref140,ref141} finds that $\delta G\approx e^{2}/h$ at $T=0$, when phase coherence is
1063: maintained over the entire sample. Since $\langle G\rangle\propto L^{-1}$, it follows that
1064: $\delta G/\langle G\rangle\propto L$ {\it increases\/} with increasing channel length; that is, there is a total
1065: absence of self-averaging.
1066:
1067: Experimentally, the large sample-to-sample conductance fluctuations
1068: predicted theoretically are difficult to study in a direct way, because of
1069: problems in the preparation of samples that differ in impurity configuration
1070: only (to allow an ensemble average). The most convenient way to study the
1071: effect is via the fluctuations in the conductance of a single sample as a
1072: function of magnetic field, because a small change in field has a similar effect
1073: on the interference pattern as a change in impurity configuration. Sections \ref{sec7c}
1074: and \ref{sec7d} deal with theoretical and experimental studies of magnetoconductance fluctuations in narrow 2DEG channels, mainly in the quasi-ballistic
1075: regime characteristic for semiconductor nanostructures. In Sections \ref{sec7a} and
1076: \ref{sec7b} we discuss the surprising universality of the conductance fluctuations at
1077: zero temperature and the finite-temperature modifications.
1078:
1079: \subsubsection{\label{sec7a} Zero-temperature conductance fluctuations}
1080:
1081: The most surprising feature of the conductance fluctuations is that their
1082: magnitude at zero temperature is of order $e^{2}/h$, regardless of the size of the
1083: sample and the degree of disorder,\cite{ref140,ref141} provided at least that $L\gg l$, so that
1084: transport through the sample is diffusive (or possibly quasi-ballistic). Lee and
1085: Stone\cite{ref141} coined the term {\it Universal Conductance Fluctuations\/} (UCF) for this
1086: effect. In this subsection we give a simplified explanation of this universality
1087: due to Lee.\cite{ref142}
1088:
1089: \begin{figure}
1090: \centerline{\includegraphics[width=8cm]{figures/fig20}}
1091: \caption{
1092: Idealized conductor connecting source (S) and drain (D) reservoirs and containing a disordered region (crosshatched). The incoming quantum channels (or transverse waveguide modes) are labeled by $\alpha$, the transmitted and back scattered channels by $\beta$.
1093: \label{fig20}
1094: }
1095: \end{figure}
1096:
1097: Consider first the classical Drude conductance (\ref{eq4.8}) for a singe spin
1098: direction (and a single valley):
1099: \be
1100: G=\frac{W}{L}\frac{e^{2}}{h}\frac{k_{\mathrm{F}}l}{2}=\frac{e^{2}}{h}\frac{\pi l}{2L} N,\;\;N \equiv\frac{k_{\mathrm{F}}W}{\pi}. \label{eq7.1}
1101: \ee
1102: The number $N$ equals the number of transverse modes, or one-dimensional
1103: subbands, that are occupied at the Fermi energy in a conductor of width $W$.
1104: We have written the conductance in this way to make contact with the
1105: Landauer approach\cite{ref4} to conduction, which relates the conductance to the
1106: transmission probabilities of modes at the Fermi energy. (A detailed
1107: discussion of this approach is given the context of quantum ballistic transport
1108: in Section \ref{sec12b}). The picture to have in mind is shown in Fig.\ \ref{fig20}. Current is
1109: passed from a source reservoir $\mathrm{S}$ to a drain reservoir $\mathrm{D}$, through a disordered
1110: region (hatched) in which only elastic scattering takes place. The two
1111: reservoirs are in thermal equilibrium and are assumed to be fully effective in
1112: randomizing the phase via inelastic scattering, so there is no phase coherence
1113: between the $N$ modes incident on the disordered region. The modes in this
1114: context are called {\it quantum channels}. If $L\gg l$, each channel has on average the
1115: same transmission probability, given by $\pi l/2L$ according to Eqs.\ (\ref{eq4.21}) and
1116: (\ref{eq7.1}). We are interested in the fluctuations around this average. The resulting
1117: fluctuations in $G$ then follow from the multichannel Landauer
1118: formula\cite{ref1,ref143,ref144}
1119: \be
1120: G= \frac{e^{2}}{h}\sum_{\alpha,\beta=1}^{N}|t_{\alpha\beta}|^{2}, \label{eq7.2}
1121: \ee
1122: where $t_{\beta\alpha}$ denotes the quantum mechanical transmission probability
1123: amplitude from the incident channel $\alpha$ to the outgoing channel $\beta$ (cf.\ Fig.\ \ref{fig20}).
1124: The ensemble averaged transmission probability $\langle|t_{\alpha\beta}|^{2}\rangle$ does not depend on
1125: $\alpha$ or $\beta$, so the correspondence between Eqs.\ (\ref{eq7.1}) and (\ref{eq7.2}) requires
1126: \be
1127: \langle|t_{\alpha\beta}|^{2}\rangle=\pi l/2NL. \label{eq7.3}
1128: \ee
1129: The magnitude of the conductance fluctuations is characterized by its
1130: variance $\mathrm{Var}\,(G)\equiv\langle(G-\langle G\rangle)^{2}\rangle$. As discussed by Lee, a difficulty arises in a
1131: direct evaluation of $\mathrm{Var}\,(G)$ from Eq.\ (\ref{eq7.2}), because the correlation in the
1132: transmission probabilities $|t_{\alpha\beta}|^{2}$ for different pairs of incident and outgoing
1133: channels $\alpha, \beta$ may not be neglected.\cite{ref142} The reason is presumably that
1134: transmission through the disordered region involves a large number of
1135: impurity collisions, so a sequence of scattering events will in general be
1136: shared by different channels. On the same grounds, it is reasonable to assume
1137: that the reflection probabilities $|r_{\alpha\beta}|^{2}$ for different pairs $\alpha\beta$ of incident and
1138: reflected channels are uncorrelated, since the reflection back into the source
1139: reservoir would seem to be dominated by only a few scattering events.\cite{ref142}
1140: (The formal diagrammatic analysis of Refs.\ \onlinecite{ref140} and \onlinecite{ref141} is required here for a
1141: convincing argument.) The reflection and transmission probabilities are
1142: related by current conservation
1143: \be
1144: \sum_{\alpha,\beta=1}^{N}|t_{\alpha\beta}|^{2}=N-\sum_{\alpha,\beta=1}^{N}|r_{\alpha\beta}|^{2}. \label{eq7.4}
1145: \ee
1146: so the variance of the conductance equals
1147: \begin{eqnarray}
1148: \mathrm{Var}\,(G)&=&\left(\frac{e^{2}}{h}\right)^{2}\mathrm{Var}\,\left(\sum|r_{\alpha\beta}|^{2}\right)\nonumber\\
1149: &=&\left(\frac{e^{2}}{h}\right)^{2}N^{2}\mathrm{Var}\,(|r_{\alpha\beta}|^{2}), \label{eq7.5}
1150: \end{eqnarray}
1151: assuming uncorrelated reflection probabilities. A large number $M$ of scattering sequences through the disordered region contributes with amplitude
1152: $A(i)$ ($i=1,2, \ldots, M$) to the reflection probability amplitude $r_{\alpha\beta}$. (The different scattering sequences can be seen as independent Feynman paths
1153: in a path integral formulation of the problem.\cite{ref142}) To calculate
1154: $\mathrm{Var}\,(|r_{\alpha\beta}|^{2})=\langle|r_{\alpha\beta}|^{4}\rangle-\langle|r_{\alpha\beta}|^{2}\rangle^{2}$, one may then write (neglecting correlations
1155: in $A(i)$ for different $i$)
1156: \begin{eqnarray}
1157: \langle|r_{\alpha\beta}|^{4}\rangle&=&\sum_{i,j,k,l=1}^{M}\langle A^{*}(i)A(j)A^{*}(k)A(l)\rangle\nonumber\\
1158: &=& \sum_{i,j,k,l=1}^{M}\left\{\langle|A(i)|^{2}\rangle\langle|A(k)|^{2}\rangle\delta_{ij}\delta_{kl}\right.\nonumber\\
1159: &&\left.\mbox{}+\langle|A(i)|^{2}\rangle\langle|A(j)|^{2}\rangle\delta_{il}\delta_{jk}\right\}\nonumber\\
1160: &=&2\langle|r_{\alpha\beta}|^{2}\rangle^{2}, \label{eq7.6}
1161: \end{eqnarray}
1162: where we have neglected terms smaller by a factor $1/M$ (assuming $M\gg 1$).
1163: One thus finds that the variance of the reflection probability is equal to the
1164: square of its average:
1165: \be
1166: \mathrm{Var}\,(|r_{\alpha\beta}|^{2})=\langle|r_{\alpha\beta}|^{2}\rangle^{2}. \label{eq7.7}
1167: \ee
1168: The average reflection probability $\langle|r_{\alpha\beta}|^{2}\rangle$ does not depend on $\alpha$ and $\beta$. Thus,
1169: from Eqs.\ (\ref{eq7.3}) and (\ref{eq7.4}) it follows that
1170: \be
1171: \langle|r_{\alpha\beta}|^{2}\rangle=N^{-1}(1- {\rm order}(l/L)). \label{eq7.8}
1172: \ee
1173: Combining Eqs.\ (\ref{eq7.5}), (\ref{eq7.7}), and (\ref{eq7.8}), one obtains the result that the zero-temperature conductance has a variance $(e^{2}/h)^{2}$, independent of $l$ or $L$ (in the
1174: diffusive limit $l\ll L$). We have discussed this argument of Lee in some detail,
1175: because no other simple argument known to us gives physical insight in this
1176: remarkable result.
1177:
1178: The numerical prefactors follow from the diagrammatic analysis.\cite{ref140,ref141,ref145,ref146} The result of Lee and Stone\cite{ref141} for the root-mean-square
1179: magnitude of the conductance fluctuations at $T=0$ can be written in the
1180: form
1181: \be
1182: \delta G\equiv[\mathrm{Var}\,(G)]^{1/2}=\frac{g_{\mathrm{s}}g_{\mathrm{v}}}{2}\beta^{-1/2}C\frac{e^{2}}{h}. \label{eq7.9}
1183: \ee
1184: Here $C$ is a constant that depends on the shape of the sample. Typically, $C$ is
1185: of order unity; for example, $C\approx 0.73$ in a narrow channel with $L\gg W$.
1186: (However, in the opposite limit $W\gg L$ of a wide and short channel, $C$ is of
1187: order $(W/L)^{1/2}$.) The parameter $\beta=1$ in a zero magnetic field when time-reversal symmetry holds; $\beta=2$ when time-reversal symmetry is broken by a
1188: magnetic field. The factor $g_{\mathrm{s}}g_{\mathrm{v}}$ assumes complete spin and valley degeneracy.
1189: If the magnetic field is sufficiently strong that the two spin directions give
1190: statistically independent contributions to the conductance, then the variances
1191: add so that the factor $g_{\mathrm{s}}$ in $\delta G$ is to be replaced by a factor $g_{\mathrm{s}}^{1/2}$. We will return
1192: to this point in Section \ref{sec7d}.
1193:
1194: \subsubsection{\label{sec7b} Nonzero temperatures}
1195:
1196: At nonzero temperatures, the magnitude of the conductance fluctuations is
1197: reduced below $\delta G\approx e^{2}/h$. One reason is the effect of a finite phase coherence
1198: length $l_{\phi}\equiv(D\tau_{\phi})^{1/2}$; another is the effect of thermal averaging, as expressed by
1199: the thermal length $l_{\mathrm{T}}\equiv(hD/k_{B}T)^{1/2}$. The effect of a finite temperature,
1200: contained in $l_{\phi}$ and $l_{\mathrm{T}}$, is to partially restore self-averaging, albeit that the
1201: suppression of the fluctuation with sample size is much weaker than would be
1202: the case classically. The theory has been presented clearly and in detail by
1203: Lee, Stone, and Fukuyama.\cite{ref145} We limit the present discussion to the 1D
1204: regime $W\ll l_{\phi}\ll L$, characteristic for narrow 2DEG channels.
1205:
1206: The effects of thermal averaging may be neglected if $l_{\phi}\ll l_{\mathrm{T}}$ (see below).
1207: The channel may then be thought to be subdivided in uncorrelated segments
1208: of length $l_{\phi}$. The conductance fluctuation of each segment individually will be
1209: of order $e^{2}/h$, as it is at zero temperature. The root-mean-square conductance
1210: fluctuation of the entire channel is easily estimated. The segments are in
1211: series, so their resistances add according to Ohm's law. We denote the
1212: resistance of a channel segment of length $l_{\phi}$ by $R_{1}$. The variance of $R_{1}$ is
1213: $\mathrm{Var}\,(R_{1})\approx\langle R_{1}\rangle^{4}\mathrm{Var}\,(R_{1}^{-1})\approx\langle R_{1}\rangle^{4}(e^{2}/h)^{2}$. The average resistance of the
1214: whole channel $\langle R\rangle=(L/l_{\phi})\langle R_{1}\rangle$ increases linearly with the number $L/l_{\phi}$ of
1215: uncorrelated channel segments, just as its variance
1216: \[
1217: \mathrm{Var}\,(R)=(L/l_{\phi})\mathrm{Var}\,(R_{1})\approx(L/l_{\phi})\langle R_{1}\rangle^{4}(e^{2}/h)^{2}.
1218: \]
1219: (The root-mean-square resistance fluctuation thus grows as $(L/l_{\phi})^{1/2}$, the square root of the number of channel
1220: segments in series.) Expressed in terms of a conductance, one thus has
1221: $\mathrm{Var}\,(G)\approx\langle R\rangle^{-4}\mathrm{Var}\,(R)\approx(l_{\phi}/L)^{3}(e^{2}/h)^{2}$, or
1222: \be
1223: \delta G= {\rm constant}
1224: \times\frac{e^{2}}{h}\left(\frac{l_{\phi}}{L}\right)^{3/2},\;\; {\rm if}\;\; l_{\phi}\ll l_{\mathrm{T}}. \label{eq7.10}
1225: \ee
1226: The constant prefactor is given in Table \ref{table3}.
1227:
1228: We now turn to the second effect of the finite temperature, which is the
1229: smearing of the fluctuations by the energy average within an interval of order
1230: $k_{\mathrm{B}}T$ around the Fermi energy $E_{\mathrm{F}}$. Note that we did not have to consider this
1231: thermal averaging in the context of the weak localization effect, since that is
1232: a systematic, rather than a fluctuating, property of the sample. Two interfering Feynman paths, traversed with an energy difference $\delta E$, have to be considered as uncorrelated after a time $t_{1}$, if the acquired phase difference $t_{1}\delta E/\hbar$
1233: is of order unity. In this time the electrons diffuse a distance
1234: $L_{1}=(Dt_{1})^{1/2}\sim(\hbar D/\delta E)^{1/2}$. One can now define a correlation energy $E_{\mathrm{c}}(L_{1})$,
1235: as the energy difference for which the phase difference following diffusion over
1236: a distance $L_{1}$ is unity:
1237: \be
1238: E_{\mathrm{c}}(L_{1})\equiv \hbar D/L_{1}^{2}. \label{eq7.11}
1239: \ee
1240: The thermal length $l_{\mathrm{T}}$ is defined such that $E_{\mathrm{c}}(l_{\mathrm{T}})\equiv k_{\mathrm{B}}T$, which implies
1241: \be
1242: l_{\mathrm{T}}\equiv(\hbar D/k_{B}T)^{1/2}. \label{eq7.12}
1243: \ee
1244:
1245: \begin{table*}
1246: \caption{Asymptotic expressions for the root-mean-square conductance fluctuations in a narrow channel.
1247: \label{table3}
1248: }
1249: \begin{ruledtabular}
1250: \begin{tabular}{cccc}
1251: &$T=0$\footnotemark[1]&\multicolumn{2}{c}{$T>0$\footnotemark[1]}\\
1252: &$l_{\rm T},l_{\phi}\gg L$&$l_{\phi}\ll L,l_{\rm T}$&$l_{\rm T}\ll l_{\phi}\ll L$\\
1253: \hline
1254: &&&\\
1255: $\displaystyle\delta G\times\frac{2}{g_{\rm s}g_{\rm v}}\beta^{1/2}$&$\displaystyle C\frac{e^{2}}{h}$&$\displaystyle C\frac{e^{2}}{h}\left(\frac{l_{\phi}}{L}\right)^{3/2}$&$\displaystyle C\frac{e^{2}}{h}\frac{l_{\rm T}l_{\phi}^{1/2}}{L^{3/2}}$\\
1256: &&&\\
1257: $C$&0.73&$\displaystyle\sqrt{12}$&$\displaystyle\left(\frac{8\pi}{3}\right)^{1/2}$
1258: \end{tabular}
1259: \end{ruledtabular}
1260: \footnotetext[1]{
1261: The results assume a narrow channel ($W\ll L$), with a 2D density of states ($W\gg\lambda_{\rm F}$), which is in the 1D limit for the conductance fluctuations ($W\ll l_{\phi}$). The expressions for $\delta G$ are from Refs.\ \onlinecite{ref140,ref141,ref145}, and \onlinecite{ref146}. The numerical prefactor $C$ for $T=0$ is from Ref.\ \onlinecite{ref141}, for $T>0$ from Ref.\ \onlinecite{ref147}. If time-reversal symmetry applies, then $\beta=1$, but in the presence of a magnetic field strong enough to suppress the cooperon contributions then $\beta=2$. If the spin degeneracy is lifted, $g_{\rm s}$ is to be replaced by $g_{\rm s}^{1/2}$.}
1262: \end{table*}
1263:
1264: (Note that this definition of $l_{\mathrm{T}}$ differs by a factor of $(2\pi)^{1/2}$ from that in Ref.\
1265: \onlinecite{ref145}.) The thermal smearing of the conductance fluctuations is of importance
1266: only if phase coherence extends beyond a length scale $l_{\mathrm{T}}$ (i.e., if $l_{\phi}\gg l_{\mathrm{T}}$). In this
1267: case the total energy interval $k_{\mathrm{B}}T$ around the Fermi level that is available for
1268: transport is divided into subintervals of width $E_{\mathrm{c}}(l_{\phi})=\hbar/\tau_{\phi}$ in which phase
1269: coherence is maintained. There is a number $N\approx k_{\mathrm{B}}T/E_{\mathrm{c}}(l_{\phi})$ of such subintervals, which we assume to be uncorrelated. The root-mean-square variation
1270: $\delta G$ of the conductance is then reduced by a factor $N^{-1/2}\approx l_{\mathrm{T}}/l_{\phi}$ with respect
1271: to the result (\ref{eq7.10}) in the absence of energy averaging. (A word of caution: as
1272: discussed in Ref.\ \onlinecite{ref145}, the assumption of $N$ uncorrelated energy intervals is
1273: valid in the 1D case $W\ll l_{\phi}$ considered here, but not in higher dimensions.)
1274: From the foregoing argument it follows that
1275: \be
1276: \delta G=\mathrm{constant}\times\frac{e^{2}}{h}\frac{l_{\mathrm{T}}l_{\phi}^{1/2}}{L^{3/2}}\;\;{\rm if}\;\;l_{\phi}\gg l_{\mathrm{T}}. \label{eq7.13}
1277: \ee
1278:
1279: The asymptotic expressions (\ref{eq7.10}) and (\ref{eq7.13}) were derived by Lee, Stone,
1280: and Fukuyama\cite{ref145} and by Al'tshuler and Khmel'nitskii\cite{ref146} up to unspecified
1281: constant prefactors. These constants have been evaluated in Ref.\ \onlinecite{ref147}, and are
1282: given in Table \ref{table3}. In that paper we also gave an interpolation formula
1283: \begin{eqnarray}
1284: \delta G&=&\frac{g_{\mathrm{s}}g_{\mathrm{v}}}{2}\beta^{-1/2}\sqrt{12}\frac{e^{2}}{h}\left(\frac{l_{\phi}}{L}\right)^{3/2}\nonumber\\
1285: &&\mbox{}\times\left[1+\frac{9}{2\pi}\left(\frac{l_{\phi}}{l_{\mathrm{T}}}\right)^{2}\right]^{-1/2}, \label{eq7.14}
1286: \end{eqnarray}
1287: with $\beta$ defined in the previous subsection. This formula is valid (within 10\%
1288: accuracy) also in the intermediate regime when $l_{\phi}\approx l_{\mathrm{T}}$, and is useful for
1289: comparison with experiments, in which generally $l_{\phi}$ and $l_{\mathrm{T}}$ are not well
1290: separated (cf.\ Table \ref{table1}).
1291:
1292: \subsubsection{\label{sec7c} Magnetoconductance fluctuations}
1293:
1294: Experimentally, one generally studies the conductance fluctuations resulting from a change in Fermi energy $E_{\mathrm{F}}$ or magnetic field $B$ rather than from a
1295: change in impurity configuration. A comparison with the theoretical ensemble average becomes possible if one assumes that, insofar as the
1296: conductance fluctuations are concerned, a sufficiently large change in $E_{\mathrm{F}}$ or $B$
1297: is equivalent to a complete change in impurity configuration (this ``ergodic
1298: hypothesis'' has been proven in Ref.\ \onlinecite{ref148}). The reason for this equivalence is
1299: that, on one hand, the conductance at $E_{\mathrm{F}}+\Delta E_{\mathrm{F}}$ and $B+\Delta B$ is uncorrelated
1300: with that at $E_{\mathrm{F}}$ and $B$, provided either $\Delta E_{\mathrm{F}}$ or $\Delta B$ is larger than a correlation
1301: energy $\Delta E_{\mathrm{c}}$ or correlation field $\Delta B_{\mathrm{c}}$. On the other hand, the correlation
1302: energies and fields are in general sufficiently small that the statistical
1303: properties of the ensemble are not modified by the increment in $E_{\mathrm{F}}$ or $B$, so
1304: one is essentially studying a new member of the same ensemble, without
1305: changing the sample.
1306:
1307: This subsection deals with the calculation of the correlation field $\Delta B_{\mathrm{c}}$. (The
1308: correlation energy is discussed in Ref.\ \onlinecite{ref145} and will not be considered here.)
1309: The magnetoconductance correlation function is defined as
1310: \be
1311: F(\Delta B)\equiv\langle[\delta G(B)-\langle G(B)\rangle][G(B+\Delta B)-\langle G(B+\Delta B)\rangle]\rangle,
1312: \label{eq7.15}
1313: \ee
1314: where the angle brackets $\langle\cdots\rangle$ denote, as before, an ensemble average. The
1315: root-mean-square variation $\delta G$ considered in the previous two subsections is
1316: equal to $F(0)^{1/2}$. The correlation field $\Delta B_{\mathrm{c}}$ is defined as the half-width at half-height $F(\Delta B_{\mathrm{c}})\equiv F(0)/2$. The correlation function $F(\Delta B)$ is determined
1317: theoretically\cite{ref141,ref145,ref146} by temporal and spatial integrals of two propagators:
1318: the {\it diffuson\/} $P_{\mathrm{d}}(\mathbf{r}, \mathbf{r}^{\prime}, t)$ and the {\it cooperon\/} $P_{\mathrm{c}}(\mathbf{r}, \mathbf{r}^{\prime}, t)$. As discussed by Chakravarty and Schmid,\cite{ref126} these propagators consist of the product of three terms:
1319: (1) the classical probability to diffuse from $\mathbf{r}$ to $\mathbf{r}^{\prime}$ in a time $t$ (independent of $B$
1320: in the field range $\omega_{\mathrm{c}}\tau\ll 1$ of interest here); (2) the relaxation factor $\exp(- t/\tau_{\phi}$),
1321: which describes the loss of phase coherence due to inelastic scattering events;
1322: (3) the average phase factor $\langle\exp(i\Delta\phi)\rangle$, which describes the loss of phase
1323: coherence due to the magnetic field. The average $\langle\cdots\rangle$ is taken over all
1324: classical trajectories that diffuse from $\mathbf{r}$ to $\mathbf{r}^{\prime}$ in a time $t$. The phase difference
1325: $\Delta\phi$ is different for a diffuson or cooperon:
1326: \begin{subequations}
1327: \label{eq7.16}
1328: \begin{eqnarray}
1329: \Delta\phi(\mathrm{diffuson})&=&\frac{e}{\hbar}\int_{\mathbf{r}}^{\mathbf{r}^{\prime}}\Delta \mathbf{A}\cdot d\mathbf{l}, \label{eq7.16a}\\
1330: \Delta\phi(\mathrm{cooperon})&=&\frac{e}{\hbar}\int_{\mathbf{r}}^{\mathbf{r}^{\prime}}(2\mathbf{A}+\Delta \mathbf{A})\cdot d\mathbf{l}, \label{eq7.16b}
1331: \end{eqnarray}
1332: \end{subequations}
1333: where the line integral is along a classical trajectory. The vector potential $\mathbf{A}$
1334: corresponds to the magnetic field $\mathbf{B}=\nabla\times \mathbf{A}$, and the vector potential
1335: increment $\Delta \mathbf{A}$ corresponds to the field increment $\Delta B$ in the correlation
1336: function $F(\Delta B)$ (according to $\Delta \mathbf{B}=\nabla\times\Delta \mathbf{A}$). An explanation of the different
1337: magnetic field dependencies of the diffuson and cooperon in terms of
1338: Feynman paths is given shortly.
1339:
1340: In Ref.\ \onlinecite{ref109} we have proven that in a narrow channel $(W\ll l_{\phi})$ the average
1341: phase factor $\langle\exp(i\Delta\phi)\rangle$ does not depend on initial and final coordinates $\mathbf{r}$
1342: and $\mathbf{r}^{\prime}$, provided that one works in the Landau gauge and that $t\gg \tau$. This is a
1343: very useful property, since it allows one to transpose the results for
1344: $\langle \exp(i\Delta\phi)\rangle$ obtained for $\mathbf{r}=\mathbf{r}^{\prime}$ in the context of weak localization to the
1345: present problem of the conductance fluctuations, where $\mathbf{r}$ can be different
1346: from $\mathbf{r}^{\prime}$. We recall that for weak localization the phase difference $\Delta\phi$ is that of
1347: the cooperon, with the vector potential increment $\Delta \mathbf{A}=0$ [cf.\ Eq.\ (\ref{eq6.5})]. The
1348: average phase factor then decays exponentially as $\langle \exp(i\Delta\phi)\rangle=\exp(-t/\tau_{B}$)
1349: [cf.\ Eq.\ (\ref{eq6.11})], with the relaxation time $\tau_{B}$ given as a function of magnetic
1350: field $B$ in Table \ref{table2}. We conclude that the same exponential decay holds for the
1351: average cooperon and diffuson phase factors after substitution of
1352: $B\rightarrow B+\Delta B/2$ and $B\rightarrow\Delta B/2$, respectively, in the expressions for $\tau_{B}$:
1353: \begin{subequations}
1354: \label{eq7.17}
1355: \begin{eqnarray}
1356: \langle e^{i\Delta\phi}\rangle(\mathrm{diffuson})&=& \exp(-t/\tau_{\Delta B/2}), \label{eq7.17a}\\
1357: \langle e^{i\Delta\phi}\rangle(\mathrm{cooperon})&=&\exp(-t/\tau_{B+\Delta B/2}). \label{eq7.17b}
1358: \end{eqnarray}
1359: \end{subequations}
1360:
1361: The cooperon is suppressed when $\tau_{B+\Delta B/2}\lesssim\tau_{\phi}$, which occurs on the same
1362: field scale as the suppression of weak localization (determined by $\tau_{B}\lesssim\tau_{\phi}$).
1363: The suppression of the cooperon can be seen as a consequence of the
1364: breaking of the time-reversal invariance by the magnetic field, similar to the
1365: suppression of weak localization. In a zero field the cooperons and the
1366: diffusons contribute equally to the variance of the conductance; therefore,
1367: when the cooperon is suppressed, $\mathrm{Var}\,(G)$ is reduced by a factor of 2. (The
1368: parameter $\beta$ in Table \ref{table3} thus changes from 1 to 2 when $B$ increases beyond
1369: $B_{\mathrm{c}}$.) In general, the magnetoconductance fluctuations are studied for $B>B_{\mathrm{c}}$
1370: (i.e., for fields beyond the weak localization peak). Then only the diffuson
1371: contributes to the conductance fluctuations, since the relaxation time of the
1372: diffuson is determined by the field {\it increment\/} $\Delta B$ in the correlation function
1373: $F(\Delta B)$, not by the magnetic field itself. This is the critical difference with weak
1374: localization: The conductance fluctuations are {\it not suppressed\/} by a weak
1375: magnetic field. The different behavior of cooperons and diffusons can
1376: be understood in terms of Feynman paths. The correlation function $F(\Delta B)$
1377: contains the product of four Feynman path amplitudes $A(i, B)$, $A^{*}(j, B)$,
1378: $A(k, B+\Delta B)$, and $A^{*}(l, B+\Delta B)$ along various paths $i, j, k, l$ from $\mathbf{r}$ to $\mathbf{r}^{\prime}$.
1379: Consider the diffuson term for which $i=l$ and $j=k$. The phase of this term
1380: $A(i, B)A^{*}(j, B)A(j, B+\Delta B)A^{*}(i, B+\Delta B)$ is
1381: \be
1382: - \frac{e}{\hbar}\oint \mathbf{A}\cdot d\mathbf{l}+\frac{e}{\hbar}\oint(\mathbf{A}+\Delta \mathbf{A})\cdot d\mathbf{l}=\frac{e}{\hbar}\Delta\Phi. \label{eq7.18}
1383: \ee
1384: where the line integral is taken along the closed loop formed by the two paths
1385: $i$ and $j$ (cf.\ Fig.\ \ref{fig21}a). The phase is thus given by the flux {\it increment\/} $\Delta\Phi\equiv S\Delta B$
1386: through this loop and does not contain the flux $\Phi\equiv SB$ itself. The fact that
1387: the magnetic relaxation time of the diffuson depends only on $\Delta B$ and not on $B$
1388: is a consequence of the cancellation contained in Eq.\ (\ref{eq7.18}). For the cooperon,
1389: the relevant phase is that of the product of Feynman path amplitudes
1390: $A_{-}(i, B)A_{-}^{*}(j, B)A_{+}(j, B+\Delta B)A_{+}^{*}(i, B+\Delta B)$, where the $-$ sign refers to a
1391: trajectory from $\mathbf{r}^{\prime}$ to $\mathbf{r}$ and the $+$ sign to a trajectory from $\mathbf{r}$ to $\mathbf{r}^{\prime}$ (see Fig.\ \ref{fig21}b).
1392: This phase is given by
1393: \be
1394: \frac{e}{\hbar}\oint \mathbf{A}\cdot d\mathbf{l}+\frac{e}{\hbar}\oint(\mathbf{A}+\Delta \mathbf{A})\cdot d\mathbf{l}=\frac{e}{\hbar}(2\Phi+\Delta\Phi). \label{eq7.19}
1395: \ee
1396: In contrast to the diffuson, the cooperon is sensitive to the flux $\Phi$ through the
1397: loop and can therefore be suppressed by a weak magnetic field.
1398:
1399: \begin{figure}
1400: \centerline{\includegraphics[width=8cm]{figures/fig21}}
1401: \caption{
1402: Illustration of the different flux sensitivity of the interference terms of diffuson type (a) and of cooperon type (b). Both contribute to the conductance fluctuations in a zero magnetic field, but the cooperons are suppressed by a weak magnetic field, as discussed in the text.
1403: \label{fig21}
1404: }
1405: \end{figure}
1406:
1407: In the following, we assume that $B>B_{\mathrm{c}}$ so that only the diffuson
1408: contributes to the magnetoconductance fluctuations. The combined effects of
1409: magnetic field and inelastic scattering lead to a relaxation rate
1410: \be
1411: \tau_{\mathrm{eff}}^{-1}=\tau_{\phi}^{-1}+\tau_{\Delta B/2}^{-1}, \label{eq7.20}
1412: \ee
1413: which describes the exponential decay of the average phase factor
1414: $\langle e^{i\Delta\phi}\rangle=\exp(- t/\tau_{\mathrm{eff}})$. Equation (\ref{eq7.20}) contains the whole effect of the
1415: magnetic field on the diffuson. Without having to do any diagrammatic
1416: analysis, we therefore conclude\cite{ref147} that the correlation function $F(\Delta B)$ can be
1417: obtained from the variance $F(0)\equiv \mathrm{Var}\,G=(\delta G)^{2}$ (given in Table \ref{table3}) by
1418: simply replacing $\tau_{\phi}$ by the effective relaxation time $\tau_{\mathrm{eff}}$ defined in Eq.\ (\ref{eq7.20}).
1419: The quantity $\tau_{\Delta B/2}$ corresponds to the magnetic relaxation time $\tau_{B}$ obtained
1420: for weak localization (see Table \ref{table2}) after substitution of $B\rightarrow\Delta B/2$. For easy
1421: reference, we give the results for the dirty and clean metal regimes
1422: explicitly:\cite{ref109,ref147}
1423: \begin{eqnarray}
1424: \tau_{\Delta B/2}&=&12\left(\frac{\hbar}{e\Delta B}\right)^{2}\frac{1}{DW^{2}},\;\;{\rm if}\;\;l\ll W, \label{eq7.21}\\
1425: \tau_{\Delta B/2}&=&4C_{1}\left(\frac{\hbar}{e\Delta B}\right)^{2}\frac{1}{v_{\mathrm{F}}W^{3}}+2C_{2}\left(\frac{\hbar}{e\Delta B}\right)\frac{l}{v_{\mathrm{F}}W^{2}},\nonumber\\
1426: &&\;\;\;\;\;\;\;{\rm if}\;\;l\gg W, \label{eq7.22}
1427: \end{eqnarray}
1428: where $C_{1}=9.5$ and $C_{2}=24/5$ for a channel with specular boundary
1429: scattering ($C_{1}=4\pi$ and $C_{2}=3$ for a channel with diffuse boundary scattering). These results are valid under the condition $W^{2}\Delta B\ll \hbar/e$, which follows
1430: from the requirement $\tau_{\mathrm{eff}}\gg \tau$ that the electronic motion on the effective phase
1431: coherence time scale $\tau_{\mathrm{eff}}$ be diffusive rather than ballistic, as well as from the
1432: requirement $(D\tau_{\mathrm{eff}})^{1/2}\gg W$ for one-dimensionality.
1433:
1434: With results (\ref{eq7.20})--(\ref{eq7.22}), the equation $F(\Delta B_{\mathrm{c}})=F(0)/2$, which defines the
1435: correlation field $\Delta B_{\mathrm{c}}$, reduces to an algebraic equation that can be solved
1436: straightforwardly. In the dirty metal regime one finds\cite{ref145}
1437: \be
1438: \Delta B_{\mathrm{c}}=2\pi C\frac{\hbar}{e}\frac{1}{Wl_{\phi}}, \label{eq7.23}
1439: \ee
1440: where the prefactor $C$ decreases from\cite{ref147} 0.95 for $l_{\phi}\gg l_{\mathrm{T}}$ to 0.42 for $l_{\phi}\ll l_{\mathrm{T}}$.
1441: Note the similarity with the result (\ref{eq6.9}) for weak localization. Just as in weak
1442: localization, one finds that the correlation field in the pure metal regime is
1443: significantly enhanced above Eq.\ (\ref{eq7.23}) due to the flux cancellation effect
1444: discussed in Section \ref{sec6c}. The enhancement factor increases from $(l/W)^{1/2}$ to
1445: $l/W$ as $l_{\phi}$ decreases from above to below the length $l^{3/2}W^{-1/2}$. The relevant
1446: expression is given in Ref.\ \onlinecite{ref147}. As an illustration, the dimensionless
1447: correlation flux $\Delta B_{\mathrm{c}}Wl_{\phi}e/h$ in the pure and dirty metal regimes is plotted as a
1448: function of $l_{\phi}/l$ in Fig.\ \ref{fig22} for $l_{\mathrm{T}}\ll l_{\phi}$.
1449:
1450: In the following discussion of the experimental situation in semiconductor
1451: nanostructures, it is important to keep in mind that the Al'tshuler-Lee-Stone theory of conductance fluctuations was formulated for an application to metals. This has justified the neglect of several possible complications,
1452: which may be important in a 2DEG. One of these is the classical curvature of
1453: the electron trajectories, which affects the conductance when $l_{\mathrm{cycl}} \lesssim\min(W, l)$.
1454: A related complication is the Landau level quantization, which in a narrow
1455: channel becomes important when $l_{\mathrm{m}}\lesssim W$. Furthermore, when $W\sim\lambda_{\mathrm{F}}$ the
1456: lateral confinement will at low fields induce the formation of 1D subbands.
1457: No quantization effects are taken into account in the theory of conductance
1458: fluctuations discussed before. Finally, the present theory is valid only in the
1459: regime of coherent diffusion $(\tau_{\phi}, \tau_{\mathrm{eff}}\gtrsim\tau)$. In high-mobility samples $\tau_{\phi}$ and $\tau$
1460: may be comparable, however, as discussed in Section \ref{sec7d}. It would be of
1461: interest to study the conductance fluctuations in this regime theoretically.
1462:
1463: \begin{figure}
1464: \centerline{\includegraphics[width=8cm]{figures/fig22}}
1465: \caption{
1466: Plot of the dimensionless correlation flux $\Phi_{\rm c}\equiv\Delta B_{\rm c}l_{\phi}We/h$ for the magnetoconductance fluctuations as a function of $l_{\phi}/l$ in the regime $l_{\rm T}\ll l_{\phi}$. The solid curve is for the case $l = 5\,W$; the dashed line is for $l\ll W$. Taken from C. W. J. Beenakker and H. van Houten, Phys.\ Rev.\ B {\bf 37}, 6544 (1988).
1467: \label{fig22}
1468: }
1469: \end{figure}
1470:
1471: In the following discussion of experimental studies of conductance
1472: fluctuations, we will have occasion to discuss briefly one further development.
1473: This is the modification of the theory\cite{ref149,ref150,ref151,ref152,ref153,ref154} to account for the differences
1474: between two- and four-terminal measurements of the conductance fluctuations, which becomes important when the voltage probes are separated by
1475: less than the phase coherence length.\cite{ref155,ref156}
1476:
1477: \subsubsection{\label{sec7d} Experiments}
1478:
1479: The experimental observation of conductance fluctuations in semiconductors has preceded the theoretical understanding of this phenomenon.
1480: Weak irregular conductance fluctuations in wide Si inversion layers were
1481: reported in 1965 by Howard and Fang.\cite{ref157} More pronounced fluctuations
1482: were found by Fowler et al.\ in narrow Si accumulation layers in the strongly
1483: localized regime.\cite{ref32} Kwasnick et al.\ made similar observations in narrow Si
1484: inversion layers in the metallic conduction regime.\cite{ref39} These fluctuations in the
1485: conductance as a function of gate voltage or magnetic field have been
1486: tentatively explained by various mechanisms.\cite{ref158} One of the explanations
1487: suggested is based on resonant tunneling,\cite{ref159} another on variable range
1488: hopping. At the 1984 conference on ``Electronic Properties of Two-Dimensional Systems'' Wheeler et al.\cite{ref161} and Skocpol et al.\cite{ref162} reported
1489: pronounced structure as a function of gate voltage in the low-temperature
1490: conductance of narrow Si inversion layers, observed in the course of their
1491: search for a quantum size effect.
1492:
1493: After the publication in 1985 of the Al'tshuler-Lee-Stone
1494: theory\cite{ref140,ref141,ref163} of universal conductance fluctuations, a consensus has
1495: rapidly developed that this theory properly accounts for the conductance
1496: fluctuations in the metallic regime, up to factor of two uncertainties in the
1497: quantitative description.\cite{ref46,ref144,ref164} Following this theoretical work, Licini et
1498: al.\cite{ref40} attributed the magnetoresistance oscillations that they observed in
1499: narrow Si inversion layers to quantum interference in a disordered conductor. Their low-temperature measurements, which we reproduce in Fig.\ \ref{fig23},
1500: show a large negative magnetoresistance peak due to weak localization at
1501: low magnetic fields, in addition to aperiodic fluctuations that persist to high
1502: fields. Such a clear weak localization peak is not found in shorter samples,
1503: where the conductance fluctuations are larger. The reason is that the
1504: magnitude of the conductance fluctuations $\Delta G$ is proportional to $(l_{\phi}/L)^{3/2}$
1505: [for $l_{\phi}\ll l_{\mathrm{T}}$, cf.\ Eq.\ (\ref{eq7.10})], while the weak localization conductance correction scales with $l_{\phi}/L$ [as discussed below Eq.\ (\ref{eq6.4})]. Weak localization thus
1506: predominates in long channels $(L\gg l_{\phi})$ where the fluctuations are relatively
1507: unimportant.
1508:
1509: \begin{figure}
1510: \centerline{\includegraphics[width=8cm]{figures/fig23}}
1511: \caption{
1512: Negative magnetoresistance and aperiodic magnetoresistance fluctuations in a narrow Si inversion layer channel for several values of the gate voltage $V_{\rm G}$. Note that the vertical offset and scale is different for each $V_{\rm G}$. Taken from J. C. Licini et al., Phys.\ Rev.\ Lett.\ {\bf 55}, 2987 (1985).
1513: \label{fig23}
1514: }
1515: \end{figure}
1516:
1517: The most extensive quantitative study of the universality of the conductance fluctuations in narrow Si inversion layers (over a wide range of
1518: channel widths, lengths, gate voltages, and temperatures) was made by
1519: Skocpol et al.\cite{ref45,ref46,ref156} In the following, we review some of these experimental
1520: results. We will not discuss the similarly extensive investigations by Webb et
1521: al.\cite{ref155,ref164,ref165} on small metallic samples, which have played an equally
1522: important role in the development of this subject. To analyze their experiments, Skocpol et al.\ estimated $l_{\phi}$ from weak localization experiments (with
1523: an estimated uncertainty of about a factor of 2). They then plotted the root-mean-square variation $\delta G$ of the conductance as a function of $L/l_{\phi}$, with $L$ the
1524: separation of the voltage probes in the channel. Their results are shown in
1525: Fig.\ \ref{fig24}. The points for $L>l_{\phi}$ convincingy exhibit for a large variety of data
1526: sets the $(L/l_{\phi})^{-3/2}$ scaling law predicted by the theory described in Section \ref{sec7c}
1527: (for $l_{\phi}<l_{\mathrm{T}}$, which is usually the case in Si inversion layers).
1528:
1529: For $L<l_{\phi}$ the experimental data of Fig.\ \ref{fig24} show a crossover to a $(L/l_{\phi})^{-2}$
1530: scaling law (dashed line), accompanied by an increase of the magnitude of the
1531: conductance fluctuations beyond the value $\delta G\approx e^{2}/h$ predicted by the
1532: Al'tshuler-Lee-Stone theory for a conductor of length $L<l_{\phi}$. A similar
1533: observation was made by Benoit et al.\cite{ref155} on metallic samples. The disagreement is explained\cite{ref155,ref156} by considering that the experimental geometry
1534: differs from that assumed in the theory discussed in Section \ref{sec7c}. Use is made
1535: of a long channel with voltage probes at different spacings. The experimental
1536: $L$ is the spacing of two voltage probes, and not the length of a channel
1537: connecting two phase-randomizing reservoirs, as envisaged theoretically. The
1538: difference is irrelevant if $L>l_{\phi}$. If the probe separation $L$ is less than the
1539: phase coherence length $l_{\phi}$, however, the measurement still probes a channel
1540: segment of length $l_{\phi}$ rather than $L$. In this sense the measurement is
1541: nonlocal.\cite{ref155,ref156} The key to the $L^{-2}$ dependence of $\delta G$ found experimentally is
1542: that the voltages on the probes fluctuate independently, implying that the
1543: {\it resistance\/} fluctuations $\delta R$ are independent of $L$ in this regime so that
1544: $\delta G\approx R^{-2}\delta R\propto L^{-2}$. This explanation is consistent with the anomalously
1545: small correlation field $B_{\mathrm{c}}$ found for $L<l_{\phi}$.\cite{ref46,ref156} One might have expected
1546: that the result $B_{\mathrm{c}}\approx h/eWl_{\phi}$ for $L>l_{\phi}$ should be replaced by the larger value
1547: $B_{\mathrm{c}}\approx h/eWL$ if $L$ is reduced below $l_{\phi}$. The smaller value found experimentally
1548: is due to the fact that the flux through parts of the channel adjacent to the
1549: segment between the voltage probes, as well as the probes themselves, has to
1550: be taken into account. These qualitative arguments\cite{ref155,ref156} are supported by
1551: detailed theoretical investigations.\cite{ref149,ref150,ref151,ref152,ref153,ref154} The important message of these
1552: theories and experiments is that the transport in a small conductor is phase
1553: coherent over large length scales and that phase randomization (due to
1554: inelastic collisions) occurs mainly as a result of the voltage probes. The
1555: Landauer-B\"{u}ttiker formalism\cite{ref4,ref5} (which we will discuss in Section \ref{sec12}) is
1556: naturally suited to study such problems theoretically. In that formalism,
1557: current and voltage contacts are modeled by phase-randomizing reservoirs
1558: attached to the conductor. We refer to a paper by B\"{u}ttiker\cite{ref149} for an
1559: instructive discussion of conductance fluctuations in a multiprobe conductor
1560: in terms of interfering Feynman paths.
1561:
1562: \begin{figure}
1563: \centerline{\includegraphics[width=8cm]{figures/fig24}}
1564: \caption{
1565: Root-mean-square amplitude $\delta g$ of the conductance fluctuations (in units of $e^2 /h$) as a function of the ratio of the distance between the voltage probes $L$ to the estimated phase coherence length $l_{\phi}$ for a set of Si inversion layer channels under widely varying experimental conditions. The solid and dashed lines demonstrate the $(L/l_{\phi})^{-3/2}$ and $(L/l_{\phi})^{-2}$ scaling of $\delta g$ in the regimes $L > l_{\phi}$ and $L < l_{\phi}$, respectively. Taken from W. J. Skocpol, Physica Scripta {\bf T19}, 95 (1987).
1566: \label{fig24}
1567: }
1568: \end{figure}
1569:
1570: Conductance fluctuations have also been observed in narrow-channel GaAs-AlGaAs
1571: heterostructures.\cite{ref166,ref167} These systems are well in the pure
1572: metal regime $(W<l)$, but unfortunately they are only marginally in the
1573: regime of coherent diffusion (characterized by $\tau_{\phi}\gg \tau$). This hampers a
1574: quantitative comparison with the theoretical results\cite{ref147} for the pure metal
1575: regime discussed in Section \ref{sec7c}. (A phenomenological treatment of conductance fluctuations in the case that $\tau_{\phi}\sim\tau$ is given in Refs.\ \onlinecite{ref168} and \onlinecite{ref169}.)
1576: The data of Ref.\ \onlinecite{ref167} are consistent with an enhancement of the correlation
1577: field due to the flux cancellation effect, but are not conclusive.\cite{ref147} We note
1578: that the flux cancellation effect can also explain the correlation field
1579: enhancement noticed in a computer simulation by Stone.\cite{ref163}
1580:
1581: In the analysis of the aforementioned experiments on magnetoconductance fluctuations, a twofold spin degeneracy has been assumed. The variance
1582: $(\delta G)^{2}$ is reduced by a factor of 2 if the spin degeneracy is lifted by a strong
1583: magnetic field $B>B_{\mathrm{c}2}$. The Zeeman energy $g\mu_{\mathrm{B}}B$ should be sufficiently large
1584: than the spin-up and spin-down electrons give statistically independent
1585: contributions to the conductance. The degeneracy factor $g_{\mathrm{s}}^{2}$ in $(\delta G)^{2}$ (introduced in Section \ref{sec7a}) should then be replaced by a factor $g_{\mathrm{s}}$, since the
1586: variances of statistically independent quantities add. Since $g_{\mathrm{s}}=2$, one
1587: obtains a factor-of-2 reduction in $(\delta G)^{2}$. Note that this reduction comes on
1588: top of the factor-of-2 reduction in $(\delta G)^{2}$ due to the breaking of time-reversal
1589: symmetry, which occurs at weak magnetic fields $B_{\mathrm{c}}$. Stone has calculated\cite{ref170}
1590: that the field $B_{\mathrm{c}2}$ in a narrow channel $(l_{\phi}\gg W)$ is given by the criterion of unit
1591: phase change $g\mu_{\mathrm{B}}B\tau_{\phi}/h$ in a coherence time, resulting in the estimate
1592: $B_{\mathrm{c}2}\approx h/g\mu_{\mathrm{B}}\tau_{\phi}$. Surprisingy, the thermal energy $k_{\mathrm{B}}T$ is irrelevant for $B_{\mathrm{c}2}$ in
1593: the 1D case $l_{\phi}\gg W$ (but not in higher dimensions\cite{ref170}).
1594:
1595: For the narrow-channel experiment of Ref.\ \onlinecite{ref167} just discussed, one finds
1596: (using the estimates $\tau_{\phi}\approx 7\,\mathrm{ps}$ and $g\approx 0.4$) a crossover field $B_{\mathrm{c}2}$ of about $2\, \mathrm{T}$,
1597: well above the field range used for the data analysis.\cite{ref147} Most importantly, no
1598: magnetoconductance fluctuations are observed if the magnetic field is applied
1599: {\it parallel\/} to the 2DEG (see Section \ref{sec9}), demonstrating that the Zeeman splitting
1600: has no effect on the conductance in this field regime. More recently, Debray et
1601: al.\cite{ref171} performed an experimental study of the reduction by a perpendicular
1602: magnetic field of the conductance fluctuations as a function of Fermi energy
1603: (varied by means of a gate). The estimated value of $\tau_{\phi}$ is larger than that of
1604: Ref.\ \onlinecite{ref167} by more than an order of magnitude. Consequently, a very small
1605: $B_{\mathrm{c}2}\approx 0.07\,\mathrm{T}$ is estimated in this experiment. The channel is relatively wide
1606: ($2\,\mu \mathrm{m}$ lithographic width), so the field $B_{\mathrm{c}}$ for time-reversal symmetry breaking
1607: is even smaller $(B_{\mathrm{c}}\approx 7\times 10^{-4}\,\mathrm{T})$. A total factor-of-4 reduction in $(\delta G)^{2}$ was
1608: found, as expected. The values of the observed crossover fields $B_{\mathrm{c}}$ and $B_{\mathrm{c}2}$
1609: also agree reasonably well with the theoretical prediction. Unfortunately, the
1610: magnetoconductance in a parallel magnetic field was not investigated by
1611: these authors, which would have provided a definitive test for the effect of
1612: Zeeman splitting on the conductance above $B_{\mathrm{c}2}$. We note that related
1613: experimental\cite{ref172,ref173} and theoretical\cite{ref174,ref175} work has been done on the
1614: reduction of {\it temporal\/} conductance fluctuations by a magnetic field.
1615:
1616: The Al'tshuler-Lee-Stone theory of conductance fluctuations ceases
1617: to be applicable when the dimensions of the sample approach the mean
1618: free path. In this ballistic regime observations of large aperiodic, as well
1619: as quasi-periodic, magnetoconductance fluctuations have been reported.\cite{ref68,ref69,ref139,ref168,ref176,ref177,ref178,ref179} Quantum interference effects in this regime are
1620: determined not by impurity scattering but by scattering off geometrical
1621: features of the device, as will be discussed in Section \ref{sec3}.
1622:
1623: \subsection{\label{sec8} Aharonov-Bohm effect}
1624:
1625: Magnetoconductance fluctuations in a channel geometry in the diffusive
1626: regime are {\it aperiodic}, since the interfering Feynman paths enclose a continuous range of magnetic flux values. A ring geometry, in contrast, encloses a
1627: well-defined flux $\Phi$ and thus imposes a fundamental periodicity
1628: \be
1629: G(\Phi)=G(\Phi+n(h/e)),\;\; n=1,2,3, \ldots, \label{eq8.1}
1630: \ee
1631: on the conductance as a function of perpendicular magnetic field $B$ (or flux
1632: $\Phi=BS$ through a ring of area $S$). Equation (\ref{eq8.1}) expresses the fact that a flux
1633: increment of an integer number of flux quanta changes by an integer multiple
1634: of $2\pi$ the phase difference between Feynman paths along the two arms of the
1635: ring. The periodicity (\ref{eq8.1}) would be an exact consequence of gauge invariance
1636: if the magnetic field were nonzero only in the interior of the ring, as in the
1637: original thought experiment of Aharonov and Bohm.\cite{ref180} In the present
1638: experiments, however, the magnetic field penetrates the arms of the ring as
1639: well as its interior so that deviations from Eq.\ (\ref{eq8.1}) can occur. Since in many
1640: situations such deviations are small, at least in a limited field range, one still
1641: refers to the magnetoconductance oscillations as an {\it Aharonov-Bohm effect}.
1642:
1643: The fundamental periodicity
1644: \be
1645: \Delta B=\frac{h}{e}\frac{1}{S} \label{eq8.2}
1646: \ee
1647: is caused by interference between trajectories that make one half-revolution
1648: around the ring, as in Fig.\ \ref{fig25}a. The first harmonic
1649: \be
1650: \Delta B=\frac{h}{2e}\frac{1}{S} \label{eq8.3}
1651: \ee
1652: results from interference after one revolution. A fundamental distinction
1653: between these two periodicities is that the phase of the $h/e$ oscillations (\ref{eq8.2}) is
1654: sample-specific, whereas the $h/2e$ oscillations (\ref{eq8.3}) contain a contribution
1655: from time-reversed trajectories (as in Fig.\ \ref{fig25}b) that has a minimum conductance at $B=0$, and thus has a sample-independent phase. Consequently,
1656: in a geometry with many rings in series (or in parallel) the $h/e$ oscillations
1657: average out, but the $h/2e$ oscillations remain. The $h/2e$ oscillations can be
1658: thought of as a periodic modulation of the weak localization effect due to
1659: coherent backscattering.
1660:
1661: \begin{figure}
1662: \centerline{\includegraphics[width=8cm]{figures/fig25}}
1663: \caption{
1664: Illustration of the Aharonov- Bohm effect in a ring geometry. Interfering trajectories responsible for the magnetoresistance oscillations with $h/e$ periodicity in the enclosed flux $\Phi$ are shown (a). (b) The pair of time-reversed trajectories lead to oscillations with $h/2e$ periodicity.
1665: \label{fig25}
1666: }
1667: \end{figure}
1668:
1669: The first observation of the Aharonov-Bohm effect in the solid state was
1670: made by Sharvin and Sharvin\cite{ref181} in a long metal cylinder. Since this is
1671: effectively a many-ring geometry, only the $h/2e$ oscillations were observed, in
1672: agreement with a theoretical prediction by Al'tshuler, Aronov, and
1673: Spivak,\cite{ref182} which motivated the experiment. (We refer to Ref.\ \onlinecite{ref125} for a simple
1674: estimate of the order of magnitude of the $h/2e$ oscillations in the dirty metal
1675: regime.) The effect was studied extensively by several groups.\cite{ref183,ref184,ref185} The $h/e$
1676: oscillations were first observed in single metal rings by Webb et al.\cite{ref186} and
1677: studied theoretically by several authors.\cite{ref1,ref144,ref187,ref188} The self-averaging of the
1678: $h/e$ oscillations has been demonstrated explicitly in experiments with a
1679: varying number of rings in series.\cite{ref189} Many more experiments have been
1680: performed on one-and two-dimensional arrays and networks, as reviewed in
1681: Refs.\ \onlinecite{ref190} and \onlinecite{ref191}.
1682:
1683: In this connection, we mention that the development of the theory of
1684: {\it aperiodic\/} conductance fluctuations (discussed in Section \ref{sec7}) has been much
1685: stimulated by their observation in metal rings by Webb et al.,\cite{ref165} in the course
1686: of their search for the Aharonov-Bohm effect. The reason that aperiodic
1687: fluctuations are observed in rings (in addition to periodic oscillations) is that
1688: the magnetic field penetrates the width of the arms of the ring and is not
1689: confined to its interior. By fabricating rings with a large ratio of radius $r$ to
1690: width $W$, researchers have proven it is possible to separate\cite{ref190} the magnetic
1691: field scales of the periodic and aperiodic oscillations (which are given by a
1692: field interval of order $h/er^{2}$ and $h/eWl_{\phi}$, respectively). The penetration of the
1693: magnetic field in the arms of the ring also leads to a broadening of the peak in
1694: the Fourier transform at the $e/h$ and $2e/h$ periodicities, associated with a
1695: distribution of enclosed flux. The width of the Fourier peak can be used as
1696: a rough estimate for the width of the arms of the ring. In addition, the
1697: nonzero field in the arms of the ring also leads to a damping of the amplitude
1698: of the ensemble-averaged $h/2e$ oscillations when the flux through the arms is
1699: sufficiently large to suppress weak localization.\cite{ref191}
1700:
1701: Two excellent reviews of the Aharonov-Bohm effect in metal rings and
1702: cylinders exist.\cite{ref190,ref191} In the following we discuss the experiments in semiconductor nanostructures in the weak-field regime $\omega_{\mathrm{c}}\tau<1$, where the effect of
1703: the Lorentz force on the trajectories can be neglected. The strong-field regime
1704: $\omega_{\mathrm{c}}\tau>1$ (which is not easily accessible in the usual polycrystalline metal
1705: rings) is only briefly mentioned; it is discussed more extensively in Section \ref{sec21}.
1706: To our knowledge, no observation of Aharonov-Bohm magnetoresistance
1707: oscillations in Si inversion layers has been reported. The first observation of
1708: the Aharonov-Bohm effect in a 2DEG ring was published by Timp et al.,\cite{ref69}
1709: who employed high-mobility GaAs-AlGaAs heterostructure material.
1710: Similar results were obtained independently by Ford et al.\cite{ref73} and Ishibashi et
1711: al.\cite{ref193} More detailed studies soon followed.\cite{ref74,ref139,ref176,ref194,ref195} A characteristic
1712: feature of these experiments is the large amplitude of the $h/e$ oscillations (up
1713: to 10\% of the average resistance), much higher than in metal rings (where the
1714: effect is at best\cite{ref192,ref196,ref197} of order 0.1\%). A similar difference in magnitude is
1715: found for the aperiodic magnetoresistance fluctuations in metals and semiconductor nanostructures. The reason is simply that the amplitude $\delta G$ of the
1716: periodic or aperiodic conductance oscillations has a maximum value of order
1717: $e^{2}/h$, so the maximum relative resistance oscillation $\delta R/R\approx R\delta G\approx Re^{2}/h$ is
1718: proportional to the average resistance $R$, which is typically much smaller in
1719: metal rings.
1720:
1721: In most studies only the $h/e$ fundamental periodicity is observed, although
1722: Ford et al.\cite{ref73,ref74} found a weak $h/2e$ harmonic in the Fourier transform of the
1723: magnetoresistance data of a very narrow ring. It is not quite clear whether
1724: this harmonic is due to the Al'tshuler-Aronov-Spivak mechanism involving
1725: the constructive interference of two time-reversed trajectories\cite{ref182} or to the
1726: random interference of two non-time-reversed Feynman paths winding
1727: around the entire ring.\cite{ref1,ref144,ref187} The relative weakness of the $h/2e$ effect in
1728: single 2DEG rings is also typical for most experiments on single metal rings
1729: (although the opposite was found to be true in the case of aluminum rings by
1730: Chandrasekhar et al.,\cite{ref197} for reasons which are not understood). This is in
1731: contrast to the case of arrays or cylinders, where, as we mentioned, the $h/2e$
1732: oscillations are predominant the $h/e$ effect being ``ensemble-averaged'' to
1733: zero because of its sample-specific phase. In view of the fact that the
1734: experiments on 2DEG rings explore the borderline between diffusive and
1735: ballistic transport, they are rather difficult to analyze quantitatively. A
1736: theoretical study of the Aharonov-Bohm effect in the purely ballistic
1737: transport regime was performed by Datta and Bandyopadhyay,\cite{ref198} in
1738: relation to an experimental observation of the effect in a double-quantum-well device.\cite{ref199} A related study was published by Barker.\cite{ref200}
1739:
1740: \begin{figure}
1741: \centerline{\includegraphics[width=8cm]{figures/fig26}}
1742: \caption{
1743: Experimental magnetoresistance of a ring of $2\,\mu{\rm m}$ diameter, defined in the 2DEG of a high-mobility GaAs-AlGaAs heterostructure ($T = 270\,{\rm mK}$). The different traces are consecutive parts of a magnetoresistance measurement from 0 to $1.4\,{\rm T}$, digitally filtered to suppress a slowly varying background. The oscillations are seen to persist for fields where $\omega_{\rm c}\tau>1$, but their amplitude is reduced substantially for magnetic fields where $2l_{\rm cycl}\ll W$. (The field value where $2l_{\rm cycl}\equiv 2r_{\rm c}=W$ is indicated). Taken from G. Timp et al., Surf.\ Sci.\ {\bf 196}, 68 (1988).
1744: \label{fig26}
1745: }
1746: \end{figure}
1747:
1748: The Aharonov-Bohm oscillations in the magnetoresistance of a small ring
1749: in a high-mobility 2DEG are quite impressive. As an illustration, we
1750: reproduce in Fig.\ \ref{fig26} the results obtained by Timp et al.\cite{ref201} Low-frequency
1751: modulations were filtered out, so that the rapid oscillations are superimposed
1752: on a constant background. The amplitude of the $h/e$ oscillations diminishes
1753: with increasing magnetic field until eventually the Aharonov-Bohm effect is
1754: completely suppressed. The reduction in amplitude is accompanied by a
1755: reduction in frequency. A similar observation was made by Ford et al.\cite{ref74} In
1756: metals, in contrast, the Aharonov-Bohm oscillations persist to the highest
1757: experimental fields, with constant frequency. The different behavior in a
1758: 2DEG is a consequence of the effect of the Lorentz force on the electrons in
1759: the ring, which is of importance when the cyclotron diameter $2l_{\mathrm{cycl}}$ becomes
1760: smaller than the width $W$ of the arm of the ring, provided $W<l$ (note that
1761: $l_{\mathrm{cycl}}=hk_{\mathrm{F}}/eB$ is much smaller in a 2DEG than in a metal, at the same
1762: magnetic field value). We will return to these effects in Section \ref{sec21}.
1763:
1764: An electrostatic potential $V$ affects the phase of the electron wave function
1765: through the term $(e/\hbar) \int Vdt$ in much the same way as a vector potential
1766: does. If the two arms of the ring have a potential difference $V$, and an
1767: electron traverses an arm in a time $t$, then the acquired phase shift would lead
1768: to oscillations in the resistance with periodicity $\Delta V=h/et$. The electrostatic
1769: Aharonov-Bohm effect has a periodicity that depends on the transit time $t$,
1770: and is not a geometrical property of the ring, as it is for the magnetic effect. A
1771: distribution of transit times could easily average out the oscillations. Note
1772: that the potential difference effectuates the phase difference by changing the
1773: wavelength of the electrons (via a change in their kinetic energy), which also
1774: distinguishes the electrostatic from the magnetic effect (where a phase shift is
1775: induced by the vector potential without a change in wavelength). An
1776: experimental search for the electrostatic Aharonov-Bohm effect in a small
1777: metal ring was performed by Washburn et al.\cite{ref202} An electric field was applied
1778: in the plane of the ring by small capacitive electrodes. They were able to shift
1779: the phase of the magnetoresistance oscillations by varying the field, but the
1780: effect was not sufficiently strong to allow the observation of purely electrostatic oscillations. Unfortunately, this experiment could not discriminate between the effect of the electric field penetrating in the arms of the ring (which
1781: could induce a phase shift by changing the trajectories) and that of the electrostatic potential. Experiments have been reported by De Vegvar et al.\cite{ref203}
1782: on the manipulation of the phase of the electrons by means of the voltage on a
1783: gate electrode positioned across one of the arms of a heterostructure ring. In
1784: this system a change in gate voltage has a large effect on the resistance of the
1785: ring, primarily because it strongy affects the local density of the electron gas.
1786: No clear periodic signal, indicative of an electrostatic Aharonov-Bohm
1787: effect, could be resolved. As discussed in Ref.\ \onlinecite{ref203}, this is not too surprising, in
1788: view of the fact that in that device 1D subband depopulation in the region
1789: under the gate occurs on the same gate voltage scale as the expected
1790: Aharonov-Bohm effect. The observation of an electrostatic Aharonov-Bohm effect thus remains an experimental challenge. A successful experiment
1791: would appear to require a ring in which only a single 1D subband is occupied,
1792: to ensure a unique transit time.\cite{ref198,ref200}
1793:
1794: \subsection{\label{sec9} Electron-electron interactions}
1795:
1796: \subsubsection{\label{sec9a} Theory}
1797:
1798: In addition to the weak localization correction to the conductivity
1799: discussed in Section \ref{sec6}, which arises from a single-electron quantum interference effect, the Coulomb interaction of the conduction electrons gives also
1800: rise to a quantum correction.\cite{ref204,ref205} In two dimensions the latter correction
1801: has a logarithmic temperature dependence, just as for weak localization [see
1802: Eq.\ (\ref{eq6.4})]. A perpendicular magnetic field can be used to distinguish the two
1803: quantum corrections, which have a different field dependence.\cite{ref118,ref204,ref205,ref206,ref207,ref208,ref209,ref210}
1804: This field of research has been reviewed in detail by Al'tshuler and
1805: Aronov,\cite{ref211} by Fukuyama,\cite{ref212} and by Lee and Ramakrishnan,\cite{ref127} with an
1806: emphasis on the theory. A broader review of electronic correlation effects in
1807: 2D systems has been given by Isihara in this series.\cite{ref213} In the present
1808: subsection we summarize the relevant theory, as a preparation for the
1809: following subsection on experimental studies in semiconductor nanostructures. We do not discuss the diagrammatic perturbation theory, since it is
1810: highly technical and does not lend itself to a discussion at the same level as for
1811: the other subjects dealt with in this review.
1812:
1813: An attempt at an intuitive interpretation of the Feynman diagrams was
1814: made by Bergmann.\cite{ref214} It is argued that one important class of diagrams may
1815: be interpreted as diffraction of one electron by the oscillations in the
1816: electrostatic potential generated by the other electrons. The Coulomb
1817: interaction between the electrons thus introduces a purely quantum mechanical correlation between their motion, which is observable in the conductivity.
1818: The diffraction of one electron wave by the interference pattern generated
1819: by another electron wave will only be of importance if their wavelength
1820: difference, and thus their energy difference, is small. At a finite temperature $T$,
1821: the characteristic energy difference is $k_{\mathrm{B}}T$. The time $\tau_{\mathrm{T}}\equiv \hbar/k_{\mathrm{B}}T$ enters as a
1822: long-time cutoff in the theory of electron-electron interactions in a disordered conductor, in the usual case\cite{ref127,ref211} $\tau_{\mathrm{T}}\lesssim\tau_{\phi}$. (Fukuyama\cite{ref212} also
1823: discusses the opposite limit $\tau_{\mathrm{T}}\gg \tau_{\phi}$.) Accordingy, the magnitude of the
1824: thermal length $l_{\mathrm{T}}\equiv(D\tau_{\mathrm{T}})^{1/2}$ compared with the width $W$ determines the
1825: dimensional crossover from 2D to 1D [for $l_{\mathrm{T}}<l_{\phi}\equiv(D\tau_{\phi})^{1/2}]$. In the
1826: expression for the conductivity correction associated with electron-electron
1827: interactions, the long-time cutoff $\tau_{\mathrm{T}}$ enters logarithmically in 2D and as a
1828: square root in 1D. These expressions thus have the same form as for weak
1829: localization, but with the phase coherence time $\tau_{\phi}$ replaced by $\tau_{\mathrm{T}}$. The origin
1830: of this difference is that a finite temperature does not introduce a long-time
1831: cutoff for the single-electron quantum interference effect responsible for weak
1832: localization, but merely induces an energy average of the corresponding
1833: conductivity correction.
1834:
1835: In terms of effective interaction parameters $g_{2\mathrm{D}}$ and $g_{1\mathrm{D}}$, the conductivity
1836: corrections due to electron-electron interactions can be written as (assuming
1837: $\tau\ll \tau_{\mathrm{T}}\ll \tau_{\phi})$
1838: \begin{subequations}
1839: \label{eq9.1}
1840: \begin{eqnarray}
1841: \delta\sigma_{\mathrm{ee}}&=&-\frac{e^{2}}{2\pi^{2}\hbar}g_{2\mathrm{D}}\ln\frac{\tau_{\mathrm{T}}}{\tau},\;\; {\rm for}\;\; l_{\mathrm{T}}\ll W, \label{eq9.1a}\\
1842: \delta\sigma_{\mathrm{ee}}&=&-\frac{e^{2}}{2^{1/2}\pi \hbar}g_{1\mathrm{D}}\frac{l_{\mathrm{T}}}{W},\;\; {\rm for}\;\; W\ll l_{\mathrm{T}}\ll L. \label{eq9.1b}
1843: \end{eqnarray}
1844: \end{subequations}
1845: Under typical experimental conditions,\cite{ref55} the constants $g_{2\mathrm{D}}$ and $g_{1\mathrm{D}}$ are
1846: positive and of order unity. Theoretically, these effective interaction parameters depend in a complicated way on the ratio of screening length to Fermi
1847: wavelength and can have either sign. We do not give the formulas here, but
1848: refer to the reviews by Al'tshuler and Aronov\cite{ref211} and Fukuyama.\cite{ref212} In 2D
1849: the interaction correction $\delta\sigma_{\mathrm{ee}}$ shares a logarithmic temperature dependence
1850: with the weak localization correction $\delta\sigma_{\mathrm{loc}}$, and both corrections are of the
1851: same order of magnitude. In 1D the temperature dependences of the two
1852: effects are different (unless $\tau_{\phi}\propto T^{-1/2}$). Moreover, in the 1D case $\delta\sigma_{\mathrm{ee}}\ll \delta\sigma_{\mathrm{loc}}$
1853: if $l_{\mathrm{T}}\ll l_{\phi}$.
1854:
1855: A weak magnetic field fully suppresses weak localization, but has only a
1856: small effect on the quantum correction from electron-electron interactions.
1857: The conductance correction $\delta G_{\mathrm{ee}}$ contains contributions of diffuson type and
1858: of cooperon type. The diffusons (which give the largest contributions to $\delta G_{\mathrm{ee}}$)
1859: are affected by a magnetic field only via the Zeeman energy, which removes
1860: the spin degeneracy when $g\mu_{\mathrm{B}}B\gtrsim k_{\mathrm{B}}T$. In the systems of interest here, spin
1861: splitting can usually be ignored below $1\mathrm{T}$, so the diffusons are insensitive to a
1862: weak magnetic field. Since the spin degeneracy is removed regardless of the
1863: orientation of the magnetic field, the $B$-dependence of the diffuson is
1864: isotropic. The smaller cooperon contributions exhibit a similar sensitivity as
1865: weak localization to a weak perpendicular magnetic field, the characteristic
1866: field being determined by $l_{\mathrm{m}}^{2}\approx l_{\mathrm{T}}^{2}$ in 2D and by $l_{\mathrm{m}}^{2}\approx Wl_{\mathrm{T}}$ in 1D (in the dirty
1867: metal regime $W\gg l$, so flux cancellation does not play a significant role). The
1868: magnetic length $l_{\mathrm{m}}\equiv(\hbar/eB_{\perp})^{1/2}$ contains only the component $B_{\perp}$ of the field
1869: perpendicular to the 2DEG, since the magnetic field affects the cooperon via
1870: the phase shift induced by the enclosed flux. The anisotropy and the small
1871: characteristic field are two ways to distinguish experimentally the cooperon
1872: contribution from that of the diffuson. It is much more difficult to distinguish
1873: the cooperon contribution to $\delta G_{\mathrm{ee}}$ from the weak localization correction,
1874: since both effects have the same anisotropy, while their characteristic fields
1875: are comparable ($l_{\mathrm{T}}$ and $l_{\phi}$ not being widely separated in the systems
1876: considered here). This complication is made somewhat less problematic by
1877: the fact that the cooperon contribution to $\delta G_{\mathrm{ee}}$ is often considerably smaller
1878: than $\delta G_{\mathrm{loc}}$, in which case it can be ignored. In 1D the reduction factor\cite{ref55,ref211} is
1879: of order $[1+\lambda\ln(E_{\mathrm{F}}/k_{\mathrm{B}}T)]^{-1}(l_{\mathrm{T}}/l_{\phi})$, with $\lambda$ a numerical coefficient of order unity.
1880:
1881: There is one additional aspect to the magnetoresistance due to electron-electron interactions that is of little experimental relevance in metals but
1882: becomes important in semiconductors in the classically strong-field regime
1883: where $\omega_{\mathrm{c}}\tau>1$ (this regime is not easily accessible in metal nanostructures
1884: because of the typically short scattering time). In such strong fields only the
1885: diffuson contributions to the conductivity corrections survive. According to
1886: Houghton et al.\cite{ref215} and Girvin et al.,\cite{ref216} the diffuson does not modify the off-diagonal elements of the conductivity tensor, but only the diagonal elements
1887: \be
1888: \delta\sigma_{xy}=\delta\sigma_{yx}=0,\;\;\delta\sigma_{xx}=\delta\sigma_{yy}\equiv\delta\sigma_{\mathrm{ee}}, \label{eq9.2}
1889: \ee
1890: where $\delta\sigma_{\mathrm{ee}}$ is approximately field-independent (provided spin splitting does
1891: not play a role). In a channel geometry one measures the longitudinal
1892: resistivity $\rho_{xx}$, which is related to the conductivity tensor elements by
1893: \begin{eqnarray}
1894: \rho_{xx}&\equiv&\frac{\sigma_{yy}}{\sigma_{xx}\sigma_{yy}+\sigma_{xy}^{2}}\nonumber\\
1895: &=&\rho_{xx}^{0}+\rho_{xx}^{0}\left(\frac{\delta\sigma_{\mathrm{ee}}}{\sigma_{xx}^{0}}-2\rho_{xx}^{0}\delta\sigma_{\mathrm{ee}}\right)+\mathrm{order}(\delta\sigma_{\mathrm{ee}})^{2}.\nonumber\\
1896: && \label{eq9.3}
1897: \end{eqnarray}
1898: Here $\rho_{xx}^{0}=\rho$ and $\sigma_{xx}^{0}=\sigma[1+(\omega_{\mathrm{c}}\tau)^{2}]^{-1}$ are the classical results (\ref{eq4.25}) and
1899: (\ref{eq4.26}). In obtaining this result the effects of Landau level quantization on the
1900: conductivity have been disregarded (see, however, Ref.\ \onlinecite{ref55}). The longitudinal
1901: resistivity thus becomes magnetic-field-dependent:
1902: \be
1903: \rho_{xx}=\rho(1+[(\omega_{\mathrm{c}}\tau)^{2}-1]\delta\sigma_{\mathrm{ee}}/\sigma). \label{eq9.4}
1904: \ee
1905: To the extent that the $B$-dependence of $\delta\sigma_{\mathrm{ee}}$ can be neglected, Eq.\ (\ref{eq9.4}) gives a
1906: parabolic negative magnetoresistance, with a temperature dependence that is
1907: that of the negative conductivity correction $\delta\sigma_{\mathrm{ee}}$. This effect can easily be
1908: studied up to $\omega_{\mathrm{c}}\tau=10$, which would imply an enhancement by a factor of
1909: 100 of the resistivity correction in zero magnetic field. (The Hall resistivity $\rho_{xy}$
1910: also contains corrections from $\delta\sigma_{\mathrm{ee}}$, but without the enhancement factor.) In
1911: 2D it is this enhancement that allows the small effect of electron-electron
1912: interactions to be observable experimentally (in as far as the effect is due to
1913: diffuson-type contributions).
1914:
1915: Experimentally, the parabolic negative magnetoresistance associated with
1916: electron-electron interactions was first identified by Paalanen et al.\cite{ref137} in
1917: high-mobility GaAs-AlGaAs heterostructure channels. A more detailed
1918: study was made by Choi et al.\cite{ref55} In that paper, as well as in Ref.\ \onlinecite{ref113}, it was
1919: found that the parabolic magnetoresistance was less pronounced in narrow
1920: channels than in wider ones. Choi et al.\ attributed this suppression to
1921: specular boundary scattering. It should be noted, however, that specular
1922: boundary scattering has no effect at all on the classical conductivity tensor $\sigma^{0}$
1923: (in the scattering time approximation; cf.\ Section \ref{sec5b}). Since the parabolic
1924: magnetoresistance results from the $(\omega_{\mathrm{c}}\tau)^{2}$ term in $1/\sigma_{xx}^{0}$ [see Eq.\ (\ref{eq9.4})], one
1925: would expect that specular boundary scattering does not suppress the
1926: parabolic magnetoresistance (assuming that the result $\delta\sigma_{xy}=\delta\sigma_{yx}=0$ still
1927: holds in the pure metal regime $l>W$). Diffuse boundary scattering does
1928: affect $\sigma^{0}$, but only for relatively weak fields such that $2l_{\mathrm{cycl}}>\sim W$ (see Section
1929: \ref{sec5}); hence, diffuse boundary scattering seems equally inadequate in explaining
1930: the observations. In the absence of a theory for electron-electron interaction
1931: effects in the pure metal regime, this issue remains unsettled.
1932:
1933: \subsubsection{\label{sec9b} Narrow-channel experiments}
1934:
1935: Wheeler et al.\cite{ref38} were the first to use magnetoresistance experiments as a
1936: tool to distinguish weak localization from electron-electron interaction
1937: effects in narrow Si MOSFETs. As in most subsequent studies, the negative
1938: magnetoresistance was entirely attributed to the suppression of weak
1939: localization; the cooperon-type contributions from electron-electron interactions were ignored. After subtraction of the weak localization correction,
1940: the remaining temperature dependence was found to differ from the simple
1941: $T^{-1/2}$ dependence predicted by the theory for $W<l_{\mathrm{T}}<l_{\phi}$ [Eq. (\ref{eq9.1b})]. This
1942: was attributed in Ref.\ \onlinecite{ref38} to temperature-dependent screening at the relatively
1943: high temperatures of the experiment. Pooke et al.\cite{ref138} found a nice $T^{-1/2}$
1944: dependence in similar experiments at lower temperatures in narrow Si
1945: accumulation layers and in GaAs-AlGaAs heterostructures.
1946:
1947: \begin{figure}
1948: \centerline{\includegraphics[width=8cm]{figures/fig27}}
1949: \caption{
1950: Negative magnetoresistance in wide and narrow GaAs-AlGaAs channels at 4.2 and 1.6 K. The temperature-independent negative magnetoresistance at low fields is a classical size effect. The temperature-dependent parabolic magnetoresistance at higher fields is a quantum interference effect associated with electron-electron interactions. Shubnikov-De Haas oscillations are visible for fields greater than about 0.3 T. Taken from K. K. Choi et al., Phys.\ Rev.\ B {\bf 33}, 8216 (1986).
1951: \label{fig27}
1952: }
1953: \end{figure}
1954:
1955: The most detailed study by far of the 2D to 1D crossover of the electron-electron interaction effect in narrow channels was made by Choi et al.\cite{ref55} in a
1956: GaAs-AlGaAs heterostructure. In Fig.\ \ref{fig27} we reproduce some of their
1957: experimental traces for channel widths from 156 to 1.1 $\mu \mathrm{m}$ and a channel
1958: length of about 300 $\mu \mathrm{m}$. The weak localization peak in the magnetoresistance
1959: is not resolved in this experiment, presumably because the channels are not in
1960: the 1D regime for this effect (the 2D weak localization peak would be small
1961: and would have a width of $10^{-4}\,\mathrm{T}$). The negative magnetoresistance that they
1962: found below $0.1-0.2\, \mathrm{T}$ in the narrowest channels is temperature-independent
1963: between 1 and $4\, \mathrm{K}$ and was therefore identified by Choi et al.\cite{ref55} as a classical
1964: size effect. The classical negative magnetoresistance extends over a field range
1965: for which $2l_{\mathrm{cycl}}\gtrsim W$. This effect has been discussed in Section \ref{sec5} in terms of
1966: reduction of backscattering by a magnetic field. The electron-electron
1967: interaction effect is observed as a (temperature-dependent) parabolic negative
1968: magnetoresistance above $0.1\, \mathrm{T}$ for the widest channel and above $0.3\, \mathrm{T}$ for the
1969: narrowest one. From the magnitude of the parabolic negative magnetoresistance, Choi et al.\cite{ref55} could find and analyze the crossover from 2D to 1D
1970: interaction effects. In addition, they investigated the cross over to 0D by
1971: performing experiments on short channels. As seen in Fig.\ \ref{fig27}, Shubnikov-De
1972: Haas oscillations are superimposed on the parabolic negative magnetoresistance at low temperatures and strong magnetic fields. It is noteworthy that
1973: stronger fields are required in narrower channels to observe the Shubnikov-De Haas oscillations, an effect discussed in terms of specular boundary
1974: scattering by Choi et al. The Shubnikov-De Haas oscillations in narrow
1975: channels are discussed further in Section \ref{sec10b}.
1976:
1977: \begin{figure}
1978: \centerline{\includegraphics[width=8cm]{figures/fig28}}
1979: \caption{
1980: Magnetoresistance at various temperatures of a GaAs-AlGaAs channel ($W = 0.12\,\mu{\rm m}$, $L = 10\,\mu{\rm m}$) defined by a shallow-mesa etch technique. The central negative magnetoresistance peak between $-0.1$ and $+0.1\,{\rm T}$ at low temperatures is due to 1D weak localization in the quasi-ballistic regime. Conductance fluctuations are seen at larger fields. The negative magnetoresistance that persists to high temperatures is a classical size effect as in Fig.\ \ref{fig27}. The temperature dependence of the resistance at $B = 0$ is due to a combination of weak localization and electron-electron interaction effects (see Fig.\ \ref{fig30}). Taken from H. van Houten et al., Appl.\ Phys.\ Lett.\ {\bf 49}, 1781 (1986).
1981: \label{fig28}
1982: }
1983: \end{figure}
1984:
1985: In Refs.\ \onlinecite{ref63,ref167}, and \onlinecite{ref27} the work by Choi et al.\cite{ref55} was extended to even
1986: narrower channels, well into the 1D pure metal regime. The results for a
1987: conducting channel width of 0.12 $\mu \mathrm{m}$ are shown in Fig.\ \ref{fig28}. The 1D weak
1988: localization peak in the magnetoresistance is quite large for this narrow
1989: channel (even at the rather high temperatures shown) and clearly visible
1990: below 0.1 T. The classical size effect due to reduction of backscattering now
1991: leads to a negative magnetoresistance on a larger field scale of about $1\, \mathrm{T}$, in
1992: agreement with the criterion $2l_{\mathrm{cycl}}\sim W$. This is best seen at temperatures
1993: above $20\, \mathrm{K}$, where the quantum mechanical effects are absent. The
1994: temperature-dependent parabolic negative magnetoresistance is no longer
1995: clearly distinguishable in the narrow channel of Fig.\ \ref{fig28}, in contrast to wider
1996: channels.\cite{ref27,ref55} The suppression of this effect in narrow channels is not yet
1997: understood (see Section \ref{sec9a}). Superimposed on the smooth classical magnetoresistance, one sees large aperiodic fluctuations on a field scale of the
1998: same magnitude as the width of the weak localization peak, in qualitative
1999: agreement with the theory of universal conductance fluctuations in the pure
2000: metal regime\cite{ref147} (see Section \ref{sec7d}). Finally, Shubnikov-De Haas oscillations
2001: are beginning to be resolved around $1.2\, \mathrm{T}$, but they are periodic in $1/B$ at
2002: stronger magnetic fields only (not shown). As discussed in Section \ref{sec10}, this
2003: anomaly in the Shubnikov-De Haas effect is a manifestation of a quantum
2004: size effect.\cite{ref167,ref217,ref218} This one figure thus summarizes the wealth of classical
2005: and quantum magnetoresistance phenomena in the quasi-ballistic transport
2006: regime.
2007:
2008: \begin{figure}
2009: \centerline{\includegraphics[width=8cm]{figures/fig29}}
2010: \caption{
2011: Angular dependence of the magnetoresistance of Fig.\ \ref{fig28}, at 4 K, proving that it has a purely orbital origin. Taken from H. van Houten et al., Superlattices and Microstructures {\bf 3}, 497 (1987).
2012: \label{fig29}
2013: }
2014: \end{figure}
2015:
2016: Essentially similar results were obtained by Taylor et al.\cite{ref219} In the field
2017: range of these experiments,\cite{ref27,ref55,ref63,ref167,ref219} the magnetoresistance is exclusively caused by the enclosed flux and the Lorentz force (so called {\it orbital\/} effects).
2018: The Zeeman energy does not play a role. This is demonstrated in Fig.\ \ref{fig29},
2019: where the magnetoresistance (obtained on the same sample as that used in
2020: Fig.\ \ref{fig28}) is shown to vanish when $B$ is in the plane of the 2DEG (similar results
2021: were obtained in Ref.\ \onlinecite{ref168}). In wide 2DEG channels a negative magnetoresistance has been found by Lin et al.\ in a parallel magnetic field.\cite{ref23} This
2022: effect has been studied in detail by Mensz and Wheeler,\cite{ref220} who attributed it
2023: to a residual orbital effect associated with deviations of the 2DEG from a
2024: perfectly flat plane. Fal'ko\cite{ref221} has calculated the effect of a magnetic field
2025: parallel to the 2DEG on weak localization, and has found a negative
2026: magnetoresistance, but only if the scattering potential does not have
2027: reflection symmetry in the plane of the 2DEG.
2028:
2029: \begin{figure}
2030: \centerline{\includegraphics[width=8cm]{figures/fig30}}
2031: \caption{
2032: Zero-field conductance (circles) and conductance corrected for the weak localization effect (squares) for the channel of Fig.\ \ref{fig28} as a function of $T^{-1/2}$, to demonstrate the $T^{-1/2}$ dependence on the temperature of the electron-electron interaction effect expected from Eq.\ (\ref{eq9.1b}). The solid and dashed lines are guides to the eye. The extrapolated value at high temperatures is the classical part of the conductance. Taken from H. van Houten et al., Acta Electronica {\bf 28}, 27 (1988).
2033: \label{fig30}
2034: }
2035: \end{figure}
2036:
2037: In Fig.\ \ref{fig30} the temperature dependence of the zero-field conductance\cite{ref27} is
2038: plotted as a function of $T^{-1/2}$, together with the conductance after subtraction of the weak localization correction. The straight line through the
2039: latter data points demonstrates that the remaining temperature dependence
2040: may, indeed, be attributed to the electron-electron interactions. A similar
2041: $T^{-1/2}$ dependence was found by Thornton et al.\cite{ref58} in a narrow GaAs-AlGaAs channel defined using the split-gate method. The slope of the straight
2042: line in Fig.\ \ref{fig30} gives $g_{1\mathrm{D}}\approx 1.5$ in Eq.\ (\ref{eq9.1b}), which is close to the value found
2043: by Choi et al.\cite{ref55} It should be noted, however, that this experiment is already in
2044: the regime where the quantum corrections are by no means small, so the
2045: perturbation theory is of questionable validity. For this reason, and also in
2046: view of other problems (such as the difficulty in determining the effective
2047: channel width, the presence of channel width variations, and a frequently
2048: observed saturation of the weak localization correction at low temperatures
2049: due to loss of phase coherence associated with external noise or radio-frequency interference), a quantitative analysis of the parameters obtained
2050: from the weak localization and electron-electron corrections in narrow
2051: channels ($\tau_{\phi}$ and $g_{1\mathrm{D}}$) is not fully warranted. Indeed, most of the narrow-channel studies available today have not been optimized for the purpose of a
2052: detailed quantitative analysis. Instead, they were primarily intended for a
2053: phenomenological exploration, and as such we feel that they have been quite
2054: successful.
2055:
2056: \subsection{\label{sec10} Quantum size effects}
2057:
2058: Quantum size effects on the resistivity result from modifications of the 2D
2059: density of states in a 2DEG channel of width comparable to the Fermi
2060: wavelength. The electrostatic lateral confinement in such a narrow channel
2061: leads to the formation of 1D subbands in the conduction band of the 2DEG
2062: (see Section \ref{sec4a}). The number $N\approx k_{\mathrm{F}}W/\pi$ of occupied 1D subbands is
2063: reduced by decreasing the Fermi energy or the channel width. This depopulation of individual subbands can be detected via the resistivity. An
2064: alternative method to depopulate the subbands is by means of a magnetic
2065: field perpendicular to the 2DEG. The magnetic field $B$ has a negligible effect
2066: on the density of states at the Fermi level if the cyclotron diameter $2l_{\mathrm{cycl}}\gg W$
2067: (i.e., for $B\ll B_{\mathrm{crit}}\equiv 2\hbar k_{\rm F}/eW$). If $B\gg B_{\mathrm{crit}}$, the electrostatic confinement can
2068: be neglected for the density of states, which is then described by Landau levels
2069: [Eq.\ (\ref{eq4.6})]. The number of occupied Landau levels $N\approx B_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}\approx k_{\mathrm{F}}l_{\mathrm{cycl}}/2$
2070: decreases linearly with $B$ for $B\gg B_{\mathrm{crit}}$. In the intermediate field range where $B$
2071: and $B_{\mathrm{crit}}$ are comparable, the electrostatic confinement and the magnetic field
2072: together determine the density of states. The corresponding {\it magnetoelectric\/}
2073: subbands are depopulated more slowly by a magnetic field than are the
2074: Landau levels, which results in an increased spacing of the Shubnikov-De
2075: Haas oscillations in the magnetoresistivity (cf.\ Section \ref{sec4c}).
2076:
2077: In the following subsection we give a more quantitative description of
2078: magnetoelectric subbands. Experiments on the electric and magnetic depopulation of subbands in a narrow channel are reviewed in Section \ref{sec9b}. We
2079: only consider here the case of a long channel $(L\gg l)$ in the quasi-ballistic
2080: regime. Quantum size effects in the fully ballistic regime $(L\lesssim l)$ are the
2081: subject of Section \ref{secIII}.
2082:
2083: \subsubsection{\label{sec10a} Magnetoelectric subbands}
2084:
2085: Consider first the case of an unbounded 2DEG in a perpendicular
2086: magnetic field $\mathbf{B}=\nabla\times \mathbf{A}$. The Hamiltonian for motion in the plane of the
2087: 2DEG is given by
2088: \be
2089: {\cal H}=\frac{(\mathbf{p}+e\mathbf{A})^{2}}{2m}, \label{eq10.1}
2090: \ee
2091: for a single spin component. In the Landau gauge ${\bf A}=(0, Bx, 0)$, with $\mathbf{B}$ in the
2092: $z$-direction, this may be written as
2093: \be
2094: {\cal H}=\frac{p_{x}^{2}}{2m}+\frac{m\omega_{\mathrm{c}}^{2}}{2}(x-x_{0})^{2}, \label{eq10.2}
2095: \ee
2096: with $\omega_{\mathrm{c}}\equiv eB/m$ and $x_{0}\equiv-p_{y}/eB$. The $y$-momentum operator $p_{y}\equiv-i\hbar\partial/\partial y$
2097: can be replaced by its eigenvalue $\hbar k_{y}$, since $p_{y}$ and ${\cal H}$ commute. The effect of
2098: the magnetic field is then represented by a harmonic oscillator potential in
2099: the $x$-direction, with center $x_{0}=-\hbar k_{y}/eB$ depending on the momentum in
2100: the $y$-direction. The energy eigenvalues $E_{n}=(n- \frac{1}{2})\hbar\omega_{\mathrm{c}}$, $n=1,2,3, \ldots$, do
2101: not depend on $k_{y}$ and are therefore highly degenerate. States with the same
2102: quantum number $n$ are referred to collectively as {\it Landau levels}.\cite{ref93} The
2103: number of Landau levels below energy $E$ is given by
2104: \be
2105: N=\mathrm{Int}[1/2+E/\hbar\omega_{\mathrm{c}}], \label{eq10.3}
2106: \ee
2107: where $\mathrm{Int}$ denotes truncation to an integer.
2108:
2109: \begin{figure}
2110: \centerline{\includegraphics[width=8cm]{figures/fig31}}
2111: \caption{
2112: Magnetic field dependence of the number $N$ of occupied subbands in a narrow channel for a parabolic confining potential according to Eq.\ (\ref{eq10.7}) (dot-dashed curve), and for a square-well confining potential according to Eq.\ (\ref{eq10.8}) (full curve). The dashed curve gives the magnetic depopulation of Landau levels in a wide 2DEG, which has a $1/B$ dependence. The calculations are done for a fixed Fermi energy and for channel width $W = W_{\rm par} = 10\pi/k_{\rm F}$.
2113: \label{fig31}
2114: }
2115: \end{figure}
2116:
2117: A narrow channel in the $y$-direction is defined by an electrostatic confining
2118: potential $V(x)$. The case of a {\it parabolic\/} confinement is easily solved analytically.\cite{ref36,ref218,ref222,ref223} Adding a term $V(x)= \frac{1}{2}m\omega_{0}^{2}x^{2}$ to the hamiltonian
2119: (\ref{eq10.2}), one finds, after a rearrangement of terms,
2120: \be
2121: {\cal H}=\frac{p_{x}^{2}}{2m}+\frac{m\omega^{2}}{2}(x-\overline{x}_{0})^{2}+\frac{\hbar^{2}k^{2}}{2M}, \label{eq10.4}
2122: \ee
2123: with $\omega\equiv(\omega_{\mathrm{c}}^{2}+\omega_{0}^{2})^{1/2}$, $\overline{x}_{0}\equiv x_{0}\omega_{\mathrm{c}}/\omega$, and $M\equiv m\omega^{2}/\omega_{0}^{2}$. The first two terms
2124: describe the motion in the $x$-direction in a harmonic potential with effective
2125: frequency $\omega\geq\omega_{0}$, and the third term describes free motion in the $y$-direction
2126: with an effective mass $M\geq m$. This last term removes the degeneracy of the
2127: Landau levels, which become 1D subbands with energy
2128: \be
2129: E_{n}(k)=(n- {\textstyle\frac{1}{2}})\hbar\omega+\hbar^{2}k^{2}/2M. \label{eq10.5}
2130: \ee
2131: The subband bottoms have energy $E_{n}=(n- \frac{1}{2})\hbar\omega$, and the number of
2132: subbands occupied at energy $E$ is $N= \mathrm{Int}[\frac{1}{2}+E/\hbar\omega]$. The quasi-1D density
2133: of states is obtained from Eq.\ (\ref{eq4.4}) on substituting $m$ for $M$. For the
2134: comparison with experiments it is useful to define an effective width for the
2135: parabolic potential. One can take the width $W_{\mathrm{par}}$ to be the separation between
2136: the equipotentials at the Fermi energy,
2137: \be
2138: W_{\mathrm{par}}\equiv 2\hbar k_{\rm F}/m\omega_{0}. \label{eq10.6}
2139: \ee
2140: (An alternative, which differs only in the numerical prefactor, is to take
2141: $W_{\mathrm{par}}\equiv n_{1\mathrm{D}}/n_{\mathrm{s}}$, with $n_{\mathrm{s}}\equiv g_{\mathrm{s}}g_{\mathrm{v}}k_{\mathrm{F}}^{2}/4\pi$ the 2D sheet density and $n_{1\mathrm{D}}$ the number
2142: of electrons per unit length in the narrow channel.\cite{ref218}) The number of
2143: occupied magnetoelectric subbands at energy $E_{\mathrm{F}}$ in a {\it parabolic\/} confining
2144: potential may then be written as
2145: \be
2146: N=\mathrm{Int}\left[
2147: \frac{1}{2}+\frac{1}{4}k_{\rm F} W_{\rm par}[1+(W_{\mathrm{par}}/2l_{\mathrm{cycl}})^{2}]^{-1/2}\right], \label{eq10.7}
2148: \ee
2149: where $l_{\mathrm{cycl}}\equiv \hbar k_{\rm F}/eB$ is the cyclotron radius at the Fermi energy. For easy
2150: reference, we also give the result for the number of occupied subbands at the
2151: Fermi energy in a {\it square-well\/} confinement potential of width $W$:
2152: \begin{widetext}
2153: \begin{subequations}
2154: \label{eq10.8}
2155: \begin{eqnarray}
2156: N&\approx& \mathrm{Int}\left[\frac{2}{\pi}\frac{E_{\mathrm{F}}}{\hbar\omega_{\mathrm{c}}}\left(\arcsin\frac{W}{2l_{\mathrm{cycl}}}+\frac{W}{2l_{\mathrm{cycl}}}\left[1-\left(\frac{W}{2l_{\mathrm{cycl}}}\right)^{2}\right]^{1/2}\right)\right],\;\;{\rm if}\;\; l_{\mathrm{cycl}}> \frac{W}{2},\label{eq10.8a}\\
2157: N&\approx& \mathrm{Int}\left[\frac{1}{2}+\frac{E_{\mathrm{F}}}{\hbar\omega_{\mathrm{c}}}\right]\;\;{\rm if}\;\;l_{\mathrm{cycl}}< \frac{W}{2}. \label{eq10.8b}
2158: \end{eqnarray}
2159: \end{subequations}
2160: \end{widetext}
2161: (This result is derived in Section \ref{sec12a} in a semiclassical approximation. The
2162: accuracy is $\pm 1$.) One easily verifies that, for $B\ll B_{\mathrm{crit}}\equiv 2\hbar k_{\rm F}/eW,$ Eq.\ (\ref{eq10.8})
2163: yields $N\approx k_{\mathrm{F}}W/\pi$. The parabolic confining potential gives $N\approx k_{\mathrm{F}}W_{\mathrm{par}}/4$ in
2164: the weak-field limit. In the strong-field limit $B\gg B_{\mathrm{crit}}$, both potentials give
2165: the result $N\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}=k_{\mathrm{F}}l_{\mathrm{cycl}}/2$ expected for pure Landau levels. In Fig.\ \ref{fig31}
2166: we compare the depopulation of Landau levels in an unbounded 2DEG with
2167: its characteristic $1/B$ dependence of $N$ (dashed curve), with the slower
2168: depopulation of magnetoelectric subbands in a narrow channel. The dash-dotted curve is for a parabolic confining potential, the solid curve for a
2169: square-well potential. These results are calculated from Eqs.\ (\ref{eq10.7}) and (\ref{eq10.8}),
2170: with $k_{\mathrm{F}}W_{\mathrm{par}}/\pi=k_{\mathrm{F}}W/\pi=10$. A $B$-independent Fermi energy was assumed in
2171: Fig.\ \ref{fig31} so that the density $n_{1\mathrm{D}}$ oscillates around its zero-field value. (For a
2172: long channel, it is more appropriate to assume that $n_{1\mathrm{D}}$ is $B$-independent, to
2173: preserve charge neutrality, in which case $E_{\mathrm{F}}$ oscillates. This case is studied in
2174: Ref.\ \onlinecite{ref218}.) Qualitatively, the two confining potentials give similar results. The
2175: numerical differences reflect the uncertainty in assigning an effective width to
2176: the parabolic potential. Self-consistent solutions of the Poisson and Schr\"{o}dinger equations\cite{ref42,ref60,ref61,ref72,ref224} for channels defined by a split gate have shown
2177: that a parabolic potential with a flat bottom section is a more realistic model.
2178: The subband depopulation for this potential has been studied in a semiclassical approximation in Ref.\ \onlinecite{ref223}. A disadvantage of this more realistic model is
2179: that an additional parameter is needed for its specification (the width of the
2180: flat section). For this practical reason, the use of either a parabolic or a
2181: square-well potential has been preferred in the analysis of most experiments.
2182:
2183: \subsubsection{\label{sec10b} Experiments on electric and magnetic depopulation of subbands}
2184:
2185: The observation of 1D subband effects unobscured by thermal smearing
2186: requires low temperatures, such that $4k_{\mathbf{B}}T\ll \Delta E$, with $\Delta E$ the energy
2187: difference between subband bottoms near the Fermi level ($4k_{\mathbf{B}}T$ being the
2188: width of the energy averaging function $df/dE_{\mathrm{F}}$; see Section \ref{sec4b}; For a square
2189: well $\Delta E\approx 2E_{\mathrm{F}}/N$, and for parabolic confinement $\Delta E\approx E_{\mathrm{F}}/N$). Moreover,
2190: the formation of subbands requires the effective mean free path (limited by
2191: impurity scattering and diffuse boundary scattering) to be much larger than
2192: $W$ (cf.\ also Ref.\ \onlinecite{ref218}). The requirement on the temperature is not difficult to
2193: meet, $\Delta E/4k_{\mathbf{B}}T$ being on the order of $50\, \mathrm{K}$ for a typical GaAs-AlGaAs
2194: channel of width $W=100\,\mathrm{nm}$, and the regime $l>W$ is also well accessible.
2195: These simple considerations seem to suggest that 1D subband effects should
2196: be rather easily observed in semiconductor nanostructures. This conclusion is
2197: misleading, however, and in reality manifestations of 1D subband structure
2198: have been elusive, at least in the quasi-ballistic regime $W<l<L$. The main
2199: reason for this is the appearance of large conductance fluctuations that mask
2200: the subband structure if the channel is not short compared with the mean free
2201: path.
2202:
2203: Calculations\cite{ref225,ref226,ref227} of the {\it average\/} conductivity of an {\it ensemble\/} of narrow
2204: channels do in fact show oscillations from the electric depopulation of
2205: subbands [resulting from the modulation of the density of states at the Fermi
2206: level, which determines the scattering time; see Eq.\ (\ref{eq4.28})]. The oscillations
2207: are not as large as the Shubnikov-De Haas oscillations from the magnetic
2208: depopulation of Landau levels or magnetoelectric subbands. One reason for
2209: this difference is that the peaks in the density of states become narrower,
2210: relative to their separation, on applying a magnetic field. (The quantum limit
2211: of a single occupied 1D subband has been studied in Refs.\ \onlinecite{ref42} and \onlinecite{ref228,ref229,ref230}.)
2212:
2213: \begin{figure}
2214: \centerline{\includegraphics[width=6cm]{figures/fig32}}
2215: \caption{
2216: (a) Dependence on the gate voltage of the current $I$ through 250 parallel narrow Si inversion layer channels at 1.2 K, showing the electric depopulation of subbands. (b) The effect is seen more clearly in the transconductance $dI/dV_{\rm G}$. Note the absence of universal conductance fluctuations, which have been averaged out by the large number of channels. Taken from A. C. Warren et al., IEEE Electron Device Lett.\ {\bf EDL-7}, 413 (1986).
2217: \label{fig32}
2218: }
2219: \end{figure}
2220:
2221: In an individual channel, aperiodic conductance fluctuations due to
2222: quantum interference (see Section 7) are the dominant cause of structure in
2223: the low-temperature conductance as a function of gate voltage (which
2224: corresponds to a variation of the Fermi energy), as was found in experiments
2225: on narrow Si inversion layers.\cite{ref46,ref161,ref162} Warren et al.\cite{ref44} were able to suppress
2226: these fluctuations by performing measurements on an array of narrow
2227: channels in a Si inversion layer. In Fig.\ \ref{fig32} we reproduce their results. The
2228: structure due to the electric depopulation of 1D subbands is very weak in the
2229: current-versus-gate-voltage plot, but a convincingly regular oscillation is
2230: seen if the derivative of the current with respect to the gate voltage is taken
2231: (this quantity is called the transconductance). Warren et al.\ pointed out that
2232: the observation of a quantum size effect in an array of 250 channels indicates
2233: a rather remarkable uniformity of the width and density of the individual
2234: channels.
2235:
2236: More recently a similar experimental study was performed by Ismail et
2237: al.\cite{ref231} on 100 parallel channels defined in the 2DEG of a GaAs-AlGaAs
2238: heterostructure. The effects were found to be much more pronounced than in
2239: the earlier experiment on Si inversion layer channels, presumably because of
2240: the much larger mean free path (estimated at 1 $\mu \mathrm{m}$), which was not much
2241: shorter than the sample length (5 $\mu \mathrm{m}$). Quantum size effects in the quantum
2242: ballistic transport regime (in particular, the conductance quantization of a
2243: quantum point contact) are discussed extensively in Section \ref{sec13}.
2244: In a wide 2DEG the minima of the Shubnikov-De Haas oscillations in the
2245: magnetoresistance are periodic in $1/B$, with a periodicity $\Delta(1/B)$ determined
2246: by the sheet density $n_{\mathrm{s}}$ according to Eq.\ (\ref{eq4.29}). In a narrow channel one
2247: observes an increase in $\Delta(1/B)$ for weak magnetic fields because the
2248: electrostatic confinement modifies the density of states, as discussed in
2249: Section \ref{sec10a}. Such a deviation is of interest as a manifestation of magnetoelectric subbands, but also because it can be used to estimate the effective
2250: channel width using the criterion $W\approx 2l_{\mathrm{cycl}}$ for the crossover field\cite{ref167} $B_{\mathrm{crit}}$
2251: (the electron density in the channel, and hence $l_{\mathrm{cycl}}$, may be estimated from
2252: the strong-field periodicity). The phenomenon has been studied in many
2253: publications.\cite{ref36,ref56,ref57,ref74,ref79,ref167,ref217,ref218,ref223,ref232,ref233}
2254:
2255: \begin{figure}
2256: \centerline{\includegraphics[width=6cm]{figures/fig33}}
2257: \caption{
2258: (a) Magnetoresistance at 2.4 K of a narrow GaAs-AlGaAs channel (as in Fig.\ \ref{fig28}). The arrows indicate magnetic field values assigned to the depopulation of magnetoelectric subbands. (b) Subband index $n \equiv N - 1$ versus inverse magnetic field (crosses). The dashed line interpolates between theoretical points for a parabolic confining potential (circles). The electrostatic confinement causes deviations from a linear dependence of $n$ on $B^{-1}$. Taken from K.-F. Berggren et al., Phys.\ Rev.\ B {\bf 37}, 10118 (1988).
2259: \label{fig33}
2260: }
2261: \end{figure}
2262:
2263: As an illustration, we reproduce in Fig.\ \ref{fig33}a an experimental magnetoresistance trace\cite{ref167,ref218} obtained for a narrow ($W\approx 140\,\mathrm{nm}$) GaAs-AlGaAs channel, defined using a shallow-mesa etch.\cite{ref63} The arrows indicate
2264: the magnetoresistance minima thought to be associated with magnetic
2265: depopulation. The assignment becomes ambiguous in weak magnetic fields,
2266: because of the presence of aperiodic conductance fluctuations. Nevertheless,
2267: the deviation from a straight line in the $N$ versus $B^{-1}$ plot in Fig.\ \ref{fig33}b is
2268: sufficiently large to be reasonably convincing. Also shown in Fig.\ \ref{fig33}b is the
2269: result of a fit to a theoretical $N(B)$ function (assuming a parabolic confining
2270: potential and a $B$-independent electron density). The parameter values found
2271: from this fit for the width and electron density are reasonable and agree with
2272: independent estimates.\cite{ref27}
2273:
2274: We have limited ourselves to a discussion of transport studies, but wish to
2275: point out that 1D subbands have been studied succesfully by capacitance\cite{ref75}
2276: measurements and by infrared\cite{ref78} spectroscopy. As mentioned earlier, the
2277: formation of 1D subbands also requires a reformulation of the theories of
2278: weak localization and conductance fluctuations in the presence of boundary
2279: scattering. Weak localization in the case of a small number of occupied
2280: subbands has been studied by Tesanovic et al.\cite{ref110,ref234} (in a zero magnetic
2281: field).
2282:
2283: We will not discuss the subject of quantum size effects further in this part
2284: of our review, since it has found more striking manifestations in the ballistic
2285: transport regime (the subject of Section \ref{secIII}), where conductance fluctuations
2286: do not play a role. The most prominent example is the conductance
2287: quantization of a point contact.
2288:
2289: \subsection{\label{sec11} Periodic potential}
2290:
2291: \subsubsection{\label{sec11a} Lateral superlattices}
2292:
2293: In a crystal, the periodic potential of the lattice opens energy gaps of zero
2294: density of electronic states. An electron with energy in a gap is Bragg-reflected and hence cannot propagate through the crystal. Esaki and Tsu\cite{ref235}
2295: proposed in 1970 that an artificial energy gap might be created by the
2296: epitaxial growth of alternating layers of different semiconductors. In such a
2297: {\it superlattice\/} a periodic potential of spacing $a$ is superimposed on the crystal
2298: lattice potential. Typically, $a\approx 10\,\mathrm{nm}$ is chosen to be much larger than the
2299: crystal lattice spacing ($0.5\, \mathrm{nm}$), leading to the formation of a large number of
2300: narrow bands within the conduction band (minibands), separated by small
2301: energy gaps (minigaps). Qualitatively new transport properties may then be
2302: expected. For example, the presence of minigaps may be revealed under
2303: strong applied voltages by a negative differential resistance phenomenon
2304: predicted by Esaki and Tsu in their original proposal and observed
2305: subsequently by Esaki and Chang.\cite{ref236,ref237} In contrast to a 3D crystal lattice, a
2306: superlattice formed by alternating layers is 1D. As a consequence of the free
2307: motion in the plane of the layers, the density of states is not zero in the
2308: minigaps, and electrons may scatter between two overlapping minibands. Of
2309: interest in the present context is the possibility of defining {\it lateral\/} superlattices\cite{ref238,ref239} by a periodic potential in the plane of a 2D electron gas. True
2310: minigaps of zero density of states may form in such a system if the potential
2311: varies periodically in two directions. Lateral superlattice effects may be
2312: studied in the linear-response regime of small applied voltages (to which we
2313: limit the discussion here) by varying $E_{\mathrm{F}}$ or the strength of the periodic
2314: potential by means of a gate voltage. The conductivity is expected to vanish if
2315: $E_{\mathrm{F}}$ is in a true minigap (so that electrons are Bragg-reflected). Calculations\cite{ref240,ref241} show pronounced minima also in the case of a 1D periodic
2316: potential.
2317:
2318: The conditions required to observe the minibands in a lateral superlattice
2319: are similar to those discussed in Section \ref{sec10} for the observation of 1D
2320: subbands in a narrow channel. The mean free path should be larger than the
2321: lattice constant $a$, and $4k_{\mathbf{B}}T$ should be less than the width of a minigap near
2322: the Fermi level. For a weak periodic potential,\cite{ref94} the $n$th minigap is
2323: approximately $\Delta E_{n}\approx 2V_{n}$, with $V_{n}$ the amplitude of the Fourier component of
2324: the potential at wave number $k_{n}=2\pi n/a$. The gap is centered at energy
2325: $E_{n}\approx(\hbar k_{n}/2)^{2}/2m$. If we consider, for example, a 1D sinusoidal potential
2326: $V(x, y)=V_{0}\sin(2\pi y/a)$, then the first energy gap $\Delta E_{1}\approx V_{0}$ occurs at
2327: $E_{1}\approx(\hbar\pi/a)^{2}/2m$. (Higher-order minigaps are much smaller.) Bragg reflection
2328: occurs when $E_{1}\approx E_{\mathrm{F}}$ (i.e., for a lattice periodicity $a\approx\lambda_{\mathrm{F}}/2$). Such a short-
2329: period modulation is not easy to achieve lithographically, however (typically
2330: $\lambda_{\mathrm{F}}=40\,\mathrm{nm})$, and the experiments on lateral superlattices discussed later are
2331: not in this regime.
2332:
2333: \begin{figure}
2334: \centerline{\includegraphics[width=8cm]{figures/fig34}}
2335: \caption{
2336: Grating gate (in black) on top of a GaAs-AlGaAs heterostructure, used to define a 2DEG with a periodic density modulation. Taken from K. Ismail et al., Appl.\ Phys.\ Lett.\ {\bf 52}, 1071 (1988).\label{fig34}
2337: }
2338: \end{figure}
2339:
2340: Warren et al.\cite{ref242} have observed a weak but regular structure in the
2341: conductance of a 1D lateral superlattice with $a=0.2\,\mu \mathrm{m}$ defined in a Si
2342: inversion layer (using the dual-gate arrangement of Fig.\ \ref{fig2}c). Ismail et al.\cite{ref62}
2343: used a grating-shaped gate on top of a GaAs-AlGaAs heterostructure to
2344: define a lateral superiattice. A schematic cross section of their device is shown in Fig.\ \ref{fig34}. The period of the grating is $0.2\,\mu{\rm m}$. One effect of the gate voltage is to change the overall carrier concentration, leading to a large but essentially smooth conductance variation (at 4.2 K). This variation proved to be essentially the same as that found for a continuous gate. As in the experiment by Warren et al., the transconductance as a function of the voltage on the grating revealed a regular oscillation. As an example, we reproduce the results of Ismail et al.\ (for various source-drain voltages) in Fig.\ \ref{fig35}. No such structure was found for control devices with a continuous, rather than a grating, gate. The observed structure is attributed to Bragg reflection in Ref.\ \onlinecite{ref62}. A 2D lateral superlattice was defined by Bernstein and Ferry,\cite{ref243} using a grid-shaped gate, but the transport properties in the linear response regime were not studied in detail. Smith et al.\cite{ref244} have used the split-gate technique to define a 2D array of 4000 dots in a high-mobility GaAs-AlGaAs heterostructure ($a = 0.5\,\mu{\rm m}$, $1= 10\,\mu{\rm m}$). When the 2DEG under the dots is depleted, a grid of conducting channels is formed. In this experiment the amplitude of
2345: the periodic potential exceeds $E_{\mathrm{F}}$. Structure in the conductance is found
2346: related to the depopulation of 1D subbands in the channels, as well as to
2347: standing waves between the dots. The analysis is thus considerably more
2348: complicated than it would be for a weak periodic potential. It becomes
2349: difficult to distinguish between the effects due to quantum interference within
2350: a single unit cell of the periodic potential and the effects due to the formation
2351: of minibands requiring phase coherence over several unit cells. Devices with a
2352: 2D periodic potential with a period comparable to the Fermi wavelength and
2353: much shorter than the mean free path will be required for the realization of
2354: true miniband effects. It appears that the fabrication of such devices will have
2355: to await further developments in the art of making nanostructures. Epitaxy
2356: on tilted surfaces with a staircase surface structure is being investigated for
2357: this purpose.\cite{ref87,ref88,ref169,ref179,ref245,ref246} Nonepitaxial growth on Si surfaces slightly
2358: tilted from (100) is known to lead to miniband formation in the inversion
2359: layer.\cite{ref20,ref247} A final interesting possibility is to use doping quantum wires, as
2360: proposed in Ref.\ \onlinecite{ref248}.
2361:
2362: \begin{figure}
2363: \centerline{\includegraphics[width=8cm]{figures/fig35}}
2364: \caption{
2365: Transconductance $g_{\rm m}\equiv \partial I/\partial V_{\rm SD}$ of the device of Fig.\ \ref{fig34} measured as a function of gate voltage for various values of the source-drain voltage. The oscillations, seen in particular at low source-drain voltages, are attributed to Bragg reflection in a periodic potential. Taken from K. Ismail et al., Appl.\ Phys.\ Lett.\ {\bf 52}, 1071 (1988).
2366: \label{fig35}
2367: }
2368: \end{figure}
2369:
2370: As mentioned, it is rather difficult to discriminate experimentally between
2371: true miniband effects and quantum interference effects occurring within one
2372: unit cell. The reason is that both phenomena give rise to structure in the
2373: conductance as a function of gate voltage with essentially the same periodicity. This difficulty may be circumvented by studying lateral superlattices with
2374: a small number of unit cells. The miniband for a finite superlattice with $P$ unit
2375: cells consists of a group of $P$ discrete states, which merge into a continuous
2376: miniband in the limit $P\rightarrow\infty$. The discrete states give rise to closely spaced
2377: resonances in the transmission probability through the superlattice as a
2378: function of energy, and may thus be observed as a series of $P$ peaks in the
2379: conductance as a function of gate voltage, separated by broad minima due to
2380: the minigaps. Such an observation would demonstrate phase coherence over
2381: the entire length $L=Pa$ of the finite superlattice and would constitute
2382: conclusive evidence of a miniband. The conductance of a finite 1D superlattice in a narrow 2DEG channel in the ballistic transport regime has been
2383: investigated theoretically by Ulloa et al.\cite{ref249} Similar physics may be studied in
2384: the quantum Hall effect regime, where the experimental requirements are
2385: considerably relaxed. A successful experiment of this type was recently
2386: performed by Kouwenhoven et al.\cite{ref250} (see Section \ref{sec22}).
2387:
2388: Weak-field magnetotransport in a 2D periodic potential (a grid) has been
2389: studied by Ferry et al.\cite{ref251,ref252} and by Smith et al.\cite{ref244} Both groups reported
2390: oscillatory structure in the magnetoconductance, suggestive of an
2391: Aharonov-Bohm effect with periodicity $\Delta B=h/eS$, where $S$ is the area of a
2392: unit cell of the ``lattice.'' In strong magnetic fields no such oscillations are
2393: found. A similar suppression of the Aharonov-Bohm effect in strong fields is
2394: found in single rings, as discussed in detail in Section \ref{sec21a}. Magnetotransport
2395: in a 1D periodic potential is the subject of the next subsection.
2396:
2397: \subsubsection{\label{sec11b} Guiding-center-drift resonance}
2398:
2399: The influence of a magnetic field on transport through layered superlattices\cite{ref253} has been studied mainly in the regime where the (first) energy gap
2400: $\Delta E\sim 100\,\mathrm{meV}$ exceeds the Landau level spacing $\hbar\omega_{\mathrm{c}}$ ($1.7\, \mathrm{meV}/\mathrm{T}$ in GaAs).
2401: The magnetic field does not easily induce transitions between different
2402: minibands in this regime. Magnetotransport through lateral superlattices is
2403: often in the opposite regime $\hbar\omega_{\mathrm{c}}\gg \Delta E$, because of the relatively large
2404: periodicity $(a\sim 300\,\mathrm{nm})$ and small amplitude $(V_{0}\sim 1\,\mathrm{meV})$ of the periodic
2405: potential. The magnetic field now changes qualitatively the structure of the
2406: energy bands, which becomes richly complex in the case of a 2D periodic
2407: potential.\cite{ref54} Much of this structure, however, is not easily observed, and the
2408: experiments discussed in this subsection involve mostly the {\it classical\/} effect of
2409: a weak periodic potential on motion in a magnetic field.
2410:
2411: \begin{figure}
2412: \centerline{\includegraphics[width=6cm]{figures/fig36a}}
2413:
2414: \centerline{\includegraphics[width=6cm]{figures/fig36b}}
2415: \caption{
2416: A brief illumination of a GaAs-AlGaAs heterostructure with an interference pattern due to two laser beams (black arrows) leads to a persistent periodic variation in the concentration of ionized donors in the AlGaAs, thereby imposing a weak periodic potential on the 2DEG. The resulting spatial variation of the electron density in the 2DEG is indicated schematically. (b) Experimental arrangement used to produce a modulated 2DEG by means of the ``holographic illumination'' of (a). The sample layout shown allows measurements of the resistivity parallel and perpendicular to the equipotentials. Taken from D. Weiss et al., in ``High Magnetic Fields in Semiconductor Physics II'' (G. Landwehr, ed.). Springer, Berlin, 1989.
2417: \label{fig36}
2418: }
2419: \end{figure}
2420:
2421: Weiss et al.\cite{ref255,ref256} used an ingenious technique to impose a weak periodic
2422: potential on a 2DEG in a GaAs-AlGaAs heterostructure. They exploit the
2423: well-known persistent ionization of donors in AlGaAs after brief illumination
2424: at low temperatures. For the illumination, two interfering laser beams are
2425: used, which generate an interference pattern with a period depending on the
2426: wavelength and on the angle of incidence of the two beams. This technique,
2427: known as {\it holographic illumination}, is illustrated in Fig.\ \ref{fig36}. The interference
2428: pattern selectively ionizes Si donors in the AlGaAs, leading to a weak
2429: periodic modulation $V(y)$ of the bottom of the conduction band in the 2DEG,
2430: which persists at low temperatures if the sample is kept in the dark. The
2431: sample layout, also shown in Fig.\ \ref{fig36}, allows independent measurements of
2432: the resistivity $\rho_{yy}(\equiv\rho_{\perp})$, perpendicular to, and $\rho_{xx}(\equiv\rho_{||})$ parallel to the
2433: grating. In Fig.\ \ref{fig37} we show experimental results of Weiss et al.\cite{ref255} for the
2434: magnetoresistivity of a 1D lateral superlattice $(a=382\,\mathrm{nm}$). In a zero
2435: magnetic field, the resistivity tensor $\mathbf{\rho}$ is approximately isotropic: $\rho_{\perp}$ and $\rho_{||}$
2436: are indistinguishable experimentally (see Fig.\ \ref{fig37}). This indicates that the
2437: amplitude of $V(y)$ is much smaller than the Fermi energy $E_{\mathrm{F}}=11\,\mathrm{meV}$. On
2438: application of a small magnetic field $B(\lesssim 0.4\,\mathrm{T})$ perpendicular to the 2DEG,
2439: a large oscillation periodic in $1/B$ develops in the resistivity $\rho_{\perp}$ for current
2440: flowing perpendicular to the potential grating. The resistivity is now strongly
2441: anisotropic, showing only weak oscillations in $\rho_{||}$ (current parallel to the
2442: potential grating). In appearance, the oscillations resemble the Shubnikov-De Haas oscillations at higher fields, but their different periodicity and much
2443: weaker temperature dependence point to a different origin.
2444:
2445: \begin{figure}
2446: \centerline{\includegraphics[width=8cm]{figures/fig37}}
2447: \caption{
2448: Solid curves: Magnetic field dependence of the resistivity $\rho_{\perp}$ for current flowing perpendicular to a potential grating. The experimental curve is the measurement of Weiss et al.\cite{ref255} the theoretical curve follows from the guiding-center-drift resonance. Note the phase shift of the oscillations, indicated by the arrows at integer $2l_{\rm cycl}/a$. The potential grating has periodicity $a =382\,{\rm nm}$ and is modeled by a sinusoidal potential with root-mean-square amplitude of $\epsilon = 1.5$\%
2449: of the Fermi energy; The mean free path in the 2DEG is $12\,\mu{\rm m}$, much larger than $a$. The dash-dotted curve is the experimental resistivity $\rho_{||}$ for current flowing parallel to the potential grating, as measured by Weiss et al. Taken from C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 62}, 2020 (1989).
2450: \label{fig37}
2451: }
2452: \end{figure}
2453:
2454: The effect was not anticipated theoretically, but now a fairly complete and
2455: consistent theoretical picture has emerged from several analyses.\cite{ref111,ref227,ref257,ref258,ref259} The strong oscillations in $\rho_{\perp}$ result from a resonance\cite{ref111}
2456: between the periodic cyclotron orbit motion and the oscillating $\mathbf{E}\times \mathbf{B}$ drift of
2457: the orbit center induced by the electric field $\mathbf{E}\equiv-\nabla V$. Such {\it guiding-center-drift resonances\/} are known from plasma physics,\cite{ref260} and the experiment by
2458: Weiss et al.\ appears to be the first observation of this phenomenon in the
2459: solid state. Magnetic quantization is not essential for these strong oscillations, but plays a role in the transition to the Shubnikov-De Haas
2460: oscillations at higher fields and in the weak oscillations in $\rho_{||}$.\cite{ref227,ref259} A
2461: simplified physical picture of the guiding-center-drift resonance can be
2462: obtained as follows.\cite{ref111}
2463:
2464: \begin{figure}
2465: \centerline{\includegraphics[width=8cm]{figures/fig38}}
2466: \caption{
2467: (a) Potential grating with a cyclotron orbit superimposed. When the electron is close to the two extremal points $Y\pm l_{\rm cycl}$, the guiding center at $Y$ acquires an ${\bf E} \times {\bf B}$ drift in the direction of the arrows. (The drift along nonextremal parts of the orbit averages out, approximately.) A resonance occurs if the drift at one extremal point reinforces the drift at the other, as shown. (b) Numerically calculated trajectories for a sinusoidal potential ($\epsilon = 0.015$). The horizontal lines are equipotentials at integer $y/a$. On resonance ($2l_{\rm cycl}/a = 6.25$) the guiding center drift is maximal; off resonance ($2_{\rm lcycl}/a = 5.75$) the drift is negligible. Taken from C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 62}, 2020 (1989).
2468: \label{fig38}
2469: }
2470: \end{figure}
2471:
2472: The guiding center $(X, Y)$ of an electron at position $(x,y)$ having velocity
2473: $(v_{x},v_{y})$ is given by $X=x-v_{y}/\omega_{\mathrm{c}}$, $Y=y+v_{x}/\omega_{\mathrm{c}}$. The time derivative of the
2474: guiding center is $x=E(y)/B,\dot{Y}=0$, so its motion is parallel to the $x$-axis.
2475: This is the ${\bf E} \times {\bf B}$ drift. In the case of a strong magnetic field and a slowly
2476: varying potential $(l_{\mathrm{cycl}}\ll a)$, one may approximate $E(y)\approx E(Y)$ to close the
2477: equations for $\dot{X}$ and $\dot{Y}$. This so-called adiabatic approximation cannot be
2478: made in the weak-field regime $(l_{\mathrm{cycl}}\gtrsim a)$ of interest here. We consider the case
2479: of a weak potential, such that $eV_{\mathrm{rms}}/E_{\mathrm{F}}\equiv\epsilon\ll 1$, with $V_{\mathrm{rms}}$ the root mean
2480: square of $V(y)$. The guiding center drift in the $x$-direction is then simply
2481: superimposed on the unperturbed cyclotron motion. Its time average $v_{\mathrm{drift}}$ is
2482: obtained by integrating the electric field along the orbit
2483: \be
2484: v_{\mathrm{drift}}(Y)=(2 \pi B)^{-1}\int_{0}^{2\pi}d\phi\,E(Y+l_{\mathrm{cycl}}\sin\phi). \label{eq11.1}
2485: \ee
2486: For $l_{\mathrm{cycl}}\gg a$ the field oscillates rapidly, so only the drift acquired close to the
2487: two extremal points $Y\pm l_{\mathrm{cycl}}$ does not average out. It follows that $v_{\mathrm{drift}}$ is
2488: large or small depending on whether $E(Y+l_{\mathrm{cycl}})$ and $E(Y-l_{\mathrm{cycl}})$ have the
2489: same sign or opposite sign (see Fig.\ \ref{fig38}). For a sinusoidal potential
2490: $V(y)=\sqrt{2}V_{\mathrm{rms}}\sin(2\pi y/a)$, one easily calculates by averaging over $Y$ that, for
2491: $l_{\mathrm{cycl}}\gg a$, the mean square drift is
2492: \be
2493: \langle v_{\mathrm{drift}}^{2}\rangle=(v_{\mathrm{F}}\epsilon)^{2}\left(\frac{l_{\mathrm{cycl}}}{a}\right)\cos^{2}\left(\frac{2\pi l_{\mathrm{cycl}}}{a}-\frac{\pi}{4}\right). \label{eq11.2}
2494: \ee
2495: The guiding center drift by itself leads, for $l_{\mathrm{cycl}}\ll l$, to 1D diffusion with
2496: diffusion coefficient $\delta D$ given by
2497: \be
2498: \delta D=\int_{0}^{\infty}\langle v_{\mathrm{drift}}^{2}\rangle \mathrm{e}{}^{-t/\mathrm{t}}dt=\tau\langle v_{\mathrm{drift}}^{2}\rangle. \label{eq11.3}
2499: \ee
2500: The term $\delta D$ is an additional contribution to the $xx$-element of the unperturbed diffusion tensor $\mathbf{D}^{0}$ given by $D_{xx}^{0}=D_{yy}^{0}=D_{0}$, $D_{xy}^{0}=-D_{xy}^{0}=-\omega_{\mathrm{c}}\tau D_{0}$, with $D_{0} \equiv\frac{1}{2}\tau v_{\mathrm{F}}^{2}[1+(\omega_{\mathrm{c}}\tau)^{2}]^{-1}$ (cf.\ Section \ref{sec4c}). At this point we
2501: assume that for $l_{\mathrm{cycl}}\ll l$ the contribution $\delta D$ from the guiding center drift is
2502: the dominant effect of the potential grating on the diffusion tensor ${\mathbf D}$. A
2503: justification of this assumption requires a more systematic analysis of the
2504: transport problem, which is given in Ref.\ \onlinecite{ref111}. Once $\mathbf{D}$ is known, the resistivity
2505: tensor $\mathbf{\rho}$ follows from the Einstein relation $\mathbf{\rho}=\mathbf{D}^{-1}/e^{2}\rho(E_{\mathrm{F}})$, with $\rho(E_{\mathrm{F}})$ the
2506: 2D density of states (cf.\ Section \ref{sec4b}). Because of the large off-diagonal
2507: components of $\mathbf{D}^{0}$, an additional contribution $\delta D$ to $D_{xx}^{0}$ modifies predominantly $\rho_{yy}\equiv\rho_{\perp}$. To leading order in $\epsilon$, one finds that
2508: \be
2509: \frac{\rho_{\perp}}{\rho_{0}}=1+2\epsilon^{2}\left(\frac{l^{2}}{al_{\mathrm{cycl}}}\right)\cos^{2}\left(2\pi\frac{l_{\mathrm{cycl}}}{a}-\frac{\pi}{4}\right), \label{eq11.4}
2510: \ee
2511: with $\rho_{0}=m/n_{\mathrm{s}}e^{2}\tau$ the unperturbed resistivity. A rigorous solution\cite{ref111} of the
2512: Boltzmann equation (for a $B$-independent scattering time) confirms this
2513: simple result in the regime $a\ll l_{\mathrm{cycl}}\ll l$ and is shown in Fig.\ \ref{fig37} to be in quite
2514: good agreement with the experimental data of Weiss et al.\cite{ref255} Similar
2515: theoretical results have been obtained by Gerhardts et al.\cite{ref257} and by Winkler
2516: et al.\cite{ref258} (using an equivalent quantum mechanical formulation; see below).
2517:
2518: As illustrated by the arrows in Fig.\ \ref{fig37}, the maxima in $\rho_{\perp}$ are not at integer
2519: $2l_{\mathrm{cycl}}/a$, but shifted somewhat toward lower magnetic fields. This phase shift is
2520: a consequence of the finite extension of the segment of the orbit around the
2521: extremal points $Y\pm l_{\mathrm{cycl}}$, which contributes to the guiding center drift
2522: $v_{\mathrm{drift}}(Y)$. Equation (\ref{eq11.4}) implies that $\rho_{\perp}$ in a sinusoidal potential grating has
2523: minima and maxima at approximately
2524: \begin{eqnarray}
2525: &&2l_{\mathrm{cycl}}/a\;\; ({\rm minima})=n - {\textstyle\frac{1}{4}},\nonumber\\
2526: &&2l_{\mathrm{cycl}}/a\;\; ({\rm maxima})=n+{\textstyle\frac{1}{4}} - {\rm order}(1/n), \label{eq11.5}
2527: \end{eqnarray}
2528: with $n$ an integer. We emphasize that the phase shift is not universal, but
2529: depends on the functional form of $V(y)$. The fact that the experimental phase
2530: shift in Fig.\ \ref{fig37} agrees so well with the theory indicates that the actual
2531: potential grating in the experiment of Weiss et al.\ is well modeled by a
2532: sinusoidal potential. The maxima in $\rho_{\perp}/\rho_{0}$ have amplitude $\epsilon^{2}(l^{2}/al_{\mathrm{cycl}})$, which
2533: for a large mean free path $l$ can be of order unity, even if $\epsilon\ll 1$. The guiding-center-drift resonance thus explains the surprising experimental finding that a
2534: periodic modulation of the Fermi velocity of order $10^{-2}$ can double the
2535: resistivity.
2536:
2537: At low magnetic fields the experimental oscillations are damped more
2538: rapidly than the theory would predict, and, moreover, an unexplained
2539: positive magnetoresistance is observed around zero field in $\rho_{\perp}$ (but not in $\rho_{||}$).
2540: Part of this disagreement may be due to nonuniformities in the potential
2541: grating, which become especially important at low fields when the cyclotron
2542: orbit overlaps many modulation periods. At high magnetic fields $B\gtrsim 0.4\,\mathrm{T}$
2543: the experimental data show the onset of Shubnikov-De Haas oscillations,
2544: which are a consequence of oscillations in the scattering time $\tau$ due to Landau
2545: level quantization (cf.\ Section \ref{sec4c}). This effect is neglected in the semiclassical
2546: analysis, which assumes a constant scattering time.
2547:
2548: The quantum mechanical $B$-dependence of $\tau$ also leads to weak-field
2549: oscillations in $\rho_{||}$, with the same periodicity as the oscillations in $\rho_{\perp}$ discussed
2550: earlier, but of much smaller amplitude and shifted in phase (see Fig.\ \ref{fig37}, where
2551: a maximum in the experimental $\sigma_{||}$ around $0.3\, \mathrm{T}$ lines up with a minimum in
2552: $\rho_{\perp})$. These small antiphase oscillations in $\rho_{||}$ were explained by Vasilopoulos
2553: and Peeters\cite{ref227} and by Gerhardts and Zhang\cite{ref259} as resulting from oscillations
2554: in $\tau$ due to the oscillatory Landau bandwidth. The Landau levels
2555: $E_{n}=(n-\frac{1}{2})\hbar\omega_{\mathrm{c}}$ broaden into a band of finite width in a periodic potential.\cite{ref261}
2556: This Landau band is described by a dispersion law $E_{n}(k)$, where the wave
2557: number $k$ is related to the classical orbit center $(X, Y)$ by $k=YeB/\hbar$ (cf.\ the
2558: similar relation in Section \ref{sec12}). The classical guiding-center-drift resonance
2559: can also be explained in these quantum mechanical terms, if one so desires, by
2560: noticing that the bandwidth of the Landau levels is proportional to the root-mean-square average of $v_{\mathrm{drift}}=dE_{n}(k)/\hbar dk$. A maximal bandwidth thus
2561: corresponds to a maximal guiding center drift and, hence, to a maximal $\rho_{\perp}$. A
2562: maximum in the bandwidth also implies a minimum in the density of states at
2563: the Fermi level and, hence, a maximum in $\tau$ [Eq.\ (\ref{eq4.28})]. A maximal
2564: bandwidth thus corresponds to a minimal $\rho_{||}$, whereas the $B$-dependence of $\tau$
2565: can safely by neglected for the oscillations in $\rho_{\perp}$ (which are dominated by the
2566: classical guiding-center-drift resonance).
2567:
2568: In a 2D periodic potential (a grid), the guiding center drift dominates the
2569: magnetoresistivity in both diagonal components of the resistivity tensor.
2570: Classically, the effect of a weak periodic potential $V(x, y)$ on $\rho_{xx}$ and $\rho_{yy}$
2571: decouples if $V(x, y)$ is separable into $V(x, y)=f(x)+g(y)$. For the 2D
2572: sinusoidal potential $V(x, y)\propto\sin(2\pi x/a)+\sin(2\pi y/b)$, one finds that the effect
2573: of the grid is simply a superposition of the effects for two perpendicular
2574: gratings of periods $a$ and $b$. (No such decoupling occurs quantum mechanically.\cite{ref254}) Experiments by Alves et al.\cite{ref262} and by Weiss et al.\cite{ref263} confirm this
2575: expectation, except for a disagreement in the phase of the oscillations. As
2576: noted, however, the phase is not universal but depends on the form of the
2577: periodic potential, which need not be sinusoidal.
2578:
2579: Because of the predominance of the classical guiding-center-drift resonance in a weak periodic potential, magnetotransport experiments are not
2580: well suited to study miniband structure in the density of states. Magnetocapacitance measurements\cite{ref256,ref264,ref265} are a more direct means of investigation, but
2581: somewhat outside the scope of this review.
2582:
2583: \section{\label{secIII} Ballistic transport}
2584:
2585: \subsection{\label{sec12} Conduction as a transmission problem}
2586:
2587: In the ballistic transport regime, it is the scattering of electrons at the
2588: sample boundaries which limits the current, rather than impurity scattering.
2589: The canonical example of a ballistic conductor is the point contact illustrated
2590: in Fig.\ \ref{fig7}c. The current $I$ through the narrow constriction in response to a
2591: voltage difference $V$ between the wide regions to the left and right is {\it finite\/}
2592: even in the absence of impurities, because electrons are scattered back at the
2593: entrance of the constriction. The {\it contact conductance\/} $G=I/V$ is proportional
2594: to the constriction width but independent of its length. One cannot therefore
2595: describe the contact conductance in terms of a local conductivity, as one can
2596: do in the diffusive transport regime. Consequently, the Einstein relation (\ref{eq4.10})
2597: between the conductivity and the diffusion constant at the Fermi level, of
2598: which we made use repeatedly in Section \ref{secII}, is not applicable in that form to
2599: determine the contact conductance. The Landauer formula is an alternative
2600: relation between the conductance and a Fermi level property of the sample,
2601: without the restriction to diffusive transport. We discuss this formulation of
2602: conduction in Section \ref{sec12b}. The Landauer formula expresses the conductance
2603: in terms of transmission probabilities of propagating modes at the Fermi
2604: level (also referred to as {\it quantum channels\/} in this context). Some elementary
2605: properties of the modes are summarized in Section \ref{sec12a}.
2606:
2607: \subsubsection{\label{sec12a} Electron waveguide}
2608:
2609: We consider a conducting channel in a 2DEG (an ``electron waveguide''),
2610: defined by a lateral confining potential $V(x)$, in the presence of a perpendicular magnetic field $B$ (in the $z$-direction). In the Landau gauge
2611: $\mathbf{A}=(0, Bx, 0)$ the hamiltonian has the form
2612: \be
2613: {\cal H}=\frac{p_{x}^{2}}{2m}+\frac{(p_{y}+eBx)^{2}}{2m}+V(x) \label{eq12.1}
2614: \ee
2615: for a single spin component (cf.\ Section \ref{sec10a}). Because the canonical
2616: momentum $p_{y}$ along the channel commutes with ${\cal H}$, one can diagonalize $p_{y}$
2617: and ${\cal H}$ simultaneously. For each eigenvalue $\hbar k$ of $p_{y}$, the hamiltonian (\ref{eq12.1})
2618: has a discrete spectrum of energy eigenvalues $E_{n}(k), n=1,2, \ldots$, corresponding to eigenfunctions of the form
2619: \be
2620: |n, k\rangle=\Psi_{n,k}(x)\mathrm{e}^{iky}. \label{eq12.2}
2621: \ee
2622: In waveguide terminology, the index $n$ labels the modes, and the dependence
2623: of the energy (or ``frequency'') $E_{n}(k)$ on the wave number $k$ is the dispersion
2624: relation of the $n$th mode. A propagating mode at the Fermi level has cutoff
2625: frequency $E_{n}(0)$ below $E_{\mathrm{F}}$. The wave function (\ref{eq12.2}) is the product of a
2626: transverse amplitude profile $\Psi_{n,k}(x)$ and a longitudinal plane wave $e^{iky}$. The
2627: average velocity $v_{n}(k)$ along the channel in state $|n, k\rangle$ is the expectation value
2628: of the $y$-component of the velocity operator $\mathbf{p}+e\mathbf{A}$:
2629: \begin{eqnarray}
2630: v_{n}(k)&\equiv&\langle n, k| \frac{p_{y}+eA_{y}}{m}|n, k\rangle\nonumber\\
2631: &=&\langle n, k| \frac{\partial{\cal H}}{\partial p_{y}}|n, k \rangle=\frac{dE_{n}(k)}{\hbar dk}. \label{eq12.3}
2632: \end{eqnarray}
2633: For a zero magnetic field, the dispersion relation $E_{n}(k)$ has the simple form
2634: (\ref{eq4.3}). The {\it group velocity\/} $v_{n}(k)$ is then simply equal to the velocity $\hbar k/m$
2635: obtained from the canonical momentum. This equality no longer holds in the
2636: presence of a magnetic field, because the canonical momentum contains an
2637: extra contribution from the vector potential. The dispersion relation in a
2638: nonzero magnetic field was derived in Section \ref{sec10a} for a parabolic confining
2639: potential $V(x)= \frac{1}{2}m\omega_{0}^{2}x^{2}$. From Eq.\ (\ref{eq10.5}) one calculates a group velocity
2640: $\hbar k/M$ that is smaller than $\hbar k/m$ by a factor of $1+(\omega_{\mathrm{c}}/\omega_{0})^{2}$.
2641:
2642: \begin{figure}
2643: \centerline{\includegraphics[width=8cm]{figures/fig39}}
2644: \caption{
2645: Energy-orbit center phase space. The two parabolas divide the space into four regions, which correspond to different types of classical trajectories in a magnetic field (clockwise from left: skipping orbits on one edge, traversing trajectories, skipping orbits on the other edge, and cyclotron orbits). The shaded region is forbidden. The region at the upper center contains traversing trajectories moving in both directions, but only one region is shown for clarity. Taken from C. W. J. Beenakker et al., Superlattices and Microstructures {\bf 5}, 127 (1989).
2646: \label{fig39}
2647: }
2648: \end{figure}
2649:
2650: Insight into the nature of the wave functions in a magnetic field can be
2651: obtained from the correspondence with classical trajectories. These are most
2652: easily visualized in a square-well confining potential, as we now discuss
2653: (following Ref.\ \onlinecite{ref266}). The position $(x, y)$ of an electron on the circle with center
2654: coordinates $(X, Y)$ can be expressed in terms of its velocity $ \mathbf{v}$ by
2655: \be
2656: x=X+v_{y}/\omega_{\mathrm{c}},\;\; y=Y-v_{x}/\omega_{\mathrm{c}}, \label{eq12.4}
2657: \ee
2658: with $\omega_{\mathrm{c}}\equiv eB/m$ the cyclotron frequency. The cyclotron radius is $(2mE)^{1/2}/eB$,
2659: with $E \equiv\frac{1}{2}mv^{2}$ the energy of the electron. Both the energy $E$ and the
2660: separation $X$ of the orbit center from the center of the channel are constants
2661: of the motion. The coordinate $Y$ of the orbit center parallel to the channel
2662: walls changes on each specular reflection. One can classify a trajectory as a
2663: cyclotron orbit, skipping orbit, or traversing trajectory, depending on
2664: whether the trajectory collides with zero, one, or both channel walls. In $(X, E)$
2665: space these three types of trajectories are separated by the two parabolas
2666: $(X\pm W/2)^{2}=2mE(eB)^{-2}$ (Fig.\ \ref{fig39}). The quantum mechanical dispersion
2667: relation $E_{n}(k)$ can be drawn into this classical ``phase diagram'' because of the
2668: correspondence $k=-XeB/h$.This correspondence exists because both $k$ and
2669: $X$ are constants of the motion and it follows from the fact that the component
2670: $\hbar k$ along the channel of the canonical momentum $\mathbf{p}=m \mathbf{v}-e\mathbf{A}$ equals
2671: \be
2672: \hbar k=mv_{y}-eA_{y}=mv_{y}-eBx=-eBx \label{eq12.5}
2673: \ee
2674: in the Landau gauge.
2675:
2676: \begin{figure}
2677: \centerline{\includegraphics[width=6cm]{figures/fig40a}}
2678:
2679: \centerline{\includegraphics[width=6cm]{figures/fig40b}}
2680: \caption{
2681: Dispersion relation $E_{n}(k)$, calculated for parameters: (a) $W = 100\,{\rm nm}$, $B = 1\,{\rm T}$; (b) $W = 200\,{\rm nm}$, $B = 1.5\,{\rm T}$. The horizontal line at 17 meV indicates the Fermi energy. The shaded area is the region of classical skipping orbits and is bounded by the two parabolas shown in Fig.\ \ref{fig39} (with the correspondence $k = - X eB/\hbar$). Note that in (a) edge states coexist at the Fermi level with states interacting with both boundaries ($B < B_{\rm crit} \equiv 2\hbar k_{\rm F}/eB$), while in (b) all states at the Fermi level interact with one boundary only ($B > B_{\rm crit}$). Taken from C. W. J. Beenakker et al., Superlattices and Microstructures {\bf 5}, 127 (1989).
2682: \label{fig40}
2683: }
2684: \end{figure}
2685:
2686: In Fig.\ \ref{fig40} we show $E_{n}(k)$ both in weak and in strong magnetic fields,
2687: calculated\cite{ref266} for typical parameter values from the Bohr-Sommerfeld
2688: quantization rule discussed here. The regions in phase space occupied by
2689: classical skipping orbits are shaded. The unshaded regions contain cyclotron
2690: orbits (at small $E$) and traversing trajectories (at larger $E$) (cf.\ Fig.\ \ref{fig39}). The
2691: {\it cyclotron orbits\/} correspond quantum mechanically to states in {\it Landau levels}.
2692: These are the flat portions of the dispersion relation at energies
2693: $E_{n}=(n- \frac{1}{2})\hbar\omega_{\mathrm{c}}$. The group velocity (\ref{eq12.3}) is zero in a Landau level, as one
2694: would expect from the correspondence with a circular orbit. The {\it traversing trajectories\/} correspond to states in {\it magnetoelectric subbands}, which interact
2695: with both the opposite channel boundaries and have a nonzero group
2696: velocity. The {\it skipping orbits\/} correspond to {\it edge states}, which interact with a
2697: single boundary only. The two sets of edge states (one for each boundary) are
2698: disjunct in $(k, E)$ space. Edge states at opposite boundaries move in opposite
2699: directions, as is evident from the correspondence with skipping orbits or by
2700: inspection of the slope of $E_{n}(k)$ in the two shaded regions in Fig.\ \ref{fig40}.
2701:
2702: If the Fermi level lies between two Landau levels, the states at the Fermi
2703: level consist only of edge states if $B>B_{\mathrm{crit}}$, as in Fig.\ \ref{fig40}b. The ``critical'' field
2704: $B_{\mathrm{crit}}=2\hbar k_{\rm F}/eW$ is obtained from the classical correspondence by requiring
2705: that the channel width $W$ should be larger than the cyclotron diameter
2706: $2\hbar k_{\rm F}/eB$ at the Fermi level. This is the same characteristic field that played a
2707: role in the discussion of magneto size effects in Sections \ref{sec5} and \ref{sec10}. At fields
2708: $B<B_{\mathrm{crit}}$, as in Fig.\ \ref{fig40}a, edge states coexist at the Fermi level with
2709: magnetoelectric subbands. In still lower fields $B<B_{\mathrm{thres}}$ {\it all\/} states at the
2710: Fermi level interact with both edges. The criterion for this is that $W$ should be
2711: smaller than the transverse wavelength\cite{ref267} $\lambda_{\mathrm{t}}=(\hbar/2k_{\mathrm{F}}eB)^{1/3}$ of the edge
2712: states, so the threshold field $B_{\mathrm{thres}}\sim \hbar/ek_{\mathrm{F}}W^{3}$. Contrary to initial expectations,\cite{ref268} this lower characteristic field does not appear to play a decisive role
2713: in magneto size effects.
2714:
2715: A quick way to arrive at the dispersion relation $E_{n}(k)$, which is sufficiently
2716: accurate for our purposes, is to apply the Bohr-Sommerfeld quantization
2717: rule\cite{ref80,ref269} to the classical motion in the $x$-direction:
2718: \be
2719: \frac{1}{\hbar}\oint p_{x}dx+\gamma=2\pi n,\;\;n=1,2, \ldots. \label{eq12.6}
2720: \ee
2721: The integral is over one period of the motion. The phase shift $\gamma$ is the sum of
2722: the phase shifts acquired at the two turning points of the projection of the
2723: motion on the $x$-axis. The phase shift upon reflection at the boundary is $\pi$, to
2724: ensure that incident and reflected waves cancel (we consider an infinite
2725: barrier potential at which the wave function vanishes). The other turning
2726: points (at which $v_{x}$ varies smoothly) have a phase shift of $-\pi/2$.\cite{ref93} Consequently, for a traversing trajectory $\gamma=\pi+\pi=0$ $(\mathrm{mod}\,2\pi)$, for a skipping
2727: orbit $\gamma=\pi-\pi/2=\pi/2$, and for a cyclotron orbit $\gamma=-\pi/2-\pi/2=\pi$
2728: $(\mathrm{mod}\,2\pi)$. In the Landau gauge one has $p_{x}=mv_{x}=eB(Y-y)$, so Eq.\ (\ref{eq12.6})
2729: takes the form
2730: \be
2731: B \oint(Y-y)dx=\frac{h}{e}(n-\frac{\gamma}{2\pi}). \label{eq12.7}
2732: \ee
2733: This quantization condition has the appealing geometrical interpretation
2734: that $n-\gamma/2\pi$ flux quanta $h/e$ are contained in the area bounded by the
2735: channel walls and a circle of cyclotron radius $(2mE)^{1/2}/eB$ centered at $X$ (cf.\
2736: Fig.\ \ref{fig41}). It is now straightforward to find for each integer $n$ and coordinate $X$
2737: the energy $E$ that satisfies this condition. The dispersion relation $E_{n}(k)$ then
2738: follows on identifying $k=-XeB/h$, as shown in Fig.\ \ref{fig40}.
2739:
2740: \begin{figure}
2741: \centerline{\includegraphics[width=6cm]{figures/fig41a}}
2742:
2743: \centerline{\includegraphics[width=6cm]{figures/fig41b}}
2744: \caption{
2745: Classical trajectories in a magnetic field. The flux through the shaded area is quantized according to the Bohr-Sommerfeld quantization rule (\ref{eq12.7}). The shaded area in (b) is bounded by the channel walls and the circle formed by the continuation (dashed) of one circular arc of the traversing trajectory.
2746: \label{fig41}
2747: }
2748: \end{figure}
2749:
2750: The total number $N$ of propagating modes at energy $E$ is determined by
2751: the maximum flux $\Phi_{\rm max}$ contained in an area bounded by the channel walls
2752: and a circle of radius $(2mE)^{1/2}/eB$, according to $N=\mathrm{Int}[e\Phi_{\rm max}/h+\gamma/2\pi]$.
2753: Note that a maximal enclosed flux is obtained by centering the circle on the
2754: channel axis. Some simple geometry then leads to the result\cite{ref80} (\ref{eq10.8}), which is
2755: plotted together with that for a parabolic confinement in Fig.\ \ref{fig31}. Equation
2756: (\ref{eq10.8}) has a discontinuity at magnetic fields for which the cyclotron diameter
2757: equals the channel width, due to the jump in the phase shift $\gamma$ as one goes
2758: from a cyclotron orbit to a traversing trajectory. This jump is an artifact of
2759: the present semiclassical approximation in which the extension of the wave
2760: function beyond the classical orbit is ignored. Since the discontinuity in $N$ is
2761: at most $\pm 1$, it is unimportant in many applications. More accurate formulas
2762: for the phase shift $\gamma$, which smooth out the discontinuity, have been derived in
2763: Ref.\ \onlinecite{ref270}. If necessary, one can also use more complicated but exact solutions
2764: of the Schr\"{o}dinger equation, which are known.\cite{ref267}
2765:
2766: \subsubsection{\label{sec12b} Landauer formula}
2767:
2768: Imagine two wide electron gas reservoirs having a slight difference $\delta n$ in
2769: electron density, which are brought into contact by means of a narrow
2770: channel, as in Fig.\ \ref{fig42}a. A diffusion current $J$ will flow in the channel, carried
2771: by electrons with energies between the Fermi energies $E_{\mathrm{F}}$ and $E_{\mathrm{F}}+\delta\mu$ in the
2772: low- and high-density regions. For small $\delta n$, one can write for the Fermi
2773: energy difference (or chemical potential difference) $\delta\mu=\delta n/\rho(E_{\mathrm{F}})$, with $\rho(E_{\mathrm{F}})$
2774: the density of states at $E_{\mathrm{F}}$ in the reservoir (cf.\ Section \ref{sec4a}). The diffusion
2775: constant (or ``diffusance'')\cite{ref4} $\tilde{D}$ is defined by $J=\tilde{D}\delta n$ and is related to the
2776: conductance $G$ by
2777: \be
2778: G=e^{2}\rho(E_{\mathrm{F}})\tilde{D}. \label{eq12.8}
2779: \ee
2780: Equation (\ref{eq12.8}) generalizes the Einstein relation (\ref{eq4.10}) and is derived in a
2781: completely analogous way [by requiring that the sum of drift current $GV/e$
2782: and diffusion current $\tilde{D}\delta n$ be zero when the sum of the electrostatic potential
2783: difference $eV$ and chemical potential difference $\delta n/\rho(E_{\mathrm{F}})$ vanishes].
2784:
2785: \begin{figure}
2786: \centerline{\includegraphics[width=6cm]{figures/fig42a}}
2787:
2788: \centerline{\includegraphics[width=6cm]{figures/fig42b}}
2789: \caption{
2790: (a) Narrow channel connecting two wide electron gas regions, having a chemical potential difference $\delta\mu$. (b) Schematic dispersion relation in the narrow channel. Left-moving states ($k > 0$) are filled up to chemical potential $E_{\rm F}$, right-moving states up to $E_{\rm F} + \delta\mu$ (solid dots). Higher-lying states are unoccupied (open dots).
2791: \label{fig42}
2792: }
2793: \end{figure}
2794:
2795: Since the diffusion current (at low temperatures) is carried by electrons
2796: within a narrow range $\delta\mu$ above $E_{\mathrm{F}}$, the diffusance can be expressed in terms
2797: of Fermi level properties of the channel (see below). The Einstein relation
2798: (\ref{eq12.8}) then yields the required Fermi level expression of the conductance. This
2799: by no means implies that the drift current induced by an electrostatic
2800: potential difference is carried entirely by electrons at the Fermi energy. To the
2801: contrary, all electrons (regardless of their energy) acquire a nonzero drift
2802: velocity in an electric field. This point has been the cause of some confusion in
2803: the literature on the quantum Hall effect, so we will return to it in Section
2804: \ref{sec18c}. In the following we will refer to electrons at the Fermi energy as
2805: ``current-carrying electrons'' and show that ``the current in the channel is
2806: shared equally among the modes at the Fermi level.'' These and similar
2807: statements should be interpreted as referring to the diffusion problem, where
2808: the current is induced by density differences without an electric field. We
2809: make no attempt here to evaluate the distribution of current in response to an
2810: electric field in a system of uniform density. That is a difficult problem, for
2811: which one has to determine the electric field distribution self-consistently
2812: from Poisson's and Boltzmann's equations. Such a calculation for a quantum
2813: point contact has been performed in Refs.\ \onlinecite{ref271} and \onlinecite{ref272}. Fortunately, there is
2814: no need to know the actual current distribution to determine the conductance, in view of the Einstein relation (\ref{eq12.8}). The distribution of current
2815: (and electric field) is of importance only beyond the regime of a linear relation
2816: between current and voltage. We will not venture beyond this linear response
2817: regime.
2818:
2819: To calculate the diffusance, we first consider the case of an ideal electron
2820: waveguide between the two reservoirs. By ``ideal'' it is meant that within the
2821: waveguide the states with group velocity pointing to the right are occupied
2822: up to $E_{\mathrm{F}}+\delta\mu$, while states with group velocity to the left are occupied up to
2823: $E_{\mathrm{F}}$ and empty above that energy (cf.\ Fig.\ \ref{fig42}b). This requires that an electron
2824: near the Fermi energy that is injected into the waveguide by the reservoir at
2825: $E_{\mathrm{F}}+\delta\mu$ propagates into the other reservoir without being reflected. (The
2826: physical requirements for this to happen will be discussed in Section \ref{sec13}.) The
2827: amount of diffusion current per energy interval carried by the right-moving
2828: states (with $k<0$) in a mode $n$ is the product of density of states $\rho_{n}^{-}$ and group
2829: velocity $v_{n}$. Using Eqs.\ (\ref{eq4.4}) and (\ref{eq12.3}), we find the total current $J_{n}$ carried by
2830: that mode to be
2831: \begin{eqnarray}
2832: J_{n}&=& \int_{E_{\rm F}}^{E_{\rm F}+\delta\mu}g_{\mathrm{s}}g_{\mathrm{v}}\left(2\pi\frac{dE_{n}(k)}{dk}\right)^{-1}\frac{dE_{n}(k)}{\hbar dk}\nonumber\\
2833: &=&\frac{g_{\mathrm{s}}g_{\mathrm{v}}}{h}\delta\mu, \label{eq12.9}
2834: \end{eqnarray}
2835: independent of mode index and Fermi energy. The current in the channel is
2836: shared equally among the $N$ modes at the Fermi level because of the
2837: cancellation of group velocity and density states. We will return to this
2838: {\it equipartition rule\/} in Section \ref{sec13}, because it is at the origin of the quantization\cite{ref6,ref7} of the conductance of a point contact.
2839:
2840: Scattering within the narrow channel may reflect part of the injected
2841: current back into the left reservoir. If a fraction $T_{n}$ of $J_{n}$ is transmitted to the
2842: reservoir at the right, then the total diffusion current in the channel becomes
2843: $J=(2/h) \delta\mu\sum_{n=1}^{N}T_{n}$. (Unless stated otherwise, the formulas in the remainder
2844: of this review refer to the case $g_{\mathrm{s}}=2, g_{\mathrm{v}}=1$ of twofold spin degeneracy and a
2845: single valley, appropriate for most of the experiments.) Using $\delta\mu=\delta n/\rho(E_{F})$,
2846: $J=\tilde{D}\delta n$, and the Einstein relation (\ref{eq12.8}), one obtains the result
2847: \begin{subequations}
2848: \label{eq12.10}
2849: \be
2850: G= \frac{2e^{2}}{h}\sum_{n=1}^{N}T_{n}, \label{eq12.10a}
2851: \ee
2852: which can also be written in the form
2853: \be
2854: G= \frac{2e^{2}}{h}\sum_{n,m=1}^{N}|t_{mn}|^{2}\equiv\frac{2e^{2}}{h}\mathrm{Tr}\,{\bf tt}^{\dagger}, \label{eq12.10b}
2855: \ee
2856: \end{subequations}
2857: where $T_{n}= \sum_{m=1}^{N}|t_{mn}|^{2}$ is expressed in terms of the matrix $\mathbf{t}$ of transmission
2858: probability amplitudes from mode $n$ to mode $m$. This relation between
2859: conductance and transmission probabilities at the Fermi energy is referred to
2860: as the {\it Landauer formula\/} because of Landauer's pioneering 1957 paper.\cite{ref4}
2861: Derivations of Eq.\ (\ref{eq12.10}) based on the Kubo formula of linear response
2862: theory have been given by several authors, both for zero\cite{ref143,ref273,ref274} and
2863: nonzero\cite{ref275,ref276} magnetic fields. The identification of $G$ as a {\it contact\/} conductance is due to Imry.\cite{ref1} In earlier work Eq.\ (\ref{eq12.10}) was considered
2864: suspect\cite{ref277,ref228,ref279} because it gives a {\it finite\/} conductance for an ideal (ballistic)
2865: conductor, and alternative expressions were proposed\cite{ref188,ref280,ref281,ref282} that were
2866: considered to be more realistic. (In one dimension these alternative formulas
2867: reduce to the original Landauer formula\cite{ref4} $G=(e^{2}/h)T(1-T)^{-1}$, which gives
2868: infinite conductance for unit transmission since the contact conductance $e^{2}/h$
2869: is ignored.\cite{ref1}) For historical accounts of this controversy, from two different
2870: points of view, we refer the interested reader to papers by Landauer\cite{ref283} and
2871: by Stone and Szafer.\cite{ref274} We have briefly mentioned this now-settled controversy, because it sheds some light onto why the quantization of the contact
2872: conductance had not been predicted theoretically prior to its experimental
2873: discovery in 1987.
2874:
2875: \begin{figure}
2876: \centerline{\includegraphics[width=8cm]{figures/fig43}}
2877: \caption{
2878: Generalization of the geometry of Fig.\ \ref{fig42}a to a multireservoir conductor.
2879: \label{fig43}
2880: }
2881: \end{figure}
2882:
2883: Equation (\ref{eq12.10}) refers to a {\it two-terminal\/} resistance measurement, in which
2884: the same two contacts (modeled by reservoirs in Fig.\ \ref{fig42}a) are used to drive a
2885: current through the system and to measure the voltage drop. More generally,
2886: one can consider a multireservoir conductor as in Fig.\ \ref{fig43} to model, for
2887: example, {\it four-terminal\/} resistance measurements in which the current source
2888: and drain are distinct from the voltage probes. The generalization of the
2889: Landauer formula (\ref{eq12.10}) to multiterminal resistances is due to B\"{u}ttiker.\cite{ref5} Let
2890: $T_{\alpha\rightarrow\beta}$ be the total transmission probability from reservoir $\alpha$ to $\beta$:
2891: \be
2892: T_{\alpha\rightarrow\beta}= \sum_{n=1}^{N_{\alpha}}\sum_{m=1}^{N_{\beta}}|t_{\beta\alpha,mn}|^{2}. \label{eq12.11}
2893: \ee
2894: Here $N_{\alpha}$ is the number of propagating modes in the channel (or ``lead'')
2895: connected to reservoir $\alpha$ (which in general may be different from the number
2896: $N_{\beta}$ in lead $\beta$), and $t_{\beta\alpha,mn}$ is the transmission probability amplitude from mode
2897: $n$ in lead $\alpha$ to mode $m$ in lead $\beta$. The leads are modeled by ideal electron
2898: waveguides, in the sense discussed before, so that the reservoir $\alpha$ at chemical
2899: potential $\mu_{\alpha}$ above $E_{\mathrm{F}}$ injects into lead $\alpha$ a (charge) current $(2e/h)N_{\alpha}\mu_{\alpha}$. A
2900: fraction $T_{\alpha\rightarrow\beta}/N_{\alpha}$ of that current is transmitted to reservoir $\beta$, and a fraction
2901: $T_{\alpha\rightarrow a}/N_{\alpha}\equiv R_{\alpha}/N_{\alpha}$ is reflected back into reservoir $\alpha$, before reaching one of the
2902: other reservoirs. The net current $I_{\alpha}$ in lead $\alpha$ is thus given by\cite{ref5}
2903: \be
2904: \frac{h}{2e}I_{\alpha}=(N_{\alpha}-R_{\alpha})\mu_{\alpha}-\sum_{\beta(\beta\neq \alpha)}T_{\rho\rightarrow \alpha}\mu_{\beta}. \label{eq12.12}
2905: \ee
2906: The chemical potentials of the reservoirs are related to the currents in the
2907: leads via a matrix of transmission and reflection coefficients. The sums of
2908: columns or rows of this matrix vanish:
2909: \begin{eqnarray}
2910: &&N_{\alpha}-R_{\alpha}- \sum_{\beta(\beta\neq \alpha)}T_{\alpha\rightarrow\beta}=0, \label{eq12.13}\\
2911: &&N_{\alpha}-R_{\alpha}- \sum_{\beta(\beta\neq \alpha)}T_{\beta\rightarrow \alpha}=0. \label{eq12.14}
2912: \end{eqnarray}
2913: Equation (\ref{eq12.13}) follows from current conservation, and Eq.\ (\ref{eq12.14}) follows
2914: from the requirement that an increase of all the chemical potentials by the
2915: same amount should have no effect on the net currents in the leads.
2916: Equation (\ref{eq12.12}) can be applied to a measurement of the four-terminal
2917: resistance $R_{\alpha\beta,\gamma\delta}=V/I$, in which a current $I$ flows from contact $\alpha$ to $\beta$ and a
2918: voltage difference $V$ is measured between contacts $\gamma$ and $\delta$. Setting
2919: $I_{\alpha}=I=-I_{\beta}$, and $I_{\alpha^{\prime}}=0$ for $\alpha^{\prime}\neq \alpha, \beta$, one can solve the set of linear
2920: equations (\ref{eq12.12}) to determine the chemical potential difference $\mu_{\gamma}-\mu_{\delta}$.
2921: (Only the {\it differences\/} in chemical potentials can be obtained from the $n$
2922: equations (\ref{eq12.12}), which are not independent in view of Eq.\ (\ref{eq12.14}). By fixing
2923: one chemical potential at zero, one reduces the number of equations to $n-1$
2924: independent ones.) The four-terminal resistance $R_{\alpha\beta,\gamma\delta}=(\mu_{\gamma}-\mu_{\delta})/eI$ is then
2925: obtained as a rational function of the transmission and reflection probabilities. We will refer to this procedure as the {\it Landauer-B\"{u}ttiker formalism}. It
2926: provides a unified description of the whole variety of electrical transport
2927: experiments discussed in this review.
2928:
2929: The transmission probabilities have the symmetry
2930: \be
2931: t_{\beta\alpha,nm}(B)=t_{\alpha\beta,mn}(-B)\Rightarrow T_{\alpha\rightarrow\beta}(B)=T_{\rho\rightarrow \alpha}(-B). \label{eq12.15}
2932: \ee
2933: Equation (\ref{eq12.15}) follows by combining the unitarity of the scattering matrix
2934: $\mathbf{t}^{\dagger}=\mathbf{t}^{-1}$, required by current conservation, with the symmetry
2935: $\mathbf{t}^{*}(-B)=\mathbf{t}^{-1}(B)$, required by time-reversal invariance ($*$ and $\dagger$ denote
2936: complex and Hermitian conjugation, respectively). As shown by B\"{u}ttiker,\cite{ref5,ref284} the symmetry (\ref{eq12.15}) of the coefficients in Eq.\ (\ref{eq12.12}) implies a
2937: {\it reciprocity relation\/} for the four-terminal resistance:
2938: \be
2939: R_{\alpha\beta,\gamma\delta}(B)=R_{\gamma\delta,\alpha\beta}(-B). \label{eq12.16}
2940: \ee
2941: The resistance is unchanged if current and voltage leads are interchanged
2942: with simultaneous reversal of the magnetic field direction. A special case of
2943: Eq.\ (\ref{eq12.16}) is that the two-terminal resistance $R_{\alpha\beta,\alpha\beta}$ is {\it even\/} in $B$. In the
2944: diffusive transport regime, the reciprocity relation for the resistance follows
2945: from the Onsager-Casimir relation\cite{ref285} $\mathbf{\rho}(B)=\mathbf{\rho}^{\mathrm{T}}(-\beta)$ for the resistivity
2946: tensor ($\mathrm{T}$ denotes the transpose). Equation (\ref{eq12.16}) holds also in cases that the
2947: concept of a local resistivity breaks down. One experimental demonstration\cite{ref80} of the validity of the reciprocity relation in the quantum ballistic
2948: transport regime will be discussed in Section \ref{sec14}. Other demonstrations have
2949: been given in Refs.\ \onlinecite{ref286,ref287,ref288,ref289}. We emphasize that strict reciprocity holds only
2950: in the linear response limit of infinitesimally small currents and voltages.
2951: Deviations from Eq.\ (\ref{eq12.16}) can occur experimentally\cite{ref290} due to nonlinearities
2952: from quantum interference,\cite{ref146,ref291} which in the case of a long phase
2953: coherence time $\tau_{\phi}$ persist down to very small voltages $V\gtrsim \hbar/e\tau_{\phi}$. Magnetic
2954: impurities can be another source of deviations from reciprocity if the applied
2955: magnetic field is not sufficiently strong to reverse the magnetic moments on
2956: field reversal. The large $\pm B$ asymmetry of the two-terminal resistance of a
2957: point contact reported in Ref.\ \onlinecite{ref292} has remained unexplained (see Section \ref{sec21}).
2958:
2959: The scattering matrix $\mathbf{t}$ in Eq.\ (\ref{eq12.15}) describes {\it elastic\/} scattering only.
2960: Inelastic scattering is assumed to take place exclusively in the reservoirs. That
2961: is a reasonable approximation at temperatures that are sufficiently low that
2962: the size of the conductor is smaller than the inelastic scattering length (or the
2963: phase coherence length if quantum interference effects play a role). Reservoirs
2964: thus play a dual role in the Landauer-B\"{u}ttiker formalism: On the one hand,
2965: a reservoir is a model for a current or voltage contact; on the other hand, a
2966: reservoir brings inelastic scattering into the system. The reciprocity relation
2967: (\ref{eq12.16}) is unaffected by adding reservoirs to the system and is not restricted to
2968: elastic scattering.\cite{ref284} More realistic methods to include inelastic scattering in
2969: a distributed way throughout the system have been proposed, but are not yet
2970: implemented in an actual calculation.\cite{ref293,ref294}
2971:
2972: \subsection{\label{sec13} Quantum point contacts}
2973:
2974: Many of the principal phenomena in ballistic transport are exhibited in
2975: the cleanest and most extreme way by quantum point contacts. These are
2976: short and narrow constrictions in a 2DEG, with a width of the order of the
2977: Fermi wavelength.\cite{ref6,ref7,ref59} The conductance of quantum point contacts is
2978: quantized in units of $2e^{2}/h$. This quantization is reminiscent of the quantization of the Hall conductance, but is measured in the absence of a magnetic
2979: field. The zero-field conductance quantization and the smooth transition to
2980: the quantum Hall effect on applying a magnetic field are essentially
2981: consequences of the equipartition of current among an integer number of
2982: propagating modes in the constriction, each mode carrying a current of $2e^{2}/h$
2983: times the applied voltage $V$. Deviations from precise quantization result from
2984: nonunit transmission probability of propagating modes and from nonzero
2985: transmission probability of evanescent (nonpropagating) modes. Experiment
2986: and theory in a zero magnetic field are reviewed in Section \ref{sec13a}. The effect of a
2987: magnetic field is the subject of Section \ref{sec13b}, which deals with depopulation of
2988: subbands and suppression of backscattering by a magnetic field, two
2989: phenomena that form the basis for an understanding of magnetotransport in
2990: semiconductor nanostructures.
2991:
2992: \subsubsection{\label{sec13a} Conductance quantization}
2993:
2994: {\bf (a) Experiments.} The study of electron transport through point contacts in
2995: metals has a long history, which goes back to Maxwell's investigations\cite{ref295} of
2996: the resistance of an orifice in the diffusive transport regime. Ballistic transport
2997: was first studied by Sharvin,\cite{ref296} who proposed and subsequently realized\cite{ref297}
2998: the injection and detection of a beam of electrons in a metal by means of point
2999: contacts much smaller than the mean free path. With the possible exception
3000: of the scanning tunneling microscope, which can be seen as a point contact on
3001: an atomic scale,\cite{ref298,ref299,ref300,ref301,ref302,ref303} these studies in metals are essentially restricted to the
3002: {\it classical\/} ballistic transport regime because of the extremely small Fermi
3003: wavelength $(\lambda_{\mathrm{F}}\approx 0.5\,\mathrm{nm})$. Point contacts in a 2DEG cannot be fabricated by
3004: simply pressing two wedge- or needle-shaped pieces of material together (as
3005: in bulk semiconductors\cite{ref304} or metals\cite{ref305}), since the electron gas is confined at
3006: the GaAs-AlGaAs interface in the interior of the heterostructure. Instead,
3007: they are defined electrostatically\cite{ref24,ref58} by means of a split gate on top of the
3008: heterostructure (a schematical cross-sectional view was given in Fig.\ \ref{fig4}b, while
3009: the micrograph Fig.\ \ref{fig5}b shows a top view of the split gate of a double-point
3010: contact device; see also the inset in Fig.\ \ref{fig44}). In this way one can define short
3011: and narrow constrictions in the 2DEG, of variable width $0\lesssim W\lesssim 250\,\mathrm{nm}$
3012: comparable to the Fermi wavelength $\lambda_{\mathrm{F}}\approx 40\,\mathrm{nm}$ and much shorter than the
3013: mean free path $l\approx 10\,\mu \mathrm{m}$.
3014:
3015: \begin{figure}
3016: \centerline{\includegraphics[width=8cm]{figures/fig44}}
3017: \caption{
3018: Point contact conductance as a function of gate voltage at 0.6 K, demonstrating the conductance quantization in units of $2e^{2}/h$. The data are obtained from the two-terminal resistance after subtraction of a background resistance. The constriction width increases with increasing voltage on the gate (see inset). Taken from B. J. van Wees et al., Phys.\ Rev.\ Lett.\ {\bf 60}, 848 (1988).
3019: \label{fig44}
3020: }
3021: \end{figure}
3022:
3023: Van Wees et al.\cite{ref6} and Wharam et al.\cite{ref7} independently discovered a sequence
3024: of steps in the conductance of such a point contact as its width was varied by
3025: means of the voltage on the split gate (see Fig.\ \ref{fig44}). The steps are near integer
3026: multiples of $2e^{2}/h\approx(13\,k\Omega)^{-1}$, after correction for a gate-voltage-independent series resistance from the wide 2DEG regions. An elementary
3027: explanation of this effect relies on the fact that each 1D subband in the
3028: constriction contributes $2e^{2}/h$ to the conductance because of the cancellation
3029: of the group velocity and the 1D density of states discussed in Section \ref{sec12}.
3030: Since the number $N$ of occupied subbands is necessarily an integer, it follows
3031: from this simple argument that the conductance $G$ is quantized,
3032: \be
3033: G=(2e^{2}/h)N, \label{eq13.1}
3034: \ee
3035: as observed experimentally. A more complete explanation requires an explicit
3036: treatment of the mode coupling at the entrance and exit of the constriction, as
3037: discussed later.
3038:
3039: The zero-field conductance quantization of a quantum point contact is not
3040: as accurate as the Hall conductance quantization in strong magnetic fields.
3041: The deviations from exact quantization are typically\cite{ref6,ref7,ref306} 1\%, while in the
3042: quantum Hall effect one obtains routinely\cite{ref97} an accuracy of 1 part in $10^{7}$. It is
3043: unlikely that a similar accuracy will be achieved in the case of the zero-field
3044: quantization, one reason being the additive contribution to the point contact
3045: resistance of a background resistance whose magnitude cannot be determined precisely. The largest part of this background resistance originates in
3046: the ohmic contacts\cite{ref307} and can thus be eliminated in a four-terminal
3047: measurement of the contact resistance. The position of the additional voltage
3048: probes on the wide 2DEG regions has to be more than an inelastic scattering
3049: length away from the point contact so that a local equilibrium is established.
3050: A residual background resistance\cite{ref307} of the order of the resistance $\rho$ of a
3051: square is therefore unavoidable. In the experiments of Refs.\ \onlinecite{ref6} and \onlinecite{ref7} one has
3052: $\rho\approx 20\,\Omega$, but lower values are possible for higher-mobility material. It would
3053: be of interest to investigate experimentally whether resistance plateaux
3054: quantized to such an accuracy are achievable. It should be noted, however,
3055: that the degree of flatness of the plateaux and the sharpness of the steps in
3056: the present experiments vary among devices of identical design, indicating
3057: that the detailed shape of the electrostatic potential defining the constriction
3058: is important. There are many uncontrolled factors affecting this shape, such
3059: as small changes in the gate geometry, variations in the pinning of the Fermi
3060: level at the free GaAs surface or at the interface with the gate metal, doping
3061: inhomogeneities in the heterostructure material, and trapping of charge in
3062: deep levels in AlGaAs.
3063:
3064: On increasing the temperature, one finds experimentally that the plateaux
3065: acquire a finite slope until they are no longer resolved.\cite{ref308} This is a
3066: consequence of the thermal smearing of the Fermi-Dirac distribution (\ref{eq4.9}). If
3067: at $T=0$ the conductance $G(E_{\mathrm{F}}, T)$ has a step function dependence on the
3068: Fermi energy $E_{\mathrm{F}}$, at finite temperatures it has the form\cite{ref309}
3069: \begin{eqnarray}
3070: G(E_{\mathrm{F}}, T)&=& \int_{0}^{\infty}G(E, 0) \frac{df}{dE_{\mathrm{F}}} dE\nonumber\\
3071: &=& \frac{2e^{2}}{h}\sum_{n=1}^{\infty}f(E_{n}-E_{\mathrm{F}}). \label{eq13.2}
3072: \end{eqnarray}
3073: Here $E_{n}$ denotes the energy of the bottom of the $n$th subband [cf.\ Eq.\ (\ref{eq4.3})].
3074: The width of the thermal smearing function $df/dE_{\mathrm{F}}$ is about $4k_{\mathbf{B}}T$, so the
3075: conductance steps should disappear for $T\gtrsim\Delta E/4k_{B}\sim 4\,\mathrm{K}$ (here $\Delta E$ is the
3076: subband splitting at the Fermi level). This is confirmed both by experiment\cite{ref308} and by numerical calculations (see below).
3077:
3078: Interestingly, it was found experimentally\cite{ref6,ref7} that in general a finite
3079: temperature yielded the most pronounced and flat plateaux as a function of
3080: gate voltage in the zero-field conductance. If the temperature is increased
3081: beyond this optimum (which is about $0.5\, \mathrm{K}$), the plateaux disappear because
3082: of the thermal averaging discussed earlier. Below this temperature, an
3083: oscillatory structure may be superimposed on the conductance plateaux. This
3084: phenomenon depends on the precise shape of the constriction, as discussed
3085: later. A small but finite voltage drop across the constriction has an effect that
3086: is qualitatively similar to that of a finite temperature.\cite{ref309} This is indeed borne
3087: out by experiment.\cite{ref308} (Experiments on conduction through quantum point
3088: contacts at larger applied voltages in the nonlinear transport regime have
3089: been reviewed in Ref.\ \onlinecite{ref307}).
3090:
3091: Theoretically, one would expect the conductance quantization to be
3092: preserved in longer channels than those used in the original experiment\cite{ref6,ref7}
3093: (in which typically $L\sim W\sim 100\,\mathrm{nm}$). Experiments on channels longer than
3094: about 1 $\mu \mathrm{m}$ did not show the quantization,\cite{ref306,ref307,ref310} however, although their
3095: length was well below the transport mean free path in the bulk (about 10 $\mu \mathrm{m}$).
3096: The lack of clear plateaux in long constrictions is presumably due to
3097: enhanced backscattering inside the constriction, either because of impurity
3098: scattering (which may be enhanced\cite{ref306,ref310} due to the reduced screening in a
3099: quasi-one-dimensional electron gas\cite{ref72}) or because of boundary scattering at
3100: channel wall irregularities. As mentioned in Section \ref{sec5}, Thornton et al.\cite{ref107}
3101: have found evidence for a small (5\%) fraction of diffuse, rather than specular,
3102: reflections at boundaries defined electrostatically by a gate. In a 200-nm-wide
3103: constriction this leads to an effective mean free path of about
3104: $200\, \mathrm{nm}/0.05\approx 4\,\mu \mathrm{m}$, comparable to the constriction length of devices that do
3105: not exhibit the conductance quantization.\cite{ref113,ref307}
3106:
3107: \begin{figure}
3108: \centerline{\includegraphics[width=6cm]{figures/fig45a}}
3109:
3110: \centerline{\includegraphics[width=6cm]{figures/fig45b}}
3111: \caption{
3112: (a) Classical ballistic transport through a point contact induced by a concentration difference $\delta n$, or electrochemical potential difference $e V$, between source (s) and drain (d). (b) The net current through a quantum point contact is carried by the shaded region in $k$-space. In a narrow channel the allowed states lie on the horizontal lines, which correspond to quantized values for $k_{y} = \pm n\pi/W$, and continuous values for $k_{x}$. The formation of these 1D subbands gives rise to a quantized conductance. Taken from H. van Houten et al., in ``Physics and Technology of Submicron Structures'' (H. Heinrich, G. Bauer, and F. Kuchar, eds.). Springer, Berlin, 1988; and in ``Nanostructure Physics and Fabrication'' (M. Reed and W. P. Kirk, eds.). Academic, New York, 1989.
3113: \label{fig45}
3114: }
3115: \end{figure}
3116:
3117: {\bf (b) Theory.} It is instructive to first consider {\it classical\/} 2D point contacts in
3118: some detail.\cite{ref31,ref311} The ballistic electron flow through a point contact is
3119: illustrated in Fig.\ \ref{fig45}a in real space, and in Fig.\ \ref{fig45}b in $k$-space, for a small
3120: excess electron density $\delta n$ at one side of the point contact. At low temperatures this excess charge moves with the Fermi velocity $v_{\mathrm{F}}$. The flux normally
3121: incident on the point contact is $\delta nv_{\mathrm{F}}\langle\cos\phi\,\theta(\cos\phi)\rangle$, where $\theta(x)$ is the unit
3122: step function and the symbol $\langle\; \rangle$ denotes an isotropic angular average (the
3123: angle $\phi$ is defined in Fig.\ \ref{fig45}a). In the ballistic limit $l\gg W$ the incident flux is
3124: fully transmitted, so the total diffusion current $J$ through the point contact is
3125: given by
3126: \be
3127: J=W \delta nv_{\mathrm{F}}\int_{-\pi/2}^{\pi/2}\cos\phi\frac{d\phi}{2\pi}=\frac{1}{\pi}Wv_{\mathrm{F}}\delta n. \label{eq13.3}
3128: \ee
3129: The diffusance $\tilde{D}\equiv J/\delta n=(1/\pi)Wv_{\mathrm{F}}$; therefore, the conductance $G=
3130: e^{2}\rho(E_{\mathrm{F}})\tilde{D}$ becomes (using the 2D density of states (\ref{eq4.2}) with the appropriate
3131: degeneracy factors $g_{\mathrm{s}}=2$, $g_{\mathrm{v}}=1$)
3132: \be
3133: G= \frac{2e^{2}}{h}\frac{k_{\mathrm{F}}W}{\pi},\;\;\mbox{in 2D}. \label{eq13.4}
3134: \ee
3135: Eq.\ (\ref{eq13.4}) is the 2D analogue\cite{ref6} of Sharvin's well-known expression\cite{ref296} for the
3136: point contact conductance in three dimensions,
3137: \be
3138: G= \frac{2e^{2}}{h}\frac{k_{\mathrm{F}}^{2}S}{4\pi},\;\;\mbox{in 3D}, \label{eq13.5}
3139: \ee
3140: where now $S$ is the area of the point contact. The number of propagating
3141: modes for a square-well lateral confining potential is $N=\mathrm{Int}[k_{\mathrm{F}}W/\pi]$ in 2D,
3142: so Eq.\ (\ref{eq13.4}) is indeed the classical limit of the quantized conductance (\ref{eq13.1}).
3143:
3144: Quantum mechanically, the current through the point contact is equipartitioned among the 1D subbands, or transverse modes, in the constriction. The
3145: equipartitioning of current, which is the basic mechanism for the conductance
3146: quantization, is illustrated in Fig.\ \ref{fig45}b for a square-well lateral confining
3147: potential of width $W$. The 1D subbands then correspond to the pairs of
3148: horizontal lines at $k_{y}=\pm n\pi/W$, with $n=1,2, \ldots, N$ and $N=\mathrm{Int}[k_{\mathrm{F}}W/\pi]$.
3149: The group velocity $v_{n}=\hbar k_{x}/m$ is proportional to $\cos\phi$ and thus decreases
3150: with increasing $n$. However, the decrease in $v_{n}$ is compensated by an increase
3151: in the 1D density of states. Since $\rho_{n}$ is proportional to the length of the
3152: horizontal lines within the dashed area in Fig.\ \ref{fig45}b, $\rho_{n}$ is proportional to
3153: $1/\cos\phi$ so that the product $v_{n}\rho_{n}$ does not depend on the subband index. We
3154: emphasize that, although the classical formula (\ref{eq13.4}) holds only for a square-well lateral confining potential, the quantization (\ref{eq13.1}) is a general result for
3155: any shape of the confining potential. The reason is simply that the
3156: fundamental cancellation of the group velocity $v_{n}=dE_{n}(k)/\hbar dk$ and the 1D
3157: density of states $\rho_{n}^{+}=(\pi dE_{n}(k)/dk)^{-1}$ holds {\it regardless\/} of the form of the
3158: dispersion relation $E_{n}(k)$. For the same reason, Eq.\ (\ref{eq13.1}) is equally applicable
3159: in the presence of a magnetic field, when magnetic edge channels at the Fermi
3160: level take over the role of 1D subbands. Equation (\ref{eq13.1}) thus implies a
3161: continuous transition from the zero-field quantization to the quantum Hall
3162: effect, as we will discuss in Section \ref{sec13b}.
3163:
3164: To analyze deviations from Eq.\ (\ref{eq13.1}) it is necessary to solve the
3165: Schr\"{o}dinger equation for the wave functions in the narrow point contact and
3166: the adjacent wide regions and to match the wave functions and their
3167: derivatives at the entrance and exit of the constriction. The resulting
3168: transmission coefficients determine the conductance via the Landauer formula (\ref{eq12.10}). This mode coupling problem has been solved numerically for
3169: point contacts of a variety of shapes\cite{ref312,ref313,ref314,ref315,ref316,ref317,ref318,ref319,ref320,ref321} and analytically in special
3170: geometries.\cite{ref322,ref323,ref324} When considering the mode coupling at the entrance and
3171: exit of the constriction, one must distinguish gradual ({\it adiabatic}) from {\it abrupt\/}
3172: transitions from wide to narrow regions.
3173:
3174: The case of an {\it adiabatic\/} constriction has been studied by Glazman et
3175: al.,\cite{ref325} by Yacoby and Imry\cite{ref326} and by Payne.\cite{ref272} If the constriction width
3176: $W(x)$ changes sufficiently gradually, the transport through the constriction is
3177: adiabatic (i.e., without intersubband scattering). The transmission coefficients
3178: then vanish, $|t_{nm}|^{2}=0$, unless $n=m\leq N_{\min}$, with $N_{\min}$ the smallest number of
3179: occupied subbands in the constriction. The conductance quantization (\ref{eq13.1})
3180: now follows immediately from the Landauer formula (\ref{eq12.10}). The criterion
3181: for adiabatic transport is\cite{ref326} $dW/dx\lesssim 1/N(x)$, with $N(x)\approx k_{\mathrm{F}}W(x)/\pi$ the local
3182: number of subbands. As the constriction widens, $N(x)$ increases and adiabaticity is preserved only if $W(x)$ increases more and more slowly. In practice,
3183: adiabaticity breaks down at a width $W_{\max}$, which is at most a factor of 2 larger
3184: than the minimum width $W_{\min}$ (cf.\ the collimated beam experiment of Ref.\
3185: \onlinecite{ref327}, discussed in Section \ref{sec15}). This does not affect the conductance of the
3186: constriction, however, if the breakdown of adiabaticity results in a mixing of
3187: the subbands without causing reflection back through the constriction. If
3188: such is the case, the total transmission probability through the constriction
3189: remains the same as in the hypothetical case of fully adiabatic transport. As
3190: pointed out by Yacoby and Imry,\cite{ref326} a relatively small adiabatic increase in
3191: width from $W_{\min}$ to $W_{\max}$ is sufficient to ensure a drastic suppression of
3192: reflections at $W_{\max}$. The reason is that the subbands with the largest reflection
3193: probability are close to cutoff, that is, they have subband index close to $N_{\max}$,
3194: the number of subbands occupied at $W_{\max}$. Because the transport is adiabatic
3195: from $W_{\min}$ to $W_{\max}$, only the $N_{\min}$ subbands with the smallest $n$ arrive at $W_{\max}$,
3196: and these subbands have a small reflection probability. In the language of
3197: waveguide transmission, one has impedance-matched the constriction to the
3198: wide 2DEG regions.\cite{ref328} The filtering of subbands by a gradually widening
3199: constriction has an interesting effect on the angular distribution of the
3200: electrons injected into the wide 2DEG. This {\it horn collimation\/} effect\cite{ref329} is
3201: discussed in Section \ref{sec15}.
3202:
3203: \begin{figure}
3204: \centerline{\includegraphics[width=8cm]{figures/fig46}}
3205: \caption{
3206: Transmission resonances exhibited by theoretical results for the conductance of a quantum point contact of abrupt (rectangular) shape. A smearing of the resonances occurs at nonzero temperatures ($T_{0}=0.02\,E_{\rm F}/k_{\rm B}\approx 2.8\,{\rm K}$). The dashed curve is an exact numerical result; the full curves are approximate. Taken from A. Szafer and A. D. Stone, Phys.\ Rev.\ Lett.\ {\bf 62}, 300 (1989).
3207: \label{fig46}
3208: }
3209: \end{figure}
3210:
3211: An adiabatic constriction improves the accuracy of the conductance
3212: quantization, but is not required to observe the effect. Calculations\cite{ref312,ref313,ref314,ref315,ref316,ref317,ref318,ref319,ref320,ref321,ref322,ref323,ref324}
3213: show that well-defined conductance plateaux persist for {\it abrupt\/} constrictions,
3214: especially if they are neither very short nor very long. The optimum length for
3215: the observation of the plateaux is given by\cite{ref313} $L_{\mathrm{opt}}\approx 0.4(W\lambda_{\mathrm{F}})^{1/2}$. In shorter
3216: constrictions the plateaux acquire a finite slope, although they do not
3217: disappear completely even at zero length. For $L>L_{\mathrm{opt}}$ the calculations
3218: exhibit regular oscillations that depress the conductance periodically below
3219: its quantized value. The oscillations are damped and have usually vanished
3220: before the next plateau is reached. As a representative illustration, we
3221: reproduce in Fig.\ \ref{fig46} a set of numerical results for the conductance as a
3222: function of width (at fixed Fermi wave vector), obtained by Szafer and
3223: Stone.\cite{ref315} Note that a finite temperature improves the flatness of the plateaux,
3224: as observed experimentally. The existence of an optimum length can be
3225: understood as follows.
3226:
3227: Because of the abrupt widening of the constriction, there is a significant
3228: probability for reflection at the exit of the constriction, in contrast to the
3229: adiabatic case considered earlier. The conductance as a function of width, or
3230: Fermi energy, is therefore not a simple step function. On the $n$th conductance
3231: plateau backscattering occurs predominantly for the $n$th subband, since it
3232: is closest to cutoff. Resonant transmission of this subband occurs if
3233: the constriction length $L$ is approximately an integer multiple of half
3234: the longitudinal wavelength $\lambda_{n}=h[2m(E_{\mathrm{F}}-E_{n})]^{-1/2}$, leading to oscillations
3235: on the conductance plateaux. These transmission resonances are
3236: damped, because the reflection probability decreases with decreasing $\lambda_{n}$. The
3237: shortest value of $\lambda_{n}$ on the $N\mathrm{th}$ conductance plateau is
3238: $h[2m(E_{N+1}-E_{N})]^{-1/2}\approx(W\lambda_{\mathrm{F}})^{1/2}$ (for a square-well lateral confining potential). The transmission resonances are thus suppressed if $L\lesssim(W\lambda_{\mathrm{F}})^{1/2}$.
3239: Transmission through evanescent modes (i.e., subbands above $E_{\mathrm{F}}$) is predominant for the $(N+1)\mathrm{th}$ subband, since it has the largest decay length
3240: $\Lambda_{N+1}=h[2m(E_{N+1}-E_{\mathrm{F}})]^{-1/2}$. The observation of that plateau requires
3241: that the constriction length exceeds this decay length at the population
3242: threshold of the $N$ th mode, or $L\gtrsim h[2m(E_{N+1}-E_{N})]^{-1/2}\approx(W\lambda_{\mathrm{F}})^{1/2}$. The
3243: optimum length\cite{ref313} $L_{\mathrm{opt}}\approx 0.4(W\lambda_{\mathrm{F}})^{1/2}$ thus separates a short constriction
3244: regime, in which transmission via evanescent modes cannot be ignored, from
3245: a long constriction regime, in which transmission resonances obscure the
3246: plateaux.
3247:
3248:
3249:
3250: \begin{figure}
3251: \centerline{\includegraphics[width=8cm]{figures/fig47}}
3252: \caption{
3253: Resistance as a function of gate voltage for an elongated quantum point contact ($L = 0.8\,\mu{\rm m}$) at temperatures of 0.2, 0.4, and 0.8 K, showing transmission resonances. Subsequent curves from the bottom are offset by 1 ${\rm k}\Omega$. Taken from R. J. Brown et al., Solid State Electron.\ {\bf 32}, 1179 (1989).
3254: \label{fig47}
3255: }
3256: \end{figure}
3257:
3258: Oscillatory structure was resolved in low-temperature experiments on the
3259: conductance quantization of one quantum point contact by van Wees et
3260: al.,\cite{ref308} but was not clearly seen in other devices. A difficulty in the
3261: interpretation of these and other experiments is that oscillations can also be
3262: caused by quantum interference processes involving impurity scattering near
3263: the constriction. Another experimental observation of oscillatory structure
3264: was reported by Hirayama et al.\cite{ref330} for short (100-nm) quantum point
3265: contacts of fixed width (defined by means of focused ion beam lithography).
3266: To observe the plateaux, they slowly varied the electron density by weakly
3267: illuminating the sample. The oscillations were quite reproducible, also after
3268: thermal cycling of the sample, but again they were found in some of the
3269: devices only (this was attributed to variations in the abruptness of the
3270: constrictions\cite{ref330,ref331}). Brown et al.\cite{ref332} have studied the conductance of split-gate constrictions of lengths $L\approx 0.3,0.8$, and 1 $\mu \mathrm{m}$, and they observed
3271: pronounced oscillations instead of the flat conductance plateaux found for
3272: shorter quantum point contacts. The observed oscillatory structure (reproduced in Fig.\ \ref{fig47}) is quite regular, and it correlates with the sequence of
3273: plateaux that is recovered at higher temperatures (around $0.8\, \mathrm{K}$). The effect
3274: was seen in all of the devices studied in Ref.\ \onlinecite{ref332}. Measurements by Timp et
3275: al.\cite{ref306} on rather similar $0.9$-$\mu \mathrm{m}$-long constrictions did not show periodic
3276: oscillations, however. Brown et al.\ conclude that their oscillations are due to
3277: transmission resonances associated with reflections at entrance and exit of
3278: the constriction. Detailed comparison with theory is difficult because the
3279: transmission resonances depend sensitively on the shape of the lateral
3280: confining potential and on the presence of a potential barrier in the
3281: constriction (see Section \ref{sec13b}). A calculation that comes close to the
3282: observation of Brown et al.\ has been published by Martin-Moreno and
3283: Smith.\cite{ref333}
3284:
3285: \begin{figure}
3286: \centerline{\includegraphics[width=8cm]{figures/fig48}}
3287: \caption{
3288: Point contact conductance (corrected for a background resistance) as a function of gate voltage for several magnetic field values, illustrating the transition from zero-field quantization to quantum Hall effect. The curves have been offset for clarity. The inset shows the device geometry. Taken from B. J. van Wees et al., Phys.\ Rev.\ B. {\bf 38}, 3625 (1988).
3289: \label{fig48}
3290: }
3291: \end{figure}
3292:
3293: \subsubsection{\label{sec13b} Depopulation of subbands and suppression of backscattering by a magnetic field}
3294:
3295: The effect of a magnetic field (perpendicular to the 2DEG) on the
3296: quantized conductance of a point contact is shown in Fig.\ \ref{fig48}, as measured by
3297: van Wees et al.\cite{ref334} First of all, Fig.\ \ref{fig48} demonstrates that the conductance
3298: quantization is conserved in the presence of a magnetic field and shows a
3299: smooth transition from zero-field quantization to quantum Hall effect. The
3300: most noticeable effect of the magnetic field is to reduce the number of
3301: plateaux in a given gate voltage interval. This provides a demonstration of
3302: depopulation of magnetoelectric subbands, which is more direct than that
3303: provided by the experiments discussed in Section \ref{sec10}. In addition, one
3304: observes that the flatness of the plateaux improves in the presence of the field.
3305: This is due to the reduction of the reflection probability at the point contact,
3306: which is revealed most clearly in a somewhat different (four-terminal)
3307: measurement configuration. These two effects of a magnetic field will be
3308: discussed separately. We will return to the magnetic suppression of back-scattering in Section \ref{sec18} in connection with the edge channel theory\cite{ref112} of the
3309: quantum Hall effect.
3310:
3311: {\bf (a) Depopulation of subbands.} Because the equipartitioning of current
3312: among the 1D subbands holds regardless of the nature of the subbands
3313: involved, one can conclude that in the presence of a magnetic field $B$ the
3314: conductance remains quantized according to $G=(2e^{2}/h)N$ (ignoring spin
3315: splitting of the subbands, for simplicity). Explicit calculations\cite{ref335} confirm this
3316: expectation. The number of occupied subbands $N$ as a function of $B$ has been
3317: studied in Sections \ref{sec10} and \ref{sec12} and is given by Eqs.\ (\ref{eq10.7}) and (\ref{eq10.8}) for a
3318: parabolic and a square-well potential, respectively. In the high-magnetic-field
3319: regime $W\gtrsim 2l_{\mathrm{cycl}}$, the number $N\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}$ is just the number of occupied
3320: Landau levels. The conductance quantization is then a manifestation of the
3321: quantum Hall effect.\cite{ref8} (The fact that $G$ is not a Hall conductance but a two-terminal conductance is not an essential distinction for this effect; see Section
3322: \ref{sec18}.) At lower magnetic fields, the conductance quantization provides a direct
3323: and extremely straightforward method to measure via $N=G(2e^{2}/h)^{-1}$ the
3324: depopulation of magnetoelectric subbands in the constriction.
3325:
3326: \begin{figure}
3327: \centerline{\includegraphics[width=8cm]{figures/fig49}}
3328: \caption{
3329: Number of occupied subbands as a function of reciprocal magnetic field for several values of the gate voltage. Data points have been obtained directly from the quantized
3330: conductance (Fig.\ \ref{fig48}); solid curves are calculated for a square-well confining potential of width $W$ and well bottom $E_{\rm c}$ as tabulated in the inset. Taken from B. J. van Wees et al., Phys.\ Rev.\ B {\bf 38}, 3625 (1988).
3331: \label{fig49}
3332: }
3333: \end{figure}
3334:
3335: Figure \ref{fig49} shows $N$ versus $B^{-1}$ for various gate voltages, as it follows from
3336: the experiment of Fig.\ \ref{fig48}. Also shown are the theoretical curves for a square-well confining potential, with the potential barrier in the constriction taken
3337: into account by replacing $E_{\mathrm{F}}$ by $E_{\mathrm{F}}-E_{\mathrm{c}}$ in Eq.\ (\ref{eq10.8}). The $B$-dependence of
3338: $E_{\mathrm{F}}$ has been ignored in the calculation. The barrier height $E_{\mathrm{c}}$ is obtained from
3339: the high-field conductance plateaux [where $N\approx(E_{\mathrm{F}}-E_{\mathrm{c}})/\hbar\omega_{\mathrm{c}}]$, and the constriction width $W$ then follows from the zero-field conductance (where
3340: $N\approx[2m(E_{\mathrm{F}}-E_{\mathrm{c}})/h^{2}]^{1/2}W/\pi)$. The good agreement found over the entire
3341: field range confirms the expectation that the quantized conductance is
3342: exclusively determined by the number of occupied subbands, irrespective of
3343: their electric or magnetic origin. The analysis in Fig.\ \ref{fig49} is for a square-well
3344: confining potential.\cite{ref334} For the narrowest constrictions a parabolic potential
3345: should be more appropriate,\cite{ref61} which has been used to analyze the data of
3346: Fig.\ \ref{fig48} in Refs.\ \onlinecite{ref336} and \onlinecite{ref308}. Wharam et al.\cite{ref337} have analyzed their
3347: depopulation data using the intermediate model of a parabolic potential
3348: with a flattened bottom (cf.\ also Ref.\ \onlinecite{ref336}). Because of the uncertainties in the
3349: actual shape of the potential, the parameter values tabulated in Fig.\ \ref{fig49} should
3350: be considered as rough estimates only.
3351:
3352: In strong magnetic fields the spin degeneracy of the energy levels is
3353: removed, and additional plateaux appear\cite{ref7,ref334} at {\it odd\/} multiples of $e^{2}/h$.
3354: Wharam et al.\cite{ref7} have demonstrated this effect in a particularly clear fashion,
3355: using a magnetic field parallel (rather than perpendicular) to the 2DEG.
3356: Rather strong magnetic fields turned out to be required to fully lift the spin
3357: degeneracy in this experiment (about $10\, \mathrm{T}$).
3358:
3359: {\bf (b) Suppression of backscattering.} Only a small fraction of the electrons
3360: injected by the current source into the 2DEG is transmitted through the
3361: point contact. The remaining electrons are scattered back into the source
3362: contact. This is the origin of the nonzero resistance of a ballistic point
3363: contact. In this subsection we shall discuss how a relatively weak magnetic
3364: field leads to a suppression of the {\it geometrical backscattering\/} caused by the
3365: finite width of the point contact, while the amount of backscattering caused
3366: by the potential barrier in the point contact remains essentially unaffected.
3367:
3368: \begin{figure}
3369: \centerline{\includegraphics[width=8cm]{figures/fig50}}
3370: \caption{
3371: Four-terminal longitudinal magnetoresistance $R_{4{\rm t}}\equiv R_{\rm L}$ of a constriction for a series of gate voltages. The negative magnetoresistance is temperature independent between 50 mK and 4 K. Solid lines are according to Eqs.\ (\ref{eq13.7}) and (\ref{eq10.8}), with the constriction width as adjustable parameter. The inset shows schematically the device geometry, with the two voltage probes used to measure $R_{\rm L}$. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 37}, 8534 (1988).
3372: \label{fig50}
3373: }
3374: \end{figure}
3375:
3376: The reduction of backscattering by a magnetic field is observed as a
3377: {\it negative\/} magnetoresistance [i.e., $R(B)-R(0)<0]$ in a {\it four-terminal\/} measurement of the longitudinal point contact resistance $R_{\mathrm{L}}$. The voltage probes in
3378: this experiment\cite{ref113} are positioned on wide 2DEG regions, well away from the
3379: constriction (see the inset in Fig.\ \ref{fig50}). This allows the establishment of local
3380: equilibrium near the voltage probes, at least in weak magnetic fields (cf.\
3381: Sections \ref{sec18} and \ref{sec19}), so that the measured four-terminal resistance does not
3382: depend on the properties of the probes. The experimental results for $R_{\mathrm{L}}$ in this
3383: geometry are plotted in Fig.\ \ref{fig50}. The negative magnetoresistance is
3384: temperature-independent (between $50\, \mathrm{mK}$ and 4 K) and is observed in weak
3385: magnetic fields once the narrow constriction is defined (for $V_{\mathrm{g}}\lesssim-0.3\,\mathrm{V}$). At
3386: stronger magnetic fields $(B>0.4\,\mathrm{T})$, a crossover is observed to a positive
3387: magnetoresistance. The zero-field resistance, the magnitude of the negative
3388: magnetoresistance, the slope of the positive magnetoresistance, as well as the
3389: crossover field, all increase with increasing negative gate voltage.
3390:
3391: The magnetic field dependence of the four-terminal resistance shown in
3392: Fig.\ \ref{fig50} is qualitatively different from that of the two-terminal resistance
3393: $R_{2\mathrm{t}}\equiv G^{-1}$ considered in the previous subsection. In fact, $R_{2\mathrm{t}}$ is approximately $B$-independent in weak magnetic fields (below the crossover fields of Fig.\
3394: \ref{fig50}). The reason is that $R_{2\mathrm{t}}$ is given by [cf.\ Eq.\ (\ref{eq13.1})]
3395: \be
3396: R_{2\mathrm{t}}= \frac{h}{2e^{2}}\frac{1}{N_{\min}}, \label{eq13.6}
3397: \ee
3398: with $N_{\min}$ the number of occupied subbands in the constriction (at the point
3399: where it has its minimum width and electron gas density). In weak magnetic
3400: fields such that $2l_{\mathrm{cycl}}>W$, the number of occupied subbands remains
3401: approximately constant [cf.\ Fig.\ \ref{fig31} or Eq.\ (\ref{eq10.8})], so $R_{2\mathrm{t}}$ is only weakly
3402: dependent on $B$ in this field regime. For stronger fields Eq.\ (\ref{eq13.6}) describes a
3403: {\it positive\/} magnetoresistance, because $N_{\min}$ decreases due to the magnetic
3404: depopulation of subbands discussed earlier. (A similar positive magnetoresistance is found in a Hall bar with a cross gate; see Ref.\ \onlinecite{ref338}.) Why then does one
3405: find a {\it negative\/} magnetoresistance in the four-terminal measurements of Fig.\
3406: \ref{fig50}? Qualitatively, the answer is shown in Fig.\ \ref{fig51}, for a constriction without a
3407: potential barrier. In a magnetic field the left-and right-moving electrons are
3408: spatially separated by the Lorentz force at opposite sides of the constriction.
3409: Quantum mechanically the skipping orbits in Fig.\ \ref{fig51} correspond to magnetic
3410: edge states (cf.\ Fig.\ \ref{fig41}). Backscattering thus requires scattering across the
3411: width of the constriction, which becomes increasingly improbable as $l_{\mathrm{cycl}}$
3412: becomes smaller and smaller compared with the width (compare Figs.\ \ref{fig51}a,b).
3413: For this reason a magnetic field suppresses the {\it geometrical\/} constriction
3414: resistance in the ballistic regime, but not the resistance associated with the
3415: constriction in energy space, which is due to the potential barrier.
3416:
3417: \begin{figure}
3418: \centerline{\includegraphics[width=8cm]{figures/fig51}}
3419: \caption{
3420: Illustration of the reduction of backscattering by a magnetic field, which is responsible for the negative magnetoresistance of Fig.\ \ref{fig50}. Shown are trajectories approaching a constriction without a potential barrier, in a weak (a) and strong (b) magnetic field. Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
3421: \label{fig51}
3422: }
3423: \end{figure}
3424:
3425: These effects were analyzed theoretically in Ref.\ \onlinecite{ref113}, with the simple result
3426: \be
3427: R_{\mathrm{L}}= \frac{h}{2e^{2}}\left(\frac{1}{N_{\min}}-\frac{1}{N_{\mathrm{wide}}}\right). \label{eq13.7}
3428: \ee
3429: Here $N_{\mathrm{wide}}$ is the number of occupied Landau levels in the wide 2DEG
3430: regions. The simplest (but incomplete) argument leading to Eq.\ (\ref{eq13.7}) is that
3431: the additivity of voltages on reservoirs (ohmic contacts) implies that the two-terminal resistance $R_{2\mathrm{t}}=(h/2e^{2})N_{\min}^{-1}$ should equal the sum of the Hall
3432: resistance $R_{\mathrm{H}}=(h/2e^{2})N_{\mathrm{wide}}^{-1}$ and the longitudinal resistance $R_{\mathrm{L}}$. This argument is incomplete because it assumes that the Hall resistance in the wide
3433: regions is not affected by the presence of the constriction. This is correct in
3434: general only if inelastic scattering has equilibrated the edge states transmitted
3435: through the constriction before they reach a voltage probe. Deviations from
3436: Eq.\ (\ref{eq13.7}) can occur in the absence of local equilibrium near the voltage
3437: probes, depending on the properties of the probes themselves. We discuss this
3438: in Section \ref{sec19}, following a derivation of Eq.\ (\ref{eq13.7}) from the Landauer-B\"{u}ttiker formalism.
3439:
3440: At small magnetic fields $N_{\min}$ is approximately constant, while
3441: $N_{\mathrm{wide}}\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}$ decreases linearly with $B$. Equation (\ref{eq13.7}) thus predicts a
3442: {\it negative\/} magnetoresistance. If the electron density in the wide and narrow
3443: regions is equal (i.e., the barrier height $E_{\mathrm{c}}=0$), then the resistance $R_{\mathrm{L}}$ {\it vanishes}
3444: for fields $B>B_{\mathrm{crit}}\equiv 2\hbar k_{\rm F}/eW$. This follows from Eq.\ (\ref{eq13.7}), because in this
3445: case $N_{\min}$ and $N_{\mathrm{wide}}$ are identical. If the electron density in the constriction is
3446: less than its value in the wide region, then Eq.\ (\ref{eq13.7}) predicts a crossover at
3447: $B_{\mathrm{crit}}$ to a strong-field regime of {\it positive\/} magnetoresistance described by
3448: \be
3449: R_{\mathrm{L}} \approx\frac{h}{2e^{2}}\left(\frac{\hbar\omega_{\mathrm{c}}}{E_{\mathrm{F}}-E_{\mathrm{c}}}-\frac{\hbar\omega_{\mathrm{c}}}{E_{\mathrm{F}}}\right)\;\;{\rm if}\;\;B>B_{\mathrm{crit}}. \label{eq13.8}
3450: \ee
3451:
3452: The experimental results are well described by the solid curves following
3453: from Eq.\ (\ref{eq13.7}) (with $N_{\min}$ given by the square-well result (\ref{eq10.8}), and with an
3454: added constant background resistance). The constriction in the present
3455: experiment is relatively long $(L\approx 3.4\,\mu \mathrm{m})$, and wide ($W$ ranging from 0.2 to
3456: 1.0 $\mu \mathrm{m}$) so that it does not exhibit quantized two-terminal conductance
3457: plateaux in the absence of a magnetic field. For this reason the discreteness of
3458: $N_{\min}$ was ignored in the theoretical curves in Fig.\ \ref{fig50}. We emphasize, however,
3459: that Eq.\ (\ref{eq13.7}) is equally applicable to the quantized case, as observed by
3460: several groups\cite{ref307,ref339,ref340,ref341,ref342} (see Section \ref{sec19}).
3461:
3462: The negative magnetoresistance (\ref{eq13.7}) due to the suppression of the
3463: contact resistance is an additive contribution to the magnetoresistance of a
3464: long and narrow channel in the quasi-ballistic regime (if the voltage probes
3465: are positioned on two wide 2DEG regions, connected by the channel). For a
3466: channel of length $L$ and a mean free path $l$ the zero-field contact resistance is a
3467: fraction $\sim l/L$ of the Drude resistance and may thus be ignored for $L\gg l$. The
3468: strong-field positive magnetoresistance (\ref{eq13.8}) resulting from a different
3469: electron density in the channel may still be important, however. The effect of
3470: the contact resistance may be suppressed to a large extent by using narrow
3471: voltage probes attached to the channel itself rather than to wide 2DEG
3472: regions. As we will see in Section \ref{sec16}, such a solution no longer works in the
3473: ballistic transport regime, because of the additional scattering induced\cite{ref289} by
3474: the voltage probes.
3475:
3476: \subsection{\label{sec14} Coherent electron focusing}
3477:
3478: A magnetic field may be used to focus the electrons injected by a point
3479: contact onto a second point contact. Electron focusing in metals was
3480: originally conceived by Sharvin\cite{ref296} as a method to investigate the shape of the
3481: Fermi surface. It has become a powerful tool in the study of surface
3482: scattering\cite{ref343} and the electron-phonon interaction,\cite{ref344} as reviewed in Refs.\
3483: \onlinecite{ref305,ref345}, and \onlinecite{ref346}. The experiment is the analogue in the solid state of
3484: magnetic focusing of electrons in vacuum. Required is a large mean free path
3485: for the carriers at the Fermi surface, to ensure ballistic motion as in vacuum.
3486: The mean free path should be much larger than the separation $L$ of the two
3487: point contacts. Moreover, $L$ should be much larger than the point contact
3488: width $W$, to achieve optimal resolution. In metals, electron focusing is
3489: essentially a {\it classical\/} phenomenon because the Fermi wavelength
3490: $\lambda_{\mathrm{F}}\sim 0.5\,\mathrm{nm}$ is much smaller than both $W\sim 1\,\mu \mathrm{m}$ and $L\sim 100\,\mu \mathrm{m}$. The
3491: ratios $\lambda_{\mathrm{F}}/L$ and $\lambda_{\mathrm{F}}/W$ are much larger in a 2DEG than in a metal, typically by
3492: factors of $10^{4}$ and $10^{2}$, respectively. {\it Coherent\/} electron focusing\cite{ref59,ref80,ref347} is
3493: possible in a 2DEG because of this relatively large value of the Fermi
3494: wavelength, and turns out to be strikingy different from classical electron
3495: focusing in metals.
3496:
3497: Electron focusing can be seen as a transmission experiment in electron
3498: optics (cf.\ Ref.\ \onlinecite{ref3} for a discussion from this point of view). An alternative point
3499: of view (emphasized in Refs.\ \onlinecite{ref80} and \onlinecite{ref348}) is that coherent electron focusing is a
3500: prototype of a nonlocal resistance measurement in the quantum ballistic
3501: transport regime, such as studied extensively in narrow-channel geometries.\cite{ref310} Longitudinal resistances that are negative (not $\pm B$ symmetric) and
3502: dependent on the properties of the current and voltage contacts as well as on
3503: their separation, periodic and aperiodic magnetoresistance oscillations,
3504: absence of local equilibrium are all characteristic features of this transport
3505: regime that appear in a most extreme and bare form in the electron focusing
3506: geometry. One reason for the simplification offered by this geometry is that
3507: the current and voltage contacts, being point contacts, are not nearly as
3508: invasive as the wide leads in a Hall bar geometry. Another reason is that the
3509: electrons interact with only one boundary (instead of two in a narrow
3510: channel).
3511:
3512: The outline of this section is as follows.In Section \ref{sec14a} the experimental
3513: results on coherent electron focusing\cite{ref59,ref80} are presented. A theoretical
3514: description\cite{ref80,ref347} is given in Section \ref{sec14b}, in terms of mode interference in the
3515: waveguide formed by the magnetic field at the 2DEG boundary. Apart from
3516: the intrinsic interest of electron focusing in a 2DEG, the experiment can also
3517: be seen as a method to study electron scattering, as in metals. Two such
3518: applications\cite{ref108,ref349} are discussed in Section \ref{sec14c}. We restrict ourselves in this
3519: section to focusing by a magnetic field. Electrostatic focusing\cite{ref350} is discussed
3520: in Section \ref{sec15b}.
3521:
3522: \subsubsection{\label{sec14a} Experiments}
3523:
3524: \begin{figure}
3525: \centerline{\includegraphics[width=8cm]{figures/fig52}}
3526: \caption{
3527: Illustration of classical electron focusing by a magnetic field. Top: Skipping orbits along the 2DEG boundary. The trajectories are drawn up to the third specular reflection. Bottom: Plot of the caustics, which are the collection of focal points of the trajectories. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
3528: \label{fig52}
3529: }
3530: \end{figure}
3531:
3532: The geometry of the experiment\cite{ref59} in a 2DEG is the transverse focusing
3533: geometry of Tsoi\cite{ref343} and consists of two point contacts on the same boundary
3534: in a perpendicular magnetic field. (In metals one can also use the geometry of
3535: Sharvin\cite{ref296} with opposite point contacts in a longitudinal field. This is not
3536: possible in two dimensions.) Two point contacts and the intermediate 2DEG
3537: boundary are created electrostatically by means of the two split gates shown
3538: in Fig.\ \ref{fig5}b. Figure \ref{fig52} illustrates electron focusing in two dimensions as it
3539: follows from the classical mechanics of electrons at the Fermi level. The
3540: injector (i) injects a divergent beam of electrons ballistically into the 2DEG.
3541: Electrons are detected if they reach the adjacent collector (c), after one or
3542: more specular reflections at the boundary connecting $\mathrm{i}$ and $\mathrm{c}$. (These are the
3543: {\it skipping orbits\/} discussed in Section \ref{sec12a}.) The focusing action of the magnetic
3544: field is evident in Fig.\ \ref{fig52} (top) from the black lines of high density of
3545: trajectories. These lines are known in optics as {\it caustics\/} and they are plotted
3546: separately in Fig.\ \ref{fig52} (bottom). The caustics intersect the 2DEG boundary at
3547: multiples of the cyclotron diameter from the injector. As the magnetic field is
3548: increased, a series of these focal points shifts past the collector. The electron
3549: flux incident on the collector thus reaches a maximum whenever its
3550: separation $L$ from the injector is an integer multiple of $2l_{\mathrm{cycl}}=2\hbar k_{\rm F}/eB$. This
3551: occurs when $B=pB_{\mathrm{focus}}$, $p=1,2, \ldots$, with
3552: \be
3553: B_{\mathrm{focus}}=2\hbar k_{\rm F}/eL. \label{eq14.1}
3554: \ee
3555: For a given injected current $I_{\mathrm{i}}$ the voltage $V_{\mathrm{c}}$ on the collector is proportional
3556: to the incident flux. The classical picture thus predicts a series of equidistant
3557: peaks in the collector voltage as a function of magnetic field.
3558:
3559: \begin{figure}
3560: \centerline{\includegraphics[width=8cm]{figures/fig53}}
3561: \caption{
3562: Bottom: Experimental electron focusing spectrum ($T= 50\,{\rm mK}$, $L= 3.0\,\mu{\rm m}$) in the generalized Hall resistance configuration depicted in the inset. The two traces $a$ and $b$ are measured with interchanged current and voltage leads, and demonstrate the injector-collector reciprocity as well as the reproducibility of the fine structure. Top: Calculated classical focusing spectrum corresponding to the experimental trace $a$ (50-nm-wide point contacts were assumed). The dashed line is the extrapolation of the classical Hall resistance seen in reverse fields. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
3563: \label{fig53}
3564: }
3565: \end{figure}
3566:
3567: In Fig.\ \ref{fig53} (top) we show such a classical focusing spectrum, calculated for
3568: parameters corresponding to the experiment discussed later $(L=3.0\,\mu \mathrm{m}$,
3569: $k_{\mathrm{F}}=1.5\times 10^{8}\,\mathrm{m}^{-1})$. The spectrum consists of equidistant focusing peaks of
3570: approximately equal magnitude superimposed on the Hall resistance (dashed
3571: line). The $p\mathrm{th}$ peak is due to electrons injected perpendicularly to the
3572: boundary that have made $p-1$ specular reflections between injector and
3573: collector. Such a classical focusing spectrum is commonly observed in
3574: metals,\cite{ref351,ref352} albeit with a decreasing height of subsequent peaks because of
3575: partially diffuse scattering at the metal surface. Note that the peaks occur in
3576: one field direction only. In reverse fields the focal points are at the wrong side
3577: of the injector for detection, and the normal Hall resistance is obtained. The
3578: experimental result for a 2DEG is shown in the bottom half of Fig.\ \ref{fig53} (trace a;
3579: trace b is discussed later). A series of five focusing peaks is evident at the
3580: expected positions. The observation of multiple focusing peaks immediately
3581: implies that the electrostatically defined 2DEG boundary scatters predominantly specularly. (This finding\cite{ref59} is supported by the magnetoresistance experiments of Thornton et a.\cite{ref107} in a narrow split-gate channel; cf.\ Section \ref{sec5}.) Figure \ref{fig53} is obtained in a measuring configuration (inset) in which an imaginary line connecting the voltage probes crosses that between the current source and drain. This is the configuration for a generalized Hall resistance measurement. If the crossing is avoided, one measures a longitudinal resistance, which shows the focusing peaks without a superimposed Hall slope. This longitudinal resistance periodically becomes negative. This is a classical result\cite{ref80} of magnetic defocusing, which causes the probability density near the point contact voltage probe to be reduced with respect to the spatially averaged probability density that determines the voltage on the wide voltage probe (cf.\ the regions of reduced density between lines of focus in Fig.\ \ref{fig52}).
3582:
3583: On the experimental focusing peaks a fine structure is resolved at low temperatures (below 1 K). The fine structure is well reproducible but sample-dependent. A nice demonstration of the reproducibility of the fine structure is obtained upon interchanging current and voltage leads, so that the injector becomes the collector, and vice versa. The resulting focusing spectrum shown in Fig.\ \ref{fig53} (trace b) is almost the precise mirror image of the original one (trace a), although this particular device had a strong asymmetry in the widths of injector and collector. The symmetry in the focusing spectra is an example of the general reciprocity relation (\ref{eq12.16}). If one applies the B\"{u}ttiker equations (\ref{eq12.12}) to the electron focusing geometry (as is done in Section \ref{sec19}), one finds that the ratio of collector voltage $V_{\rm c}$ to injector current $I_{\rm i}$ is given by
3584: \be
3585: \frac{V_{\rm c}}{I_{\rm i}}=\frac{2e^{2}}{h}\frac{T_{{\rm i}\rightarrow{\rm c}}}{G_{\rm i}G_{\rm c}},\label{eq14.2}
3586: \ee
3587: where $T_{{\rm i}\rightarrow{\rm c}}$ is the transmission probability from injector to collector, and $G_{\rm i}$ and $G_{\rm c}$ are the conductances of the injector and collector point contact. Since $T_{{\rm i}\rightarrow{\rm c}}(B)=T_{{\rm c}\rightarrow{\rm i}}(-B)$ and $G(B) = G( - B)$, this expression for the focusing spectrum is manifestly symmetric under interchange of injector and collector with reversal of the magnetic field.
3588:
3589:
3590: \begin{figure}
3591: \centerline{\includegraphics[width=8cm]{figures/fig54}}
3592: \caption{
3593: Experimental electron focusing spectrum over a larger field range and for very narrow point contacts (estimated width 20--40 nm; $T= 50\,{\rm mK}$, $L= 1.5\,\mu{\rm m}$). The inset gives the Fourier transform for $B\geq 0.8\,{\rm T}$. The high-field oscillations have the same dominant periodicity as the low-field focusing peaks, but with a much larger amplitude. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
3594: \label{fig54}
3595: }
3596: \end{figure}
3597:
3598: The fine structure on the focusing peaks in Fig.\ \ref{fig53} is the first indication that electron focusing in a 2DEG is qualitatively different from the corresponding experiment in metals. At higher magnetic fields the resemblance to the classical focusing spectrum is lost; see Fig.\ \ref{fig54}. A Fourier transform of the spectrum for $B \geq 0.8\,{\rm T}$ (inset in Fig.\ \ref{fig54}) shows that the large-amplitude highfield oscillations have a dominant periodicity of 0.1 T, which is approximately the same as the periodicity $B_{\rm focus}$ of the much smaller focusing peaks at low magnetic fields ($B_{\rm focus}$ in Fig.\ \ref{fig54} differs from Fig.\ \ref{fig53} because of a smaller
3599: $L=1.5\,\mu \mathrm{m})$. This dominant periodicity can be explained in terms of quantum
3600: interference between the different skipping orbits from injector to collector or
3601: in terms of interference of coherently excited edge channels, as we discuss in
3602: the following subsection. The experimental implication is that the injector
3603: acts as a {\it coherent\/} point source with the coherence maintained over a distance
3604: of several microns to the collector.
3605:
3606: \subsubsection{\label{sec14b} Theory}
3607:
3608: To explain the characteristic features of the coherent electron focusing
3609: experiments we have described, we must go beyond the classical description.\cite{ref80,ref347} As discussed in Section \ref{sec12}, quantum ballistic transport along
3610: the 2DEG boundary in a magnetic field takes place via magnetic edge states,
3611: which form the propagating modes at the Fermi level. Since the injector has a
3612: width below $\lambda_{\mathrm{F}}$, it excites these modes coherently. For $k_{\mathrm{F}}L\gg 1$ the interference of modes at the collector is dominated by their rapidly varying phase
3613: factors $ \exp(ik_{n}L)$. The wave number $k_{n}$ corresponds classically to the
3614: separation of the center of the cyclotron orbit from the 2DEG boundary [Eq.\
3615: (\ref{eq12.5})]. In the Landau gauge ${\bf A} =(0, Bx, 0)$ (with the axis chosen as in Fig.\ \ref{fig52})
3616: one has $k_{n}=k_{\mathrm{F}}\sin\alpha_{n}$, where $\alpha$ is the angle with the $x$-axis under which the
3617: cyclotron orbit is reflected from the boundary. The quantized values $\alpha_{n}$ follow
3618: in this semiclassical description from the Bohr-Sommerfeld quantization
3619: rule (\ref{eq12.6}) that the flux enclosed by the cyclotron orbit and the boundary
3620: equals $(n- \frac{1}{4})h/e$ [the phase shift $\gamma$ in Eq.\ (\ref{eq12.6}) equals $\pi/2$ for an edge state at
3621: an infinite barrier potential]. Simple geometry shows that this requires that
3622: \be
3623: \frac{\pi}{2}-\alpha_{n}-\frac{1}{2}\sin 2\alpha_{n}=\frac{2\pi}{k_{\mathrm{F}}l_{\mathrm{cycl}}}\left(n-\frac{1}{4}\right),\;\; n=1,2, \ldots, N. \label{eq14.3}
3624: \ee
3625:
3626: \begin{figure}
3627: \centerline{\includegraphics[width=8cm]{figures/fig55}}
3628: \caption{
3629: Phase $k_{n}L$ of the edge channels at the collector, calculated from Eq.\ (\ref{eq14.3}). Note the domain of approximately linear $n$-dependence of the phase, responsible for the oscillations with $B_{\rm focus}$-periodicity. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
3630: \label{fig55}
3631: }
3632: \end{figure}
3633:
3634: As plotted in Fig.\ \ref{fig55}, the dependence on $n$ of the phase $k_{n}L$ is close to
3635: linear in a broad interval. This also follows from expansion of Eq.\ (\ref{eq14.3})
3636: around $\alpha_{n}=0$, which gives
3637: \be
3638: k_{n}L= \mathrm{constant}-2\pi n\frac{B}{B_{\mathrm{focus}}}+k_{\mathrm{F}}L\times {\rm order} \left( \frac{N-2n}{N}\right)^{3}. \label{eq14.4}
3639: \ee
3640: If $B/B_{\mathrm{focus}}$ is an integer, a fraction of order $(1/k_{\mathrm{F}}L)^{1/3}$ of the $N$ edge states
3641: interfere constructively at the collector. Because of the $1/3$ power, this is a
3642: substantial fraction even for the large $k_{\mathrm{F}}L\sim 10^{2}$ of the experiment. The
3643: resulting mode interference oscillations with $B_{\mathrm{focus}}$-periodicity can become
3644: much larger than the classical focusing peaks. This has been shown in Refs.\
3645: \onlinecite{ref347} and \onlinecite{ref80}, where the transmission probability $T_{\mathrm{i}\rightarrow \mathrm{c}}$ was calculated in the
3646: WKB approximation with neglect of the finite width of the injector and
3647: detector. From Eq.\ (\ref{eq14.2}) the focusing spectrum is then obtained in the form
3648: \be
3649: \frac{V_{\mathrm{c}}}{I_{\mathrm{i}}}=\frac{h}{2e^{2}}\left|\frac{1}{N}\sum_{n=1}^{N}\mathrm{e}^{ik_{n}L}\right|^{2}, \label{eq14.5}
3650: \ee
3651: which is plotted in Fig.\ \ref{fig56} for parameter values corresponding to the
3652: experimental Fig.\ \ref{fig54}. The inset shows the Fourier transform for $B\geq 0.8\,{\rm T}$.
3653:
3654: \begin{figure}
3655: \centerline{\includegraphics[width=8cm]{figures/fig56}}
3656: \caption{
3657: Focusing spectrum calculated from Eq.\ (\ref{eq14.5}), for parameters corresponding to the experimental Fig.\ \ref{fig54}. The inset shows the Fourier transform for $B\geq 0.8\,{\rm T}$. Infinitesimally small point contact widths are assumed in the calculation. Taken from C. W. J. Beenakker et al., Festk\"{o}rperprobleme {\bf 9}, 299 (1989).
3658: \label{fig56}
3659: }
3660: \end{figure}
3661:
3662: There is no detailed one-to-one correspondence between the experimental
3663: and theoretical spectra. No such correspondence was to be expected in view
3664: of the sensitivity of the experimental spectrum to small variations in the
3665: voltage on the gate defining the point contacts and the 2DEG boundary.
3666: Those features of the experimental spectrum that are insensitive to the precise
3667: measurement conditions are, however, well reproduced by the calculation:
3668: We recognize in Fig.\ \ref{fig56} the low-field focusing peaks and the large-amplitude
3669: high-field oscillations with the same $B_{\mathrm{focus}}$-periodicity. The high-field oscillations range from about 0 to $10\, \mathrm{k}\Omega$ in both theory and experiment. The
3670: maximum amplitude is not far below the theoretical upper bound of
3671: $h/2e^{2}\approx 13\,\mathrm{k}\Omega$, which follows from Eq.\ (\ref{eq14.5}) if we assume that {\it all\/} the modes
3672: interfere constructively. This indicates that a {\it maximal phase coherence\/} is
3673: realized in the experiment and implies that the experimental injector and
3674: collector point contacts resemble the idealized point source detector in the
3675: calculation.
3676:
3677: \subsubsection{\label{sec14c} Scattering and electron focusing}
3678:
3679:
3680: \begin{figure}
3681: \centerline{\includegraphics[width=8cm]{figures/fig57}}
3682: \caption{
3683: Experimental electron focusing spectra (in the generalized longitudinal resistance configuration) at 0.3 K for five different injector-collector separations in a very high mobility material. The vertical scale varies among the curves. Taken from J. Spector et al., Surf.\ Sci.\ {\bf 228}, 283 (1990).
3684: \label{fig57}
3685: }
3686: \end{figure}
3687:
3688: Scattering events other than specular boundary scattering can be largely
3689: ignored for the relatively small point contact separations $L\leq 3\,\mu \mathrm{m}$ in the
3690: experiments discussed earlier.\cite{ref59,ref80} (any other inelastic or elastic scattering
3691: events would have been detected as a reduction of the oscillations with $B_{\mathrm{focus}}$-periodicity below the theoretical estimate). Spector et al.\cite{ref349} have repeated the
3692: experiments for larger $L$ to study scattering processes in an ultrahigh
3693: mobility\cite{ref353,ref354} 2DEG ($\mu_{\mathrm{e}}=5.5\times 10^{6}\,\mathrm{cm}^{2}/\mathrm{V}\mathrm{s}$). They used relatively wide
3694: point contacts (about 1 $\mu \mathrm{m}$) so that electron focusing was in the classical
3695: regime. In Fig.\ \ref{fig57} we reproduce their experimental results for point contact
3696: separations up to 64 $\mu \mathrm{m}$. The peaks in the focusing spectrum for a given $L$
3697: have a roughly constant amplitude, indicating that scattering at the boundary is mostly specular rather than diffusive --- in agreement with the experiments of Ref.\ \onlinecite{ref59}. Spector et al.\cite{ref349} find that the amplitude of the focusing
3698: peaks decreases exponentially with increasing $L$, due to scattering in the
3699: electron gas (see Fig.\ \ref{fig58}). The decay $\exp(- L/L_{0})$ with $L_{0}\approx 10\,\mu \mathrm{m}$ implies an
3700: effective mean free path (measured along the arc of the skipping orbits) of
3701: $L_{0}\pi/2\approx 15\,\mu \mathrm{m}$. This is smaller than the transport mean free path derived
3702: from the conductivity by about a factor of 2, which may point to a greater
3703: sensitivity of electron focusing to forward scattering.
3704:
3705:
3706: \begin{figure}
3707: \centerline{\includegraphics[width=8cm]{figures/fig58}}
3708: \caption{
3709: Exponential decay of the oscillation amplitude of the collector voltage (normalized by the injector voltage) as a function of injector-collector separation $d$ (denoted by $L$ in the text). Taken from J. Spector et al., Surf.\ Sci.\ {\bf 228}, 283 (1990).
3710: \label{fig58}
3711: }
3712: \end{figure}
3713:
3714: Electron focusing by a magnetic field may also play a role in geometries
3715: other than the double-point contact geometry of Fig.\ \ref{fig52}. One example is
3716: mentioned in the context of junction scattering in a cross geometry in Section
3717: \ref{sec16}. Another example is the experiment by Nakamura et al.\cite{ref108} on the
3718: magnetoresistance of equally spaced narrow channels in parallel (see Fig.\ \ref{fig59}).
3719: Resistance peaks occur in this experiment when electrons that are transmitted through one of the channels are focused back through another
3720: channel. The resistance peaks occur at $B=(n/m)B_{\mathrm{focus}}$, where $B_{\mathrm{focus}}$ is given
3721: by Eq.\ (\ref{eq14.1}) with $L$ the spacing of adjacent channels. The identification of the
3722: various peaks in Fig.\ \ref{fig59} is given in the inset. Nakamura et al.\cite{ref108} conclude
3723: from the rapidly diminishing height of consecutive focusing peaks (which
3724: require an increasing number of specular reflections) that there is a large
3725: probability of diffuse boundary scattering. The reason for the difference with
3726: the experiments discussed previously is that the boundary in the experiment
3727: of Fig.\ \ref{fig59} is defined by focused ion beam lithography, rather than electrostatically by means of a gate. As discussed in Section \ref{sec5}, the former technique may
3728: introduce a considerable boundary roughness.
3729:
3730: \begin{figure}
3731: \centerline{\includegraphics[width=8cm]{figures/fig59}}
3732: \caption{
3733: Magnetoresistance of $N$ constrictions in parallel at 1.3 K. The arrows indicate the oscillations due to electron focusing, according to the mechanisms illustrated in the inset. The resistance scale is indicated by $10\,\Omega$ bars. Taken from K. Nakamura et al., Appl.\ Phys.\ Lett.\ {\bf 56}, 385 (1990).
3734: \label{fig59}
3735: }
3736: \end{figure}
3737:
3738: Electron focusing has been used by Williamson et al.\cite{ref355} to study scattering
3739: processes for ``hot'' electrons, with an energy in excess of the Fermi energy,
3740: and for ``cool'' holes, or empty states in the conduction band below the Fermi
3741: level (see Ref.\ \onlinecite{ref307} for a review). An interesting aspect of hot-electron focusing
3742: is that it allows a measurement of the local electrostatic potential drop across
3743: a current-carrying quantum point contact,\cite{ref355} something that is not possible
3744: using conventional resistance measurements, where the sum of electrostatic
3745: and chemical potentials is measured. The importance of such alternative
3746: techniques to study electrical conduction has been stressed by Landauer.\cite{ref356}
3747:
3748: \subsection{\label{sec15} Collimation}
3749:
3750: The subject of this section is the collimation of electrons injected by a
3751: point contact\cite{ref329} and its effect on transport measurements in geometries
3752: involving two opposite point contacts.\cite{ref327,ref357} Collimation (i.e., the narrowing
3753: of the angular injection distributions) follows from the constraints on the
3754: electron momentum imposed by the potential barrier in the point contact
3755: ({\it barrier collimation}), and by the gradual widening of the point contact at its
3756: entrance and exit ({\it horn collimation}). We summarize the theory in Section \ref{sec15a}.
3757: The effect was originally proposed\cite{ref329} to explain the remarkable observation
3758: of Wharam et al.\cite{ref357} that the series resistance of two opposite point contacts is
3759: considerably less than the sum of the two individual resistances (Section \ref{sec15c}).
3760: A direct experimental proof of collimation was provided by Molenkamp et
3761: al.,\cite{ref327} who measured the deflection of the injected beam of electrons in a
3762: magnetic field (Section \ref{sec15b}). A related experiment by Sivan et al.,\cite{ref350} aimed at
3763: the demonstration of the focusing action of an electrostatic lens, is also
3764: discussed in this subsection. The collimation effect has an importance in
3765: ballistic transport that goes beyond the point contact geometry. It will be
3766: shown in Section \ref{sec16} that the phenomenon is at the origin of a variety of
3767: magnetoresistance anomalies in narrow multiprobe conductors.\cite{ref358,ref359,ref360}
3768:
3769: \subsubsection{\label{sec15a} Theory}
3770:
3771: \begin{figure}
3772: \centerline{\includegraphics[width=8cm]{figures/fig60}}
3773: \caption{
3774: Illustration of the collimation effect for an abrupt constriction (a) containing a potential barrier of height $E_{\rm c}$ and for a horn-shaped constriction (b) that is flared from a width $W_{\rm min}$ to $W_{\rm max}$. The dash-dotted trajectories approaching at an angle $\alpha$ outside the injection/acceptance cone are reflected. Taken from H. van Houten and C. W. J. Beenakker, in ``Nanostructure Physics and Fabrication'' (M. Reed and W. P. Kirk, eds.). Academic, New York, 1989.
3775: \label{fig60}
3776: }
3777: \end{figure}
3778:
3779: Since collimation follows from classical mechanics, a semiclassical theory
3780: is sufficient to describe the essential phenomena, as we now discuss (following
3781: Refs.\ \onlinecite{ref329} and \onlinecite{ref311}). Semiclassically, collimation results from the adiabatic
3782: invariance of the product of channel width $W$ and absolute value of the
3783: transverse momentum $\hbar k_{y}$ (this product is proportional to the action for
3784: motion transverse to the channel\cite{ref361}). Therefore, if the electrostatic potential
3785: in the point contact region is sufficiently smooth, the quantity $S=|k_{y}|W$ is
3786: approximately constant from point contact entrance to exit. Note that $S/\pi$
3787: corresponds to the quantum mechanical 1D subband index $n$. The quantum
3788: mechanical criterion for adiabatic transport was derived by Yacoby and
3789: Imry\cite{ref326} (see Section \ref{sec13}). As was discussed there, adiabatic transport breaks
3790: down at the exit of the point contact, where it widens abruptly into a 2DEG
3791: of essentially infinite width. Collimation reduces the {\em injection/acceptance cone\/}
3792: of the point contact from its original value of $\pi$ to a value of $2\alpha_{\max}$. This effect
3793: is illustrated in Fig.\ \ref{fig60}. Electrons incident at an angle $|\alpha|>\alpha_{\max}$ from normal
3794: incidence are reflected. (The geometry of Fig.\ \ref{fig60}b is known in optics as a
3795: {\it conical\/} reflector.\cite{ref362}.) Vice versa, all electrons leave the constriction at an angle
3796: $|\alpha|<\alpha_{\max}$ (i.e., the injected electrons form a collimated beam of angular
3797: opening $2\alpha_{\max}$).
3798:
3799: To obtain an analytic expression for the collimation effect, we describe the
3800: shape of the potential in the point contact region by three parameters: $W_{\min}$,
3801: $W_{\max}$, and $E_{\mathrm{c}}$ (see Fig.\ \ref{fig60}). We consider the case that the point contact has its
3802: minimal width $W_{\min}$ at the point where the barrier has its maximal height $E_{\mathrm{c}}$
3803: above the bottom of the conduction band in the broad regions. At that point
3804: the largest possible value of $S$ is
3805: \[
3806: S_{1}\equiv(2m/\hbar^{2})^{1/2}(E_{\mathrm{F}}-E_{\mathrm{c}})^{1/2}W_{\min}.
3807: \]
3808: We assume that adiabatic transport (i.e., $S=$ constant) holds up to a point of
3809: zero barrier height and maximal width $W_{\max}$. The abrupt separation of
3810: adiabatic and nonadiabatic regions is a simplification that can be, and has
3811: been, tested by numerical calculations (see below). At the point contact exit,
3812: the largest possible value of $S$ is
3813: \[
3814: S_{2}\equiv(2m/\hbar^{2})^{1/2}(E_{\mathrm{F}})^{1/2}\sin\alpha_{\max}\,W_{\max}.
3815: \]
3816: The invariance of $S$ implies that $S_{1}=S_{2}$; hence,
3817: \be
3818: \alpha_{\max}=\arcsin\left(\frac{1}{f}\right);\;\; f \equiv\left(\frac{E_{\mathrm{F}}}{E_{\mathrm{F}}-E_{\mathrm{c}}}\right)^{1/2}\frac{W_{\max}}{W_{\min}}. \label{eq15.1}
3819: \ee
3820: The {\it collimation factor\/} $f\geq 1$ is the product of a term describing the
3821: collimating effect of a barrier of height $E_{\mathrm{c}}$ (barrier collimation) and a term
3822: describing collimation due to a gradual widening of the point contact width
3823: from $W_{\min}$ to $W_{\max}$ (horn collimation). In the adiabatic approximation, the
3824: angular injection distribution $P(\alpha)$ is proportional to $\cos\alpha$ with an abrupt
3825: truncation at $\pm\alpha_{\max}$. The cosine angular dependence follows from the cosine
3826: distribution of the incident flux in combination with time-reversal symmetry
3827: and is thus not affected by the reduction of the injection/acceptance cone.
3828: We therefore conclude that in the adiabatic approximation $P(\alpha)$ (normalized
3829: to unity) is given by
3830: \be
3831: P( \alpha)=\left\{\begin{array}{ll}
3832: \frac{1}{2}f\cos\alpha&{\rm if}\;\;|\alpha|<\arcsin(1/f),\\
3833: =0,&{\rm otherwise}.
3834: \end{array}\right. \label{eq15.2}
3835: \ee
3836: We defer to Section \ref{sec15b} a comparison of the analytical result (\ref{eq15.2}) with a
3837: numerical calculation.
3838:
3839: Barrier collimation does not require adiabaticity. For an abrupt barrier,
3840: collimation simply results from transverse momentum conservation, as in
3841: Fig.\ \ref{fig60}a, leading directly to Eq.\ (\ref{eq15.2}). (The total external reflection at an
3842: abrupt barrier for trajectories outside the collimation cone is similar to the
3843: optical effect of total internal reflection at a boundary separating a region of
3844: high refractive index from a region of small refractive index; see the end of
3845: Section \ref{sec15b}.) A related collimation effect resulting from transverse momentum conservation occurs if electrons tunnel through a potential barrier.
3846: Since the tunneling probability through a high potential barrier is only
3847: weakly dependent on energy, it follows that the strongest collimation is to be
3848: expected if the barrier height equals the Fermi energy. On lowering the
3849: barrier below $E_{\mathrm{F}}$ ballistic transport over the barrier dominates, and the
3850: collimation cone widens according to Eq.\ (\ref{eq15.2}). A quantum mechanical
3851: calculation of barrier collimation may be found in Ref.\ \onlinecite{ref363}.
3852:
3853: The injection distribution (\ref{eq15.2}) can be used to obtain (in the semiclassical
3854: limit) the direct transmission probability $T_{\mathrm{d}}$ between two opposite identical
3855: point contacts separated by a large distance $L$. To this end, first note that
3856: $T_{\mathrm{d}}/N$ is the fraction of the injected current that reaches the opposite point
3857: contact (since the transmission probability through the first point contact is
3858: $N$, for $N$ occupied subbands in the point contact). Electrons injected within a
3859: cone of opening angle $W_{\max}/L$ centered at $\alpha=0$ reach the opposite point
3860: contact and are transmitted. If this opening angle is much smaller than the
3861: total opening angle $2\alpha_{\max}$ of the beam, then the distribution function $P(\alpha)$ can
3862: be approximated by $P(0)$ within this cone. This approximation requires
3863: $W_{\max}/L\ll 1/f$, which is satisfied experimentally in devices with a sufficiently
3864: large point contact separation. We thus obtain $T_{\mathrm{d}}/N=P(0)W_{\max}/L$, which,
3865: using Eq.\ (\ref{eq15.2}), can be written as\cite{ref329}
3866: \be
3867: T_{\mathrm{d}}=f(W_{\max}/2L)N. \label{eq15.3}
3868: \ee
3869: This simple analytical formula can be used to describe the experiments on
3870: transport through identical opposite point contacts in terms of one empirical
3871: parameter $f$, as discussed in the following subsections.
3872:
3873: \subsubsection{\label{sec15b} Magnetic deflection of a collimated electron beam}
3874:
3875: \begin{figure}
3876: \centerline{\includegraphics[width=8cm]{figures/fig61}}
3877: \caption{
3878: Detection of a collimated electron beam over a distance of $4\,\mu{\rm m}$. In this four-terminal measurement, two ohmic contacts to the 2DEG region between the point contacts are used: One of these acts as a drain for the current $I_{\rm i}$ through the injector, and the other is used as a zeroreference for the voltage $V_{\rm c}$ on the collector. The drawn curve is the experimental data at $T = 1.8\,{\rm K}$. The black dots are the result of a semiclassical simulation, using a hard-wall potential with contours as shown in the inset. The dashed curve results from a simulation without collimation (corresponding to rectangular corners in the potential contour). Taken from L. W. Molenkamp et al., Phys.\ Rev.\ B {\bf 41}, 1274 (1990).
3879: \label{fig61}
3880: }
3881: \end{figure}
3882:
3883: A method\cite{ref311,ref329} to sensitively detect the collimated electron beam
3884: injected by a point contact is to sweep the beam past a second opposite point
3885: contact by means of a magnetic field. The geometry is shown in Fig.\ \ref{fig61} (inset).
3886: The current $I_{\mathrm{i}}$ through the injecting point contact is drained to ground at one
3887: or two (the difference is not essential) ends of the 2DEG channel separating
3888: the point contacts. The opposite point contact, the collector, serves as a
3889: voltage probe (with the voltage $V_{\mathrm{c}}$ being measured relative to ground). In the
3890: case that both ends of the 2DEG channel are grounded, the collector voltage
3891: divided by the injected current is given by
3892: \be
3893: \frac{V_{\mathrm{c}}}{I_{\mathrm{i}}}=\frac{1}{G}\frac{T_{\mathrm{d}}}{N},\;\; T_{\mathrm{d}}\ll N, \label{eq15.4}
3894: \ee
3895: with $G=(2e^{2}/h)N$ the two-terminal conductance of the individual point
3896: contact (both point contacts are assumed to be identical) and $T_{\mathrm{d}}$ the direct
3897: transmission probability between the two point contacts calculated in
3898: Section \ref{sec15a}. Equation (\ref{eq15.4}) can be obtained from the Landauer-B\"{u}ttiker
3899: formalism (as done in Ref.\ \onlinecite{ref311}) or simply by noting that the current $I_{\mathrm{i}}T_{\mathrm{d}}/N$
3900: incident on the collector has to be counterbalanced by an equal outgoing
3901: current $GV_{\mathrm{c}}$. In the absence of a magnetic field, we obtain [using Equation
3902: (\ref{eq15.3}) for the direct transmission probability]
3903: \be
3904: \frac{V_{\mathrm{c}}}{I_{\mathrm{i}}}=\frac{h}{2e^{2}}f^{2}\frac{\pi}{2k_{\mathrm{F}}L}, \label{eq15.5}
3905: \ee
3906: where $k_{\mathrm{F}}$ is the Fermi wave vector in the region between the point contacts. In
3907: an experimental situation $L$ and $k_{\mathrm{F}}$ are known, so the collimation factor $f$ can
3908: be directly determined from the collector voltage by means of Eq.\ (\ref{eq15.5}).
3909:
3910: The result (\ref{eq15.5}) holds in the absence of a magnetic field. A small magnetic
3911: field $B$ will deflect the collimated electron beam past the collector. Simple
3912: geometry leads to the criterion $L/2l_{\mathrm{cycl}}=\alpha_{\max}$ for the cyclotron radius at
3913: which $T_{\mathrm{d}}$ is reduced to zero by the Lorentz force (assuming that $L\gg W_{\max}$).
3914: One would thus expect to see in $V_{\mathrm{c}}/I_{\mathrm{i}}$ a peak around zero field, of height given
3915: by Eq.\ (\ref{eq15.5}) and of width
3916: \be
3917: \Delta B=(4\hbar k_{\rm F}/eL)\arcsin(1/f), \label{eq15.6}
3918: \ee
3919: according to Eq.\ (\ref{eq15.1}).
3920:
3921: In Fig.\ \ref{fig61} this collimation peak is shown (solid curve), as measured by
3922: Molenkamp et al.\cite{ref327} at $T=1.2\,\mathrm{K}$ in a device with a $L=4.0$-$\mu \mathrm{m}$ separation
3923: between injector and collector. In this measurement only one end of the
3924: region between the point contacts was grounded --- a measurement configuration referred to in narrow Hall bar geometries as a {\it bend resistance\/}
3925: measurement\cite{ref289,ref364} (cf.\ Section \ref{sec16}). One can show, using the Landauer-B\"{u}ttiker formalism,\cite{ref5} that the height of the collimation peak is still given by
3926: Eq.\ (\ref{eq15.5}) if one replaces\cite{ref327} $f^{2}$ by $f^{2}- \frac{1}{2}$. The expression (\ref{eq15.6}) for the width
3927: is not modified. The experimental result in Fig.\ \ref{fig61} shows a peak height of
3928: $\approx 150\,\Omega$ (measured relative to the background resistance at large magnetic
3929: fields). Using $L=4.0\,\mu \mathrm{m}$ and the value $k_{\mathrm{F}}=1.1\times 10^{8}\,\mathrm{m}^{-1}$ obtained from
3930: Hall resistance measurements in the channel between the point contacts, one
3931: deduces a collimation factor $f\approx 1.85$. The corresponding opening angle of the
3932: injection/acceptance cone is $2\alpha_{\max}\approx 65^{\circ}$. The calculated value of $f$ would
3933: imply a width $\Delta B\approx 0.04\,\mathrm{T}$, which is not far from the measured full width at
3934: half maximum of $\approx 0.03\,{\rm T}$.
3935:
3936: \begin{figure}
3937: \centerline{\includegraphics[width=8cm]{figures/fig62}}
3938: \caption{
3939: Calculated angular injection distributions in zero magnetic field. The solid histogram is the result of a simulation of the classical trajectories at the Fermi energy in the geometry shown in the inset of Fig.\ \ref{fig61}. The dotted curve follows from the adiabatic approximation (\ref{eq15.2}), with the experimental collimation factor $f = 1.85$. The dashed curve is the cosine distribution in the absence of any collimation. Taken from L. W. Molenkamp et al., Phys.\ Rev.\ B. {\bf 41}, 1274 (1990).
3940: \label{fig62}
3941: }
3942: \end{figure}
3943:
3944: The experimental data in Fig.\ \ref{fig61} are compared with the result\cite{ref327} from a
3945: numerical simulation of classical trajectories of the electrons at the Fermi
3946: level (following the method of Ref.\ \onlinecite{ref329}). This semiclassical calculation was
3947: performed in order to relax the assumption of adiabatic transport in the point
3948: contact region, and of small $T_{\mathrm{d}}/N$, on which Eqs.\ (\ref{eq15.3}) and (\ref{eq15.5}) are based.
3949: The dashed curve is for point contacts defined by hardwall contours with
3950: straight corners (no collimation); the dots are for the smooth hardwall
3951: contours shown in the inset, which lead to collimation via the horn effect (cf.\
3952: Fig.\ \ref{fig60}b; the barrier collimation of Fig.\ \ref{fig60}a is presumably unimportant at the
3953: small gate voltage used in the experiment and is not taken into account in the
3954: numerical simulation). The angular injection distributions $P(\alpha)$ that follow
3955: from these numerical simulations are compared in Fig.\ \ref{fig62} (solid histogram)
3956: with the result (\ref{eq15.2}) from the adiabatic approximation for $f=1.85$ (dotted
3957: curve). The uncollimated distribution $P(\alpha)=(\cos\alpha)/2$ is also shown for
3958: comparison (dashed curve). Taken together, Figs.\ \ref{fig61} and \ref{fig62} unequivocally
3959: demonstrate the importance of collimation for the transport properties, as
3960: well as the adequateness of the adiabatic approximation as an estimator of
3961: the collimation cone.
3962:
3963: Once the point contact width becomes less than a wavelength, diffraction
3964: inhibits collimation of the electron beam. In the limit $k_{\mathrm{F}}W\ll 1$, the injection
3965: distribution becomes proportional to $\cos^{2}\alpha$ for all $\alpha$, independent of the
3966: shape of the potential in the point contact region.\cite{ref80,ref313} The coherent electron
3967: focusing experiments\cite{ref59,ref80} discussed in Sections \ref{sec14a} and \ref{sec14b} were performed
3968: in this limit.
3969:
3970: \begin{figure}
3971: \centerline{\includegraphics[width=6cm]{figures/fig63}}
3972: \caption{
3973: Electrostatic focusing onto a collector ($C_{2}$) of an injected electron beam (at i) by means of a concave lens corresponding to a region of reduced electron density. Focusing in such an arrangement was detected experimentally.\cite{ref350}
3974: \label{fig63}
3975: }
3976: \end{figure}
3977:
3978: We conclude this subsection by briefly discussing an alternative way to
3979: increase the transmission probability between two opposite point contacts,
3980: which is {\it focusing\/} of the injected electron beam onto the collector. Magnetic
3981: focusing, discussed in Section \ref{sec14} for adjacent point contacts, cannot be used
3982: for opposite point contacts in two dimensions (unlike in three dimensions,
3983: where a magnetic field along the line connecting the point contacts will focus
3984: the beam\cite{ref296}). A succesful demonstration of {\it electrostatic\/} focusing was recently
3985: reported by Sivan et al.\ and by Spector et al.\cite{ref350} The focusing is achieved by
3986: means of a potential barrier of a concave shape, created as a region of reduced
3987: density in the 2DEG by means of a gate between the injector and the collector
3988: (see Fig.\ \ref{fig63}). A focusing lens for electrons is {\it concave\/} because electrons
3989: approaching a potential barrier are deflected in a direction perpendicular to
3990: the normal. This is an amusing difference with light, which is deflected toward
3991: the normal on entering a more dense medium, so an optical focusing lens is
3992: {\it convex}. The different dispersion laws are the origin of this different behavior
3993: of light and electrons.\cite{ref350}
3994:
3995: \subsubsection{\label{sec15c} Series resistance}
3996:
3997: The first experimental study of ballistic transport through two opposite
3998: point contacts was carried out by Wharam et al.,\cite{ref357} who discovered that the
3999: series resistance is considerably less than the sum of the two individual
4000: resistances. Sugsequent experiments confirmed this result.\cite{ref365,ref366} The
4001: theoretical explanation\cite{ref329} of this observation is that collimation of the
4002: electrons injected by a point contact enhances the direct transmission
4003: probability through the opposite point contact, thereby significantly reducing the series resistance below its ohmic value. We will discuss the transport
4004: and magnetotransport in this geometry. We will not consider the alternative
4005: geometry of two adjacent point contacts in parallel (studied in Refs.\ \onlinecite{ref367,ref368,ref369}).
4006: In that geometry the collimation effect cannot enhance the coupling of
4007: the two point contacts, so only small deviations from Ohm's law are to be
4008: expected.
4009:
4010: \begin{figure}
4011: \centerline{\includegraphics[width=8cm]{figures/fig64}}
4012: \caption{
4013: Magnetic field dependence of the series conductance of two opposite point contacts (measured as shown in the inset; the point contact separation is $L = 1.0\,\mu{\rm m}$) for three different values of the gate voltage (solid curves) at $T = 100\,{\rm mK}$. For clarity, subsequent curves from bottom to top are offset by $0.5 \times 10^{-4}\,\Omega^{-1}$, with the lowest curve shown at its actual value. The dotted curves are calculated from Eqs.\ (\ref{eq15.10}) and (\ref{eq10.8}), with the point contact width as adjustable parameter. Taken from A. A. M. Staring et al., Phys.\ Rev.\ B. {\bf 41}, 8461 (1990).
4014: \label{fig64}
4015: }
4016: \end{figure}
4017:
4018: The expression for the two-terminal series resistance of two identical
4019: opposite point contacts in terms of the direct transmission probability can be
4020: obtained from the Landauer-B\"{u}ttiker formalism,\cite{ref5} as was done in Ref.\ \onlinecite{ref329}.
4021: We give here an equivalent, somewhat more intuitive derivation. Consider
4022: the geometry shown in Fig.\ \ref{fig64} (inset). A fraction $T_{\mathrm{d}}/N$ of the current $GV$
4023: injected through the first point contact by the current source is directly
4024: transmitted through the second point contact (and then drained to ground).
4025: Here $G=(2e^{2}/h)N$ is the conductance of the individual point contact, and $V$
4026: is the source-drain voltage. The remaining fraction $1 - T_{\mathrm{d}}/N$ equilibrates in
4027: the region between the point contacts, as a result of inelastic scattering (elastic
4028: scattering is sufficient if phase coherence does not play a role). Since that
4029: region cannot drain charge (the attached contacts are not connected to
4030: ground), these electrons will eventually leave via one of the two point
4031: contacts. For a symmetric structure we may assume that the fraction
4032: $\frac{1}{2}(1-T_{\mathrm{d}}/N)$ of the injected current $GV$ is transmitted through the second
4033: point contact after equilibration. The total source-drain current $I$ is the sum
4034: of the direct and indirect contributions:
4035: \be
4036: I={\textstyle\frac{1}{2}}(1+T_{\mathrm{d}}/N)GV. \label{eq15.7}
4037: \ee
4038: The series conductance $G_{\mathrm{series}}=I/V$ becomes
4039: \be
4040: G_{\mathrm{series}}= {\textstyle\frac{1}{2}}G(1+T_{\mathrm{d}}/N). \label{eq15.8}
4041: \ee
4042: In the absence of direct transmission $(T_{\mathrm{d}}=0)$, one recovers the ohmic
4043: addition law for the resistance, as expected for the case of complete
4044: intervening equilibration (cf.\ the related analysis by B\"{u}ttiker of tunneling in
4045: series barriers\cite{ref370,ref371}). At the opposite extreme, if all transmission is
4046: direct ($T_{\mathrm{d}}=N$), the series conductance is identical to that of the single point
4047: contact. Substituting (\ref{eq15.3}) into Eq.\ (\ref{eq15.8}), we obtain the result\cite{ref329} for small
4048: but nonzero direct transmission:
4049: \be
4050: G_{\mathrm{series}}= {\textstyle\frac{1}{2}}G(1+f(W_{\max}/2L)). \label{eq15.9}
4051: \ee
4052:
4053: The quantized plateaus in the series resistance, observed experimentally,\cite{ref357} are of course not obtained in the semiclassical calculation leading to
4054: Eq.\ (\ref{eq15.9}). However, since the nonadditivity is essentially a semiclassical
4055: collimation effect, the present analysis should give a reasonably reliable
4056: estimate of deviations from additivity for not too narrow point contacts. For
4057: a comparison with experiments we refer to Refs.\ \onlinecite{ref307} and \onlinecite{ref329}. A fully
4058: quantum mechanical calculation of the series resistance has been carried out
4059: numerically by Baranger (reported in Ref.\ \onlinecite{ref306}) for two closely spaced
4060: constrictions.
4061:
4062: So far we have only considered the case of a zero magnetic field. In a weak
4063: magnetic field $(2l_{\mathrm{cycl}}>L)$ the situation is rather complicated. As discussed in
4064: detail in Ref.\ \onlinecite{ref329}, there are two competing effects in weak fields: On the one
4065: hand, the deflection of the electron beam by the Lorentz force reduces the
4066: direct transmission probability, with the effect of decreasing the series
4067: conductance. On the other hand, the magnetic field enhances the indirect
4068: transmission, with the opposite effect. The result is an initial {\it decrease\/} in the
4069: series conductance for small magnetic fields in the case of strong collimation
4070: and an {\it increase\/} in the case of weak collimation. This is expected to be a
4071: relatively small effect compared with the effects at stronger fields that are
4072: discussed below.
4073:
4074: In stronger fields $(2l_{\mathrm{cycl}}<L)$, the direct transmission probability vanishes,
4075: which greatly simplifies the situation. If we assume that all transmission
4076: between the opposite point contacts is with intervening equilibration, then
4077: the result is\cite{ref329}
4078: \be
4079: G_{\mathrm{series}}= \frac{2e^{2}}{h}\left(\frac{2}{N}-\frac{1}{N_{\mathrm{wide}}}\right)^{-1}. \label{eq15.10}
4080: \ee
4081: Here $N$ is the ($B$-dependent) number of occupied subbands in the point
4082: contacts, and $N_{\mathrm{wide}}$ is the number of occupied Landau levels in the 2DEG
4083: between the point contacts. The physical origin of the simple addition rule
4084: (\ref{eq15.10}) is additivity of the four-terminal longitudinal resistance (\ref{eq13.7}). From
4085: this additivity it follows that for $n$ different point contacts in series, Eq.\ (\ref{eq15.10})
4086: generalizes to
4087: \be
4088: \frac{1}{G_{\mathrm{series}}}-\frac{h}{2e^{2}}\frac{1}{N_{\mathrm{wide}}}=\sum_{i=1}^{n}R_{\mathrm{L}}(i), \label{eq15.11}
4089: \ee
4090: where
4091: \be
4092: R_{\mathrm{L}}(i)=\left( \frac{h}{2e^{2}}\right)\left(\frac{1}{N_{i}}-\frac{1}{N_{\mathrm{wide}}}\right) \label{eq15.12}
4093: \ee
4094: is the four-terminal longitudinal resistance of point contact $i$. Equation
4095: (\ref{eq15.10}) predicts a nonmonotonic $B$-dependence for $G_{\mathrm{series}}$. This can most
4096: easily be seen by disregarding the discreteness of $N$ and $N_{\mathrm{wide}}$. We then have
4097: $N_{\mathrm{L}}\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}$, while the magnetic field dependence of $N$ (for a square-well
4098: confining potential in the point contacts) is given by Eq.\ (\ref{eq10.8}). The resulting
4099: $B$-dependence of $G_{\mathrm{series}}$ is shown in Fig.\ \ref{fig64} (dotted curves). The nonmonotonic behavior is due to the delayed depopulation of subbands in the point
4100: contacts compared with the broad 2DEG. While the number of occupied
4101: Landau levels $N_{\mathrm{wide}}$ in the region between the point contacts decreases
4102: steadily with $B$ for $2l_{\mathrm{cycl}}<L$, the number $N$ of occupied subbands in the
4103: point contacts remains approximately constant until $2l_{\mathrm{c},\min}\approx W_{\min}$, with
4104: $l_{\mathrm{c},\min}\equiv l_{\mathrm{cycl}}(1-E_{\mathrm{c}}/E_{\mathrm{F}})^{1/2}$ denoting the cyclotron radius in the point contact
4105: region. In this field interval $G_{\mathrm{series}}$ increases with $B$, according to Eq.\ (\ref{eq15.10}).
4106: For stronger fields, depopulation in the point contacts begins to dominate
4107: $G_{\mathrm{series}}$, leading finally to a decreasing conductance (as is the rule for single
4108: point contacts; see Section \ref{sec13b}). The peak in $G_{\mathrm{series}}$ thus occurs at
4109: $2l_{\mathrm{c},\min}\approx W_{\min}$.
4110:
4111: The remarkable camelback shape of $G_{\mathrm{series}}$ versus $B$ predicted by Eq.\
4112: (\ref{eq15.10}) has been observed experimentally by Staring et al.\cite{ref372} The data are
4113: shown in Fig.\ \ref{fig64} (solid curves) for three values of the gate voltage $V_{\mathrm{g}}$ at
4114: $T=100\,\mathrm{mK}$. The measurement configuration is as shown in the inset, with a
4115: point contact separation $L=1.0\,\mu \mathrm{m}$. The dotted curves in Fig.\ \ref{fig64} are the
4116: result of a one-parameter fit to the theoretical expression. It is seen that Eq.\
4117: (\ref{eq15.10}) provides a good description of the overall magnetoresistance behavior
4118: from low magnetic fields up to the quantum Hall effect regime. The
4119: additional structure in the experimental curves has several different origins,
4120: for which we refer to the paper by Staring et al.\cite{ref372} Similar structure in the
4121: two-terminal resistance of a single point contact will be discussed in detail in
4122: Section \ref{sec21}.
4123:
4124: We emphasize that Eq.\ (\ref{eq15.10}) is based on the assumption of complete
4125: equilibration of the current-carrying edge states in the region between the
4126: point contacts. In a quantizing magnetic field, local equilibrium is reached by
4127: inter-Landau level scattering. If the potential landscape (both in the point
4128: contacts themselves and in the 2DEG region in between) varies by less than
4129: the Landau level separation $\hbar\omega_{\mathrm{c}}$ on the length scale of the magnetic length
4130: $(\hbar/eB)^{1/2}$, then inter-Landau level scattering is suppressed in the absence of
4131: other scattering mechanisms (see Section \ref{sec18}). This means that the transport
4132: from one point contact to the other is adiabatic. The series conductance is
4133: then simply $G_{\mathrm{series}}=(2e^{2}/h)N$ for two identical point contacts
4134: [$N \equiv\min(N_{1}, N_{2})$ for two different point contacts in series]. This expression
4135: differs from Eq.\ (\ref{eq15.10}) if a barrier is present in the point contacts, since that
4136: causes the number $N$ of occupied Landau levels in the point contact to be less
4137: than the number $N_{\mathrm{wide}}$ of occupied levels in the wide 2DEG. [In a strong
4138: magnetic field, $N\approx(E_{\mathrm{F}}-E_{\mathrm{c}})/\hbar\omega_{\mathrm{c}}$, while $N_{\mathrm{wide}}\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}.$] Adiabatic transport in a magnetic field through two point contacts in series has been studied
4139: experimentally by Kouwenhoven et al.\cite{ref373} and by Main et al.\cite{ref374}
4140:
4141: \subsection{\label{sec16} Junction scattering}
4142:
4143: In the regime of diffusive transport, the Hall bar geometry (a straight
4144: current-carrying channel with small side contacts for voltage drop measurements) is very convenient, since it allows an independent determination of the
4145: various components of the resistivity tensor. A downscaled Hall bar was
4146: therefore a natural first choice as a geometry to study ballistic transport in a
4147: 2DEG.\cite{ref67,ref68,ref74,ref139,ref178,ref364} The resistances measured in narrow-channel
4148: geometries are mainly determined by scattering at the junctions with the side
4149: probes.\cite{ref289} These scattering processes depend strongly on the junction shape.
4150: This is in contrast to the point contact geometry; compare the very similar
4151: results of van Wees et al.\cite{ref6} and Wharam et al.\cite{ref7} on the quantized conductance
4152: of point contacts of a rather different design. The strong dependence of the
4153: low-field Hall resistance on the junction shape was demonstrated theoretically by Baranger and Stone\cite{ref358} and experimentally by Ford et al.\cite{ref77} and Chang
4154: et al.\cite{ref375} These results superseded many earlier attempts (listed in Ref.\ \onlinecite{ref360}) to
4155: explain the discovery by Roukes et al.\cite{ref67} of the {\it quenching of the Hall effect\/}
4156: without modeling the shape of the junction realistically. Baranger and
4157: Stone\cite{ref358} argued that the rounded corners (present in a realistic situation) at
4158: the junction between the main channel and the side branches lead to a
4159: suppression (quenching) of the Hall resistance at low magnetic fields as a
4160: consequence of the horn collimation effect discussed in Section \ref{sec15a}. A Hall
4161: bar with straight corners, in contrast, does not show a generic suppression of
4162: the Hall resistance,\cite{ref376,ref377,ref378} although quenching can occur for special parameter values if only a few subbands are occupied in the channel.
4163:
4164: The quenched Hall effect\cite{ref67,ref77,ref375,ref379} is just one of a whole variety of
4165: magnetoresistance anomalies observed in narrow Hall bars. Other anomalies
4166: are the {\it last Hall plateau},\cite{ref67,ref68,ref77,ref139,ref178,ref379} reminiscent of quantum Hall
4167: plateaus, but occurring at much lower fields; the {\it negative Hall resistance},\cite{ref77} as
4168: if the carriers were holes rather than electrons; the {\it bend resistance},\cite{ref289,ref306,ref364,ref380} a longitudinal resistance associated with a current bend,
4169: which is negative at small magnetic fields and zero at large fields, with an
4170: overshoot to a positive value at intermediate fields; and more.
4171:
4172: In Refs.\ \onlinecite{ref359} and \onlinecite{ref360} we have shown that all these phenomena can be
4173: qualitatively explained in terms of a few simple semiclassical mechanisms
4174: (reviewed in Section \ref{sec16a}). The effects of quantum interference and of
4175: quantization of the lateral motion in the narrow conductor are not essential.
4176: These magnetoresistance anomalies can thus be characterized as classical
4177: magneto size effects in the ballistic regime. In Section \ref{sec5}, we have discussed
4178: classical size effects in the quasi-ballistic regime, where the mean free path is
4179: larger than the channel width but smaller than the separation between the
4180: voltage probes. In that regime, the size effects found in a 2DEG were known
4181: from work on metal films and wires. These earlier investigations had not
4182: anticipated the diversity of magnetoresistance anomalies due to junction
4183: scattering in the ballistic regime. That is not surprising, considering that the
4184: theoretical formalism to describe a resistance measurement within a mean
4185: free path had not been developed in that context. Indeed, this Landauer-B\"{u}ttiker formalism (described in Section \ref{sec12}) found one of its earliest
4186: applications\cite{ref268} in the context of the quenching of the Hall effect, and the
4187: success with which the experimental magnetoresistance anomalies can be
4188: described by means of this formalism forms strong evidence for its validity.
4189:
4190: \subsubsection{\label{sec16a} Mechanisms}
4191:
4192: \begin{figure}
4193: \centerline{\includegraphics[width=8cm]{figures/fig65}}
4194: \caption{
4195: Classical trajectories in an electron billiard, illustrating the collimation (a), scrambling (b), rebound (c), magnetic guiding (d), and electron focusing (e) effects. Taken from C. W. J. Beenakker and H. van Houten, in ``Electronic Properties of Multilayers and Low-Dimensional Semiconductor Structures'' (J. M. Chamberlain, L. Eaves, and J. C. Portal, eds.). Plenum, London, 1990.
4196: \label{fig65}
4197: }
4198: \end{figure}
4199:
4200: The variety of magnetoresistance anomalies mentioned can be understood
4201: in terms of a few simple characteristics of the curved trajectories of electrons
4202: in a classical billiard in the presence of a perpendicular magnetic field.\cite{ref359,ref360}
4203: At very small magnetic fields, {\it collimation\/} and {\it scrambling\/} are the key concepts.
4204: The gradual widening of the channel on approaching thejunction reduces the
4205: injection/acceptance cone, which is the cone of angles with the channel axis
4206: within which an electron is injected into the junction or within which an
4207: electron can enter the channel coming from the junction. This is the horn
4208: collimation effect\cite{ref329} discussed in Section \ref{sec15a} (see Fig.\ \ref{fig65}a). If the injection/acceptance cone is smaller than $90^{\circ}$, then the cones of two channels at right
4209: angles do not overlap. That means that an electron approaching the side
4210: probe coming from the main channel will be reflected (Fig.\ \ref{fig65}a) and will then
4211: typically undergo multiple reflections in the junction region (Fig.\ \ref{fig65}b). The
4212: trajectory is thus scrambled, whereby the probability for the electron to enter
4213: the left or right side probe in a weak magnetic field is equalized. This
4214: suppresses the Hall voltage. This ``scrambling'' mechanism for the quenching
4215: of the Hall effect requires a weaker collimation than the ``nozzle'' mechanism
4216: put forward by Baranger and Stone\cite{ref358} (we return to both these mechanisms
4217: in Section \ref{sec16c}). Scrambling is not effective in the geometry shown in Fig.\ \ref{fig65}c,
4218: in which a large portion of the boundary in the junction is oriented at
4219: approximately $45^{\circ}$ with the channel axis. An electron reflected from a side
4220: probe at this boundary has a large probability of entering the opposite side
4221: probe. This is the ``rebound'' mechanism for a negative Hall resistance
4222: proposed by Ford et al.\cite{ref77}
4223:
4224: At somewhat larger magnetic fields, {\it guiding\/} takes over. As illustrated in
4225: Fig.\ \ref{fig65}d, the electron is guided by the magnetic field along equipotentials
4226: around the corner. Guiding is fully effective when the cyclotron radius $l_{\mathrm{cycl}}$
4227: becomes smaller than the minimal radius of curvature $r_{\min}$ of the corner ---
4228: that is, for magnetic fields greater than the guiding field $B_{\mathrm{g}}\equiv \hbar k_{\rm F}/er_{\min}$. In the
4229: regime $B\gtrsim B_{\mathrm{g}}$ the junction cannot scatter the electron back into the channel
4230: from which it came. The absence of backscattering in this case is an entirely
4231: classical, weak-field phenomenon (cf.\ Section \ref{sec13b}). Because of the absence of
4232: backscattering, the longitudinal resistance vanishes, and the Hall resistance
4233: $R_{\mathrm{H}}$ becomes equal to the contact resistance of the channel, just as in the
4234: quantum Hall effect, but without quantization of $R_{\mathrm{H}}$. The contact resistance
4235: $R_{\mathrm{contact}}\approx(h/2e^{2})(\pi/k_{\mathrm{F}}W)$ is approximately independent of the magnetic field
4236: for fields such that the cyclotron diameter $2l_{\mathrm{cycl}}$ is greater than the channel
4237: width $W$, that is, for fields below $B_{\mathrm{crit}}\equiv 2\hbar k_{\rm F}/eW$ (see Sections \ref{sec12} and \ref{sec13}).
4238: This explains the occurrence of the ``last plateau'' in $R_{\mathrm{H}}$ for $B_{\mathrm{g}}\sim<B\lesssim B_{\mathrm{crit}}$ as a
4239: classical effect. At the low-field end of the plateau, the Hall resistance is
4240: sensitive to {\it geometrical resonances\/} that increase the fraction of electrons
4241: guided around the corner into the side probe. Figure \ref{fig65}e illustrates the
4242: occurrence of one such a geometrical resonance as a result of the magnetic
4243: focusing of electrons into the side probe, at magnetic fields such that the
4244: separation of the two perpendicular channels is an integer multiple of the
4245: cyclotron diameter. This is in direct analogy with electron focusing in a
4246: double-point contact geometry (see Section \ref{sec14}) and leads to periodic
4247: oscillations superimposed on the Hall plateau. Another geometrical resonance with similar effect is discussed in Ref.\ \onlinecite{ref360}.
4248:
4249: These mechanisms for oscillations in the resistance depend on a commensurability between the cyclotron radius and a characteristic dimension of
4250: the junction, but do not involve the wavelength of the electrons as an
4251: independent length scale. This distinguishes these geometrical resonances
4252: conceptually from the quantum resonances due to bound states in the
4253: junction considered in Refs.\ \onlinecite{ref376,ref377}, and \onlinecite{ref380,ref381,ref382}.
4254:
4255: \subsubsection{\label{sec16b} Magnetoresistance anomalies}
4256:
4257: In this subsection we compare, following Ref.\ \onlinecite{ref360}, the semiclassical theory
4258: with representative experiments on laterally confined two-dimensional
4259: electron gases in high-mobility GaAs-AlGaAs heterostructures. The calculations are based on a simulation of the classical trajectories of a large
4260: number (typically $10^{4}$) of electrons with the Fermi energy, to determine the
4261: classical transmission probabilities. The resistances then follow from the
4262: B\"{u}ttiker formula (\ref{eq12.12}). We refer to Refs.\ \onlinecite{ref359} and \onlinecite{ref360} for details on the
4263: method of calculation. We first discuss the Hall resistance $R_{\mathrm{H}}$.
4264:
4265: \begin{figure}
4266: \centerline{\includegraphics[width=8cm]{figures/fig66}}
4267: \caption{
4268: Hall resistance as measured (solid curve) by Simmons et al.,\cite{ref178} and as calculated (dashed curve) for the hard-wall geometry in the inset ($W= 0.8\,\mu{\rm m}$ and $E_{\rm F} = 14\,{\rm meV}$). The dotted line is $R_{\rm H}$ in a bulk 2DEG. Taken from C. W. J. Beenakker and H. van Houten, in ``Electronic Properties of Multilayers and Low-Dimensional Semiconductor Structures'' (J. M. Chamberlain, L. Eaves, and J. C. Portal, eds.). Plenum, London, 1990.
4269: \label{fig66}
4270: }
4271: \end{figure}
4272:
4273: Figure \ref{fig66} shows the precursor of the classical Hall plateau (the ``last
4274: plateau'') in a relatively wide Hall cross. The experimental data (solid curve) is
4275: from a paper by Simmons et al.\cite{ref178} The semiclassical calculation (dashed
4276: curve) is for a square-well confining potential of channel width $W=0.8\,\mu \mathrm{m}$
4277: (as estimated in the experimental paper) and with the relatively sharp corners
4278: shown in the inset. The Fermi energy used in the calculation is $E_{\mathrm{F}}=14\,\mathrm{meV}$,
4279: which corresponds (via $n_{\mathrm{s}}=E_{\mathrm{F}}m/\pi \hbar^{2}$) to a sheet density in the channel of
4280: $n_{\mathrm{s}}=3.9\times 10^{15}\,\mathrm{m}^{-2}$, somewhat below the value of $4.9\times 10^{15}\,\mathrm{m}^{-2}$ of the bulk
4281: material in the experiment. Good agreement between theory and experiment
4282: is seen in Fig.\ \ref{fig66}. Near zero magnetic field, the Hall resistance in this
4283: geometry is close to the linear result $R_{\mathrm{H}}=B/en_{\mathrm{s}}$ for a bulk 2DEG (dotted
4284: line). The corners are sufficiently smooth to generate a Hall plateau via the
4285: guiding mechanism discussed in Section \ref{sec16a}. The horn collimation effect,
4286: however, is not sufficiently large to suppress $R_{\mathrm{H}}$ at small $B$. Indeed, the
4287: injection/acceptance cone for this junction is considerably wider (about
4288: $115^{\circ})$ than the maximal angular opening of $90^{\circ}$ required for quenching of the
4289: Hall effect via the scrambling mechanism described in Section \ref{sec16a}.
4290:
4291: \begin{figure}
4292: \centerline{\includegraphics[width=6cm]{figures/fig67a}}
4293:
4294: \centerline{\includegraphics[width=6cm]{figures/fig67b}}
4295: \caption{
4296: Hall resistance as measured (a) by Ford et al.\cite{ref77} and as calculated (b). In (a) as well as in (b), the solid curve corresponds to the geometry in the upper left inset, the dotted curve to the geometry in the lower right inset. The insets in (a) indicate the shape of the gates, not the actual confining potential. The insets in (b) show equipotentials of the confining potential at $E_{\rm F}$ (thick contour) and 0 (thin contour). The potential rises parabolically from 0 to $E_{\rm F}$ and vanishes in the diamond-shaped region at the center of the junction. Taken from C. W. J. Beenakker and H. van Houten, in ``Electronic Properties of Multilayers and Low-Dimensional Semiconductor Structures'' (J. M. Chamberlain, L. Eaves, and J. C. Portal, eds.). Plenum, London, 1990.
4297: \label{fig67}
4298: }
4299: \end{figure}
4300:
4301: The low-field Hall resistance changes drastically if the channel width
4302: becomes smaller, relative to the radius of curvature of the corners. Figure \ref{fig67}a
4303: shows experimental data by Ford et al.\cite{ref77} The solid and dotted curves are for
4304: the geometries shown respectively in the upper left and lower right insets of
4305: Fig.\ \ref{fig67}a. Note that these insets indicate the gates with which the Hall crosses
4306: are defined electrostatically. The equipotentials in the 2DEG will be
4307: smoother than the contours of the gates. The experiment shows a well-developed Hall plateau with superimposed fine structure. At small positive
4308: fields $R_{\mathrm{H}}$ is either quenched or negative, depending on the geometry. The
4309: geometry is seen to affect also the width of the Hall plateau but not the height.
4310: In Fig.\ \ref{fig67}b we give the results of the semiclassical theory for the two
4311: geometries in the insets, which should be reasonable representations of the
4312: confining potential induced by the gates in the experiment. In the theoretical
4313: plot the resistance and the magnetic field are given in units of
4314: \be
4315: R_{0} \equiv\frac{h}{2e^{2}}\frac{\pi}{k_{\mathrm{F}}W},\;\;B_{0} \equiv\frac{\hbar k_{\rm F}}{eW}, \label{eq16.1}
4316: \ee
4317: where the channel width $W$ for the parabolic confinement used is defined as
4318: the separation of the equipotentials at the Fermi energy ($W_{\mathrm{par}}$ in Section \ref{sec10}).
4319: The experimental estimates $W\approx 90\,\mathrm{nm}$, $n_{\mathrm{s}}\approx 1.2\times 10^{15}\,\mathrm{m}^{-2}$ imply
4320: $R_{0}=5.2\,\mathrm{k}\Omega$, $B_{0}=0.64\,{\rm T}$. With these parameters the calculated resistance
4321: and field scales do not agree well with the experiment, which may be due in
4322: part to the uncertainties in the modeling of the shape of the experimental
4323: confining potential. The $\pm B$ asymmetry in the experimental plot is undoubtedly due to asymmetries in the cross geometry [in the calculation the
4324: geometry has fourfold symmetry, which leads automatically to
4325: $R_{\mathrm{H}}(B)=-R_{\mathrm{H}}(-B)]$. Apart from these differences, there is agreement in all
4326: the important features: the appearance of quenched and negative Hall
4327: resistances, the independence of the height of the last Hall plateau on the
4328: smoothness of the corners, and the shift of the onset of the last plateau to
4329: lower fields for smoother corners. The oscillations on the last plateau in the
4330: calculation (which, as we discussed in Section \ref{sec16a}, are due to geometrical
4331: resonances) are also quite similar to those in the experiment, indicating that
4332: these are classical rather than quantum resonances.
4333:
4334: \begin{figure}
4335: \centerline{\includegraphics[width=6cm]{figures/fig68a}}
4336:
4337: \centerline{\includegraphics[width=6cm]{figures/fig68b}}
4338: \caption{
4339: Hall resistance $R_{\rm H} \equiv R_{13,24}$ (a) and bend resistance $R_{\rm B} \equiv R_{12,43}$ (b), as measured (solid curves) by Timp et al.\cite{ref306} and as calculated (dashed curves) for the geometry in the inset (consisting of a parabolic confining potential with the equipotentials at $E_{\rm F}$ and 0 shown respectively as thick and thin contours; the parameters are $W = 100\,{\rm nm}$ and $E_{\rm F} = 3.9\,{\rm meV}$). The dotted line in (a) is $R_{\rm H}$ in a bulk 2DEG. Taken from C. W. J. Beenakker and H. van Houten, in ``Electronic Properties of Multilayers and Low-Dimensional Semiconductor Structures'' (J. M. Chamberlain, L. Eaves, and J. C. Portal, eds.). Plenum, London, 1990.
4340: \label{fig68}
4341: }
4342: \end{figure}
4343:
4344: We now turn to the bend resistance $R_{\mathbf{B}}$. In Fig.\ \ref{fig68} we show experimental
4345: data by Timp et al.\cite{ref306} (solid curves) on $R_{\mathbf{B}}\equiv R_{12,43}$ and $R_{\mathrm{H}}\equiv R_{13,24}$
4346: measured in the same Hall cross (defined by gates of a shape similar to that in
4347: the lower right inset of Fig.\ \ref{fig67}a; see the inset of Fig.\ \ref{fig68}a for the numbering of
4348: the channels). The dashed curves are calculated for a parabolic confining
4349: potential in the channels (with the experimental values $W=100\,\mathrm{nm}$,
4350: $E_{\mathrm{F}}=3.9\,\mathrm{meV})$ and with corners as shown in the inset of Fig.\ \ref{fig68}a. The
4351: calculated quenching of the Hall resistance and the onset of the last plateau
4352: are in good agreement with the experiment, and also the observed overshoot
4353: of the bend resistance around $0.2\, \mathrm{T}$ as well as the width of the negative peak in
4354: $R_{\mathbf{B}}$ around zero field are well described by the calculation. The calculated
4355: height of the negative peak, however, is too small by more than a factor of 2.
4356: We consider this disagreement to be significant in view of the quantitative
4357: agreement with the other features in both $R_{\mathbf{B}}$ and $R_{\mathrm{H}}$. The negative peak in $R_{\mathbf{B}}$
4358: is due to the fact that the collimation effect couples the current source 1 more
4359: strongly to voltage probe 3 than to voltage probe 4, so $R_{\mathbf{B}}\propto V_{4}-V_{3}$ is
4360: negative for small magnetic fields (at larger fields the Lorentz force destroys
4361: collimation by bending the trajectories, so $R_{\mathbf{B}}$ shoots up to a positive value
4362: until guiding takes over and brings $R_{\mathbf{B}}$ down to zero by eliminating
4363: backscattering at the junction). The discrepancy in Fig.\ \ref{fig68}b thus seems to
4364: indicate that the semiclassical calculation underestimates the collimation
4365: effect in this geometry. The positive overshoot of $R_{\mathbf{B}}$ seen in Fig.\ \ref{fig68}b is found
4366: only for rounded corners. This explains the near absence of the effect in the
4367: calculation of Kirczenow\cite{ref381} for a junction with straight corners.
4368:
4369: For a discussion of the temperature dependence of the magnetoresistance
4370: anomalies, we refer to Ref.\ \onlinecite{ref360}. Here it suffices to note that the experiments
4371: discussed were carried out at temperatures around $1\, \mathrm{K}$, for which we expect
4372: the zero-temperature semiclassical calculation to be appropriate. At lower
4373: temperatures the effects of quantum mechanical phase coherence that have
4374: been neglected will become more important. At higher temperatures the
4375: thermal average smears out the magnetoresistance anomalies and eventually
4376: inelastic scattering causes a transition to the diffusive transport regime in
4377: which the resistances have their normal $B$-dependence.
4378:
4379: \subsubsection{\label{sec16c} Electron waveguide versus electron billiard}
4380:
4381: The overall agreement between the experiments and the semiclassical
4382: calculations is remarkable in view of the fact that the channel width in the
4383: narrowest structures considered is comparable to the Fermi wavelength.
4384: When the first experiments on these ``electron waveguides'' appeared, it was
4385: expected that the presence of only a small number of occupied transverse
4386: waveguide modes would fundamentally alter the nature of electron transport.\cite{ref68} The results of Refs.\ \onlinecite{ref359} and \onlinecite{ref356} show instead that the modal structure
4387: plays only a minor role and that the magnetoresistance anomalies observed
4388: are characteristic for the {\it classical\/} ballistic transport regime. The reason that a
4389: phenomenon such as the quenching of the Hall effect has been observed only
4390: in Hall crosses with narrow channels is simply that the radius of curvature of
4391: the corners at the junction is too small compared with the channel width in
4392: wider structures. This is not an essential limitation, and the various
4393: magnetoresistance anomalies discussed here should be observable in macrocopic Hall bars with artificially smoothed corners, provided of course that
4394: the dimensions of the junction remain well below the mean free path. Ballistic
4395: transport is essential, but a small number of occupied modes is not.
4396:
4397: Although we believe that the characteristic features of the magnetoresistance anomalies are now understood, several interesting points of disagreement between theory and experiment remain that merit further investigation.
4398: One of these is the discrepancy in the magnitude of the negative bend
4399: resistance at zero magnetic field noted before. The disappearance of a region
4400: of quenched Hall resistance at low electron density is another unexpected
4401: observation by Chang et al.\cite{ref375} and Roukes et al.\cite{ref383} The semiclassical theory
4402: predicts a universal behavior (for a given geometry) if the resistance and
4403: magnetic field are scaled by $R_{0}$ and $B_{0}$ defined in Eq.\ (\ref{eq16.1}). For a square-well
4404: confining potential the channel width $W$ is the same at each energy, and since
4405: $B_{0}\propto k_{\mathrm{F}}$ one would expect the field region of quenched Hall resistance to vary
4406: with the electron density as $\sqrt{n_{\mathrm{s}}}$. For a more realistic smooth confining
4407: potential, $W$ depends on $E_{\mathrm{F}}$ and thus on $n_{\mathrm{s}}$ as well, in a way that is difficult to
4408: estimate reliably. In any case, the experiments point to a systematic
4409: disappearance of the quench at the lowest densities, which is not accounted
4410: for by the present theory (and has been attributed by Chang et al.\cite{ref375} to
4411: enhanced diffraction at low electron density as a result of the increase in the
4412: Fermi wavelength). For a detailed investigation of departures from classical
4413: scaling, we refer to a paper by Roukes et al.\cite{ref384} As a third point, we mention
4414: the curious density dependence of the quenching observed in approximately
4415: straight junctions by Roukes et al.,\cite{ref383} who find a low-field suppression of $R_{\mathrm{H}}$
4416: that occurs only at or near certain specific values of the electron density. The
4417: semiclassical model applied to a straight Hall cross (either defined by a
4418: square well or by a parabolic confining potential) gives a low-field slope of $R_{\mathrm{H}}$
4419: close to its bulk 2D value. The fully quantum mechanical calculations for a
4420: straight junction\cite{ref376,ref381} do give quenching at special parameter values, but
4421: not for the many-mode channels in this experiment (in which quenching
4422: occurs with as many as 10 modes occupied, whereas in the calculations a
4423: straight cross with more than 3 occupied modes in the channel does not show
4424: a quench).
4425:
4426: In addition to the points of disagreement discussed, there are fine details in
4427: the measured magnetoresistances, expecially at the lowest temperatures
4428: (below $100\, \mathrm{mK}$), which are not obtained in the semiclassical approximation.
4429: The quantum mechanical calculations\cite{ref358,ref376,ref377,ref381} show a great deal of fine
4430: structure due to interference of the waves scattered by the junction. The fine
4431: structure in most experiments is not quite as pronounced as in the
4432: calculations presumably partly as a result of a loss of phase coherence after
4433: many multiple scatterings in the junction. The limited degree of phase
4434: coherence in the experiments and the smoothing effect of a finite temperature
4435: help to make the semiclassical model work so well even for the narrowest
4436: channels. We draw attention to the fact that classical {\it chaotic\/} scattering can
4437: also be a source of irregular resistance fluctuations (see Ref.\ \onlinecite{ref360}).
4438:
4439: Some of the most pronounced features in the quantum mechanical
4440: calculations are due to transmission resonances that result from the presence
4441: of bound states in the junction.\cite{ref376,ref377,ref380,ref381,ref382} In Section \ref{sec16a} we have
4442: discussed a different mechanism for transmission resonances that has a
4443: classical, rather than a quantum mechanical, origin. As mentioned in Section
4444: \ref{sec16b}, the oscillations on the last Hall plateau observed experimentally are
4445: quite well accounted for by these geometrical resonances. One way to
4446: distinguish experimentally between these resonance mechanisms is by means
4447: of the temperature dependence, which should be much weaker for the
4448: classical than for the quantum effect. One would thus conclude that the
4449: fluctuations in Fig.\ \ref{fig67}a, measured by Ford et al.\cite{ref77} at $4.2\, \mathrm{K}$, have a classical
4450: origin, while the fine structure that Ford et al.\cite{ref385} observe only at $\mathrm{mK}$
4451: temperatures (see below) is intrinsically quantum mechanical.
4452:
4453: The differences between the semiclassical and the quantum mechanical
4454: models may best be illustrated by considering once again the quenching of
4455: the Hall effect, which has the most subtle explanation and is the most
4456: sensitive to the geometry among the magnetoresistance anomalies observed
4457: in the ballistic regime. The classical scrambling of the trajectories after
4458: multiple reflections suppresses the asymmetry between the transmission
4459: probabilities $t_{\rm l}$ and $t_{\rm r}$ to enter the left or right voltage probe, and without this
4460: transmission asymmetry there can be no Hall voltage. We emphasize that this
4461: {\it scrambling mechanism\/} is consistent with the original findings of Baranger and
4462: Stone\cite{ref358} that quenching requires collimation. The point is that the collimation effect leads to nonoverlapping injection/acceptance cones of two perpendicular channels, which ensures that electrons cannot enter the voltage
4463: probe from the current source directly, but rather only after multiple
4464: reflections (cf.\ Section \ref{sec16a}). In this way a rather weak collimation to within
4465: an injection/acceptance cone of about $90^{\circ}$ angular opening is sufficient to
4466: induce a suppression of the Hall resistance via the scrambling mechanism.
4467:
4468: Collimation can also suppress $R_{\mathrm{H}}$ directly by strongly reducing $t_{\rm l}$ and $t_{\rm r}$
4469: relative to $t_{\mathrm{s}}$ (the probability for transmission straight through the junction).
4470: This {\it nozzle mechanism}, introduced by Baranger and Stone,\cite{ref358} requires a
4471: strong collimation of the injected beam in order to affect $R_{\mathrm{H}}$ appreciably. In
4472: the geometries considered here, we find that quenching of $R_{\mathrm{H}}$ is due
4473: predominantly to scrambling and not to the nozzle mechanism ($t_{\rm l}$ and $t_{\rm r}$ each
4474: remain more than 30\% of $t_{\mathrm{s}}$), but data by Baranger and Stone\cite{ref358} show that
4475: both mechanisms can play an important role.
4476:
4477: \begin{figure}
4478: \centerline{\includegraphics[width=8cm]{figures/fig69}}
4479: \caption{
4480: Measured Hall resistance in an abrupt (a) and in a widened (b) cross as a function of $B$ in the strong field regime. Large fluctuations are resolved at the low temperature of $22\,{\rm mK}$. The dotted curves indicate the reproducibility of the measurement. Taken from C. J. B. Ford et al., Surf.\ Sci.\ {\bf 229}, 298 (1990).
4481: \label{fig69}
4482: }
4483: \end{figure}
4484:
4485: There is a third proposed mechanism for the quenching of the Hall
4486: effect,\cite{ref376,ref377} which is the reduction of the transmission asymmetry due to a
4487: bound state in the junction. The {\it bound state mechanism\/} is purely quantum
4488: mechanical and does not require collimation (in contrast to the classical
4489: scrambling and nozzle mechanisms). Numerical calculations have shown that
4490: it is only effective in straight Hall crosses with very narrow channels (not
4491: more than three modes occupied), and even then for special values of the
4492: Fermi energy only. Although this mechanism cannot account for the
4493: experiments performed thus far, it may become of importance in future work.
4494: A resonant suppression of the Hall resistance may also occur in strong
4495: magnetic fields, in the regime where the Hall resistance in wide Hall crosses
4496: would be quantized. Such an effect is intimately related to the high-field
4497: Aharonov-Bohm magnetoresistance oscillations in a singly connected
4498: geometry (see Section \ref{sec21}). Ford et al.\cite{ref385} have observed oscillations superimposed on quantized Hall plateaux at low temperatures in very narrow crosses
4499: of two different shapes (see Fig.\ \ref{fig69}). The strong temperature dependence
4500: indicates that these oscillations are resonances due to the formation of bound
4501: states in the cross.\cite{ref306,ref385,ref386}
4502:
4503: \subsection{\label{sec17} Tunneling}
4504:
4505: In this section we review recent experiments on tunneling through
4506: potential barriers in a two-dimensional electron gas. Subsection \ref{sec17a} deals
4507: with {\it resonant\/} tunneling through a bound state in the region between two
4508: barriers. Resonant tunneling has previously been studied extensively in
4509: layered semiconductor heterostructures for transport perpendicular to the
4510: layers.\cite{ref387,ref388,ref389} For example, a thin AlGaAs layer embedded between two
4511: GaAs layers forms a potential barrier, whose height and width can be tailored
4512: with great precision by means of advanced growth techniques (such as
4513: molecular beam epitaxy). Because of the free motion in the plane of the layers,
4514: one can only realize bound states with respect to one direction. Tunneling
4515: resonances are consequently smeared out over a broad energy range. A
4516: 2DEG offers the possibility of confinement in all directions and thus of a
4517: sharp resonance. A gate allows one to define potential barriers of adjustable
4518: height in the 2DEG. In contrast, the heterostructure layers form fixed
4519: potential barriers, so one needs to study a current-voltage characteristic to
4520: tune the system through a resonance (observable as a peak in the $I-V$ curve).
4521: The gate-induced barriers in a 2DEG offer a useful additional degree of
4522: freedom, allowing a study of resonant tunneling in the linear response regime
4523: of small applied voltages (to which we limit the discussion in this review). A
4524: drawback of these barriers is that their shape cannot be precisely controlled,
4525: or modeled, so that a description of the tunneling process will of necessity be
4526: qualitative.
4527:
4528: Subsection \ref{sec17b} deals with the effects of Coulomb repulsion on tunneling
4529: in a 2DEG. The electrostatic effects of charge buildup in the 1D potential well
4530: formed by heterostructure layers have received considerable attention in
4531: recent years.\cite{ref389,ref390} Because of the large capacitance of the potential well in
4532: this case (resulting from the large surface area of the layers) these are
4533: macroscopic effects, involving a large number of electrons. The 3D potential
4534: well in a 2DEG nanostructure, in contrast, can have a very small capacitance
4535: and may contain a few electrons only. The tunneling of a single electron into
4536: the well will then have a considerable effect on the electrostatic potential
4537: difference with the surrounding 2DEG. For a small applied voltage this effect
4538: of the Coulomb repulsion can completely suppress the tunneling current. In
4539: metals this ``Coulomb blockade'' of tunneling has been studied extensively.\cite{ref391} In those systems a semiclassical description suffices. The large Fermi
4540: wavelength in a 2DEG should allow the study of quantum mechanical effects
4541: on the Coulomb blockade or, more generally, of the interplay between
4542: electron-electron interactions and resonant tunneling.\cite{ref318,ref392,ref393}
4543:
4544: \subsubsection{\label{sec17a} Resonant tunneling}
4545:
4546: The simplest geometry in which one might expect to observe transmission
4547: resonances is formed by a single potential barrier across a 2DEG channel.
4548: Such a geometry was studied by Washburn et al.\cite{ref394} in a GaAs-AlGaAs
4549: heterostructure containing a 2-$\mu \mathrm{m}$-wide channel with a 45-nm-long gate on
4550: top of the heterostructure. At low temperatures (around $20\, \mathrm{mK}$) an irregular
4551: set of peaks was found in the conductance as a function of gate voltage in the
4552: region close to the depletion threshold. The amplitude of the peaks was on
4553: the order of $e^{2}/h$. The origin of the effect could not be pinned down. The
4554: authors examine the possibility that transmission resonances associated with
4555: a square potential barrier are responsible for the oscillations in the conductance, but also note that the actual barrier is more likely to be smooth on
4556: the scale of the wavelength. For such a smooth barrier the transmission
4557: probability as a function of energy does not show oscillations. It seems most
4558: likely that the effect is disorder-related. Davies and Nixon\cite{ref395} have suggested
4559: that some of the structure observed in this experiment could be due to
4560: potential fluctuations in the region under the gate. These fluctuations can be
4561: rather pronounced close to the depletion threshold, due to the lack of
4562: screening in the low-density electron gas. A quantum mechanical calculation
4563: of transmission through such a fluctuating barrier has not been performed.
4564: As discussed below, conductance peaks of order $e^{2}/h$ occur in the case of
4565: resonant tunneling via localized states in the barrier (associated with
4566: impurities), a mechanism that might well play a role in the experiment of
4567: Washburn et al.\cite{ref394}
4568:
4569: \begin{figure}
4570: \centerline{\includegraphics[width=8cm]{figures/fig70}}
4571: \caption{
4572: Resistance versus gate voltage of a cavity (defined by gates on top of a GaAs-AlGaAs heterostructure; see inset), showing plateau like features (for $R\lesssim h/2e^{2}$) and tunneling resonances (for $R\gtrsim h/2e^{2}$). The left- and right-hand curves refer to the adjacent resistance scales. Taken from C. G. Smith et al., Surf.\ Sci.\ {\bf 228}, 387 (1990).
4573: \label{fig70}
4574: }
4575: \end{figure}
4576:
4577: In pursuit of resonant tunneling in a 2DEG, Chou et al.\cite{ref396} have fabricated
4578: double-barrier devices involving two closely spaced short gates across a wide
4579: GaAs-AlGaAs heterostructure. Both the spacing and the length of the gates
4580: were $100\, \mathrm{nm}$. They observed a peak in the transconductance (the derivative of
4581: the channel current with respect to the gate voltage), which was attributed to
4582: resonant tunneling through a quasi-bound state in the 2D potential well
4583: between the barriers. Palevski et al.\cite{ref397} have also investigated transport
4584: through two closely spaced potential barriers in a double-gate structure, but
4585: they did not find evidence for transmission resonances.
4586: A 3D potential well has truly bound states and is expected to show
4587: the strongest transmission resonances. Transport through such a cavity or
4588: ``quantum box'' has been studied theoretically by several
4589: authors.\cite{ref318,ref333,ref382,ref398} Experiments have been performed by Smith et
4590: al.\cite{ref399,ref400,ref401} Their device is based on a quantum point contact, but contains two
4591: potential barriers that separate the constriction from the wide 2DEG regions
4592: (see the inset of Fig.\ \ref{fig70}). As the negative gate voltage is increased, a potential
4593: well is formed between the two barriers, resulting in confinement in all
4594: directions. The tunneling regime corresponds to a resistance $R$ that is greater
4595: than $h/2e^{2}$. It is also possible to study the ballistic regime $R<h/2e^{2}$ when the
4596: height of the potential barriers is less than the Fermi energy. In this regime
4597: the transmission resonances are similar to the resonances in long quantum
4598: point contacts (these are determined by an interplay of tunneling through
4599: evanescent modes and reflection at the entrance and exit of the point contact;
4600: cf.\ Section \ref{sec13}). Results of Smith et al.\cite{ref399,ref400,ref401} for the resistance as a function of
4601: gate voltage at $330\, \mathrm{mK}$ are reproduced in Fig.\ \ref{fig70}. In the tunneling regime
4602: ($R>h/2e^{2}$) giant resistance oscillations are observed. A regular series of
4603: smaller resistance peaks is found in the ballistic regime ($R<h/2e^{2}$). Martin-Moreno and Smith\cite{ref333} have modeled the electrostatic potential in the device
4604: of Refs.\ \onlinecite{ref399,ref400,ref401} and have performed a quantum mechanical calculation of
4605: the resistance. Very reasonable agreement with the experimental data in the
4606: ballistic regime was obtained. The tunneling regime was not compared in
4607: detail with the experimental data. The results were found to depend rather
4608: critically on the assumed chape of the potential, in particular on the rounding
4609: of the tops of the potential barriers. Martin-Moreno and Smith also
4610: investigated the effects of asymmetries in the device structure on the tunneling
4611: resonances and found in particular that small differences in the two barrier
4612: heights (of order 10\%) lead to a sharp suppression of the resonances, a finding
4613: that sheds light on the fact that they were observed in certain devices only.
4614: Experimentally, the effect of a magnetic field on the oscillations in the
4615: resistance versus gate voltage was also investigated.\cite{ref399,ref400,ref401} A strong suppression of the peaks was found in relatively weak magnetic fields (of about
4616: $0.3\, \mathrm{T}$).
4617:
4618:
4619: \begin{figure}
4620: \centerline{\includegraphics[width=8cm]{figures/fig71}}
4621: \caption{
4622: Schematic diagram of a Si MOSFET with a split gate (a), which creates a potential barrier in the inversion layer (b). Taken from T. E. Kopley et al. Phys.\ Rev.\ Lett.\ {\bf 61}, 1654 (1988).
4623: \label{fig71}
4624: }
4625: \end{figure}
4626:
4627: \begin{figure}
4628: \centerline{\includegraphics[width=8cm]{figures/fig72}}
4629: \caption{
4630: Oscillations in the conductance as a function of gate voltage at 0.5 K are attributed to resonant tunneling through localized states in the potential barrier. A second trace is shown for a magnetic field of 6 T (with a horizontal offset of $- 0.04\,{\rm V}$). The inset is a close-up of the largest peak at 6 T, together with a Lorentzian fit. Taken from T. E. Kopley et al. Phys.\ Rev.\ Lett.\ {\bf 61}, 1654 (1988).
4631: \label{fig72}
4632: }
4633: \end{figure}
4634:
4635: Tunneling through a cavity, as in the experiment by Smith et al.,\cite{ref399,ref400,ref401} is
4636: formally equivalent to tunneling through an impurity state (see, e.g., Refs.\ \onlinecite{ref402}
4637: and \onlinecite{ref403}). The dramatic subthreshold structure found in the conductance of
4638: quasi-one-dimensional MOSFETs has been interpreted in terms of resonant
4639: tunneling through a series of localized states.\cite{ref32,ref35,ref36,ref37} Kopley et al.\cite{ref404} have
4640: observed large conductance peaks in a MOSFET with a split gate (see Fig.\
4641: \ref{fig71}). Below the 200-nm-wide slot in the gate, the inversion layer is interrupted
4642: by a potential barrier. Pronounced conductance peaks were seen at $0.5\, \mathrm{K}$ as
4643: the gate voltage was varied in the region close to threshold (see Fig.\ \ref{fig72}). No
4644: clear correlation was found between the channel width and the peak spacing
4645: or amplitude. The peaks were attributed to resonant transmission through
4646: single localized states associated with bound states in the Si band gap in the
4647: noninverted region under the gate.
4648:
4649: The theory of resonant tunneling of noninteracting electrons through
4650: localized states between two-dimensional reservoirs was developed by Xue
4651: and Lee\cite{ref405} (see also Refs.\ \onlinecite{ref159} and \onlinecite{ref406}). If the resonances are well separated in
4652: energy, a single localized state will give the dominant contribution to the
4653: transmission probability. The maximum conductance on resonance is then
4654: $e^{2}/h$ (for one spin direction), regardless of the number of channels $N$ in the
4655: reservoirs.\cite{ref405,ref406} This maximum (which may be interpreted as a contact
4656: resistance, similar to that of a quantum point contact) is attained if the
4657: localized state has identical leak rates $\Gamma_{\mathrm{L}}/\hbar$ and $\Gamma_{\mathrm{R}}/\hbar$ to the left and right
4658: reservoirs. Provided these leak rates are small (cf.\ Section \ref{sec21}) the conductance
4659: $G$ as a function of Fermi energy $E_{\mathrm{F}}$ is a Lorentzian centered around the
4660: resonance energy $E_{0}$:
4661: \be
4662: G(E_{\mathrm{F}})= \frac{e^{2}}{h}\frac{\Gamma_{\mathrm{L}}\Gamma_{\mathrm{R}}}{(E_{\mathrm{F}}-E_{0})^{2}+\frac{1}{4}(\Gamma_{\mathrm{L}}+\Gamma_{\mathrm{R}})^{2}}. \label{eq17.1}
4663: \ee
4664: This is the Breit-Wigner formula of nuclear physics.\cite{ref93} For an asymmetrically placed impurity the peak height is reduced below $e^{2}/h$ (by up to a factor
4665: $4\Gamma_{\mathrm{R}}/\Gamma_{\mathrm{L}}$ if $\Gamma_{\mathrm{L}}\gg \Gamma_{\mathrm{R}}$).
4666:
4667: The amplitudes of the peaks observed by Kopley et al.\cite{ref404} were found to be
4668: in agreement with this prediction, while the line shape of an isolated peak
4669: could be well described by a Lorentzian (see inset of Fig.\ \ref{fig72}). (Most of the
4670: peaks overlapped, hampering a line-shape analysis). In addition, they studied
4671: the effect of a strong magnetic field on the conductance peaks and found that
4672: the amplitudes of most peaks were substantially suppressed. This was
4673: interpreted as a reduction of the leak rates because of a reduced overlap
4674: between the wave functions on the impurity and the reservoirs. The
4675: amplitude of one particular peak was found to be unaffected by the field,
4676: indicative of a symmetrically placed impurity in the barrier $(\Gamma_{\mathrm{R}}=\Gamma_{\mathrm{L}})$, while
4677: the width of that peak was reduced, in agreement with Eq.\ (\ref{eq17.1}). This study
4678: therefore exhibits many characteristic features of resonant tunneling through
4679: a single localized state.
4680:
4681: \begin{figure}
4682: \centerline{\includegraphics[width=8cm]{figures/fig73}}
4683: \caption{
4684: Conductance as a function of gate voltage for a quantum point contact at 0.55 K. The inset is a close-up of the low-conductance regime, showing peaks attributed to transmission resonances associated with impurity states in the constriction. Taken from P. L. McEuen et al., Surf.\ Sci. {\bf 229}, 312 (1990).
4685: \label{fig73}
4686: }
4687: \end{figure}
4688:
4689: Transmission resonances due to an impurity in a quantum point contact
4690: or narrow channel have been studied theoretically in Refs.\ \onlinecite{ref241,ref407}, and \onlinecite{ref408}.
4691: In an experiment it may be difficult to distinguish these resonances from
4692: those associated with reflection at the entrance and exit of the quantum point
4693: contact (discussed in Section \ref{sec13}). A conductance peak associated with
4694: resonant tunneling through an impurity state in a quantum point contact was
4695: reported by McEuen et al.\cite{ref409} The experimental results are shown in Fig.\ \ref{fig73}.
4696: The resonant tunneling peak is observed near the onset of the first
4697: conductance plateau, where $G<2e^{2}/h$. A second peak seen in Fig.\ \ref{fig73} was
4698: conjectured to be a signature of resonant scattering, in analog with similar
4699: processes known in atomic physics.\cite{ref410}
4700:
4701: \begin{figure}
4702: \centerline{\includegraphics[width=8cm]{figures/fig74a}}
4703:
4704: \centerline{\includegraphics[width=8cm]{figures/fig74b}}
4705: \caption{
4706: (a) Schematic diagram of a constriction with two adjustable external reflectors defined by gates on top of a GaAs-AlGaAs heterostructure. (b) Plot of the constriction resistance as a function of gate voltage with the external reflector gates (Y1, Y2) grounded. Inset: Fabry-Perot-type transmission resonances due to a variation of the gate voltage on the reflectors (Y1, Y2) (bottom panel), and Fourier power spectrum (top panel). Taken from C. G. Smith et al., Surf.\ Sci. {\bf 228}, 387 (1990).
4707: \label{fig74}
4708: }
4709: \end{figure}
4710:
4711: We want to conclude this subsection on transmission resonances by
4712: discussing an experiment by Smith et al.\cite{ref401,ref411} on what is essentially a
4713: Fabry-Perot interferometer. The device consists of a point contact with
4714: external reflectors in front of its entrance and exit. The reflectors are potential
4715: barriers erected by means of two additional gate electrodes (see Fig.\ \ref{fig74}a). By
4716: varying the gate voltage on the external reflectors of this device, Smith et al.\
4717: could tune the effective cavity length without changing the width of the
4718: narrow section. This experiment is therefore more controlled than the
4719: quantum dot experiment\cite{ref399,ref400,ref401} discussed earlier. The resulting periodic
4720: transmission resonances are reproduced in Fig.\ \ref{fig74}b. A new oscillation
4721: appears each time the separation between the reflectors increases by $\lambda_{\mathrm{F}}/2$. A
4722: numerical calculation for a similar geometry was performed by Avishai et
4723: al.\cite{ref412} The significance of this experiment is that it is the first clear realization
4724: of an electrostatically tuned electron interferometer. Such a device has
4725: potential transistor applications. Other attempts to fabricate an electrostatic
4726: interferometer have been less succesful. The electrostatic Aharonov-Bohm
4727: effect in a ring was discussed in Section \ref{sec8}. The solid-state analogue of the
4728: microwave stub tuner (proposed by Sols et al.\cite{ref413} and by Datta\cite{ref414}) was
4729: studied experimentally by Miller et al.\cite{ref415} The idea is to modify the
4730: transmission through a narrow channel by changing the length of a side
4731: branch (by means of a gate across the side branch). Miller et al.\ have
4732: fabricated such a $\mathrm{T}$-shaped conductor and found some evidence for the
4733: desired effect. Much of the structure was due, however, to disorder-related
4734: conductance fluctuations. The electrostatic Aharonov-Bohm effect had
4735: similar problems. Transport in a long and narrow channel is simply not fully
4736: ballistic, because of partially diffuse boundary scattering and impurity
4737: scattering. The device studied by Smith et al.\ worked because it made use of a
4738: very short constriction (a quantum point contact), while the modulation of
4739: the interferometer length was done externally in the wide 2DEG, where the
4740: effects of disorder are much less severe (in high-mobility material).
4741:
4742: \subsubsection{\label{sec17b} Coulomb blockade}
4743:
4744: In this subsection we would like to speculate on the effects of electron-electron interactions on tunneling through impurities in narrow semiconductor channels, in relation to a recent paper in which Scott-Thomas et al.\cite{ref416}
4745: announced the discovery of conductance oscillations periodic in the density
4746: of a narrow Si inversion layer. The device features a continuous gate on top of
4747: a split gate, as illustrated schematically in Fig.\ \ref{fig75}. In the experiment, the
4748: voltage on the upper gate is varied while the split-gate voltage is kept
4749: constant. Figure \ref{fig76} shows the conductance as a function of gate voltage at
4750: $0.4\, \mathrm{K}$, as well as a set of Fourier power spectra obtained for devices of
4751: different length. A striking pattern of rapid periodic oscillations is seen. No
4752: correlation is found between the periodicity of the oscillations and the
4753: channel length, in contrast to the transmission resonances in ballistic
4754: constrictions discussed in Sections \ref{sec13} and \ref{sec17a}. The oscillations die out as the
4755: channel conductance increases toward $e^{2}/h\approx 4\times 10^{-5}\,\Omega^{-1}$. The conductance peaks are relatively insensitive to a change in temperature, while the
4756: minima depend exponentially on temperature as $\exp(- E_{\mathrm{a}}/k_{\mathbf{B}}T)$, with an
4757: activation energy $E_{\mathrm{a}}\approx 50\,\mu \mathrm{eV}$. Pronounced nonlinearities occur in the
4758: current as a function of source-drain voltage. An interpretation in terms of
4759: pinned charge density waves was suggested,\cite{ref416} based on a model due to
4760: Larkin and Lee\cite{ref417} and Lee and Rice.\cite{ref418} In such a model, one expects the
4761: conductance to be thermally activated, because of the pinning of the charge
4762: density wave by impurities in the one-dimensional channel. The activation
4763: energy is determined by the most strongly pinned segment in the channel, and
4764: periodic oscillations in the conductance as a function of gate voltage
4765: correspond to the condition that an integer number of electrons is contained
4766: between the two impurities delimiting that specific segment. The same
4767: interpretation has been given to a similar effect observed in a narrow channel
4768: in a GaAs-AlGaAs heterostructure by Meirav et al.\cite{ref85}
4769:
4770: \begin{figure}
4771: \centerline{\includegraphics[width=6cm]{figures/fig75a}}
4772:
4773: \centerline{\includegraphics[width=6cm]{figures/fig75b}}
4774: \caption{
4775: Schematic cross sectional (a) and top (b) view of a double-gate Si MOSFET device. The lower split gate is at a negative voltage, confining the inversion layer (due to the positive voltage on the upper gate) to a narrow channel. Taken from J. H. F. Scott-Thomas et al., Phys.\ Rev.\ Lett.\ {\bf 62}, 583 (1989).
4776: \label{fig75}
4777: }
4778: \end{figure}
4779:
4780: \begin{figure}
4781: \centerline{\includegraphics[width=8cm]{figures/fig76}}
4782: \caption{
4783: Top panel: Periodic oscillations in the conductance versus gate voltage at 0.4 K for a 10-$\mu{\rm m}$-long inversion channel. Next three panels: Fourier power spectra of this curve and of data obtained for 2- and 1-$\mu{\rm m}$-long channels. Bottom panel: Fourier spectrum for the 1-$\mu{\rm m}$-long device in a magnetic field of 6 T. Taken from J. H. F. Scott-Thomas et al., Phys.\ Rev.\ Lett.\ {\bf 62}, 583 (1989).
4784: \label{fig76}
4785: }
4786: \end{figure}
4787:
4788: We have proposed\cite{ref419} an alternative {\it single-electron\/} explanation of the
4789: remarkable effect discovered by Scott-Thomas et al.,\cite{ref416} based upon the
4790: concept of the Coulomb blockade of tunneling mentioned at the beginning of
4791: this section. Likharev\cite{ref391} and Mullen et al.\cite{ref420} have studied theoretically the
4792: possibility of removing the Coulomb blockade by capacitive charging (by
4793: means of a gate electrode) of the region between two tunnel barriers. They
4794: found that the conductance of this system exhibits periodic peaks as a
4795: function of gate voltage, due to the modulation of the net charge (mod $e$) on
4796: the interbarrier region. Following the theoretical papers,\cite{ref391,ref420} the authors
4797: in Ref.\ \onlinecite{ref419} proposed that the current through the channel in the experiment
4798: of Scott-Thomas et al.\cite{ref416} is limited by tunneling through potential barriers
4799: constituted by two dominant scattering centers that delimit a segment of the
4800: channel (see Fig.\ \ref{fig77}). Because the number of electrons localized in the region
4801: between the two barriers is necessarily an integer, a charge imbalance, and
4802: hence an electrostatic potential difference, arises between this region and the
4803: adjacent regions connected to wide electron gas reservoirs. As the gate
4804: voltage is varied, the resulting Fermi level difference $\Delta E_{\mathrm{F}}$ oscillates in a
4805: sawtooth pattern between $\pm e\Delta$, where $\Delta=e/2C$ and $C=C_{1}+C_{2}$ is the
4806: effective capacitance of the region between the two barriers. The single-electron charging energy $e^{2}/2C$ maintains the Fermi level difference until
4807: $\Delta E_{\mathrm{F}}=\pm e\Delta$ (this is the Coulomb blockade). When $\Delta E_{\mathrm{F}}=\pm e\Delta$, the energy
4808: required for the transfer of a single electron to (or from) the region between
4809: the two barriers vanishes so that the Coulomb blockade is removed. The
4810: conductance then shows a maximum at low temperatures $T$ and source-drain voltages $V$ ($k_{B}T/e, V\lesssim\Delta$). We note that in the case of very different
4811: tunneling rates through the two barriers, one would expect steps in the
4812: current as a function of source-drain voltage, which are not observed in the
4813: experiments.\cite{ref85,ref416} For two similar barriers this ``Coulomb staircase'' is
4814: suppressed.\cite{ref420} The oscillation of the Fermi energy as the gate voltage is
4815: varied thus leads to a sequence of conductance peaks. The periodicity of the
4816: oscillations corresponds to the addition of a single electron to the region
4817: between the two scattering centers forming the tunnel barriers, so the
4818: oscillations are periodic in the density, as in the experiment. This single-electron tunneling mechanism also explains the observed activation of the
4819: conductance minima and the insensitivity to a magnetic field.\cite{ref85,ref416} The
4820: capacitance associated with the region between the scattering centers is hard
4821: to ascertain. The experimental value of the activation energy $E_{\mathrm{a}}\approx 50\,\mu \mathrm{eV}$
4822: would imply $C\approx e^{2}/2E_{\mathrm{a}}\approx 10^{-15}\,{\rm F}$. Kastner et al.\cite{ref421} argue that the
4823: capacitance in the device is smaller than this amount by an order of
4824: magnitude (the increase in the effective capacitance due to the presence of the
4825: gate electrodes is taken into account in their estimate). In addition, they point
4826: to a discrepancy between the value for the Coulomb blockade inferred from
4827: the nonlinear conductance and that from the thermal activation energy. The
4828: temperature dependence of the oscillatory conductance was found to be
4829: qualitatively different in the experiment by Meirav et al.\cite{ref85} At elevated
4830: temperatures an exponential $T$-dependence was found, but at low temperatures the data suggest a much weaker $T$-dependence. It is clear that more
4831: experimental and theoretical work is needed to arrive at a definitive
4832: interpretation of this intriguing phenomenon.
4833:
4834: \begin{figure}
4835: \centerline{\includegraphics[width=8cm]{figures/fig77}}
4836: \caption{
4837: Schematic representation of the bottom of the conduction band $E_{\rm c}$, and Fermi energy $E_{\rm F}$ in the device of Fig.\ \ref{fig76} along the channel. The band bending at the connections of the narrow channel to the wide source S and drain D regions arises from the higher threshold for the electrostatic creation of a narrow inversion layer by a gate (shaded part). Tunnel barriers associated with two scattering centers are shown. The maximum Fermi energy difference sustainable by the Coulomb blockade, $\Delta E_{\rm F}=\pm e\Delta$ (where $\Delta=e/2C$, with $C = C_{1} + C_{2}$), is indicated. Taken from H. van Houten and C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 63}, 1893 (1989).
4838: \label{fig77}
4839: }
4840: \end{figure}
4841:
4842: It would be of interest to study the effects of the Coulomb blockade of
4843: tunneling in a more controlled fashion in a structure with two adjustable
4844: potential barriers. Such an experiment was proposed by Glazman and
4845: Shekter,\cite{ref422} who studied theoretically a system similar to the cavity of the
4846: experiments by Smith et al.\cite{ref399,ref400,ref401} (discussed in Section \ref{sec17a}). A difficulty
4847: with this type of device is, as pointed out in Ref.\ \onlinecite{ref422}, that a variation in gate
4848: voltage affects the barrier height (and thus their transparency) as well as the
4849: charge in the cavity. This is expected to lead to an exponential damping of the
4850: oscillations due to the Coulomb blockade.\cite{ref391,ref420} A characteristic feature of
4851: these oscillations is their insensitivity to an applied magnetic field, which can
4852: serve to distinguish the effect from oscillations due to resonant tunneling
4853: (Section \ref{sec17a}). The field dependence of the peaks observed by Smith et
4854: al.\cite{ref399,ref400,ref401} in the tunneling regime was not reported, so the question of
4855: whether or not the Coulomb oscillations are observed in their experiment
4856: remains unanswered. In our opinion, substantial progress could be made
4857: with the development of thin tunnel barriers of larger height, which would be
4858: less sensitive to the application of an external gate voltage. If our interpretation of the experiments by Scott-Thomas et al.\cite{ref416} and Meirav et al.\cite{ref85} is
4859: correct, such tunneling barriers might be formed by the incorporation of
4860: negatively charged impurities (e.g., ionized acceptors) in a narrow electron
4861: gas channel. This speculation is based on the fact that such acceptor
4862: impurities are present in the Si inversion layers of the experiment of Scott-Thomas et al.,\cite{ref416} as well as in the {\it p}-{\it n\/} junctions employed for lateral
4863: confinement by Meirav et al.\cite{ref85}
4864:
4865: As we were completing this review, we learned of several experiments that
4866: demonstrate the Coulomb blockade in split-gate confined GaAs-AlGaAs
4867: heterostructures.\cite{ref423,ref424,ref425} These experiments should open the way for the
4868: controlled study of the effects of Coulomb interactions on tunneling in
4869: semiconductor nanostructures.
4870:
4871: \section{\label{secIV} Adiabatic transport}
4872: \subsection{\label{sec18} Edge channels and the quantum Hall effect}
4873:
4874: In this section we give an overview of the characteristics of adiabatic
4875: transport via edge channels in the regime of the quantum Hall effect as a
4876: background to the following sections. We restrict ourselves here to the {\it integer}
4877: quantum Hall effect, where the edge channels can be described by single-electron states. Recent developments on adiabatic transport in the regime of
4878: the {\it fractional\/} quantum Hall effect (which is fundamentally a many-body
4879: effect) will be considered in Section \ref{sec20}.
4880:
4881: \subsubsection{\label{sec18a} Introduction}
4882:
4883: Both the quantum Hall effect (QHE) and the quantized conductance of a
4884: ballistic point contact are described by the same relation, $G=Ne^{2}/h$,
4885: between the conductance $G$ and the number $N$ of propagating modes at the
4886: Fermi level (counting both spin directions separately). The smooth transition
4887: from zero-field quantization to QHE that follows from this relation is evident
4888: from Fig.\ \ref{fig48}. The nature of the modes is very different, however, in weak and
4889: strong magnetic fields. As we discussed in Section \ref{sec12a}, the propagating
4890: modes in a strong magnetic field consist of edge states, which interact with
4891: one of the sample edges only. Edge states with the same mode index are
4892: referred to collectively as an {\it edge channel}. Edge channels at opposite edges
4893: propagate in opposite directions. In a weak magnetic field, in contrast, the
4894: modes consist of magnetoelectric subbands that interact with both edges. In
4895: that case there is no spatial separation of modes propagating in opposite
4896: directions.
4897:
4898: The different spatial extension of edge channels and magnetoelectric
4899: subbands leads to an entirely different sensitivity to scattering processes in
4900: weak and strong magnetic fields. Firstly, the zero-field conductance
4901: quantization is destroyed by a small amount of elastic scattering (due to
4902: impurities or roughness of the channel boundaries; cf.\ Refs.\ \onlinecite{ref313,ref316,ref317,ref407},
4903: and \onlinecite{ref408}), while the QHE is robust to scattering.\cite{ref97} This difference is a
4904: consequence of the {\it suppression of backscattering\/} by a magnetic field discussed
4905: in Section \ref{sec13b}, which itself follows from the spatial separation at opposite
4906: edges of edge channels moving in opposite directions. Second, the spatial
4907: separation of edge channels at the {\it same\/} edge in the case of a smooth confining
4908: potential opens up the possibility of {\it adiabatic transport\/} (i.e., the full
4909: suppression of interedge channel scattering). In weak magnetic fields,
4910: adiabaticity is of importance within a point contact, but not on longer length
4911: scales (cf.\ Sections \ref{sec13a} and \ref{sec15a}). In a wide 2DEG region, scattering among
4912: the modes in weak fields establishes local equilibrium on a length scale given
4913: by the inelastic scattering length (which in a high-mobility GaAs-AlGaAs
4914: heterostructure is presumably not much longer than the elastic scattering
4915: length $l\sim 10\,\mu \mathrm{m}$). The situation is strikingly different in a strong magnetic
4916: field, where the {\it selective\/} population and detection of edge channels observed
4917: by van Wees et al.\cite{ref426} has demonstrated the persistence of adiabaticity outside
4918: the point contact.
4919:
4920: In the absence of interedge channel scattering the various edge channels at
4921: the same boundary can be occupied up to different energies and consequently
4922: carry different amounts of current. The electron gas at the edge of the sample
4923: is then not in {\it local\/} equilibrium. Over some long distance (which is not yet
4924: known precisely) adiabaticity breaks down, leading to a partial equilibration
4925: of the edge channels. However, as demonstrated by Komiyama et al.\cite{ref427} and
4926: by others,\cite{ref307,ref428,ref429,ref430} local equilibrium is not fully established even on
4927: macroscopic length scales exceeding $0.25\, \mathrm{mm}$. Since local equilibrium is a
4928: prerequisite for the use of a local resistivity tensor, these findings imply a
4929: nonlocality of the transport that had not been anticipated in theories of the
4930: QHE (which are commonly expressed in terms of a local resistivity).\cite{ref97}
4931:
4932: A theory of the QHE that is able to explain anomalies resulting from the
4933: absence of local equilibrium has to take into account the properties of the
4934: current and voltage contacts used to measure the Hall resistance. That is not
4935: necessary if local equilibrium is established at the voltage contacts, for the
4936: fundamental reason that two systems in equilibrium that are in contact have
4937: identical electrochemical potentials. In the Landauer-B\"{u}ttiker formalism
4938: described in Section \ref{sec12b}, the contacts are modeled by electron gas reservoirs
4939: and the resistances are expressed in terms of transmission probabilities of
4940: propagating modes at the Fermi level from one reservoir to the other. This
4941: formalism is not restricted to zero or weak magnetic fields, but can equally
4942: well be applied to the QHE, where edge channels form the modes. In this way
4943: B\"{u}ttiker could show\cite{ref112} that the nonideality of the coupling of the reservoirs
4944: to the conductor affects the accuracy of the QHE in the absence of local
4945: equilibrium. An {\it ideal\/} contact in the QHE is one that establishes an
4946: equilibrium population among the outgoing edge channels by distributing
4947: the injected current equally among these propagating modes (this is the
4948: equipartitioning of current discussed for an ideal electron waveguide in
4949: Section \ref{sec12b}). A quantum point contact that selectively populates certain
4950: edge channels\cite{ref426} can thus be seen as an extreme example of a nonideal, or
4951: {\it disordered}, contact.
4952:
4953: \subsubsection{\label{sec18b} Edge channels in a disordered conductor}
4954:
4955: After this general introduction, let us now discuss in some detail how edge
4956: channels are formed at the boundary of a 2DEG in a strong magnetic field. In
4957: Section \ref{sec12a} we discussed the edge states in the case of a narrow channel
4958: without disorder, relevant for the point contact geometry. Edge states were
4959: seen to originate from Landau levels, which in the bulk lie below the Fermi
4960: level but rise in energy on approaching the sample boundary (cf.\ Fig.\ \ref{fig40}b).
4961: The point of intersection of the $n$th Landau level $(n=1,2, \ldots)$ with the Fermi
4962: level forms the site of edge states belonging to the $n$th edge channel. The
4963: number $N$ of edge channels at $E_{\mathrm{F}}$ is equal to the number of bulk Landau
4964: levels below $E_{\mathrm{F}}$. This description can easily be generalized to the case of a
4965: slowly varying potential energy landscape $V(x, y)$ in the 2DEG, in which case
4966: a semiclassical analysis can be applied.\cite{ref431} The energy $E_{\mathrm{F}}$ of an electron at the
4967: Fermi level in a strong magnetic field contains a part $(n-\frac{1}{2})\hbar\omega_{\mathrm{c}}$ due to the
4968: quantized cyclotron motion and a part $\pm\frac{1}{2} g\mu_{\mathbf{B}}B$ (depending on the spin
4969: direction) from spin splitting. The remainder is the energy $E_{\mathrm{G}}$ due to the
4970: electrostatic potential
4971: \be
4972: E_{\mathrm{G}}=E_{\mathrm{F}}-(n- {\textstyle\frac{1}{2}})\hbar\omega_{\mathrm{c}}\pm{\textstyle\frac{1}{2}}g\mu_{\mathbf{B}}B. \label{eq18.1}
4973: \ee
4974: The cyclotron orbit center $\mathbf{R}$ is guided along equipotentials of $V$ at the
4975: guiding center energy $E_{\mathrm{G}}$. As derived in Section \ref{sec11b}, the drift velocity $\mathbf{v}_{\mathrm{drift}}$ of
4976: the orbit center (known as the guiding center drift or $\mathbf{E}\times \mathbf{B}$ drift) is given by
4977: \be
4978: \mathbf{v}_{\mathrm{drift}}(\mathbf{R})=\frac{1}{eB^{2}}\nabla V(\mathbf{R})\times \mathbf{B}, \label{eq18.2}
4979: \ee
4980: which indeed is parallel to the equipotentials. An important distinction with
4981: the weak-field case of Section \ref{sec11b} is that the spatial extension of the
4982: cyclotron orbit can now be neglected, so $V$ is evaluated at the position of the
4983: orbit center in Eq.\ (\ref{eq18.2}) [compared with Eq.\ (\ref{eq11.1})]. The guiding center drift
4984: contributes a kinetic energy
4985: $\frac{1}{2}mv_{\mathrm{drift}}^{2}$ to the energy of the electron, which is
4986: small for large $B$ and smooth $V$. (More precisely,
4987: $\frac{1}{2}mv_{\mathrm{drift}}^{2}\ll \hbar\omega_{\mathrm{c}}$ if
4988: $|\nabla V|\ll \hbar\omega_{\mathrm{c}}/l_{\mathrm{m}}$, with $l_{\mathrm{m}}$ the magnetic length defined as $l_{\mathrm{m}}\equiv(h/eB)^{1/2}.)$ This
4989: kinetic energy term has therefore not been included in Eq.\ (\ref{eq18.1}).
4990:
4991: The simplicity of the guiding center drift along equipotentials has been
4992: originally used in the percolation theory\cite{ref432,ref433,ref434} of the QHE, soon after its
4993: experimental discovery.\cite{ref8} In this theory the existence of edge states is ignored,
4994: so the Hall resistance is not expressed in terms of equilibrium properties of
4995: the 2DEG (in contrast to the edge channel formulation that will be discussed).
4996: The physical requirements on the smoothness of the disorder potential have
4997: received considerable attention\cite{ref435,ref436} in the context of the percolation
4998: theory and, more recently,\cite{ref437,ref438,ref439} in the context of adiabatic transport in
4999: edge channels. Strong potential variations should occur on a spatial scale
5000: that is large compared with the magnetic length $l_{\mathrm{m}}$ ($l_{\mathrm{m}}$ corresponds to the
5001: cyclotron radius in the QHE, $l_{\mathrm{cycl}}\equiv l_{\mathrm{m}}(2n-1)^{1/2}\approx l_{\mathrm{m}}$ if the Landau level
5002: index $n\approx 1$). More rapid potential fluctuations may be present provided their
5003: amplitude is much less than $\hbar\omega_{\mathrm{c}}$ (the energy separation of Landau levels).
5004:
5005: \begin{figure}
5006: \centerline{\includegraphics[width=8cm]{figures/fig78}}
5007: \caption{
5008: Formation of edge channels in a disordered potential, from various viewpoints discussed in the text.
5009: \label{fig78}
5010: }
5011: \end{figure}
5012:
5013: In Fig.\ \ref{fig78} we have illustrated the formation of edge channels in a smooth
5014: potential energy landscape from various viewpoints. The wave functions of
5015: states at the Fermi level are extended along equipotentials at the guiding
5016: center energy (\ref{eq18.1}), as shown in Fig.\ \ref{fig78}a (for Landau level index $n=1,2,3$
5017: and a single spin direction). One can distinguish between {\it extended\/} states near
5018: the sample boundaries and {\it localized\/} states encircling potential maxima and
5019: minima in the bulk. The extended states at the Fermi level form the edge
5020: channels. The edge channel with the smallest index $n$ is closest to the sample
5021: boundary, because it has the largest $E_{\mathrm{G}}$ [Eq.\ (\ref{eq18.1})]. This is seen more clearly
5022: in the cross-sectional plot of $V(x, y)$ in Fig.\ \ref{fig78}b (along the line connecting the
5023: two arrows in Fig.\ \ref{fig78}a). The location of the states at the Fermi level is
5024: indicated by dots and crosses (depending on the direction of motion). The
5025: value of $E_{\mathrm{G}}$ for each $n$ is indicated by the dashed line. If the peaks and dips of
5026: the potential in the bulk have amplitudes below $\hbar\omega_{\mathrm{c}}/2$, then only states with
5027: highest Landau level index can exist in the bulk at the Fermi level. This is
5028: obvious from Fig.\ \ref{fig78}c, which shows the total energy of a state
5029: $E_{\mathrm{G}}+(n- \frac{1}{2})\hbar\omega_{\mathrm{c}}$ along the same cross section as Fig.\ \ref{fig78}b. If one identifies
5030: $k=-xeB/\hbar$, this plot can be compared with Fig.\ \ref{fig40}b of the dispersion
5031: relation $E_{n}(k)$ for a disorder-free electron waveguide in strong magnetic field.
5032:
5033: \begin{figure}
5034: \centerline{\includegraphics[width=8cm]{figures/fig79}}
5035: \caption{
5036: Measurement configuration for the two-terminal resistance $R_{2{\rm t}}$, the four-terminal Hall resistance $R_{\rm H}$, and the longitudinal resistance $R_{\rm L}$. The edge channels at the Fermi level are indicated; arrows point in the direction of motion of edge channels filled by the source contact at chemical potential $E_{\rm F} + \delta\mu$. The current is equipartitioned among the edge channels at the upper edge, corresponding to the case of local equilibrium.
5037: \label{fig79}
5038: }
5039: \end{figure}
5040:
5041: A description of the QHE based on extended edge states and localized
5042: bulk states, as in Fig.\ \ref{fig78}, was first put forward by Halperin\cite{ref440} and further
5043: developed by several authors.\cite{ref441,ref442,ref443,ref444} In these papers a local equilibrium is
5044: assumed at each edge. In the presence of a chemical potential difference $\delta\mu$
5045: between the edges, each edge channel carries a current $(e/h)\delta\mu$ and thus
5046: contributes $e^{2}/h$ to the Hall conductance (cf.\ the derivation of Landauer's
5047: formula in Section \ref{sec12b}). In this case of local equilibrium the two-terminal
5048: resistance $R_{2\mathrm{t}}$ of the Hall bar is the same as the four-terminal Hall resistance
5049: $R_{\mathrm{H}}=R_{2\mathrm{t}}=h/e^{2}N$ (see Fig.\ \ref{fig79}). The longitudinal resistance vanishes, $R_{\mathrm{L}}=0$.
5050: The distinction between a longitudinal and Hall resistance is topological: A
5051: four-terminal resistance measurement gives $R_{\mathrm{H}}$ if current and voltage
5052: contacts alternate along the boundary of the conductor, and $R_{\mathrm{L}}$ if that is not
5053: the case. There is no need to further characterize the contacts in the case of
5054: local equilibrium at the edge.
5055:
5056: If the edges are not in local equilibrium, the measured resistance depends
5057: on the properties of the contacts. Consider, for example, a situation in which
5058: the edge channels at the lower edge are in equilibrium at chemical potential
5059: $E_{\mathrm{F}}$, while the edge channels at the upper edge are not in local equilibrium.
5060: The current at the upper edge is then not equipartitioned among the $N$
5061: modes. Let $f_{n}$ be the fraction of the total current $I$ that is carried by states
5062: above $E_{\mathrm{F}}$ in the $n$th edge channel at the upper edge, $I_{n}=f_{n}I$. The voltage
5063: contact at the lower edge measures a chemical potential $E_{\mathrm{F}}$ regardless of its
5064: properties. The voltage contact at the upper edge, however, will measure a
5065: chemical potential that depends on how it couples to each of the edge
5066: channels. The transmission probability $T_{n}$ is the fraction of $I_{n}$ that is
5067: transmitted through the voltage probe to a reservoir at chemical potential
5068: $E_{\mathrm{F}}+\delta\mu$. The incoming current
5069: \be
5070: I_{\mathrm{in}}= \sum_{n=1}^{N}T_{n}f_{n}I,\;\;{\rm with}\;\;
5071: \sum_{n=1}^{N}f_{n}=1, \label{eq18.3}
5072: \ee
5073: has to be balanced by an outgoing current
5074: \be
5075: I_{\mathrm{out}}= \frac{e}{h}\delta\mu(N-R)=\frac{e}{h}\delta\mu\sum_{n=1}^{N}T_{n} \label{eq18.4}
5076: \ee
5077: of equal magnitude, so that the voltage probe draws no net current. (In Eq.\
5078: (\ref{eq18.4}) we have applied Eq.\ (\ref{eq12.14}) to identify the total transmission probability $N-R$ of outgoing edge channels with the sum of transmission
5079: probabilities $T_{n}$ of incoming edge channels.) The requirement $I_{\mathrm{in}}=I_{\mathrm{out}}$
5080: determines $\delta\mu$ and hence the Hall resistance $R_{\mathrm{H}}=\delta\mu/eI$:
5081: \be
5082: R_{\mathrm{H}}= \frac{h}{e^{2}}\left(\sum_{n=1}^{N}T_{n}f_{n}\right)\left(\sum_{n=1}^{N}T_{n}\right)^{-1}. \label{eq18.5}
5083: \ee
5084: The Hall resistance has its regular quantized value $R_{\mathrm{H}}=h/e^{2}N$ only if {\it either}
5085: $f_{n}=1/N$ {\it or\/} $T_{n}=1$, for $n=1,2, \ldots, N$. The first case corresponds to local
5086: equilibrium (the current is equipartitioned among the modes), the second case
5087: to an ideal contact (all edge channels are fully transmitted). The Landauer-B\"{u}ttiker formalism discussed in Section \ref{sec12b} forms the basis on which
5088: anomalies in the QHE due to the absence of local equilibrium in combination
5089: with nonideal contacts can be treated theoretically.\cite{ref112}
5090:
5091: A nonequilibrium population of the edge channels is generally the result of
5092: {\it selective backscattering}. Because edge channels at opposite edges of the
5093: sample move in opposite directions, backscattering requires scattering from
5094: one edge to the other. Selective backscattering of edge channels with $n\geq n_{0}$ is
5095: induced by a potential barrier across the sample,\cite{ref113,ref339,ref340,ref427} if its height is
5096: between the guiding center energies of edge channel $n_{0}$ and $n_{0}-1$ (note that
5097: the edge channel with a larger index $n$ has a smaller value of $E_{\mathrm{G}}$). The
5098: anomalous Shubnikov-De Haas effect,\cite{ref428} to be discussed in Section \ref{sec19}, has
5099: demonstrated that selective backscattering can also occur {\it naturally\/} in the
5100: absence of an imposed potential barrier. The edge channel with the highest
5101: index $n=N$ is selectively backscattered when the Fermi level approaches the
5102: energy $(N- \frac{1}{2})\hbar\omega_{\mathrm{c}}$ of the $N$th bulk Landau level. The guiding center energy of
5103: the $N$th edge channel then approaches zero, and backscattering either by
5104: tunneling or by thermally activated processes becomes effective, but for that
5105: edge channel only, which remains almost completely decoupled from the
5106: other $N-1$ edge channels over distances as large as 250 $\mu \mathrm{m}$ (although on
5107: that length scale the edge channels with $n\leq N-1$ have equilibrated to a
5108: large extent).\cite{ref429}
5109:
5110: \subsubsection{\label{sec18c} Current distribution}
5111:
5112: The edge channel theory has been criticized on the grounds that experiments measure a nonzero current in the bulk of a Hall bar.\cite{ref445} In this
5113: subsection we want to point out that a measurement of the current
5114: distribution cannot be used to prove or disprove the edge channel formulation of the QHE.
5115:
5116: The fact that the Hall resistance can be expressed in terms of the
5117: transmission probabilities of edge states at the Fermi level does {\it not\/} imply that
5118: these few states carry a macroscopic current, {\it nor\/} does it imply that the
5119: current flows at the edges. A determination of the spatial current distribution
5120: $i(\mathbf{r})$, rather than just the total current $I$, requires consideration of all the states
5121: below the Fermi level, which acquire a net drift velocity because of the Hall
5122: field. As we discussed in Section \ref{sec12b}, knowledge of $i(\mathbf{r})$ is not necessary to
5123: know the resistances in the regime of linear response, because the Einstein
5124: relation allows one to obtain the resistance from the diffusion constant. Edge
5125: channels tell you where the current flows if the electrochemical potential
5126: difference $\delta\mu$ is entirely due to a density difference, relevant for the diffusion
5127: problem. Edge channels have nothing to say about where the current flows if
5128: $\delta\mu$ is mainly of electrostatic origin, relevant for the problem of electrical
5129: conduction. The ratio $\delta\mu/I$ is the same for both problems, but $i(\mathbf{r})$ is not.
5130:
5131: With this in mind, it remains an interesting problem to find out just how
5132: the current is distributed in a Hall bar, or, alternatively, what is the
5133: electrostatic potential profile. This problem has been treated theoretically in
5134: many papers.\cite{ref446,ref447,ref448,ref449,ref450,ref451,ref452,ref453,ref454,ref455} In the case of a 3D conductor, a linearly varying
5135: potential and uniform current density are produced by a surface charge. As
5136: noted by MacDonald et al.,\cite{ref446} the electrostatics is qualitatively different in
5137: the 2D case because an edge charge $\delta(x-W/2)$ produces a potential
5138: proportional to $\ln|x-W/2|$, which is weighted toward the edge, and hence a
5139: concentration of current at the edge.
5140:
5141: Experiments aimed at measuring the electrostatic potential distribution
5142: were originally carried out by attaching contacts to the interior of the Hall
5143: bar and measuring the voltage differences between adjacent contacts.\cite{ref456,ref457,ref458,ref459,ref460}
5144: It was learned from these studies that relatively small inhomogeneities in the
5145: density of the 2DEG have a large effect on these voltage differences in the
5146: QHE regime. The main difficulty in the interpretation of such experiments is
5147: that the voltage difference measured between two contacts is the difference in
5148: electrochemical potential, not the line integral of the electric field. B\"{u}ttiker\cite{ref461}
5149: has argued that the voltage measured at an interior contact can exhibit large
5150: variations for a small increase in magnetic field without an appreciable
5151: change in the current distribution. Contactless measurements of the QHE
5152: from the absorption of microwave radiation\cite{ref462} are one alternative to interior
5153: contacts, which might be used to determine the potential (or current)
5154: distribution.
5155:
5156: \begin{figure}
5157: \centerline{\includegraphics[width=8cm]{figures/fig80}}
5158: \caption{
5159: Electrostatic potential $V_{\rm H}$ induced by passing a current through a Hall bar. The sample edges are at $x = \pm 1\,{\rm mm}$. The data points are from the experiment of Fontein et a1.,\cite{ref463} at two magnetic field values on the $R_{\rm H} = h/4e^{2}$ quantized Hall plateau (triangles: $B = 5\,{\rm T}$; crosses: $B = 5.25\,{\rm T}$). The solid curve is calculated from Eq.\ (\ref{eq18.9}), assuming an impurity-free Hall bar with four filled Landau levels. The theory contains no adjustable parameters.
5160: \label{fig80}
5161: }
5162: \end{figure}
5163:
5164: Fontein et al.\cite{ref463} have used the birefringence of GaAs induced by an
5165: electric field to perform a contactless measurement of the electrostatic
5166: potential distribution in a Hall bar. They measure the Hall potential profile
5167: $V_{\mathrm{H}}(x)$ as a change in the local electrostatic potential if a current is passed
5168: through the Hall bar. The data points shown in Fig.\ \ref{fig80} were taken at $1.5\, \mathrm{K}$
5169: for two magnetic field values on the plateau of quantized Hall resistance at
5170: $\frac{1}{4}h/e^{2}$. The potential varies steeply at the edges (at $x=\pm 1\,\mathrm{mm}$ in Fig.\ \ref{fig80}) and
5171: is approximately linear in the bulk. The spatial resolution of the experiment
5172: was 70 $\mu \mathrm{m}$, limited by the laser beam used to measure the birefringence. The
5173: current distribution is not directly measured, but can be estimated from the
5174: guiding center drift (\ref{eq18.2}) (this assumes a slowly varying potential). The
5175: nonequilibrium current density $i(x)$ along the Hall bar is then given by
5176: \be
5177: i( x)=\frac{en_{\mathrm{s}}}{B}\frac{dV_{\rm H}(x)}{dx}. \label{eq18.6}
5178: \ee
5179: Fontein et al.\ thus estimate that under the conditions of their experiment two
5180: thirds of the total imposed current $I=5\,\mu{\rm A}$ flows within 70 $\mu \mathrm{m}$ from the
5181: edges while the remainder is uniformly distributed in the bulk.
5182:
5183: This experimental data can be modeled\cite{ref464} by means of an integral
5184: equation derived by MacDonald et al.\cite{ref446} for the self-consistent potential
5185: profile in an ideal impurity-free sample with $N$ completely filled (spin-split)
5186: Landau levels. The electron charge density $\rho_{\mathrm{e}}(x)$ in the 2DEG is given by
5187: \be
5188: \rho_{\mathrm{e}}(x)=-en_{\mathrm{s}}\left[1-\frac{el_{\mathrm{m}}^{2}}{\hbar\omega_{\mathrm{c}}}V_{\mathrm{H}}^{\prime\prime}(x)\right]. \label{eq18.7}
5189: \ee
5190: This equation follows from the Schr\"{o}dinger equation in a smoothly varying
5191: electrostatic potential, so the factor between brackets is close to unity.
5192: Substitution of the net charge density $en_{\mathrm{s}}+\rho_{\mathrm{e}}(x)$ into the Poisson equation
5193: gives\cite{ref446}
5194: \be
5195: V_{\mathrm{H}}( x)=-\xi\int_{-W/2}^{+W/2}dx^{\prime}\ln\left(\frac{2}{W}|x-x^{\prime}|\right)V_{\mathrm{H}}^{\prime\prime}(x^{\prime}). \label{eq18.8}
5196: \ee
5197: The characteristic length $\xi\equiv Nl_{\mathrm{m}}^{2}/\pi a^{*}$ is defined in terms of the magnetic
5198: length $l_{\mathrm{m}}$ and the effective Bohr radius $a^{*}\equiv\epsilon \hbar^{2}/me^{2}$ (with $\epsilon$ the dielectric
5199: constant).
5200:
5201: The integral equation (\ref{eq18.8}) was solved numerically by MacDonald et
5202: al.\cite{ref446} and analytically by means of the Wiener-Hopf technique by Thouless.\cite{ref448} Here we describe a somewhat simpler approach,\cite{ref464} which is sufficiently accurate for the present purpose. For magnetic field strengths in the
5203: QHE regime the length $\xi$ is very small. For example, if $N=4$, $l_{\mathrm{m}}=11.5\,\mathrm{nm}$
5204: (for $B=5\,\mathrm{T}$), $a^{*}=10\,\mathrm{nm}$ (for GaAs with $\epsilon=13\,\epsilon_{0}$ and $m=0.067\,m_{\mathrm{e}}$), then
5205: $\xi=17\,\mathrm{nm}$. It is therefore meaningful to look for a solution of Eq.\ (\ref{eq18.8}) in the
5206: limit $\xi\ll W$. The result is that $V_{\mathrm{H}}(x)=\mathrm{constant}\times\ln|(x-W/2)/(x+W/2)|$ if
5207: $|x|\leq W/2-\xi$, with a linear extrapolation from $|x|=W/2-\xi$ to $|x|=W/2$.
5208: One may verify that this is indeed the answer, by substituting the preceding
5209: expression into Eq.\ (\ref{eq18.8}) and performing one partial integration. The
5210: arbitrary constant in the expression for $V_{\mathrm{H}}$ may be eliminated in favor of the
5211: total current $I$ flowing through the Hall bar, by applying Eq.\ (\ref{eq18.6}) to the case
5212: of $N$ filled spin-split Landau levels. This gives the final answer
5213: \begin{eqnarray}
5214: V_{\mathrm{H}}( x)&=&\frac{1}{2}IR_{\mathrm{H}}\left(1+\ln\frac{W}{\xi}\right)^{-1}\ln\left|\frac{x-W/2}{x+W/2}\right|\nonumber\\
5215: &&{\rm if}\;\;|x| \leq\frac{W}{2}-\xi, \label{eq18.9}
5216: \end{eqnarray}
5217: with a linear extrapolation of $V_{\mathrm{H}}$ to $\pm\frac{1}{2}IR_{\mathrm{H}}$ in the interval within $\xi$ from the
5218: edge. The Hall resistance is $R_{\mathrm{H}}=h/Ne^{2}$. The approximation (\ref{eq18.9}) is
5219: equivalent for small $\xi$ to the analytical solution of Thouless, and is close to
5220: the numerical solutions given by MacDonald et al., even for a relatively large
5221: value $\xi/W=0.1$.
5222:
5223: In Fig.\ \ref{fig80} the result (\ref{eq18.9}) has been plotted (solid curve) for the parameters
5224: of the experiment by Fontein et al.\ ($\xi/W=0.85\times 10^{-5}$ for $N=4$, $B=5\,\mathrm{T}$,
5225: and $W=2\,\mathrm{mm}$). The agreement with experiment is quite satisfactory in view
5226: of the fact that the theory contains {\it no\/} adjustable parameters. The theoretical
5227: profile is steeper at the edges than in the experiment, which may be due to the
5228: limited experimental resolution of 70 $\mu \mathrm{m}$. The total voltage drop between the
5229: two edges in the calculation ($hI/Ne^{2}\approx 32\,\mathrm{mV}$ for $I=5\,\mu \mathrm{A}$ and $N=4$) agrees
5230: with the measured Hall voltage of $\approx 30\,\mathrm{mV}$, but the optically determined
5231: value of $40\, \mathrm{mV}$ is somewhat larger for a reason that we do not understand.
5232:
5233: We have discussed this topic of the current distribution in the QHE in
5234: some detail to convince the reader that the concentration of the potential
5235: drop (and hence of the current) near the edges can be understood from the
5236: electrostatics of edge {\it charges}, but cannot be used to test the validity of a
5237: linear response formulation of the QHE in terms of edge {\it states}. Indeed, edge
5238: states were completely neglected in the foregoing theoretical analysis, which
5239: nonetheless captures the essential features of the experiment.
5240:
5241: \subsection{\label{sec19} Selective population and detection of edge channels}
5242:
5243: The absence of local equilibrium at the current or voltage contacts leads to
5244: anomalies in the quantum Hall effect, unless the contacts are ideal (in the
5245: sense that each edge channel at the Fermi level is transmitted through the
5246: contact with probability 1). Ideal versus disordered contacts are dealt with in
5247: Sections \ref{sec19a} and \ref{sec19b}. A quantum point contact can be seen as an extreme
5248: example of a disordered contact, as discussed in Section \ref{sec19c}. Anomalies in
5249: the Shubnikov-De Haas effect due to the absence of local equilibrium are the
5250: subject of Section \ref{sec19d}.
5251:
5252: \subsubsection{\label{sec19a} Ideal contacts}
5253:
5254: In a two-terminal measurement of the quantum Hall effect the contact
5255: resistances of the current source and drain are measured in series with the
5256: Hall resistance. For this reason precision measurements of the QHE are
5257: usually performed in a four-terminal measurement configuration, in which
5258: the voltage contacts do not carry a current.\cite{ref445} Contact resistances then do
5259: not play a role, provided that local equilibrium is established near the voltage
5260: contacts [or, by virtue of the reciprocity relation (\ref{eq12.16}), near the current
5261: contacts]. As we mentioned in Section \ref{sec18}, local equilibrium can be grossly
5262: violated in the QHE. Accurate quantization then requires that either the
5263: current or the voltage contacts are {\it ideal}, in the sense that the edge states at
5264: the Fermi level have unit transmission probability through the contacts.
5265: In this subsection we return to the four-terminal measurements on a
5266: quantum point contact considered in Section \ref{sec13b}, but now in the QHE
5267: regime where the earlier assumption of local equilibrium near the voltage
5268: contacts is no longer applicable in general. We assume strong magnetic fields
5269: so that the four-terminal longitudinal resistance $R_{\mathrm{L}}$ of the quantum point
5270: contact is determined by the potential barrier in the constriction (rather than
5271: by its width).
5272:
5273: \begin{figure}
5274: \centerline{\includegraphics[width=8cm]{figures/fig81}}
5275: \caption{
5276: Motion along equipotentials in the QHE regime, in a four-terminal geometry with a saddle-shaped potential formed by a split gate (shaded). Ideal contacts are assumed. The thin lines indicate the location of the edge channels at the Fermi level, with the arrows pointing in the direction of motion of edge channels that are populated by the contacts (crossed squares). Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
5277: \label{fig81}
5278: }
5279: \end{figure}
5280:
5281: Let us apply the Landauer-B\"{u}ttiker formalism to the geometry of Fig.\ \ref{fig81}.
5282: As in Section \ref{sec13b}, the number of spin-degenerate edge channels in the wide
5283: 2DEG and in the constriction are denoted by $N_{\mathrm{wide}}$ and $N_{\min}$, respectively.
5284: An ideal contact to the wide 2DEG perfectly transmits $N_{\mathrm{wide}}$ channels,
5285: whereas the constriction transmits only $N_{\min}$ channels. The remaining
5286: $N_{\mathrm{wide}}-N_{\min}$ channels are reflected back along the opposite 2DEG boundary
5287: (cf.\ Fig.\ \ref{fig81}). We denote by $\mu_{\mathrm{l}}$ and $\mu_{\mathrm{r}}$ the chemical potentials of adjacent
5288: voltage probes to the left and to the right of the constriction. The current
5289: source is at $\mu_{\mathrm{s}}$, and the drain at $\mu_{\mathrm{d}}$. Applying Eq.\ (\ref{eq12.12}) to this case, using
5290: $I_{\mathrm{s}}=-I_{\mathrm{d}}\equiv I$, $I_{\mathrm{r}}=I_{\mathrm{l}}=0$, one finds for the magnetic field direction indicated in Fig.\ \ref{fig81},
5291: \begin{subequations}
5292: \label{eq19.1}
5293: \begin{eqnarray}
5294: (h/2e)I&=&N_{\mathrm{wide}}\mu_{\mathrm{s}}-(N_{\mathrm{wide}}-N_{\min})\mu_{\mathrm{l}}, \label{eq19.1a}\\
5295: 0&=&N_{\mathrm{wide}}\mu_{\mathrm{l}}-N_{\mathrm{wide}}\mu_{\mathrm{s}}, \label{eq19.1b}\\
5296: 0&=&N_{\mathrm{wide}}\mu_{\mathrm{r}}-N_{\min}\mu_{\mathrm{l}}. \label{eq19.1c}
5297: \end{eqnarray}
5298: \end{subequations}
5299: We have used the freedom to choose the zero level of chemical potential by
5300: fixing $\mu_{\mathrm{d}}=0$, so we have three independent (rather than four dependent)
5301: equations. The two-terminal resistance $R_{2\mathrm{t}}\equiv\mu_{\mathrm{s}}/eI$ following from Eq.\ (\ref{eq19.1})
5302: is
5303: \be
5304: R_{2\mathrm{t}}= \frac{h}{2e^{2}}\frac{1}{N_{\min}}, \label{eq19.2}
5305: \ee
5306: unaffected by the presence of the additional voltage probes in Fig.\ \ref{fig81}. The
5307: four-terminal longitudinal resistance $R_{\mathrm{L}}\equiv(\mu_{\mathrm{l}}-\mu_{\mathrm{r}})/eI$ is
5308: \be
5309: R_{\mathrm{L}}= \frac{h}{2e^{2}}\left(\frac{1}{N_{\min}}-\frac{1}{N_{\mathrm{wide}}}\right). \label{eq19.3}
5310: \ee
5311: In the reversed field direction the same result is obtained. Equation (\ref{eq19.3}),
5312: derived for ideal contacts without assuming local equilibrium near the
5313: contacts, is identical to Eq.\ (\ref{eq13.7}), derived for the case of local equilibrium.
5314:
5315: \begin{figure}
5316: \centerline{\includegraphics[width=8cm]{figures/fig82}}
5317: \caption{
5318: Perspective view of a six-terminal Hall bar containing a point contact, showing the various two- and four-terminal resistances mentioned in the text. Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
5319: \label{fig82}
5320: }
5321: \end{figure}
5322:
5323: In a six-terminal measurement geometry (see Fig.\ \ref{fig82}), one can also
5324: measure the Hall resistance in the wide regions, which is simply
5325: $R_{\mathrm{H}}=R_{2\mathrm{t}}-R_{\mathrm{L}}$ or
5326: \be
5327: R_{\mathrm{H}}= \frac{h}{2e^{2}}\frac{1}{N_{\mathrm{wide}}}, \label{eq19.4}
5328: \ee
5329: which is unaffected by the presence of the constriction. This is a consequence
5330: of our assumption of ideal voltage probes. One can also measure the two
5331: four-terminal {\it diagonal\/} resistances $R_{\mathrm{D}}^{+}$ and $R_{\mathrm{D}}^{-}$ across the constriction in such
5332: a way that the two voltage probes are on opposite edges of the 2DEG, on
5333: either side of the constriction (see Fig.\ \ref{fig82}). Additivity of voltages on contacts
5334: tells us that $R_{\mathrm{D}}^{\pm}=R_{\mathrm{H}}\pm R_{\mathrm{L}}$ (for the magnetic field direction of Fig.\ \ref{fig82}); thus,
5335: \be
5336: R_{\mathrm{D}}^{+}= \frac{h}{2e^{2}}\frac{1}{N_{\min}};\;\;R_{\mathrm{D}}^{-}= \frac{h}{2e^{2}}\left(\frac{2}{N_{\mathrm{wide}}}-\frac{1}{N_{\min}}\right). \label{eq19.5}
5337: \ee
5338: On field reversal, $R_{\mathrm{D}}^{+}$ and $R_{\mathrm{D}}^{-}$ are interchanged. Thus, a four-terminal
5339: resistance [$R_{\mathrm{D}}^{+}$ in Eq.\ (\ref{eq19.5})] can in principle be equal to the two-terminal
5340: resistance [$R_{2\mathrm{t}}$ in Eq.\ (\ref{eq19.2})]. The main difference between these two
5341: quantities is that an additive contribution of the ohmic contact resistance
5342: (and of a part of the diffusive background resistance in weak magnetic fields)
5343: is eliminated in the four-terminal resistance measurement.
5344:
5345: The fundamental reason that the assumption of local equilibrium made in
5346: Section \ref{sec13b} (appropriate for weak magnetic fields) and that of ideal contacts
5347: made in this section (for strong fields) yield identical answers is that an ideal
5348: contact attached to the wide 2DEG regions {\it induces\/} a local equilibrium by
5349: equipartitioning the outgoing current among the edge channels. (This is
5350: illustrated in Fig.\ \ref{fig81}, where the current entering the voltage probe to the right
5351: of the constriction is carried by a singe edge channel, while the equally large
5352: current flowing out of that probe is equipartitioned over the two edge
5353: channels available for transport in the wide region.) In weaker magnetic
5354: fields, when the cyclotron radius exceeds the width of the narrow 2DEG
5355: region connecting the voltage probe to the Hall bar, not all edge channels in
5356: the wide 2DEG region are transmitted into the voltage probe (even if it does
5357: not contain a potential barrier). This probe is then not effective in equipartitioning the current. That is the reason that the weak-field analysis in Section
5358: \ref{sec13b} required the assumption of a local equilibrium in the wide 2DEG near
5359: the contacts.
5360:
5361: \begin{figure}
5362: \centerline{\includegraphics[width=8cm]{figures/fig83}}
5363: \caption{
5364: ``Fractional'' quantization in the integer QHE of the four-terminal longitudinal conductance $R_{\rm L}^{-1}$ of a point contact in a magnetic field of 1.4 T at $T = 0.6\,{\rm K}$. The solid horizontal lines indicate the quantized plateaus predicted by Eq.\ (\ref{eq19.3}), with $N_{\rm wide} = 5$ and $N_{\rm min} = 1,2,3,4$. The dashed lines give the location of the spin-split plateaux, which are not well resolved at this magnetic field value. Taken from L. P. Kouwenhoven, Master's thesis, Delft University of Technology, 1988.
5365: \label{fig83}
5366: }
5367: \end{figure}
5368:
5369: We now discuss some experimental results, which confirm the behavior
5370: predicted by Eq.\ (\ref{eq19.3}) in the QHE regime, to complement the weak-field
5371: experiments discussed in Section \ref{sec13b}. Measurements on a quantum point
5372: contact by Kouwenhoven et al.\cite{ref307,ref465} in Fig.\ \ref{fig83} show the quantization of the
5373: longitudinal conductance $R_{\mathrm{L}}^{-1}$ in {\it fractions\/} of $2e^{2}/h$ (for unresolved spin
5374: degeneracy). The magnetic field is kept fixed at $1.4\, \mathrm{T}$ (such that $N_{\mathrm{wide}}=5$) and
5375: the gate voltage is varied (such that $N_{\min}$ ranges from 1 to 4). Conductance
5376: plateaux close to 5/4, 10/3, 15/2, and $20\times(2e^{2}/h)$ (solid horizontal lines) are
5377: observed, in accord with Eq.\ (\ref{eq19.3}). Spin-split plateaux (dashed lines) are
5378: barely resolved at this rather low magnetic field. Similar data were reported
5379: by Snell et al.\cite{ref342} Observations of such a ``fractional'' quantization due to the
5380: integer QHE were made before on wide Hall bars with regions of different
5381: electron density in series,\cite{ref466,ref467} but the theoretical explanation\cite{ref468} given at
5382: that time was less straightforward than Eq.\ (\ref{eq19.3}).
5383:
5384: In the high-field regime the point contact geometry of Fig.\ \ref{fig81} is essentially
5385: equivalent to a geometry in which a potential barrier is present across the
5386: entire width of the Hall bar (created by means of a narrow continuous gate).
5387: The latter geometry was studied by Haug et al.\cite{ref340} and by Washburn et al.\cite{ref339}
5388: The geometries of both experiments\cite{ref339,ref340} are the same (see Figs.\ \ref{fig84} and \ref{fig85}),
5389: but the results exhibit some interesting differences because of the different
5390: dimensions of gate and channel. Hauge et al.\cite{ref340} used a sample of macroscopic dimensions, the channel width being 100 $\mu \mathrm{m}$ and the gate length 10
5391: and 20 $\mu \mathrm{m}$. Results are shown in Fig.\ \ref{fig84}. As the gate voltage is varied, a
5392: quantized plateau at $h/2e^{2}$ is seen in the longitudinal resistance at fixed
5393: magnetic field, in agreement with Eq.\ (\ref{eq19.3}) (the plateau occurs for two spin-split Landau levels in the wide region and one spin-split level under the gate).
5394: A qualitatively different aspect of the data in Fig.\ \ref{fig84}, compared with Fig.\ \ref{fig83},
5395: is the presence of a resistance minimum. Equation (\ref{eq19.3}), in contrast, predicts
5396: that $R_{\mathrm{L}}$ varies monotonically with barrier height, and thus with gate voltage.
5397: A model for the effect has been proposed in a different paper by Haug et
5398: al.,\cite{ref341} based on a competition between backscattering and tunneling through
5399: localized states in the barrier region. They find that edge states that are
5400: totally reflected at a given barrier height may be partially transmitted if the
5401: barrier height is further increased. The importance of tunneling is consistent
5402: with the increase of the amplitude of the dip as the gate length is reduced from
5403: 20 to 10 $\mu \mathrm{m}$. A related theoretical study was performed by Zhu et al.\cite{ref469}
5404:
5405: \begin{figure}
5406: \centerline{\includegraphics[width=6cm]{figures/fig84a}}
5407:
5408: \centerline{\includegraphics[width=6cm]{figures/fig84b}}
5409: \caption{
5410: (a) Schematic view of a wide Hall bar containing a potential barrier imposed by a gate electrode of length $b_{\rm g}$. (b) Longitudinal resistance as a function of gate voltage in the QHE regime (two spin-split Landau levels are occupied in the unperturbed electron gas regions). The plateau shown is at $R_{\rm L} = h/2e^{2}$, in agreement with Eq.\ (\ref{eq19.3}). Results for $b_{\rm g} = 10\,\mu{\rm m}$ and $20\,\mu{\rm m}$ are compared. A pronounced dip develops in the device with the shortest gate length. Taken from R. J. Haug et al., Phys.\ Rev.\ B {\bf 39}, 10892 (1989).
5411: \label{fig84}
5412: }
5413: \end{figure}
5414:
5415: \begin{figure}
5416: \centerline{\includegraphics[width=8cm]{figures/fig85}}
5417: \caption{
5418: (a) Schematic view of a $2$-$\mu{\rm m}$-wide channel containing a potential barrier imposed by a $0.1$-$\mu{\rm m}$-long gate. (b) Top: diagonal resistance $R_{13,42} \equiv R_{\rm D}^{+}$ and longitudinal resistance $R_{12,43} \equiv R_{\rm L}$ as a function of gate voltage in a strong magnetic field ($B = 5.2\,{\rm T}$), showing a quantized plateau in agreement with Eqs.\ (\ref{eq19.5}) and (\ref{eq19.3}), respectively. For comparison also the two zero-field traces are shown, which are almost identical. Bottom: Difference $R_{\rm D}^{+}- R_{\rm L} = R_{\rm H}$ at 5.2 T. A normal quantum Hall plateau is found, with oscillatory structure superimposed in gate voltage regions where $R_{\rm D}^{+}$ and $R_{\rm L}$ are not quantized. Taken from S. Washburn et al., Phys.\ Rev.\ Lett.\ {\bf 61}, 2801 (1988).
5419: \label{fig85}
5420: }
5421: \end{figure}
5422:
5423: Washburn et al.\cite{ref339} studied the longitudinal resistance of a barrier defined
5424: by a $0.1$-$\mu \mathrm{m}$-long gate across a $2$-$\mu \mathrm{m}$-wide channel. The relevant dimensions
5425: are thus nearly two orders of magnitude smaller than in the experiment of
5426: Haug et al. Again, the resistance is studied as a function of gate voltage at
5427: fixed magnetic field. The longitudinal $(R_{\mathrm{L}}\equiv R_{12,43})$ and diagonal
5428: $(R_{\mathrm{D}}^{+}\equiv R_{13,42})$ resistances are shown in Fig.\ \ref{fig85}, as well as their difference
5429: [which according to Eqs.\ (\ref{eq19.3}) and (\ref{eq19.5}) would equal the Hall resistance
5430: $R_{\mathrm{H}}$]. In this small sample the quantized plateaux predicted by Eq.\ (\ref{eq19.3}) are
5431: clearly seen, but the resistance dips of the large sample of Haug et al.\ are not.
5432: We recall that resistance dips were not observed in the quantum point
5433: contact experiment of Fig.\ \ref{fig83} either. The model of Haug et al.\cite{ref341} would imply
5434: that localized states do not form in barriers of small area. Washburn et al.\
5435: find weak resistance fluctuations in the gate voltage intervals between
5436: quantized plateaux. These fluctuations are presumably due to some form of
5437: quantum interference, but have not been further identified.
5438: Related experiments on the quantum Hall effect in a 2DEG with a
5439: potential barrier have been performed by Hirai et al.\ and by Komiyama et
5440: al.\cite{ref427,ref470,ref471,ref472} These studies have focused on the role of nonideal contacts in
5441: the QHE, which is the subject of the next subsection.
5442:
5443: \subsubsection{\label{sec19b} Disordered contacts}
5444:
5445: The validity of Eqs.\ (\ref{eq19.2}--\ref{eq19.5}) in the QHE regime breaks down for
5446: nonideal contacts if local equilibrium near the contacts is not established. The
5447: treatment of Section \ref{sec19a} for ideal contacts implies that the Hall voltage over
5448: the wide 2DEG regions adjacent to the constriction is {\it unaffected\/} by the
5449: presence of the constriction or potential barrier. Experiments by Komiyama
5450: et al.\cite{ref427,ref472} have demonstrated that this is no longer true if one or more
5451: contacts are disordered. The analysis of their experiments is rather involved,\cite{ref472} which is why we do not give a detailed discussion here. Instead we
5452: review a different experiment,\cite{ref113} which shows a deviating Hall resistance in a
5453: sample with a constriction and a singe disordered contact. This experiment
5454: can be analyzed in a relatively simple way,\cite{ref307} following the work of
5455: B\"{u}ttiker\cite{ref112} and Komiyama et al.\cite{ref427,ref470,ref471,ref472}
5456:
5457: \begin{figure}
5458: \centerline{\includegraphics[width=8cm]{figures/fig86}}
5459: \caption{
5460: Nonvanishing Shubnikov-De Haas minima in the longitudinal resistance $R_{\rm L}$ and anomalous quantum Hall resistance $R_{\rm H}$, measured in the point contact geometry of Fig.\ \ref{fig82} at 50 mK. These experimental results are extensions to higher fields of the weak-field traces shown in Fig.\ \ref{fig50}. The Hall resistance has been measured across the wide region, more than $100\,\mu{\rm m}$ away from the constriction, yet $R_{\rm H}$ is seen to increase if the gate voltage is raised from $- 0.3\,{\rm V}$ to $-2.5\,{\rm V}$. The magnitude at $B = 2.2\,{\rm T}$ of the deviation in $R_{\rm H}$ and of the Shubnikov-De Haas minimum in $R_{\rm L}$ are indicated by arrows, which both for $R_{\rm H}$ and $R_{\rm L}$ have a length of $(h/2e^{2})(\frac{1}{2}-\frac{1}{3})$, in agreement with the analysis given in the text. Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
5461: \label{fig86}
5462: }
5463: \end{figure}
5464:
5465: The sample geometry is that of Fig.\ \ref{fig82}. In Fig.\ \ref{fig86} the four-terminal
5466: longitudinal resistance $R_{\mathrm{L}}$ and Hall resistance $R_{\mathrm{H}}$ are shown for both a small
5467: voltage ($-0.3$ V) and a large voltage ($-2.5$ V) on the gate defining the
5468: constriction. The longitudinal resistance decreases in weak fields because of
5469: reduction of backscattering, as discussed in Section \ref{sec13b}. At larger fields
5470: Shubnikov-De Haas oscillations develop. The data for $V_{\mathrm{g}}=-0.3\,\mathrm{V}$ exhibit
5471: zero minima in the Shubnikov-De Haas oscillations in $R_{\mathrm{L}}$ and the normal
5472: quantum Hall resistance $R_{\mathrm{H}}=(h/2e^{2})N_{\mathrm{wide}}^{-1}$, determined by the number of
5473: Landau levels occupied in the wide regions ($N_{\mathrm{wide}}$ can be obtained from the
5474: quantum Hall effect measured in the absence of the constriction or from the
5475: periodicity of the Shubnikov-De Haas oscillations).
5476:
5477: At the higher gate voltage $V_{\mathrm{g}}=-2.5\,\mathrm{V}$, nonvanishing minima in $R_{\mathrm{L}}$ are
5478: seen in Fig.\ \ref{fig86} as a result of the formation of a potential barrier in the
5479: constriction. At the minima, $R_{\mathrm{L}}$ has the fractional quantization predicted by
5480: Eq.\ (\ref{eq19.3}). For example, the plateau in $R_{\mathrm{L}}$ around $2.2\, \mathrm{T}$ for $V_{\mathrm{g}}=-2.5\,\mathrm{V}$ is
5481: observed to be at $R_{\mathrm{L}}=2.1\,\mathrm{k}\Omega\approx(h/2e^{2})\times(\frac{1}{2}-\frac{1}{3})$, in agreement with the fact
5482: that the two-terminal resistance yields $N_{\min}=2$ and the number of Landau
5483: levels in the wide regions $N_{\mathrm{wide}}=3$. In spite of this agreement, and in
5484: apparent conflict with Eq.\ (\ref{eq19.4}), the Hall resistance $R_{\mathrm{H}}$ has {\it increased\/} over its
5485: value for small gate voltages. Indeed, around $2.2\, \mathrm{T}$ a Hall plateau at
5486: $R_{\mathrm{H}}=6.3\, \mathrm{k}\Omega\approx(h/2e^{2})\times\frac{1}{2}$ is found for $V_{\mathrm{g}}=-2.5\,\mathrm{V}$, as if the number of
5487: occupied Landau levels was given by $N_{\min}=2$ rather than by $N_{\mathrm{wide}}=3$. This
5488: unexpected deviation was noted in Ref.\ \onlinecite{ref113}, but was not understood at the
5489: time. At higher magnetic fields (not shown in Fig.\ \ref{fig86}) the $N=1$ plateau is
5490: reached, and the deviation in the Hall resistance vanishes.
5491:
5492: As pointed out in Ref.\ \onlinecite{ref307}, the likely explanation of the data of Fig.\ \ref{fig86} is
5493: that one of the ohmic contacts used to measure the Hall voltage is {\it disordered}
5494: in the sense of B\"{u}ttiker\cite{ref112} that not all edge channels have unit transmission probability into the voltage probe. The disordered contact can be
5495: modeled by a potential barrier in the lead with a height not below that of the
5496: barrier in the constriction, as illustrated in Fig.\ \ref{fig87}. A net current $I$ flows
5497: through the constriction, determined by its two-terminal resistance according
5498: to $I=(2e/h)N_{\min}\mu_{\mathrm{s}}$, with $\mu_{\mathrm{s}}$ the chemical potential of the source reservoir (the
5499: chemical potential of the drain reservoir $\mu_{\mathrm{d}}$ is taken as a zero reference).
5500: Equation (\ref{eq12.12}) applied to the two opposite Hall probes $l_{1}$ and $l_{2}$ in Fig.\ \ref{fig87}
5501: takes the form (using $I_{l_{1}}=I_{l_{2}}=0$, $\mu_{\mathrm{s}}=(h/2e)I/N_{\min}$, and $\mu_{\mathrm{d}}=0$)
5502: \begin{subequations}
5503: \label{eq19.6}
5504: \begin{eqnarray}
5505: 0&=&N_{\mathrm{wide}} \mu_{l_{1}}-T_{\mathrm{s}\rightarrow l_{1}}\frac{h}{2e}\frac{I}{N_{\min}}-T_{l_{2}\rightarrow l_{1}}\mu_{l_{2}}, \label{eq19.6a}\\
5506: 0&=&N_{l_{2}} \mu_{l_{2}}-T_{\mathrm{s}\rightarrow l_{2}}\frac{h}{2e}\frac{I}{N_{\min}}-T_{l_{1}\rightarrow l_{2}}\mu_{l_{1}}, \label{eq19.6b}
5507: \end{eqnarray}
5508: \end{subequations}
5509: where we have assumed that the disordered Hall probe $l_{2}$ transmits only
5510: $N_{l_{2}}<N_{\mathrm{wide}}$ edge channels because of the barrier in the lead. For the field
5511: direction shown in Fig.\ \ref{fig87} one has, under the assumption of no inter-edge-channel scattering from constriction to probe $l_{2}$, $T_{\mathrm{s}\rightarrow l_{1}}=N_{\mathrm{wide}}$,
5512: $T_{\mathrm{s}\rightarrow l_{2}}=T_{l_{2}\rightarrow l_{1}}=0$, and $T_{l_{1}\rightarrow l_{2}}= \max(0, N_{l_{2}}-N_{\min})$. Equation (\ref{eq19.6}) then
5513: leads to a Hall resistance $R_{\mathrm{H}}\equiv(\mu_{l_{1}}-\mu_{l_{2}})/eI$ given by
5514: \be
5515: R_{\mathrm{H}}=\frac{h}{2e^{2}}\frac{1}{\max(N_{l_{2}}, N_{\min})}.\label{eq19.7}
5516: \ee
5517: In the opposite field direction the normal Hall resistance $R_{\mathrm{H}}=(h/2e^{2})N_{\mathrm{wide}}^{-1}$ is
5518: recovered.
5519:
5520: \begin{figure}
5521: \centerline{\includegraphics[width=8cm]{figures/fig87}}
5522: \caption{
5523: Illustration of the flow of edge channels along equipotentials in a sample with a constriction (defined by the shaded gates) and a disordered voltage probe (a potential barrier in the probe is indicated by the shaded bar). Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
5524: \label{fig87}
5525: }
5526: \end{figure}
5527:
5528: The assumption of a single disordered probe, plus absence of interedge
5529: channel scattering from constriction to probe, thus explains the observation
5530: in Fig.\ \ref{fig86} of an anomalously high quantum Hall resistance for large gate
5531: voltages, such that $N_{\min}<N_{\mathrm{wide}}$. Indeed, the experimental Hall resistance for
5532: $V_{\mathrm{g}}=-2.5\,\mathrm{V}$ has a plateau around $2.2\, \mathrm{T}$ close to the value $R_{\mathrm{H}}=(h/2e^{2})N_{\min}^{-1}$
5533: (with $N_{\min}=2$), in agreement with Eq.\ (\ref{eq19.7}) if $N_{l_{2}}\leq N_{\min}$ at this gate voltage.
5534: This observation demonstrates the absence of interedge channel scattering
5535: over 100 $\mu \mathrm{m}$ (the separation of constriction and probe), but only between the
5536: highest-index channel (with index $n=N_{\mathrm{wide}}=3$) and the two lower-index
5537: channels. Since the $n=1$ and $n=2$ edge channels are either both empty or
5538: both filled (cf.\ Fig.\ \ref{fig87}, where these two edge channels lie closest to the sample
5539: boundary), any scattering between $n=1$ and 2 would have no measurable
5540: effect on the resistances. As discussed in Section \ref{sec19c}, we know from the work
5541: of Alphenaar et al.\cite{ref429} that (at least in the present samples) the edge channels
5542: with $n\leq N_{\mathrm{wide}}-1$ do in fact equilibrate to a large extent on a length scale of
5543: $100\,\mu \mathrm{m}$.
5544:
5545: In the absence of a constriction, or at small gate voltages (where the
5546: constriction is just defined), one has $N_{\min}=N_{\mathrm{wide}}$ so that the normal Hall
5547: effect is observed in both field directions. This is the situation realized in the
5548: experimental trace for $V_{\mathrm{g}}=-0.3\,\mathrm{V}$ in Fig.\ \ref{fig86}. In very strong fields such that
5549: $N_{\min}=N_{l_{2}}=N_{\mathrm{wide}}=1$ (still assuming nonresolved spin splitting), the
5550: normal result $R_{\mathrm{H}}=h/2e^{2}$ would follow even if the contacts contain a
5551: potential barrier, in agreement with the experiment (not shown in Fig.\ \ref{fig86}).
5552: This is a more general result, which holds also for a barrier that only partially
5553: transmits the $n=1$ edge channel.\cite{ref112,ref308,ref472,ref473,ref474,ref475}
5554:
5555: A similar analysis as the foregoing predicts that the longitudinal resistance
5556: measured on the edge of the sample that contains ideal contacts retains its
5557: regular value (\ref{eq19.3}). On the opposite sample edge the measurement would
5558: involve the disordered contact, and one finds instead
5559: \be
5560: R_{\mathrm{L}}= \frac{h}{2e^{2}}\left(\frac{1}{N_{\min}}-\frac{1}{\max(N_{l_{2}},N_{\min})}\right) \label{eq19.8}
5561: \ee
5562: for the field direction shown in Fig.\ \ref{fig87}, while Eq.\ (\ref{eq19.3}) is recovered for the
5563: other field direction. The observation in the experiment of Fig.\ \ref{fig86} for
5564: $V_{\mathrm{g}}=-2.5\,\mathrm{V}$ of a regular longitudinal resistance [in agreement with Eq.\
5565: (\ref{eq19.3})], along with an anomalous quantum Hall resistance is thus consistent
5566: with this analysis.
5567:
5568: The experiments\cite{ref426,ref429} discussed in the following subsection are topologically equivalent to the geometry of Fig.\ \ref{fig87}, but involve quantum point
5569: contacts rather than ohmic contacts. This gives the possibility of populating
5570: and detecting edge channels selectively, thereby enabling a study of the effects
5571: of a nonequilibrium population of edge channels in a controlled manner.
5572:
5573: \subsubsection{\label{sec19c} Quantum point contacts}
5574:
5575: \begin{figure}
5576: \centerline{\includegraphics[width=6cm]{figures/fig88a}}
5577:
5578: \centerline{\includegraphics[width=6cm]{figures/fig88b}}
5579: \caption{
5580: (a) Schematic potential landscape, showing the 2DEG boundary and the saddleshaped injector and collector point contacts. In a strong magnetic field the edge channels are extended along equipotentials at the guiding center energy, as indicated here for edge channels with index $n = 1,2$ (the arrows point in the direction of motion). In this case a Hall conductance of ($2e^{2}/h)N$ with $N = 1$ would be measured by the point contacts, in spite of the presence of two occupied spin-degenerate Landau levels in the bulk 2DEG. Taken from C. W. J. Beenakker et al., Festk\"{o}rperprobleme {\bf 29}, 299 (1989). (b) Three-terminal conductor in the electron focusing geometry. Taken from H. van Houten et al., Phys.\ Rev.\ B {\bf 39}, 8556 (1989).
5581: \label{fig88}
5582: }
5583: \end{figure}
5584:
5585: In Section \ref{sec14} we have seen how a quantum point contact can inject a
5586: {\it coherent\/} superposition of edge channels at the 2DEG boundary, in the
5587: coherent electron focusing experiment.\cite{ref59} In that section we restricted
5588: ourselves to weak magnetic fields. Here we discuss the experiment by van
5589: Wees et al.,\cite{ref426} which shows how in the QHE regime the point contacts can be
5590: operated in a different way as {\it selective\/} injectors (and detectors) of edge
5591: channels. We recall that electron focusing can be measured as a generalized
5592: Hall resistance, in which case the pronounced peaked structure due to mode
5593: interference is superimposed on the weak-field Hall resistance (cf.\ Fig.\ \ref{fig53}). If
5594: the weak-field electron-focusing experiments are extended to stronger magnetic fields, a transition is observed to the quantum Hall effect, provided the
5595: injecting and detecting point contacts are not too strongy pinched off.\cite{ref59} The
5596: oscillations characteristic of mode interference disappear in this field regime,
5597: suggesting that the coupling of the edge channels (which form the propagating modes from injector to collector) is suppressed, and adiabatic transport is
5598: realized. It is now no longer sufficient to model the point contacts by a point
5599: source-detector of infinitesimal width (as was done in Section \ref{sec14}), but a
5600: somewhat more detailed description of the electrostatic potential $V(x, y)$
5601: defining the point contacts and the 2DEG boundary between them is
5602: required. Schematically, $V(x, y)$ is represented in Fig.\ \ref{fig88}a. Fringing fields from
5603: the split gate create a potential barrier in the point contacts, so $V$ has a saddle
5604: form as shown. The heights of the barriers $E_{\mathrm{i}}, E_{\mathrm{c}}$ in the injector and collector
5605: are separately adjustable by means of the voltages on the split gates and can
5606: be determined from the two-terminal conductances of the individual point
5607: contacts. The point contact separation in the experiment of Ref.\ \onlinecite{ref426} is small
5608: (1.5 $\mu \mathrm{m}$), so one can assume fully adiabatic transport from injector to
5609: collector in strong magnetic fields. As discussed in Section \ref{sec18}, adiabatic
5610: transport is along equipotentials at the guiding center energy $E_{\mathrm{G}}$. Note that
5611: the edge channel with the smallest index $n$ has the largest guiding center
5612: energy [according to Eq.\ (\ref{eq18.1})]. In the absence of inter-edge-channel
5613: scattering, edge channels can only be transmitted through a point contact if
5614: $E_{\mathrm{G}}$ exceeds the potential barrier height (disregarding tunneling through the
5615: barrier). The injector thus injects $N_{\mathrm{i}}\approx(E_{\mathrm{F}}-E_{\mathrm{i}})/\hbar\omega_{\mathrm{c}}$ edge channels into the
5616: 2DEG, while the collector is capable of detecting $N_{\mathrm{c}}\approx(E_{\mathrm{F}}-E_{\mathrm{c}})/\hbar\omega_{\mathrm{c}}$
5617: channels. Along the boundary of the 2DEG, however, a larger number of
5618: $N_{\mathrm{wide}}\approx E_{\mathrm{F}}/\hbar\omega_{\mathrm{c}}$ edge channels, equal to the number of occupied bulk Landau
5619: levels in the 2DEG, are available for transport at the Fermi level. The
5620: selective population and detection of Landau levels leads to deviations from
5621: the normal Hall resistance.
5622:
5623: These considerations can be put on a theoretical basis by applying the
5624: Landauer-B\"{u}ttiker formalism discussed in Section \ref{sec12} to the electron-focusing geometry.\cite{ref80} We consider a three-terminal conductor as shown in
5625: Fig.\ \ref{fig88}b, with point contacts in two of the probes (injector $\mathrm{i}$ and collector $\mathrm{c}$),
5626: and a wide ideal drain contact $\mathrm{d}$. The collector acts as a voltage probe,
5627: drawing no net current, so that $I_{\mathrm{c}}=0$ and $I_{\mathrm{d}}=-I_{\mathrm{i}}$. The zero of energy is
5628: chosen such that $\mu_{\mathrm{d}}=0$. One then finds from Eq.\ (\ref{eq12.12}) the two equations
5629: \begin{subequations}
5630: \label{eq19.9}
5631: \begin{eqnarray}
5632: 0&=&(N_{\mathrm{c}}-R_{\mathrm{c}})\mu_{\mathrm{c}}-T_{\mathrm{i}\rightarrow \mathrm{c}}\mu_{\mathrm{i}}, \label{19.9a}\\
5633: (h/2e)I_{\mathrm{i}}&=&(N_{\mathrm{i}}-R_{\mathrm{i}})\mu_{\mathrm{i}}-T_{\mathrm{c}\rightarrow \mathrm{i}}\mu_{\mathrm{c}}, \label{eq19.9b}
5634: \end{eqnarray}
5635: \end{subequations}
5636: and obtains for the ratio of collector voltage $V_{\mathrm{c}}=\mu_{\mathrm{c}}/e$ (measured relative to
5637: the voltage of the current drain) to injected current $I_{\mathrm{i}}$ the result
5638: \be
5639: \frac{V_{\mathrm{c}}}{I_{\mathrm{i}}}=\frac{2e^{2}}{h}\frac{T_{\mathrm{i}\rightarrow \mathrm{c}}}{G_{\mathrm{i}}G_{\mathrm{c}}-\delta}. \label{eq19.10}
5640: \ee
5641: Here $\delta\equiv(2e^{2}/h)^{2}T_{\mathrm{i}\rightarrow \mathrm{c}}T_{\mathrm{c}\rightarrow \mathrm{i}}$, and $G_{\mathrm{i}}\equiv(2e^{2}/h)(N_{\mathrm{i}}-R_{\mathrm{i}})$, $G_{\mathrm{c}}\equiv(2e^{2}/h)(N_{\mathrm{c}}-R_{\mathrm{c}})$
5642: denote the conductances of injector and collector point contact.
5643:
5644: \begin{figure}
5645: \centerline{\includegraphics[width=8cm]{figures/fig89}}
5646: \caption{
5647: Experimental correlation between the conductances $G_{\rm i}$, $G_{\rm c}$ of injector and collector, and the Hall conductance $G_{\rm H}\equiv I_{\rm i}/V_{\rm c}$, shown to demonstrate the validity of Eq.\ (\ref{eq19.11}) ($T = 1.3\,{\rm K}$, point contact separation is $1.5\,\mu{\rm m}$). The magnetic field was kept fixed (top: $B = 2.5\,{\rm T}$, bottom: $B = 3.8\,{\rm T}$, corresponding to a number of occupied bulk Landau levels $N = 3$ and $2$, respectively). By increasing the gate voltage on one half of the split-gate defining the injector, $G_{\rm i}$ was varied at constant $G_{\rm c}$. Taken from B. J. van Wees et al., Phys.\ Rev.\ Lett.\ {\bf 62}, 1181 (1989).
5648: \label{fig89}
5649: }
5650: \end{figure}
5651:
5652: For the magnetic field direction indicated in Fig.\ \ref{fig88}, the term $\delta$ in Eq.\
5653: (\ref{eq19.10}) can be neglected since $T_{\mathrm{c}\rightarrow \mathrm{i}}\approx 0$ [the resulting Eq.\ (\ref{eq14.2}) was used in
5654: Section \ref{sec14}]. An additional simplification is possible in the adiabatic transport
5655: regime. We consider the case that the barrier in one of the two point contacts
5656: is sufficiently higher than in the other, to ensure that electrons that are
5657: transmitted over the highest barrier will have a negligible probability of being
5658: reflected at the lowest barrier. Then $T_{\mathrm{i}\rightarrow \mathrm{c}}$ is dominated by the transmission
5659: probability over the highest barrier, $T_{\mathrm{i}\rightarrow \mathrm{c}} \approx\min(N_{\mathrm{i}}-R_{\mathrm{i}}, N_{\mathrm{c}}-R_{\mathrm{c}})$. Substitution in Eq.\ (\ref{eq19.10}) gives the remarkable result\cite{ref426} that the {\it Hall conductance\/}
5660: $G_{\mathrm{H}}\equiv I_{\mathrm{i}}/V_{\mathrm{c}}$ measured in the electron focusing geometry can be expressed
5661: entirely in terms of the {\it contact conductances\/} $G_{\mathrm{i}}$ and $G_{\mathrm{c}}$:
5662: \be
5663: G_{\mathrm{H}} \approx\max(G_{\mathrm{i}}, G_{\mathrm{c}}). \label{eq19.11}
5664: \ee
5665: Equation (\ref{eq19.11}) tells us that quantized values of $G_{\mathrm{H}}$ occur not at
5666: $(2e^{2}/h)N_{\mathrm{wide}}$, as one would expect from the $N_{\mathrm{wide}}$ populated Landau levels in
5667: the 2DEG but at the smaller value of $(2e^{2}/h) \max(N_{\mathrm{i}}, N_{\mathrm{c}})$. As shown in Fig.\ \ref{fig89}
5668: this is indeed observed experimentally.\cite{ref426} Notice in particular how any
5669: deviation from quantization in
5670: $\max(G_{\mathrm{i}}, G_{\mathrm{c}})$ is faithfully reproduced in $G_{\mathrm{H}}$, in
5671: complete agreement with Eq.\ (\ref{eq19.11}).
5672:
5673: \begin{figure}
5674: \centerline{\includegraphics[width=8cm]{figures/fig90}}
5675: \caption{
5676: Results of an experiment similar to that of Fig.\ \ref{fig89}, but with a much larger separation of $80\,\mu{\rm m}$ between injector and collector. Shown are $R_{\rm i} = G_{\rm i}^{-1}$, $R_{\rm c} = G_{\rm c}^{-1}$, and $R_{\rm H} = G_{\rm H}^{-1}$, as a function of the gate voltage on the collector. ($T = 0.45\,{\rm K}$, $B = 2.8\,{\rm T}$; the normal quantized Hall resistance is $\frac{1}{3}(h/2e^{2})$.) Regimes I, II, and III are discussed in the text. Taken from B. W. Alphenaar et al., Phys.\ Rev.\ Lett.\ {\bf 64}, 677 (1990).
5677: \label{fig90}
5678: }
5679: \end{figure}
5680:
5681: The experiment of Ref.\ \onlinecite{ref426} was repeated by Alphenaar et al.\cite{ref429} for much
5682: larger point contact separations $(\approx 100\,\mu \mathrm{m})$, allowing a study of the length
5683: scale for equilibration of edge channels at the 2DEG boundary. Even after
5684: such a long distance, no complete equilibration of the edge channels was
5685: found, as manifested by a dependence of the Hall resistance on the gate
5686: voltage used to vary the number of edge channels transmitted through the
5687: point contact voltage probe (see Fig.\ \ref{fig90}). As discussed in Section \ref{sec18b}, a
5688: dependence of the resistance on the properties of the contacts is only possible
5689: in the absence of local equilibrium. In contrast to the experiment by van Wees
5690: et al.,\cite{ref426} and in disagreement with Eq.\ (\ref{eq19.11}), the Hall resistance in Fig.\ \ref{fig90}
5691: does not simply follow the smallest of the contact resistances of current and
5692: voltage probe. This implies that the assumption of fully adiabatic transport
5693: has broken down on a length scale of $100\,\mu \mathrm{m}$.
5694:
5695: In the experiment a magnetic field was applied such that three edge
5696: channels were available at the Fermi level. The contact resistance of the
5697: injector was adjusted to $R_{i}=h/2e^{2}$, so current was injected in a single edge
5698: channel $(n=1)$ only. The gate voltage defining the collector point contact
5699: was varied. In Fig.\ \ref{fig90} the contact resistances of injector $(R_{\mathrm{i}})$ and collector $(R_{\mathrm{c}})$
5700: are plotted as a function of this gate voltage, together with the Hall resistance
5701: $R_{\mathrm{H}}$. At zero gate voltage the Hall resistance takes on its normal quantized
5702: value [$R_{\mathrm{H}}= \frac{1}{3}(h/2e^{2})$]. On increasing the negative gate voltage three regions
5703: of interest are traversed (labeled III to I in Fig.\ \ref{fig90}). In region III edge
5704: channels 1 and 2 are completely transmitted through the collector, but the
5705: $n=3$ channel is partially reflected. In agreement with Eq.\ (\ref{eq19.11}), $R_{\mathrm{H}}$
5706: increases following $R_{\mathrm{c}}$. As region II is entered, $R_{\mathrm{H}}$ levels off while $R_{\mathrm{c}}$ continues
5707: to increase up to the $\frac{1}{2}(h/2e^{2})$ quantized value. The fact that $R_{\mathrm{H}}$ stops slightly
5708: short of this value proves that some scattering between the $n=3$ and $n=1,2$
5709: channels has occurred. On increasing the gate voltage further, $R_{\mathrm{c}}$ rises to
5710: $h/2e^{2}$ in region I. However, $R_{\mathrm{H}}$ shows hardly any increase with respect to its
5711: value in region II. This demonstrates that the $n=2$ and $n=1$ edge channels
5712: have almost fully equilibrated. A quantitative analysis\cite{ref429} shows that, in fact,
5713: $92\%$ of the current originally injected into the $n=1$ edge channel is
5714: redistributed equally over the $n=1$ and $n=2$ channels, whereas only $8\%$ is
5715: transferred to the $n=3$ edge channel. The suppression of scattering between
5716: the highest-index $n=N$ edge channel and the group of edge channels with
5717: $n\leq N-1$ was found to exist only if the Fermi level lies in (or near) the $N\mathrm{th}$
5718: bulk Landau level. As a qualitative explanation it was suggested\cite{ref429,ref476} that
5719: the $N\mathrm{th}$ edge channel hybridizes with the $N\mathrm{th}$ bulk Landau level when both
5720: types of states coexist at the Fermi level. Such a coexistence does not occur
5721: for $n\leq N-1$ if the potential fluctuations are small compared with $\hbar\omega_{\mathrm{c}}$ (cf.\
5722: Fig.\ \ref{fig78}). The spatial extension of the wave functions of the edge channels is
5723: illustrated in Fig.\ \ref{fig91} (shaded ellipsoids) for various values of the Fermi level
5724: between the $n=3$ and $n=4$ bulk Landau levels. As the Fermi level
5725: approaches the $n=3$ bulk Landau level, the corresponding edge channel
5726: penetrates into the bulk, so the overlap with the wave functions of lower-index edge channels decreases. This would explain the decoupling of the
5727: $n=3$ and $n=1,2$ edge channels.
5728:
5729: \begin{figure}
5730: \centerline{\includegraphics[width=8cm]{figures/fig91}}
5731: \caption{
5732: Illustration of the spatial extension (shaded ellipsoids) of edge channels for four different values of the Fermi energy. The $n = 3$ edge channel can penetrate into the bulk by hybridizing with the $n = 3$ bulk Landau level, coexisting at the Fermi level. This would explain the absence of equilibration between the $n = 3$ and $n = 1,2$ edge channels. The penetration depth $l_{\rm loc}$ and the magnetic length are indicated. Taken from B. W. Alphenaar et al., Phys.\ Rev.\ Lett.\ {\bf 64}, 677 (1990).
5733: \label{fig91}
5734: }
5735: \end{figure}
5736:
5737: These experiments thus point the way in which the transition from
5738: microscopic to macroscopic behavior takes place in the QHE, while they also
5739: demonstrate that quite large samples will be required before truly macroscopic behavior sets in.
5740:
5741: \subsubsection{\label{sec19d} Suppression of the Shubnikov-De Haas oscillations}
5742:
5743: Shubnikov-De Haas magnetoresistance oscillations were discussed in
5744: Sections \ref{sec4c} and \ref{sec10}. In weak magnetic fields, where a theoretical description
5745: in terms of a local resistivity tensor applies, a satisfactory agreement between
5746: theory and experiment is obtained.\cite{ref20} As we now know, in strong magnetic
5747: fields the concept of a local resistivity tensor may break down entirely
5748: because of the absence of local equilibrium. A theory of the Shubnikov-De
5749: Haas effect then has to take into account explicitly the properties of the
5750: contacts used for the measurement. The resulting anomalies are considered in
5751: this subsection.
5752:
5753: \begin{figure}
5754: \centerline{\includegraphics[width=8cm]{figures/fig92}}
5755: \caption{
5756: Illustration of the mechanism for the suppression of Shubnikov-De Haas oscillations due to selective detection of edge channels. The black area denotes the split-gate point contact in the voltage probe, which is at a distance of $250\,\mu{\rm m}$ from the drain reservoir. Dashed arrows indicate symbolically the selective back scattering in the highest-index edge channel, via states in the highest bulk Landau level that coexist at the Fermi level. Taken from H. van Houten et al., in Ref.\ \onlinecite{ref9}.
5757: \label{fig92}
5758: }
5759: \end{figure}
5760:
5761: Van Wees et al.\cite{ref428} found that the amplitude of the high-field Shubnikov-De Haas oscillations was suppressed if a quantum point contact was used as
5762: a voltage probe. To discuss this anomalous Shubnikov-De Haas effect, we
5763: consider the three-terminal geometry of Fig.\ \ref{fig92}, where a single voltage
5764: contact is present on the boundary between source and drain contacts. (An
5765: alternative two-terminal measurement configuration is also possible; see Ref.\
5766: \onlinecite{ref428}.) The voltage probe $\mathrm{p}$ is formed by a quantum point contact, while source
5767: $\mathrm{s}$ and drain $\mathrm{d}$ are normal ohmic contacts. (Note that {\it two\/} special contacts were
5768: required for the anomalous quantum Hall effect of Section \ref{sec19c}.) One
5769: straightforwardly finds from Eq.\ (\ref{eq12.12}) that the three-terminal resistance
5770: $R_{3\mathrm{t}}\equiv(\mu_{\mathrm{p}}-\mu_{\mathrm{d}})/eI$ measured between point contact probe and drain is given
5771: by
5772: \be
5773: R_{3\mathrm{t}}= \frac{h}{2e^{2}}\frac{T_{\mathrm{s}\rightarrow \mathrm{p}}}{(N_{\mathrm{s}}-R_{\mathrm{s}})(N_{\mathrm{p}}-R_{\mathrm{p}})-T_{\mathrm{p}\rightarrow \mathrm{s}}T_{\mathrm{s}\rightarrow \mathrm{p}}}. \label{eq19.12}
5774: \ee
5775: This three-terminal resistance corresponds to a generalized {\it longitudinal\/}
5776: resistance if the magnetic field has the direction of Fig.\ \ref{fig92}. In the absence of
5777: backscattering in the 2DEG, one has $T_{\mathrm{s}\rightarrow \mathrm{p}}=0$, so $R_{3\mathrm{t}}$ vanishes, as it should
5778: for a longitudinal resistance in a strong magnetic field.
5779:
5780: \begin{figure}
5781: \centerline{\includegraphics[width=8cm]{figures/fig93}}
5782: \caption{
5783: Measurement of the anomalous Shubnikov-De Haas oscillations in the geometry of Fig.\ \ref{fig92}. The plotted longitudinal resistance is the voltage drop between contacts p and d divided by the current from s to d. At high magnetic fields the oscillations are increasingly suppressed as the point contact in the voltage probe is pinched off by increasing the negative gate voltage. The number of occupied spin-split Landau levels in the bulk is indicated at several of the Shubnikov-De Haas maxima. Taken from B. J. van Wees et al., Phys.\ Rev.\ B {\bf 39}, 8066 (1989).
5784: \label{fig93}
5785: }
5786: \end{figure}
5787:
5788: Shubnikov-De Haas oscillations in the longitudinal resistance arise when
5789: backscattering leads to $T_{\mathrm{s}\rightarrow \mathrm{p}}\neq 0$. The resistance reaches a maximum when
5790: the Fermi level lies in a bulk Landau level, corresponding to a maximum
5791: probability for backscattering (which requires scattering from one edge to the
5792: other across the bulk of the sample, as indicated by the dashed lines in Fig.\
5793: \ref{fig92}). From the preceding discussion of the anomalous quantum Hall effect, we
5794: know that the point contact voltage probe in a high magnetic field functions
5795: as a selective detector of edge channels with index $n$ less than some value
5796: determined by the barrier height in the point contact. If backscattering itself
5797: occurs selectively for the channel with the highest index $n=N$, and if the edge
5798: channels with $n\leq N-1$ do not scatter to that edge channel, then a
5799: suppression of the Shubnikov-De Haas oscillations is to be expected when
5800: $R_{3\mathrm{t}}$ is measured with a point contact containing a sufficiently high potential
5801: barrier. This was indeed observed experimentally,\cite{ref428} as shown in Fig.\ \ref{fig93}. The
5802: Shubnikov-De Haas maximum at $5.2\, \mathrm{T}$, for example, is found to disappear at
5803: gate voltages such that the point contact conductance is equal to, or smaller
5804: than $2e^{2}/h$, which means that the point contact only transmits two spin-split
5805: edge channels. The number of occupied spin-split Landau levels in the bulk at
5806: this magnetic field value is 3. This experiment thus demonstrates that the
5807: Shubnikov-De Haas oscillations result from the highest-index edge channel
5808: only, presumably because that edge channel can penetrate into the bulk via
5809: states in the bulk Landau level with the same index that coexist at the Fermi
5810: level (cf.\ Section \ref{sec19c}). Moreover, it is found that this edge channel does not
5811: scatter to the lower-index edge channels over the distance of 250 $\mu \mathrm{m}$ from
5812: probe $\mathrm{p}$ to drain $\mathrm{d}$, consistent with the experiment of Alphenaar et al.\cite{ref429}
5813:
5814: In Section \ref{sec19a} we discussed how an ``ideal'' contact at the 2DEG
5815: boundary {\it induces\/} a local equilibrium by equipartitioning the outgoing
5816: current equally among the edge channels. The anomalous Shubnikov-De
5817: Haas effect provides a direct way to study this contact-induced equilibration
5818: by means of a second point contact between the point contact voltage probe
5819: $\mathrm{p}$ and the current drain $\mathrm{d}$ in Fig.\ \ref{fig92}. This experiment was also carried out by
5820: van Wees et al., as described in Ref.\ \onlinecite{ref308}. Once again, use was made of the
5821: double-split-gate point contact device (Fig.\ \ref{fig5}b), in this case with a $1.5$-$\mu \mathrm{m}$
5822: separation between point contact $\mathrm{p}$ and the second point contact. It is found
5823: that the Shubnikov-De Haas oscillations in $R_{3\mathrm{t}}$ are suppressed only if the
5824: second point contact has a conductance of $(2e^{2}/h)(N_{\mathrm{wide}}-1)$ or smaller. At
5825: larger conductances the oscillations in $R_{3\mathrm{t}}$ return, because this point contact
5826: can now couple to the highest-index edge channel and distribute the
5827: backscattered electrons over the lower-index edge channels. The point
5828: contact positioned between contacts $\mathrm{p}$ and $\mathrm{d}$ thus functions as a controllable
5829: ``edge channel mixer.''
5830:
5831: The conclusions of the previous paragraph have interesting implications
5832: for the Shubnikov-De Haas oscillations in the strong-field regime even if
5833: measured with contacts that do {\it not\/} selectively detect certain edge channels
5834: only.\cite{ref307} Consider again the geometry of Fig.\ \ref{fig92}, in the low-gate voltage limit
5835: where the point contact voltage probe transmits all edge channels with unit
5836: probability. (This is the case of an ``ideal'' contact; cf.\ Section \ref{sec18b}.) To simplify
5837: expression (\ref{eq19.12}) for the three-terminal longitudinal resistance $R_{3\mathrm{t}}$, we use
5838: the fact that the transmission and reflection probabilities $T_{\mathrm{s}\rightarrow \mathrm{p}}$, $R_{\mathrm{s}}$, and $R_{\mathrm{p}}$
5839: refer to the highest-index edge channel only (with index $n=N$), under the
5840: assumptions of selective backscattering and absence of scattering to lower-index edge channels discussed earlier. As a consequence, $T_{\mathrm{s}\rightarrow \mathrm{p}}, R_{\mathrm{s}}$, and $R_{\mathrm{p}}$ are
5841: each at most equal to 1; thus, up to corrections smaller by a factor $N^{-1}$, we
5842: may put these terms equal to zero in the denominator on the right-hand side
5843: of Eq.\ (\ref{eq19.12}). In the numerator, the transmission probability $T_{\mathrm{s}\rightarrow \mathrm{p}}$ may be
5844: replaced by the backscattering probability $t_{\mathrm{bs}}\leq 1$, which is the probability
5845: that the highest-index edge channel injected by the source contact reaches the
5846: point contact probe following scattering across the wide 2DEG (dashed lines
5847: in Fig.\ \ref{fig92}). With these simplifications Eq.\ (\ref{eq19.12}) takes the form (assuming
5848: spin degeneracy)
5849: \be
5850: R_{3\mathrm{t}}= \frac{h}{2e^{2}}\frac{t_{\mathrm{bs}}}{N^{2}}\times(1+ {\rm order}\,N^{-1}). \label{eq19.13}
5851: \ee
5852: Only if $t_{\mathrm{bs}}\ll 1$ may the backscattering probability be expected to scale
5853: linearly with the separation of the two contacts $\mathrm{p}$ and $\mathrm{d}$ (between which the
5854: voltage drop is measured). If $t_{\mathrm{bs}}$ is not small, then the upper limit $t_{\mathrm{bs}^{<}}1$ leads
5855: to the prediction of a {\it maximum\/} possible amplitude\cite{ref307}
5856: \be
5857: R_{\max}=\frac{h}{2e^{2}}\frac{1}{N^{2}}\times(1+{\rm order}\,N^{-1}) \label{eq19.14}
5858: \ee
5859: of the Shubnikov-De Haas resistance oscillations in a given large magnetic
5860: field, independently of the length of the segment over which the voltage drop
5861: is measured, provided equilibration does not occur on this segment. Equilibration might result, for example, from the presence of additional contacts
5862: between the voltage probes, as discussed before. One easily verifies that the
5863: high-field Shubnikov-De Haas oscillations in Fig.\ \ref{fig93} at $V_{\mathrm{g}}=-0.6\,\mathrm{V}$ (when
5864: the point contact is just defined, so that the potential barrier is small) lie well
5865: below the upper limit (\ref{eq19.14}). For example, the peak around $2\, \mathrm{T}$ corresponds
5866: to the case of four occupied spin-degenerate Landau levels, so the theoretical
5867: upper limit is $(h/2e^{2}) \times\frac{1}{16}\approx 800\,\Omega$, well above the observed peak value of
5868: about $350\,\Omega$. The prediction of a maximum longitudinal resistance implies
5869: that the linear scaling of the amplitude of the Shubnikov-De Haas oscillations with the distance between voltage probes found in the weak-field
5870: regime, and expected on the basis of a description in terms of a local
5871: resistivity tensor,\cite{ref20} breaks down in strong magnetic fields. Anomalous
5872: scaling of the Shubnikov-De Haas effect has been observed experimentally\cite{ref457,ref460,ref466} and has recently also been interpreted\cite{ref430} in terms of a
5873: nonequilibrium between the edge channels. A quantitative experimental and
5874: theoretical investigation of these issues has now been carried out by McEuen
5875: et al.\cite{ref477}
5876:
5877: Selective backscattering and the absence of local equilibrium have
5878: consequences as well for the two-terminal resistance in strong magnetic
5879: fields.\cite{ref307} In weak fields one usually observes in two-terminal measurements a
5880: superposition of the Shubnikov-De Haas longitudinal resistance oscillations
5881: and the quantized Hall resistance. This superposition shows up as a
5882: characteristic ``overshoot'' of the two-terminal resistance as a function of the
5883: magnetic field as it increases from one quantized Hall plateau to the next (the
5884: plateaux coincide with minima of the Shubnikov-De Haas oscillations). In
5885: the strong-field regime (in the absence of equilibration between source and
5886: drain contacts), no such superposition is to be expected. Instead, the two-terminal resistance would increase monotonically from $(h/2e^{2})N^{-1}$ to
5887: $(h/2e^{2})(N-1)^{-1}$ as the transmission probability from source to drain
5888: decreases from $N$ to $N-1$. We are not aware of an experimental test of this
5889: prediction.
5890:
5891: The foregoing analysis assumes that the length $L$ of the conductor is much
5892: greater than its width $W$, so edge channels are the only states at the Fermi
5893: level that extend from source to drain. If $L\ll W$, additional extended states
5894: may appear in the bulk of the 2DEG, whenever the Fermi level lies in a bulk
5895: Landau level. An experiment by Fang et al.\ in this short-channel regime, to
5896: which our analysis does not apply, is discussed by B\"{u}ttiker.\cite{ref386}
5897:
5898: \subsection{\label{sec20} Fractional quantum Hall effect}
5899:
5900: Microscopically, quantization of the Hall conductance $G_{\mathrm{H}}$ in fractional
5901: multiples of $e^{2}/h$ is entirely different from quantization in integer multiples.
5902: While the {\it integer\/} quantum Hall effect\cite{ref8} can be explained satisfactorily in
5903: terms of the states of noninteracting electrons in a magnetic field (see Section
5904: \ref{sec18}), the {\it fractional\/} quantum Hall effect\cite{ref478} exists only because of electron-electron interactions.\cite{ref479} Phenomenologically, however, the two effects are
5905: quite similar. Several experiments on edge channel transport in the integer
5906: QHE,\cite{ref339,ref340,ref426} reviewed in Section \ref{sec19} have been repeated\cite{ref480,ref481} for the
5907: fractional QHE with a similar outcome. The interpretation of Section \ref{sec19} in
5908: terms of selective population and detection of edge channels cannot be
5909: applied in that form to the fractional QHE. Edge channels in the integer
5910: QHE are defined in one-to-one correspondence to bulk Landau levels
5911: (Section \ref{sec18b}). The fractional QHE requires a generalization of the concept of
5912: edge channels that allows for independent current channels within the same
5913: Landau level. Two recent papers have addressed this problem\cite{ref482,ref483} and
5914: have obtained different answers. The present status of theory and experiment
5915: on transport in ``fractional'' edge channels is reviewed in Section \ref{sec20b},
5916: preceded by a brief introduction to the fractional QHE.
5917:
5918: \subsubsection{\label{sec20a} Introduction}
5919:
5920: Excellent high-level introductions to the fractional QHE in an unbounded
5921: 2DEG can be found in Refs.\ \onlinecite{ref97} and \onlinecite{ref484}. The following is an oversimplification of Laughlin's theory\cite{ref479} of the effect and is only intended to introduce the
5922: reader to some of the concepts that play a role in edge channel transport in
5923: the fractional QHE.
5924:
5925: It is instructive to first consider the motion of two interacting electrons in a
5926: strong magnetic field.\cite{ref485} The dynamics of the relative coordinate $\mathbf{r}$ decouples
5927: from that of the center of mass. Semiclassically, $\mathbf{r}$ moves along equipotentials
5928: of the Coulomb potential $e^{2}/\epsilon r$ (this is the guiding center drift discussed in
5929: Section \ref{sec18b}). The relative coordinate thus executes a circular motion around
5930: the origin, corresponding to the two electrons orbiting around their center of
5931: mass. The phase shift acquired on one complete revolution,
5932: \be
5933: \Delta\phi=\frac{e}{\hbar}\oint d{\bf l}\cdot {\bf A} = \frac{e}{\hbar}B\pi r^{2}, \label{eq20.1}
5934: \ee
5935: should be an integer multiple of $2\pi$ so that
5936: \be
5937: r=l_{\mathrm{m}}\sqrt{2q},\;\;q=1,2, \ldots. \label{eq20.2}
5938: \ee
5939: The interparticle separation in units of the magnetic length $l_{\mathrm{m}}\equiv(\hbar/eB)^{1/2}$ is
5940: quantized. In the field regime where the fractional QHE is observed, only one
5941: spin-split Landau level is occupied in general. If the electrons have the same
5942: spin, the wave function should change sign when two coordinates are
5943: interchanged. In the case considered here of two electrons, an interchange of
5944: the coordinates is equivalent to $\mathbf{r}\rightarrow-\mathbf{r}$. A change of sign is then obtained if
5945: the phase shift for one half revolution is an odd multiple of $\pi$ (i.e., for $\Delta\phi$ an
5946: odd multiple of $2\pi$). The Pauli principle thus restricts the integer $q$ in Eq.\
5947: (\ref{eq20.2}) to {\it odd\/} values.
5948:
5949: The interparticle separation of a system of more than two electrons is not
5950: quantized. Still, one might surmise that the energy at densities $n_{\mathrm{s}}\approx 1/\pi \bar{r}^{2}$
5951: corresponding to an average separation $\bar{r}$ in accord with Eq.\ (\ref{eq20.2}) would be
5952: particularly low. This occurs when the Landau level filling factor $\nu\equiv hn_{\mathrm{s}}/eB$
5953: equals $\nu\approx 1/q$. Theoretical work by Laughlin, Haldane, and Halperin\cite{ref479,ref486,ref487} shows that the energy density $u(\nu)$ of a uniform 2DEG in a
5954: strong magnetic field has downward {\it cusps\/} at these values of $\nu$ as well as at
5955: other fractions, given generally by
5956: \be
5957: \nu=p/q, \label{eq20.3}
5958: \ee
5959: with $p$ and $q$ mutually prime integers and $q$ odd. The cusp in $u$ at {\it integer\/} $\nu$ is a
5960: consequence solely of Landau level quantization, according to
5961: \be
5962: du/dn_{\mathrm{s}}=( \mathrm{Int}[\nu]+{\textstyle\frac{1}{2}})\hbar\omega_{\mathrm{c}}. \label{eq20.4}
5963: \ee
5964: Because of the cusp in $u$, the chemical potential $du/dn_{\mathrm{s}}$ has a discontinuity
5965: $\Delta\mu=\hbar\omega_{\mathrm{c}}$ at integer $\nu$. At these values of the filling factor an infinitesimal
5966: increase in electron density costs a finite amount of energy, so the electron gas
5967: can be said to be {\it incompressible}. The cusp in $u$ at {\it fractional\/} $\nu$ exists because of
5968: the Coulomb interaction. The discontinuity $\Delta\mu$ is now approximately
5969: $\Delta\mu\approx e^{2}/\epsilon l_{\mathrm{m}}\propto \sqrt{B}$, which at a typical field of $6\, \mathrm{T}$ in GaAs is $10\, \mathrm{meV}$, of the
5970: same magnitude as the Landau level separation $\hbar\omega_{\mathrm{c}}\propto B$.
5971:
5972: The incompressibility of the 2DEG at $\nu=p/q$ implies that a nonzero
5973: minimal energy is required to add charge to the system. An important
5974: consequence of Laughlin's theory is that charge can be added only in the
5975: form of quasiparticle excitations of {\it fractional\/} charge $e^{*}=e/q$. The discontinuity $\Delta\mu$ in the chemical potential equals the energy that it costs to
5976: create $p$ pairs of oppositely charged quasiparticles (widely separated from
5977: each other), $\Delta\mu=p\times 2\Delta$ with $\Delta$ the quasiparticle creation energy.
5978:
5979: The fractional QHE in a disordered macroscopic sample occurs because
5980: the quasiparticles are localized by potential fluctuations in the bulk of the
5981: 2DEG. A variation of the filling factor $\nu=p/q+\delta \nu$ in an interval around the
5982: fractional value changes the density of localized quasiparticles without
5983: changing the Hall conductance, which retains the value $G_{\mathrm{H}}=(p/q)e^{2}/h$. The
5984: precision of the QHE has been explained by Laughlin\cite{ref488} in terms of the
5985: quantization of the quasiparticle charge $e^{*}$, which is argued to imply
5986: quantization of $G_{\mathrm{H}}$ at integer multiples of $ee^{*}/h$.
5987:
5988: \subsubsection{\label{sec20b} Fractional edge channels}
5989:
5990: In a small sample the fractional QHE can occur in the absence of disorder
5991: and can show deviations from precise quantization. Moreover, in special
5992: geometries\cite{ref481} $G_{\mathrm{H}}$ can take on quantized values that are not simply related to
5993: $e^{*}$. These observations cannot be easily understood within the conventional
5994: description of the fractional QHE, as outlined in the previous subsection. An
5995: approach along the lines of the edge channel formulation of the integer QHE
5996: (Sections \ref{sec18} and \ref{sec19}) seems more promising. In Ref.\ \onlinecite{ref482} the concept of an edge
5997: channel was generalized to the fractional QHE, and a generalized Landauer
5998: formula relating the conductance to the transmission probabilities of the edge
5999: channels was derived. We review this theory and the application to experiments. A different edge channel theory by MacDonald\cite{ref483} is discussed toward
6000: the end of this subsection.
6001:
6002: The edge channels for the conductance in the linear transport regime are
6003: defined in terms of properties of the equilibrium state of the system. If the
6004: electrostatic potential energy $V(x, y)$ varies slowly in the 2DEG, then the
6005: equilibrium density distribution $n(x, y)$ follows by requiring that the local
6006: electrochemical potential $V(\mathbf{r})+du/dn$ has the same value $\mu$ at each point $\mathbf{r}$ in
6007: the 2DEG. Here $du/dn$ is the chemical potential of the {\it uniform\/} 2DEG with
6008: density $n(\mathbf{r})$. As discussed in Section \ref{sec20a}, the internal energy density $u(n)$ of a
6009: uniform interacting 2DEG in a strong magnetic field has downward cusps at
6010: densities $n=\nu_{p}Be/h$ corresponding to certain fractional filling factors $\nu_{p}$. As a
6011: result, the chemical potential $du/dn$ has a discontinuity (an energy gap) at
6012: $\nu=\nu_{p}$, with $du_{p}^{+}/dn$ and $du_{p}^{-}/dn$ the two limiting values as $\nu\rightarrow \nu_{p}$. As noted
6013: by Halperin,\cite{ref489} when $\mu-V$ lies in the energy gap the filling factor is pinned
6014: at the value $\nu_{p}$. The equilibrium electron density is thus given by\cite{ref489}
6015: \be
6016: n=\left\{\begin{array}{l}
6017: \nu_{p}Be/h,\;\;{\rm if}\;\;du_{p}^{-}/dn<\mu-V<du_{p}^{+}/dn,\\
6018: du/dn+V(\mathbf{r})=\mu,\;\;{\rm otherwise}.
6019: \end{array}\right.\label{eq20.5}
6020: \ee
6021: Note that $V(\mathbf{r})$ itself depends on $n(\mathbf{r})$ and thus has to be determined self-consistently from Eq.\ (\ref{eq20.5}), taking the electrostatic screening in the 2DEG
6022: into account. We do not need to solve explicitly for $n(\mathbf{r})$, but we can identify
6023: the edge channels from the following general considerations.\cite{ref482}
6024:
6025: \begin{figure}
6026: \centerline{\includegraphics[width=8cm]{figures/fig94}}
6027: \caption{
6028: Schematic drawing of the variation in filling factor $\nu$, electrostatic potential $V$, and chemical potential $du/dn$, at a smooth boundary in a 2DEG. The dashed line in the bottom panel denotes the constant electrochemical potential $\mu=V+du/dn$. The dotted intervals indicate a discontinuity (energy gap) in $du/dn$ and correspond in the top panel to regions of constant fractional filling factor $\nu_{p}$ that spatially separate the edge channels. The width of the edge channel regions shrinks to zero in the integer QHE, since the compressibility $\chi$ of these regions is infinitely large in that case. Taken from C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 64}, 216 (1990).
6029: \label{fig94}
6030: }
6031: \end{figure}
6032:
6033: At the edge of the 2DEG, the electron density decreases from its bulk value
6034: to zero. Eq.\ (\ref{eq20.5}) implies that this decrease is stepwise, as illustrated in Fig.\ \ref{fig94}. The requirement on the smoothness of $V$ for the appearance of a well-defined region at the edge in which $\nu$ is pinned at the fractional value $\nu_{p}$ is that
6035: the change in $V$ within the magnetic length $l_{\mathrm{m}}$ is small compared with the
6036: energy gap $du_{p}^{+}/dn-du_{p}^{-}/dn$. This ensures that the width of this region is
6037: large compared with $l_{\mathrm{m}}$, which is a necessary (and presumably sufficient)
6038: condition for the formation of the incompressible state. Depending on the
6039: smoothness of $V$, one thus obtains a series of steps at $\nu=\nu_{p}$ ($p=1,2, \ldots, P$) as
6040: one moves from the edge toward the bulk. The series terminates in the filling
6041: factor $\nu_{P}=\nu_{\mathrm{bulk}}$ of the bulk, assuming that in the bulk the chemical potential
6042: $\mu-V$ lies in an energy gap. The regions of constant $\nu$ at the edge form bands
6043: extending along the wire. These {\it incompressible bands\/} [in which the compressibility $\chi\equiv(n^{2}d^{2}u/dn^{2})^{-1}=0]$ alternate with bands in which $\mu-V$ does not lie
6044: in an energy gap. The latter compressible bands (in which $\chi>0$) may be
6045: identified as the {\it edge channels\/} of the transport problem, as will be discussed
6046: later. To resolve a misunderstanding,\cite{ref490} we note that the particular potential
6047: and density profile illustrated in Fig.\ \ref{fig94} (in which the edge channels have a
6048: nonzero width) assumes that the compressibility of the edge channels is not
6049: infinitely large, but the subsequent analysis is independent of this assumption
6050: (requiring only that the edge channels are flanked by bands of zero
6051: compressibility). Indeed, the analysis is applicable also to the integer QHE,
6052: where the edge channels have an infinitely large compressibility and hence an
6053: infinitesimally small width (limited only by the magnetic length).
6054:
6055: \begin{figure}
6056: \centerline{\includegraphics[width=8cm]{figures/fig95}}
6057: \caption{
6058: Schematic drawing of the incompressible bands (hatched) of fractional filling factor $\nu_{p}$, alternating with the edge channels (arrows indicate the direction of electron motion in each channel). (a) A uniform conductor. (b) A conductor containing a barrier of reduced filling factor. Taken from C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 64}, 216 (1990).
6059: \label{fig95}
6060: }
6061: \end{figure}
6062:
6063: The conductance is calculated by bringing one end of the conductor in
6064: contact with a reservoir at a slightly higher electrochemical potential $\mu+\Delta\mu$
6065: without changing $V$ (as in the derivation of the usual Landauer formula; cf.\
6066: Section \ref{sec12b}). The resulting change $\Delta n$ in electron density is
6067: \be
6068: \Delta n=\left(\frac{\delta n}{\delta\mu}\right)_{V}\Delta\mu=-\left(\frac{\delta n}{\delta V}\right)_{\mu}\Delta\mu, \label{eq20.6}
6069: \ee
6070: where $\delta$ denotes a functional derivative. In the second equality in Eq.\ (\ref{eq20.6}),
6071: we used the fact that $n$ is a functional of $\mu-V$, by virtue of Eq.\ (\ref{eq20.5}). In a
6072: strong magnetic field, this excess density moves along equipotentials with the
6073: guiding-center-drift velocity $E/B$ ($\mathbf{E}\equiv\partial V/e\partial \mathbf{r}$ being the electric field). The
6074: component $v_{\mathrm{drift}}$ of the drift velocity in the $y$-direction (along the conductor) is
6075: \be
6076: v_{\mathrm{drift}}= \hat{\mathbf{y}}\cdot\left(\mathbf{E}\times\frac{\mathbf{B}}{B^{2}}\right)=-\frac{1}{eB}\frac{\partial V}{\partial x}. \label{eq20.7}
6077: \ee
6078: The current density $j=-e\Delta nv_{\mathrm{drift}}$ becomes simply
6079: \be
6080: j=- \frac{e}{h}\Delta\mu\frac{\partial v}{\partial x}. \label{eq20.8}
6081: \ee
6082: It follows from Eq.\ (\ref{eq20.8}) that the incompressible bands of constant $\nu=\nu_{p}$
6083: do not contribute to $j$. The reservoir injects the current into the compressible
6084: bands at one edge of the conductor only (for which the sign of $\partial \nu/\partial x$ is such
6085: that $j$ moves away from the reservoir). The edge channel with index $p=1,2,
6086: \ldots, P$ is defined as that compressible band that is flanked by incompressible
6087: bands at filling factors $\nu_{p}$ and $\nu_{p-1}$. The outermost band from the center of the
6088: conductor, which is the $p=1$ edge channel, is included by defining formally
6089: $\nu_{0}\equiv 0$. The arrangement of alternating edge channels and compressible
6090: bands is illustrated in Fig.\ \ref{fig95}a. Note that different edges may have a different
6091: series of edge channels at the same magnetic field value, depending on the
6092: smoothness of the potential $V$ at the edge (which, as discussed before,
6093: determines the incompressible bands that exist at the edge). This is in contrast
6094: to the situation in the integer QHE, where a one-to-one correspondence
6095: exists between edge channels and bulk Landau levels (Section \ref{sec18b}). In the
6096: fractional QHE an infinite hierarchy of energy gaps exists, in principle,
6097: corresponding to an infinite number of possible edge channels, of which only
6098: a small number (corresponding to the largest energy gaps) will be realized in
6099: practice.
6100:
6101: The current $I_{p}=(e/h)\Delta\mu(\nu_{p}-\nu_{p-1})$ injected into edge channel $p$ by the
6102: reservoir follows directly from Eq.\ (\ref{eq20.8}) on integration over $x$. The total
6103: current $I$ through the wire is $I= \sum_{p=1}^{P}I_{p}T_{p}$, if a fraction $T_{p}$ of the injected
6104: current $I_{p}$ is transmitted to the reservoir at the other end of the wire (the
6105: remainder returning via the opposite edge). For the conductance $G\equiv eI/\Delta\mu$,
6106: one thus obtains the generalized Landuer formula for a two-terminal
6107: conductor,\cite{ref482}
6108: \be
6109: G= \frac{e^{2}}{h}\sum_{p=1}^{P}T_{p}\Delta \nu_{p}, \label{eq20.9}
6110: \ee
6111: which differs from the usual Landauer formula by the presence of the
6112: fractional weight factors $\Delta \nu_{p}\equiv \nu_{p}-\nu_{p-1}$. In the integer QHE, $\Delta \nu_{p}=1$ for all
6113: $p$ so that the usual Landauer formula with unit weight factor is recovered.
6114:
6115: A multiterminal generalization of Eq.\ (\ref{eq20.9}) for a two-terminal conductor
6116: is easily constructed, following B\"{u}ttiker\cite{ref5} (cf.\ Section \ref{sec12b}):
6117: \begin{subequations}
6118: \label{eq20.10}
6119: \begin{eqnarray}
6120: I_{\alpha}&=& \frac{e}{h}\nu_{\alpha}\mu_{\alpha}-\frac{e}{h}\sum_{\beta}T_{\alpha\beta}\mu_{\beta}, \label{eq20.10a}\\
6121: T_{\alpha\beta}&=&\sum_{p=1}^{P_{\beta}}T_{p,\alpha\beta}\Delta \nu_{p}. \label{eq20.10b}
6122: \end{eqnarray}
6123: \end{subequations}
6124: Here $I_{\alpha}$ is the current in lead $\alpha$ connected to a reservoir at electrochemical
6125: potential $\mu_{\alpha}$ and fractional filling factor $\nu_{\alpha}$. Equation (\ref{eq20.10b}) defines the
6126: transmission probability $T_{\alpha\beta}$ from reservoir $\beta$ to reservoir $\alpha$ (or the reflection
6127: probability for $\alpha=\beta$) in terms of a sum over the generalized edge channels in
6128: lead $\beta$. The contribution from each edge channel $p=1,2, \ldots ,P_{\beta}$ contains the
6129: weight factor $\Delta \nu_{p}\equiv \nu_{p}-\nu_{p-1}$ and the fraction $T_{p,\alpha\beta}$ of the current injected by
6130: reservoir $\beta$ into the $p\mathrm{th}$ edge channel of lead $\beta$ that reaches reservoir $\alpha$. Apart
6131: from the fractional weight factors, the structure of Eq.\ (\ref{eq20.10}) is the same as
6132: that of the usual B\"{u}ttiker formula (\ref{eq12.12}).
6133:
6134: Applying the generalized Landauer formula (\ref{eq20.9}) to the ideal conductor
6135: in Fig.\ \ref{fig95}a, where $T_{p}=1$ for all $p$, one finds the quantized two-terminal
6136: conductance
6137: \be
6138: G= \frac{e^{2}}{h}\sum_{p=1}^{P}\Delta \nu_{p}=\frac{e^{2}}{h}\nu_{P}. \label{eq20.11}
6139: \ee
6140: The four-terminal Hall conductance $G_{\mathrm{H}}$ has the same value, because each
6141: edge is in local equilibrium. In the presence of disorder this edge channel
6142: formulation of the fractional QHE is generalized in an analogous way as in
6143: the integer QHE by including localized states in the bulk. In a smoothly
6144: varying disorder potential, these localized states take the form of circulating
6145: edge channels, as in Figs.\ \ref{fig78} and \ref{fig79}. In this way the filling factor of the bulk
6146: can locally deviate from $\nu_{P}$ without a change in the Hall conductance, leading
6147: to the formation of a plateau in the magnetic field dependence of $G_{\mathrm{H}}$. In a
6148: narrow channel, localized states are not required for a finite plateau width
6149: because the edge channels make it possible for the chemical potential to lie in
6150: an energy gap for a finite-magnetic-field interval. The Hall conductance then
6151: remains quantized at $\nu_{P}(e^{2}/h)$ as long as $\mu-V$ in the bulk lies between
6152: $du_{P}^{+}/dn$ and $du_{P}^{-}/dn$.
6153:
6154: We now turn to a discussion of experiments on the fractional QHE in
6155: semiconductor nanostructures. Timp et al.\cite{ref491} have measured the fractionally
6156: quantized four-terminal Hall conductance $G_{\mathrm{H}}$ in a narrow cross geometry
6157: (defined by two sets of split gates). The channel width $W\approx 90\,\mathrm{nm}$ is greater
6158: than, but comparable to, the correlation length $l_{\mathrm{m}}$ of the incompressible state
6159: in this experiment ($l_{\mathrm{m}}\approx 9\,\mathrm{nm}$ at $B=8\,\mathrm{T}$), so one may expect the fractional
6160: QHE to be modified by the lateral confinement.\cite{ref492} Timp et al.\ find, in
6161: addition to quantized plateaux near $\frac{1}{3}$, $\frac{2}{5}$, and $\frac{2}{3}\times e^{2}/h$, a plateau-like feature
6162: around $\frac{1}{2}\times e^{2}/h$. This even-denominator fraction is not observed as a Hall
6163: plateau in a bulk 2DEG.\cite{ref493} The plateaux in $G_{\mathrm{H}}$ correlate with dips in a four-terminal longitudinal resistance (the bend resistance defined in Section \ref{sec16}).
6164:
6165: \begin{figure}
6166: \centerline{\includegraphics[width=8cm]{figures/fig96}}
6167: \caption{
6168: Two-terminal conductance of a constriction containing a potential barrier, as a function of the voltage on the split gate defining the constriction, at a fixed magnetic field of 7 T. The conductance is quantized according to Eq.\ (\ref{eq20.12}). Taken from L. P. Kouwenhoven et al., unpublished.
6169: \label{fig96}
6170: }
6171: \end{figure}
6172:
6173: Consider now a conductor containing a potential barrier. The potential
6174: barrier corresponds to a region of reduced filling factor $\nu_{P_{\min}}\equiv \nu_{\min}$ separating
6175: two regions of filling factor $\nu_{P_{\mathrm{m}}}\equiv \nu_{\max}$. The arrangement of edge channels
6176: and incompressible bands is illustrated in Fig.\ \ref{fig95}b. We assume that the
6177: potential barrier is sufficiently smooth that scattering between the edge
6178: channels at opposite edges can be neglected. All transmission probabilities
6179: are then either 0 or 1: $T_{p}=1$ for $1\leq p\leq P_{\min}$, and $T_{p}=0$ for
6180: $P_{\min}<p\leq P_{\max}$. Equation (\ref{eq20.9}) then tells us that the two-terminal conductance is
6181: \be
6182: G=(e^{2}/h)\nu_{\min}. \label{eq20.12}
6183: \ee
6184: In Fig.\ \ref{fig96} we show experimental data by Kouwenhoven et al.\cite{ref481} of the
6185: fractionally quantized two-terminal conductance of a constriction containing
6186: a potential barrier. The constriction (or point contact) is defined by a split
6187: gate on top of a GaAs-AlGaAs heterostructure. The conductance in Fig.\ \ref{fig96}
6188: is shown for a fixed magnetic field of $7\, \mathrm{T}$ as a function of the gate voltage.
6189: Increasing the negative gate voltage increases the barrier height, thereby
6190: reducing $G$ below the Hall conductance corresponding to $\nu_{\max}=1$ in the wide
6191: 2DEG. The curve in Fig.\ \ref{fig96} shows plateaux corresponding to $\nu_{\min}=1$, $\frac{2}{3}$, and
6192: $\frac{1}{3}$ in Eq.\ (\ref{eq20.12}). The $\frac{2}{3}$ plateau is not exactly quantized, but is too low by a few
6193: percent. The constriction width on this plateau is estimated\cite{ref481} at $500\, \mathrm{nm}$,
6194: which is a factor of 50 larger than the magnetic length at $B=7$ T. It would
6195: seem that scattering between fractional edge channels at opposite edges
6196: (necessary to reduce the conductance below its quantized value) can only
6197: occur via states in the bulk for this large ratio of $W/l_{\mathrm{m}}$.
6198:
6199: \begin{figure}
6200: \centerline{\includegraphics[width=8cm]{figures/fig97}}
6201: \caption{
6202: Four-terminal resistances of a 2DEG channel containing a potential barrier, as a function of the gate voltage ($B = 0.114\,{\rm T}$, $T= 70\,{\rm mK}$). The current flows from contact 1 to contact 5 (see inset), the resistance curves are labeled by the contacts $i$ and $j$ between which the voltage is measured. (The curves for $i,j = 2,4$ and $8,6$ are identical.) The magnetic field points outward. This measurement corresponds to the case $\nu_{\rm max} = 1$ and $\nu_{\rm min}=\nu_{\rm b}$ varying from 1 at $V_{\rm g}\geq-10\,{\rm mV}$ to $2/3$ at $V_{\rm g}\approx-90\,{\rm mV}$ (arrow). The resistances $R_{\rm L} \equiv R_{2,4} = R_{8,6}$ and $R_{\rm D}^{+} \equiv R_{2,6}$ are quantized according to Eqs.\ (\ref{eq20.13}) and (\ref{eq20.14}), respectively. The resistances $R_{3,7}$ and $R_{2,8}$ are the Hall resistances in the gated and ungated regions, respectively. From Eq.\ (\ref{eq20.10}) one can also derive that $R_{8,7} = R_{3,4} = R_{\rm L}$ and $R_{2,3} = R_{7,6} = 0$ on the quantized plateaux, as observed experimentally. Taken from A. M. Chang and J. E. Cunningham, Surf.\ Sci.\ {\bf 229}, 216 (1990).
6203: \label{fig97}
6204: }
6205: \end{figure}
6206:
6207: A four-terminal measurement of the fractional QHE in a conductor
6208: containing a potential barrier can be analyzed by means of Eq.\ (\ref{eq20.10}),
6209: analogously to the case of the integer QHE discussed in Section \ref{sec19}. The four-terminal longitudinal resistance $R_{\mathrm{L}}$ (in the geometry of Fig.\ \ref{fig82}) is given by the
6210: analog of Eq.\ (\ref{eq19.3}),
6211: \be
6212: R_{\mathrm{L}}= \frac{h}{e^{2}}\left(\frac{1}{\nu_{\min}}-\frac{1}{\nu_{\max}}\right), \label{eq20.13}
6213: \ee
6214: provided that {\it either\/} the edge channels transmitted across the barrier have
6215: equilibrated with the extra edge channels available outside the barrier region
6216: {\it or\/} the voltage contacts are ideal; that is, they have unit transmission
6217: probability for all fractional edge channels. Similarly, the four-terminal
6218: diagonal resistances $R_{\mathrm{D}}^{\pm}$ defined in Fig.\ \ref{fig82} are given by [cf.\ Eq.\ (\ref{eq19.5})]
6219: \be
6220: R_{\mathrm{D}}^{+}= \frac{h}{e^{2}}\frac{1}{v_{\min}};\;\;R_{\mathrm{D}}^{-}= \frac{h}{e^{2}}\left(\frac{2}{v_{\max}}-\frac{1}{v_{\min}}\right). \label{eq20.14}
6221: \ee
6222: Chang and Cunningham\cite{ref480} have measured $R_{\mathrm{L}}$ and $R_{\mathrm{D}}$ in the fractional QHE,
6223: using a $1.5$-$\mu \mathrm{m}$-wide 2DEG channel with a gate across a segment of the
6224: channel (the gate length is also approximately 1.5 $\mu \mathrm{m}$). Ohmic contacts to the
6225: gated and ungated regions allowed $\nu_{\min}$ and $\nu_{\max}$ to be determined independently. Equations (\ref{eq20.13}) and (\ref{eq20.14}) were found to hold to within $0.5\%$
6226: accuracy. This is illustrated in Fig.\ \ref{fig97} for the case that $\nu_{\max}=1$ and $\nu_{\min}$
6227: varying from 1 to 2/3 on increasing the negative gate voltage (at a fixed
6228: magnetic field of $0.114\,\mathrm{T}$). Similar results were obtained\cite{ref480} for the case that
6229: $\nu_{\max}= \frac{2}{3}$ and $\nu_{\min}$ varies from $\frac{2}{3}$ to $\frac{1}{3}$.
6230:
6231: \begin{figure}
6232: \centerline{\includegraphics[width=8cm]{figures/fig98}}
6233: \caption{
6234: (a) Schematic of the experimental geometry of Kouwenhoven et al.\cite{ref481} The crossed squares are contacts to the 2DEG. One current lead and one voltage lead contain a barrier (shaded), of which the height can be adjusted by means of a gate (not drawn). The current $I$ flows between contacts 1 and 3; the voltage $V$ is measured between contacts 2 and 4. (b) Arrangement of incompressible bands (hatched) and edge channels near the two barriers. In the absence of scattering between the two fractional edge channels, one would measure a Hall conductance $G_{\rm H} \equiv I/V$ that is fractionally quantized at $\frac{2}{3}\times e^{2}/h$, although the bulk has unit filling factor. Taken from C. W. J. Beenakker, Phys.\ Rev.\ Lett.\ {\bf 64}, 216 (1990).
6235: \label{fig98}
6236: }
6237: \end{figure}
6238:
6239: Adiabatic transport in the fractional QHE can be studied by the selective
6240: population and detection of fractional edge channels, achieved by means of
6241: barriers in two closely separated current and voltage contacts (Fig.\ \ref{fig98}a). The
6242: analysis using Eq.\ (\ref{eq20.10}) is completely analogous to the analysis of the
6243: experiment in the integer QHE,\cite{ref426} discussed in Section \ref{sec19}. Figure \ref{fig98}b
6244: illustrates the arrangement of edge channels and incompressible bands for the
6245: case that the chemical potential lies in an energy gap for the bulk 2DEG (at
6246: $\nu=\nu_{\mathrm{bulk}})$, as well as for the two barriers (at $\nu_{\mathrm{I}}$ and $\nu_{\mathrm{V}}$ for the barrier in the
6247: current and voltage lead, respectively). Adiabatic transport is assumed over
6248: the barrier, as well as from barrier I to barrier V (for the magnetic field
6249: direction indicated in Fig.\ \ref{fig98}). Equation (\ref{eq20.10}) for this case reduces to
6250: \be
6251: I= \frac{e}{h}\nu_{\mathrm{I}}\mu_{\mathrm{I}},\;\;
6252: 0= \frac{e}{h}\nu_{\mathrm{V}}\mu_{\mathrm{V}}-\frac{e}{h}\min(\nu_{\mathrm{I}}, \nu_{\mathrm{V}})\mu_{\mathrm{I}}, \label{eq20.15}
6253: \ee
6254: so the Hall conductance $G_{\mathrm{H}}=eI/\mu_{\mathrm{V}}$ becomes
6255: \be
6256: G_{\mathrm{H}}= \frac{e^{2}}{h}\max(\nu_{\mathrm{I}}, \nu_{\mathrm{V}}) \leq\frac{e^{2}}{h}\nu_{\mathrm{bulk}}.
6257: \label{eq20.16}
6258: \ee
6259: The quantized Hall plateaux are determined by the fractional filling factors of
6260: the current and voltage leads, not of the bulk 2DEG. Kouwenhoven et al.\cite{ref481}
6261: have demonstrated the selective population and detection of fractional edge
6262: channels in a device with a $2$-$\mu \mathrm{m}$ separation of the gates in the current and
6263: voltage leads. The gates extended over a length of 40 $\mu \mathrm{m}$ along the 2DEG
6264: boundary. In Fig.\ \ref{fig99} we reproduce one of the experimental traces of
6265: Kouwenhoven et al.\ The Hall conductance is shown for a fixed magnetic field
6266: of $7.8\, \mathrm{T}$ as a function of the gate voltage (all gates being at the same voltage).
6267: As the barrier heights in the two leads are increased, the Hall conductance
6268: decreases from the bulk value $1\times e^{2}/h$ to the value $\frac{2}{3}\times e^{2}/h$ determined by the
6269: leads, in accord with Eq.\ (\ref{eq20.16}). A more general formula for $G_{\mathrm{H}}$ valid also in
6270: between the quantized plateaux is shown in Ref.\ \onlinecite{ref481} to be in quantitative
6271: agreement with the experiment.
6272:
6273: \begin{figure}
6274: \centerline{\includegraphics[width=8cm]{figures/fig99}}
6275: \caption{
6276: Anomalously quantized Hall conductance in the geometry of Fig.\ \ref{fig98}, in accord with Eq.\ (\ref{eq20.16}) ($\nu_{\rm bulk} = 1$, $\nu_{\rm I} = \nu_{\rm V}$ decreases from 1 to $2/3$ as the negative gate voltage is increased). The temperature is 20 mK. The rapidly rising part (dotted) is an artifact due to barrier pinch-off. Taken from L. P. Kouwenhoven et al., Phys.\ Rev.\ Lett.\ {\bf 64}, 685 (1990).
6277: \label{fig99}
6278: }
6279: \end{figure}
6280:
6281: MacDonald has, independent of Ref.\ \onlinecite{ref482}, proposed a different generalized
6282: Landauer formula for the fractional QHE.\cite{ref483} The difference with Eq.\ (\ref{eq20.9}) is
6283: that the weight factors in MacDonald's formula can take on both positive {\it and negative\/} values (corresponding to electron and hole channels). In the case of
6284: local equilibrium at the edge, the sum of weight factors is such that the two
6285: formulations give identical results. The results differ in the absence of local
6286: equilibrium if fractional edge channels are selectively populated and detected.
6287: For example, MacDonald predicts a {\it negative\/} longitudinal resistance in a
6288: conductor at filling factor $\nu= \frac{2}{3}$ containing a segment at $\nu=1$. Another
6289: implication of Ref.\ \onlinecite{ref483} is that the two-terminal conductance $G$ of a conductor
6290: at $\nu_{\max}=1$ containing a potential barrier at filling factor $\nu_{\min}$ is reduced to $\frac{1}{3}\times e^{2}/h$ if $\nu_{\min}=\frac{1}{3}$ [in accord with Eq.\ (\ref{eq20.12})], but remains at $1\times e^{2}/h$ if
6291: $\nu_{\min}=2/3$. That this is not observed experimentally (cf.\ Fig.\ \ref{fig96}) could be due
6292: to interedge channel scattering, as argued by MacDonald. The experiment by
6293: Kouwenhoven et al.\cite{ref481} (Fig.\ \ref{fig99}), however, is apparently in the adiabatic
6294: regime, and was interpreted in Fig.\ \ref{fig98} in terms of an edge channel of weight $\frac{1}{3}$
6295: at the edge of a conductor at $\nu=1$. In MacDonald's formulation, the
6296: conductor at $\nu=1$ has only a singe edge channel of weight 1. This would
6297: need to be reconciled with the experimental observation of quantization of
6298: the Hall conductance at $\frac{2}{3}\times e^{2}/h$.
6299:
6300: We conclude this section by briefly addressing the question: What charge
6301: does the resistance measure? The fractional quantization of the conductance
6302: in the experiments discussed is understood as a consequence of the fractional
6303: weight factors in the generalized Landauer formula (\ref{eq20.9}). These weight
6304: factors $\Delta \nu_{p}=\nu_{p}-\nu_{p-1}$ are {\it not\/} in general equal to $e^{*}/e$, with $e^{*}$ the fractional
6305: charge of the quasiparticle excitations of Laughlin's incompressible state (cf.\
6306: Section \ref{sec20a}). The reason for the absence of a one-to-one correspondence
6307: between $\Delta \nu_{p}$ and $e^{*}$ is that the edge channels themselves are not incompressible.\cite{ref482} The transmission probabilities in Eq.\ (\ref{eq20.9}) refer to charged ``gapless''
6308: excitations of the edge channels, which are not identical to the charge $e^{*}$
6309: excitations above the energy gap in the incompressible bands (the latter
6310: charge might be obtained from thermal activation measurements; cf.\ Ref.\
6311: \onlinecite{ref494}). It is an interesting and (to date) unsolved problem to determine the
6312: charge of the edge channel excitations. Kivelson and Pokrovsky\cite{ref495} have
6313: suggested performing tunneling experiments in the fractional QHE regime
6314: for such a purpose, by using the charge dependence of the magnetic length
6315: $(\hbar/eB)^{1/2}$ (which determines the penetration of the wave function in a tunnel
6316: barrier and, hence, the transmission probability through the barrier). Alternatively, one could use the $h/e$ periodicity of the Aharanov-Bohm magnetoresistance oscillations as a measure of the edge channel charge. Simmons
6317: et al.\cite{ref496} find that the characteristic field scale of quasiperiodic resistance
6318: fluctuations in a $2$-$\mu \mathrm{m}$-wide Hall bar increases from $0.016\,{\rm T}\pm 30\%$ near
6319: $\nu=1,2,3,4$ to $0.05\, \mathrm{T}\pm 30\%$ near $\nu= \frac{1}{3}$. This is suggestive of a reduction in
6320: charge from $e$ to $e/3$, but not conclusive since the area for the Aharonov-
6321: Bohm effect is not well defined in a Hall bar (cf.\ Section \ref{sec21}).
6322:
6323: \subsection{\label{sec21} Aharonov-Bohm effect in strong magnetic fields}
6324:
6325: As mentioned briefly in Section \ref{sec8}, the Aharonov-Bohm oscillations in the
6326: magnetoresistance of a ring are gradually suppressed in strong magnetic
6327: fields. This suppression provides additional support for edge channel trans-
6328: port in the quantum Hall effect regime (Section \ref{sec21a}). Entirely new mechanisms for the Aharonov-Bohm effect become operative in strong magnetic
6329: fields. These mechanisms, resonant tunneling and resonant reflection of edge
6330: channels, do not require a ring geometry. Theory and experiments on
6331: Aharonov-Bohm oscillations in singly connected geometries are the subject
6332: of Section \ref{sec21b}.
6333:
6334: \subsubsection{\label{sec21a} Suppression of the Aharonov-Bohm effect in a ring}
6335:
6336: \begin{figure}
6337: \centerline{\includegraphics[width=6cm]{figures/fig100}}
6338: \caption{
6339: Illustration of a localized edge channel circulating along the inner perimeter of a ring, and of extended edge channels on the
6340: leads and on the outer perimeter. No Aharonov-Bohm magnetoresistance oscillations can occur in the absence of scattering between these two types of edge channels.
6341: \label{fig100}
6342: }
6343: \end{figure}
6344:
6345: In Section \ref{sec8} we have seen how the quantum interference of clockwise and
6346: counterclockwise trajectories in a ring in the diffusive transport regime leads
6347: to magnetoresistance oscillations with two different periodicities: the fundamental Aharonov-Bohm effect with $\Delta B=(h/e)S^{-1}$ periodicity, and the
6348: harmonic with $\Delta B=(h/2e)S^{-1}$ periodicity, where $S$ is the area of the ring. In
6349: arrays of rings only the $h/2e$ effect is observable, since the $h/e$ effect has a
6350: sample specific phase and is averaged to zero. In experiments by Timp et al.\cite{ref69}
6351: and by Ford et al.\cite{ref74} on single rings in the 2DEG of high-mobility GaAs-AlGaAs heterostructures, the $h/e$ effect was found predominantly. The
6352: amplitude of these oscillations is strongly reduced\cite{ref69,ref74,ref195,ref497} by a large
6353: magnetic field (cf.\ the magnetoresistance traces shown in Fig.\ \ref{fig26}). This
6354: suppression was found to occur for fields such that $2l_{\mathrm{cycl}}<W$, where $W$ is the
6355: width of the arms of the ring. The reason is that in strong magnetic fields the
6356: states at the Fermi level that can propagate through the ring are edge states at
6357: the outer perimeter. These states do not complete a revolution around the
6358: ring (see Fig.\ \ref{fig100}). Scattering between opposite edges is required to complete
6359: a revolution, but such backscattering would also lead to a nonzero longitudinal resistance. This argument\cite{ref112,ref498} explains the absence of Aharonov-Bohm oscillations on the quantized Hall plateaux, where the longitudinal
6360: resistance is zero. Magnetoresistance oscillations return between the plateaux
6361: in the Hall resistance, but at a larger value of $\Delta B$ than in weak fields. Timp et
6362: al.\cite{ref497} have argued that the Aharonov-Bohm oscillations in a ring in strong
6363: magnetic fields are associated with scattering from the outer edge to edge
6364: states circulating along the inner perimeter of the ring. The smaller area
6365: enclosed by the inner perimeter explains the increase in $\Delta B$. This interpretation is supported by numerical calculations.\cite{ref497}
6366:
6367: \subsubsection{\label{sec21b} Aharonov-Bohm effect in singly connected geometries}
6368:
6369: \begin{figure}
6370: \centerline{\includegraphics[width=8cm]{figures/fig101}}
6371: \caption{
6372: Two-terminal magnetoresistance of a point contact for a series of gate voltages at $T= 50\,{\rm mK}$, showing oscillations that are periodic in $B$ between the quantum Hall plateaux. The second, third, and fourth curves from the bottom have offsets of, respectively, 5, 10, and 15 ${\rm k}\Omega$. The rapid oscillations below 1 T are Shubnikov-De Haas oscillations periodic in $1/B$, originating from the wide 2DEG regions. The sharp peak around $B = 0\,{\rm T}$ originates from the ohmic contacts. Taken from P. H. M. van Loosdrecht et al., Phys.\ Rev.\ B {\bf 38}, 10162 (1988).
6373: \label{fig101}
6374: }
6375: \end{figure}
6376:
6377: \begin{figure}
6378: \centerline{\includegraphics[width=8cm]{figures/fig102}}
6379: \caption{
6380: Curves a and b are close-ups of the curve for $V_{\rm g} = -1.7\,{\rm V}$ in Fig.\ \ref{fig101}. Curve c is a separate measurement on the same device (note the different field scale due to a change in electron density in the constriction). Taken from P. H. M. van Loosdrecht et al., Phys.\ Rev.\ B {\bf 38}, 10162 (1988).
6381: \label{fig102}
6382: }
6383: \end{figure}
6384:
6385: {\bf (a) Point contact.} Aharonov-Bohm oscillations in the magnetoresistance of
6386: a quantum point contact were discovered by van Loosdrecht et al.\cite{ref292} The
6387: magnetic field dependence of the two-terminal resistance is shown in Fig.\ \ref{fig101},
6388: for various gate voltages. The periodic oscillations occur predominantly
6389: between quantum Hall plateaux, in a limited range of gate voltages, and only
6390: at low temperatures (in Fig.\ \ref{fig101}, $T=50\,\mathrm{mK}$; the effect has disappeared at
6391: $1\, \mathrm{K}$). The fine structure is very well reproducible if the sample is kept in the
6392: cold, but changes after cycling to room temperature. As one can see from the
6393: enlargements in Fig.\ \ref{fig102}, a splitting of the peaks occurs in a range of magnetic
6394: fields, presumably as spin splitting becomes resolved. A curious aspect of the
6395: effect (which has remained unexplained) is that the oscillations have a much
6396: larger amplitude in one field direction than in the other (see Fig.\ \ref{fig101}), in
6397: apparent conflict with the $\pm B$ symmetry of the two-terminal resistance
6398: required by the reciprocity relation (\ref{eq12.16}) in the absence of magnetic
6399: impurities. Other devices of the same design did not show oscillations of well-defined periodicity and had a two-terminal resistance that was approximately
6400: $\pm B$ symmetric.
6401:
6402: \begin{figure}
6403: \centerline{\includegraphics[width=8cm]{figures/fig103}}
6404: \caption{
6405: Equipotentials at the guiding center energy in the saddle-shaped potential created by a split gate (shaded). Aharonov-Bohm oscillations in the point contact magnetoresistance result from the interference of tunneling paths ab and adcb. Tunneling from $a$ to $b$ may be assisted by an impurity at the entrance of the constriction. Taken from P. H. M. van Loosdrecht et al., Phys.\ Rev.\ B {\bf 38}, 10162 (1988).
6406: \label{fig103}
6407: }
6408: \end{figure}
6409:
6410: Figure \ref{fig103} illustrates the tunneling mechanism for the periodic magnetoresistance oscillations as it was originally proposed\cite{ref292} to explain the
6411: observations. Because of the presence of a barrier in the point contact, the
6412: electrostatic potential has a saddle form. Equipotentials at the guiding center
6413: energy (\ref{eq18.1}) are drawn schematically in Fig.\ \ref{fig103} (arrows indicate the
6414: direction of motion along the equipotential). An electron that enters the
6415: constriction at $a$ can be reflected back into the broad region by tunneling to
6416: the opposite edge, either at the potential step at the entrance of the
6417: constriction (from $a$ to $b$) or at its exit (from $d$ to $c$). These two tunneling paths
6418: acquire an Aharonov-Bohm phase difference\cite{ref499} of $eBS/\hbar$ (were $S$ is the
6419: enclosed area {\it abcd}), leading to periodic magnetoresistance oscillations. (Note
6420: that the periodicity $\Delta B$ may differ\cite{ref438,ref500} somewhat from the usual expression
6421: $\Delta B=h/eS$, since $S$ itself is $B$-dependent due to the $B$-dependence of the
6422: guiding center energy.) This mechanism shows how an Aharonov-Bohm
6423: effect is possible in principle in a singly connected geometry: The point
6424: contact behaves as if it were multiply connected, by virtue of the spatial
6425: separation of edge channels moving in opposite directions. (Related mechanisms, based on circulating edge currents, have been considered for
6426: Aharonov-Bohm effects in small conductors.\cite{ref473,ref474,ref501,ref502,ref503}) The oscillations
6427: periodic in $B$ are only observed at large magnetic fields (above about $1\, \mathrm{T}$; the
6428: oscillations at lower fields are Shubnikov-De Haas oscillations periodic in
6429: $1/B$, due to the series resistance of the wide 2DEG regions). At low magnetic
6430: fields the spatial separation of edge channels responsible for the Aharanov-Bohm effect is not yet effective. The spatial separation can also be destroyed
6431: by a large negative gate voltage (top curve in Fig.\ \ref{fig101}), when the width of the
6432: point contact becomes so small that the wave functions of edge states at
6433: opposite edges overlap.
6434:
6435: Although the mechanism illustrated in Fig.\ \ref{fig103} is attractive because it is
6436: an intrinsic consequence of the point contact geometry, the observed well-defined periodicity of the magnetoresistance oscillations requires that the
6437: potential induced by the split gate varies rapidly over a short distance (in
6438: order to have a well-defined area $S$). A smooth saddle potential seems more
6439: realistic. Moreover, one would expect the periodicity to vary more strongly
6440: with gate voltage than the small $10\%$ variation observed experimentally as $V_{\mathrm{g}}$
6441: is changed from $-1.4$ to $-1.7$ V. Glazman and Jonson\cite{ref438} have proposed
6442: that one of the two tunneling processes (from $a$ to $b$ in Fig.\ \ref{fig103}) is mediated
6443: by an impurity outside but close to the constriction. The combination of
6444: impurity and point contact introduces a well-defined area even for a smooth
6445: saddle potential, which moreover will not be strongly gate-voltage-dependent. Such an impurity-assisted Aharonov-Bohm effect in a quantum
6446: point contact has been reported by Wharam et al.\cite{ref504} In order to study the
6447: Aharonov-Bohm effect due to interedge channel tunneling under more
6448: controlled conditions, a double-point contact device is required, as discussed
6449: below.
6450:
6451: \begin{figure}
6452: \centerline{\includegraphics[width=8cm]{figures/fig104}}
6453: \caption{
6454: Cavity (of $1.5\,\mu{\rm m}$ diameter) defined by a double set of split gates A and B. For large negative gate voltages the 2DEG region under the narrow gap between gates A and B is fully depleted, while transmission remains possible over the potential barrier in the wider openings at the left and right of the cavity. Taken from B. J. van Wees et al., Phys.\ Rev.\ Lett.\ {\bf 62}, 2523 (1989).
6455: \label{fig104}
6456: }
6457: \end{figure}
6458:
6459: \begin{figure}
6460: \centerline{\includegraphics[width=8cm]{figures/fig105}}
6461: \caption{
6462: Magnetoconductance experiments on the device of Fig.\ \ref{fig104} at 6 mK, for a fixed gate voltage of $-0.35\,{\rm V}$. (a) Conductance of point contact A, measured with gate B grounded. (b) Conductance of point contact B (gate A grounded). (c) Measured conductance of the entire cavity. (d) Calculated conductance of the cavity, obtained from Eqs.\ (\ref{eq21.1}) and (\ref{eq21.2}) with the measured $G_{\rm A}$ and $G_{\rm B}$ as input. Taken from B. J. van Wees et al., Phys.\ Rev.\ Lett.\ {\bf 62}, 2523 (1989).
6463: \label{fig105}
6464: }
6465: \end{figure}
6466:
6467: {\bf (b) Cavity.} Van Wees et al.\cite{ref500} performed magnetoresistance experiments in
6468: a geometry shown schematically in Fig.\ \ref{fig104}. A cavity with two opposite point
6469: contact openings is defined in the 2DEG by split gates. The diameter of the
6470: cavity is approximately $1.5\,\mu \mathrm{m}$. The conductances $G_{\mathrm{A}}$ and $G_{\mathrm{B}}$ of the two point
6471: contacts A and B can be measured independently (by grounding one set of
6472: gates), with the results plotted in Fig.\ \ref{fig105}a,b (for $V_{\mathrm{g}}=-0.35\,\mathrm{V}$ on either gate
6473: A or B). The conductance $G_{\mathrm{C}}$ of the cavity (for $V_{\mathrm{g}}=-0.35\,\mathrm{V}$ on both the split
6474: gates) is plotted in Fig.\ \ref{fig105}c. A long series of periodic oscillations is observed
6475: between two quantum Hall plateaux. Similar series of oscillations (but with a
6476: different periodicity) have been observed between other quantum Hall
6477: plateaux. The oscillations are suppressed on the plateaux themselves. The
6478: amplitude of the oscilIations is comparable to that observed in the experiment on a single point contact\cite{ref292} (discussed before), but the period is much
6479: smaller (consistent with a larger effective area in the double-point contact
6480: device), and no splitting of the peaks is observed (presumably due to a fully
6481: resolved spin degeneracy). No gross $\pm B$ asymmetries were found in the
6482: present experiment, although an accurate test of the symmetry on field
6483: reversal was not possible because of difficulties with the reproducibility. The
6484: oscillations are quite fragile, disappearing when the temperature is raised
6485: above $200\, \mathrm{mK}$ or when the voltage across the device exceeds $40\,\mu \mathrm{V}$ (the data
6486: in Fig.\ \ref{fig105} were taken at $6\, \mathrm{mK}$ and 6 $\mu \mathrm{V}$). The experimental data are well
6487: described by resonant transmission through a circulating edge state in the
6488: cavity,\cite{ref500} as illustrated in Fig.\ \ref{fig106}a and described in detail later. Aharonov-Bohm oscillations due to resonant transmission through a similar structure
6489: have been reported by Brown et al.\cite{ref505} and analyzed theoretically by
6490: Yosephin and Kaveh.\cite{ref506}
6491:
6492: \begin{figure}
6493: \centerline{\includegraphics[width=8cm]{figures/fig106}}
6494: \caption{
6495: Illustration of mechanisms leading to Aharonov-Bohm oscillations in singly connected geometries. (a) Cavity containing a circulating edge state. Tunneling through the left and right barriers (as indicated by dashed lines) occurs with transmission probabilities $T_{\rm A}$ and $T_{\rm B}$. On increasing the magnetic field, resonant tunneling through the cavity occurs periodically each time the flux $\Phi$ enclosed by the circulating edge state increases by one flux quantum $h/e$. (b) A circulating edge state bound on a local potential maximum causes resonant backscattering, rather than resonant transmission.
6496: \label{fig106}
6497: }
6498: \end{figure}
6499:
6500: {\bf (c) Resonant transmission and reflection of edge channels.} The electrostatic
6501: potential in a point contact has a saddle shape (cf.\ Fig.\ \ref{fig103}), due to the
6502: combination of the lateral confinement and the potential barrier. The height
6503: of the barrier can be adjusted by means of the gate voltage. An edge state with
6504: a guiding center energy below the barrier height is a bound state in the cavity
6505: formed by two opposite point contacts, as is illustrated in Fig.\ \ref{fig106}a.
6506: Tunneling of edge channels through the cavity via this bound state occurs
6507: with transmission probability $T_{\mathrm{AB}}$, which for a singe edge channel is given
6508: by\cite{ref474,ref498}
6509: \begin{eqnarray}
6510: T_{\mathrm{AB}}&=&\left| \frac{t_{\mathrm{A}}t_{\mathrm{B}}}{1-r_{\mathrm{A}}r_{\mathrm{B}} \exp(\mathrm{i}\Phi e/h)}\right|^{2}\nonumber\\
6511: &=&\frac{T_{\mathrm{A}}T_{\mathrm{B}}}{1+R_{\mathrm{A}}R_{\mathrm{B}}-2(R_{\mathrm{A}}R_{\mathrm{B}})^{1/2}\cos(\phi_{0}+\Phi e/\hbar)}.\nonumber\\
6512: &&
6513: \label{eq21.1}
6514: \end{eqnarray}
6515: Here $t_{\mathrm{A}}$ and $r_{\mathrm{A}}$ are the transmission and reflection probability amplitudes
6516: through point contact $\mathrm{A}$, $T_{\mathrm{A}}\equiv|t_{\mathrm{A}}|^{2}$, and $R_{\mathrm{A}}\equiv|r_{\mathrm{A}}|^{2}=1-T_{\mathrm{A}}$ are the transmission and reflection probabilities, and $t_{\mathrm{B}}, r_{\mathrm{B}}, T_{\mathrm{B}}, R_{\mathrm{B}}$ denote the corresponding quantities for point contact B. In Eq.\ (\ref{eq21.1}) the phase acquired by the
6517: electron on one revolution around the cavity is the sum of the phase $\phi_{0}$ from
6518: the reflection probability amplitudes (which can be assumed to be only
6519: weakly $B$-dependent) and of the Aharonov-Bohm phase $\Phi\equiv BS$, which
6520: varies rapidly with $B$ ($\Phi$ is the flux through the area $S$ enclosed by the
6521: equipotential along which the circulating edge state is extended). Resonant
6522: transmission occurs periodically with $B$, whenever $\phi_{0}+\Phi e/\hbar$ is a multiple of
6523: $2\pi$. In the weak coupling limit ($T_{\mathrm{A}}, T_{\mathrm{B}}\ll 1$), Eq.\ (\ref{eq21.1}) is equivalent to the
6524: Breit-Wigner resonant tunneling formula (\ref{eq17.1}). This equivalence has been
6525: discussed by B\"{u}ttiker,\cite{ref386} who has also pointed out that the Breit-Wigner
6526: formula is more generally applicable to the case that several edge channels
6527: tunnel through the cavity via the same bound state.
6528:
6529: In the case that only a single (spin-split) edge channel is occupied in the
6530: 2DEG, the conductance $G_{\mathrm{C}}=(e^{2}/h)T_{\mathrm{AB}}$ of the cavity follows directly from Eq.\
6531: (\ref{eq21.1}). The transmission and reflection probabilities can be determined
6532: independently from the individual point contact conductances $G_{\mathrm{A}}=(e^{2}/h)T_{\mathrm{A}}$
6533: (and similarly for $G_{\mathrm{B}}$), at least if one may assume that the presence of the
6534: cavity has no effect on $T_{\mathrm{A}}$ and $T_{\mathrm{B}}$ itself (but only on the total transmission
6535: probability $T_{\mathrm{AB}}$). If $N>1$ spin-split edge channels are occupied and the
6536: $N-1$ lowest-index edge channels are fully transmitted, one can write
6537: \begin{eqnarray}
6538: &&G_{\mathrm{C}}= \frac{e^{2}}{h}(N-1+T_{\mathrm{AB}}),\;\;G_{\mathrm{A}}= \frac{e^{2}}{h}(N-1+T_{\mathrm{A}}),\nonumber\\
6539: && G_{\mathrm{B}}= \frac{e^{2}}{h}(N-1+T_{\mathrm{B}}).
6540: \label{eq21.2}
6541: \end{eqnarray}
6542: Van Wees et al.\cite{ref500} have compared this simple model with their experimental
6543: data, as shown in Fig.\ \ref{fig105}. The trace in Fig.\ \ref{fig105}d has been calculated from
6544: Eqs.\ (\ref{eq21.1}) and (\ref{eq21.2}) by using the individual point contact conductances in
6545: Fig.\ \ref{fig105}a,b as input for $T_{\mathrm{A}}$ and $T_{\mathrm{B}}$. The flux $\Phi$ has been adjusted to the
6546: experimental periodicity of $3\, \mathrm{mT}$, and the phase $\phi_{0}$ in Eq.\ (\ref{eq21.1}) has been
6547: ignored (since that would only amount to a phase shift of the oscillations).
6548: Energy averaging due to the finite temperature and voltage has been taken
6549: into account in the calculation. The agreement with experimental trace (Fig.\
6550: \ref{fig105}c) is quite satisfactory.
6551:
6552: Resonant reflection of an edge channel can occur in addition to the
6553: resonant transmission already considered. Aharonov-Bohm oscillations due
6554: to interference of the reflections at the entrance and exit of a point contact,
6555: illustrated in Fig.\ \ref{fig103}, are one example of resonant reflection.\cite{ref292}
6556: Jain\cite{ref498} has
6557: considered resonant reflection via a localized state circulating around a
6558: potential maximum, as in Fig.\ \ref{fig106}b. Such a maximum may result naturally
6559: from a repulsive scatterer or artificially in a ring geometry (cf.\ Fig.\ \ref{fig100}).
6560: Tunneling of an edge state at each of the channel boundaries through the
6561: localized state occurs with probabilities $T_{\mathrm{A}}$ and $T_{\mathrm{B}}$. The reflection probability
6562: of the edge channel is still given by $T_{\mathrm{AB}}$ in Eq.\ (\ref{eq20.1}), but the channel
6563: conductance $G_{\mathrm{C}}$ is now a decreasing function of $T_{\mathrm{AB}}$, according to
6564: \be
6565: G_{\mathrm{C}}= \frac{e^{2}}{h}(N-T_{\mathrm{AB}}). \label{eq21.3}
6566: \ee
6567: Quasi-periodic magnetoresistance oscillations have been observed in narrow
6568: channels by several groups.\cite{ref70,ref496,ref507} These may occur by resonant reflection
6569: via one or more localized states in the channel, as in Fig.\ \ref{fig106}b.
6570:
6571: \subsection{\label{sec22} Magnetically induced band structure}
6572:
6573: The one-dimensional nature of edge channel transport has recently been
6574: exploited in an innovative way by Kouwenhoven et al.\cite{ref250} to realize a one-dimensional superlattice exhibiting band structure in strong magnetic fields.
6575: The one-dimensionality results because only the highest-index edge channel
6576: (with the smallest guiding center energy) has an appreciable backscattering
6577: probability. The $N-1$ lower-index edge channels propagate adiabatically,
6578: with approximately unit transmission probability. One-dimensionality in
6579: zero magnetic fields cannot be achieved with present techniques. That is one
6580: important reason why the zero-field superlattice experiments described in
6581: Section \ref{sec11} could not provide conclusive evidence for a bandstructure effect.
6582: The work by Kouwenhoven et al.\cite{ref250} is reviewed in Section \ref{sec22a}. The
6583: magnetically induced band structure differs in an interesting way from the
6584: zero-field band structure familiar from solid-state textbooks, as we show in
6585: Section \ref{sec22b}.
6586:
6587: \subsubsection{\label{sec22a} Magnetotransport through a one-dimensional superlattice}
6588:
6589: \begin{figure}
6590: \centerline{\includegraphics[width=8cm]{figures/fig107}}
6591: \caption{
6592: Inset: Corrugated gate used to define a narrow channel with a one-dimensional periodic potential (the total number of barriers is 16, corresponding to 15 unit cells). Plotted is the conductance in a magnetic field of 2 T as a function of the voltage on the smooth gate at 10 mK. The deep conductance minima (marked by $+$ and $*$) are attributed to minigaps, and the 15 enclosed maxima to discrete states in the miniband. Taken from L. P. Kouwenhoven et al., Phys.\ Rev.\ Lett.\ {\bf 65}, 361 (1990).
6593: \label{fig107}
6594: }
6595: \end{figure}
6596:
6597: The device studied by Kouwenhoven et al.\cite{ref250} is shown in the inset of Fig.\ \ref{fig107}. A narrow channel is defined in the 2DEG of a GaAs-AlGaAs
6598: heterostructure by two opposite gates. One of the gates is corrugated with
6599: period $a=200\,\mathrm{nm}$, to introduce a periodic modulation of the confining
6600: potential. At large negative gate voltages the channel consists of 15 cavities
6601: [as in Section \ref{sec21b}(b)] coupled in series. The conductance of the channel was
6602: measured at $10\, \mathrm{mK}$ in a fixed magnetic field of $2\, \mathrm{T}$, as a function of the voltage
6603: on the gate that defines the smooth channel boundary. The results, reproduced in Fig.\ \ref{fig107}, show two pronounced conductance dips (of magnitude
6604: $0.1\,e^{2}/h)$, with 15 oscillations in between of considerably smaller amplitude.
6605: The two deep and widely spaced dips are attributed to minigaps, the more
6606: rapid oscillations to discrete states in the miniband.
6607:
6608: \begin{figure}
6609: \centerline{\includegraphics[width=8cm]{figures/fig108}}
6610: \caption{
6611: Top: Calculated transmission probability $T_{N}$ of an edge channel through a periodic potential of $N = 15$ periods as a function of the Aharonov-Bohm phase $eBS/\hbar$ (with $S$ the area of one unit cell). The transmission probability through a single barrier is varied as shown in the bottom panel. Taken from L. P. Kouwenhoven et al., Phys.\ Rev.\ Lett.\ {\bf 65}, 361 (1990).
6612: \label{fig108}
6613: }
6614: \end{figure}
6615:
6616: This interpretation is supported in Ref.\ \onlinecite{ref250} by a calculation of the
6617: transmission probability amplitude $t_{n}$ through $n$ cavities in series, given by
6618: the recursion formula
6619: \be
6620: t_{n}= \frac{tt_{n-1}}{1-rr_{n-1} \exp(\mathrm{i}\phi)}. \label{eq22.1}
6621: \ee
6622: Here $t$ and $r$ are transmission and reflection probability amplitudes of the
6623: barrier separating two cavities (all cavitities are assumed to be identical), and
6624: $\phi=eBS/h$ is the Aharonov-Bohm phase for a circulating edge state
6625: enclosing area $S$. Equation (\ref{eq22.1}) is a generalization of Eq.\ (\ref{eq21.1}) for a single
6626: cavity. The dependence on $\phi$ of $T_{n}=|t_{n}|^{2}$ shown in Fig.\ \ref{fig108} is indeed
6627: qualitatively similar to the experiment. Deep minima in the transmission
6628: probability occur with periodicity $\Delta\phi=2\pi$. Experimentally (where $S$ is
6629: varied via the gate voltage at constant $B$) this would correspond to
6630: oscillations with periodicity $\Delta S=h/eB$ of Aharonov-Bohm oscillations in a
6631: single cavity. The 15 smaller oscillations between two deep minima have the
6632: periodicity of Aharonov-Bohm oscillations in the entire area covered by the
6633: 15 cavities. The observation of such faster oscillations shows that phase
6634: coherence is maintained in the experiment throughout the channel and
6635: thereby provides conclusive evidence for band structure in a lateral
6636: superlattice.
6637:
6638: \subsubsection{\label{sec22b} Magnetically induced band structure}
6639:
6640: {\bf (a) Skew minibands.} The band structure in the experiment of Kouwenhoven
6641: et al.\cite{ref250} is present only in the quantum Hall effect regime and can thus be said
6642: to be {\it magnetically induced}. The magnetic field breaks time-reversal symmetry.
6643: Let us see what consequences that has for the band structure.
6644:
6645: The hamiltonian in the Landau gauge ${\bf A}=(0, Bx, 0)$ is
6646: \be
6647: {\cal H}=\frac{p_{x}^{2}}{2m}+\frac{(p_{y}+eBx)^{2}}{2m}+V(x, y),\;\;V(x, y+a)=V(x, y), \label{eq22.2}
6648: \ee
6649: where $V$ is the periodically modulated confining potential. Bloch's theorem is
6650: not affected by the presence of the magnetic field, since ${\cal H}$ remains periodic in
6651: $y$ (in the Landau gauge). The eigenstates $\Psi$ have the form
6652: \be
6653: \Psi_{nk}(x, y)=\mathrm{e}^{i\mathrm{k}y}f_{nk}(x, y),\;\;f_{nk}(x, y+a)=f_{nk}(x, y), \label{eq22.3}
6654: \ee
6655: where the function $f$ is a solution periodic in $y$ of the eigenvalue problem
6656: \begin{eqnarray}
6657: \left( \frac{p_{x}^{2}}{2m}+\frac{(p_{y}+\hbar k+eBx)^{2}}{2m}+V(x, y)\right)f_{nk}(x, y)\nonumber\\
6658: =E_{n}(k, B)f_{nk}(x, y). \label{eq22.4}
6659: \end{eqnarray}
6660: If the wave number $k$ is restricted to the first Brillouin zone $|k|<\pi/a$, the
6661: index $n$ labels both the subbands from the lateral confinement and the
6662: minibands from the periodic modulation. Since $E$ and $V$ are real, one finds by
6663: taking the complex conjugate of Eq.\ (\ref{eq22.4}) that
6664: \be
6665: E_{n}(k, B)=E_{n}(-k, -B). \label{eq22.5}
6666: \ee
6667: In zero magnetic fields the energy $E$ is an even function of $k$, regardless of the
6668: symmetry of the potential $V$. This can be viewed as a consequence of time-reversal symmetry.\cite{ref508} In nonzero magnetic fields, however, $E$ is only even in
6669: $k$ if the lateral confinement is symmetric:
6670: \be
6671: E_{n}(k, B)=E_{n}(-k, B)\;;\mbox{only if}\;\; V(x, y)=V(-x, y). \label{eq22.6}
6672: \ee
6673:
6674: \begin{figure}
6675: \centerline{\includegraphics[width=8cm]{figures/fig109}}
6676: \caption{
6677: Illustration of magnetically induced band structure in a narrow channel with a weak periodic modulation of the confining potential $V(x)$ (for the case $V(x)\neq V(-x)$). The dashed curves represent the unperturbed dispersion relation (\ref{eq22.7}) for a single Landau level. Skew minibands result from the broken time-reversal symmetry in a magnetic field.
6678: \label{fig109}
6679: }
6680: \end{figure}
6681:
6682: To illustrate the formation of {\it skew\/} minibands in a magnetically induced
6683: band structure, we consider the case of a weak periodic modulation $V_{1}(y)$ of
6684: the confining potential $V(x, y)=V_{0}(x)+V_{1}(x, y)$. The dispersion relation
6685: $E_{n}^{0}(k)$ in the absence of the periodic modulation can be approximated by
6686: \be
6687: E_{n}^{0}(k)=(n- {\textstyle\frac{1}{2}})\hbar\omega_{\mathrm{c}}+V_{0}(x=-kl_{m}^{2}). \label{eq22.7}
6688: \ee
6689: The index $n$ labels the Landau levels, and the wave number $k$ runs from $-\infty$
6690: to $+\infty$. The semiclassical approximation (\ref{eq22.7}) is valid if the confining
6691: potential $V_{0}$ is smooth on the scale of the magnetic length $l_{\mathrm{m}}\equiv(\hbar/eB)^{1/2}$.
6692: [Equation (\ref{eq22.7}) follows from the guiding center energy (\ref{eq18.1}), using the
6693: identity $x\equiv-k\hbar/eB$ between the guiding center coordinate and the wave
6694: number; cf.\ Section \ref{sec12a}] For simplicity we restrict ourselves to the strictly
6695: one-dimensional case of one Landau level and suppress the Landau level
6696: index in what follows. To first order in the amplitude of the periodic
6697: modulation $V_{1}$, the zeroth-order dispersion relation is modified only near the
6698: points of degeneracy $K_{p}$ defined by
6699: \be
6700: E^{0}[K_{p}-p(2\pi/a)]=E^{0}(K_{p}),\;\;p=\pm 1, \pm 2, \ldots. \label{eq22.8}
6701: \ee
6702: A gap opens near $K_{p}$, leading to the formation of a band structure as
6703: illustrated in Fig.\ \ref{fig109}. The gaps do not occur at multiples of $\pi/a$, as in a
6704: conventional 1D band structure. Moreover, the maxima and minima of two
6705: subsequent bands occur at different $k$-values. This implies {\it indirect\/} optical
6706: transitions between the bands if the Fermi level lies in the gap.
6707:
6708: It is instructive to consider the special case of a parabolic confining
6709: potential $V_{0}( x)=\frac{1}{2}m\omega_{0}^{2}x^{2}$ in more detail, for which the zeroth-order dispersion relation can be obtained exactly (Section \ref{sec10}). Since the confinement is
6710: symmetric in $x$, the minigaps in this case occur at the Brillouin zone
6711: boundaries $k=p\pi/a$. Other gaps at points where the periodic modulation
6712: induces transitions between different 1D subbands are ignored for simplicity.
6713: From Eq.\ (\ref{eq10.5}) one then finds that the Fermi energy lies in a minigap when
6714: \be
6715: E_{\mathrm{F}}=(n- {\textstyle\frac{1}{2}})\hbar\omega+\frac{\hbar^{2}}{2M}\left(\frac{p\pi}{a}\right)^{2}, \label{eq22.9}
6716: \ee
6717: with the definitions $\omega\equiv(\omega_{\mathrm{c}}^{2}+\omega_{0}^{2})^{1/2}$, $M\equiv m\omega^{2}/\omega_{0}^{2}$. In the limiting case
6718: $B=0,$ Eq.\ (\ref{eq22.9}) reduces to the usual condition\cite{ref249} that Bragg reflection
6719: occurs when the longitudinal momentum $mv_{y}$ is a multiple of $\hbar\pi/a$. In the
6720: opposite limit of strong magnetic fields $(\omega_{\mathrm{c}}\gg \omega_{0}),$ Eq.\ (\ref{eq22.9}) becomes
6721: \be
6722: aW_{\mathrm{eff}}B=p \frac{h}{e},\;\;W_{\mathrm{eff}} \equiv 2\left(\frac{2E_{\mathrm{G}}}{m\omega_{0}^{2}}\right)^{1/2}. \label{eq22.10}
6723: \ee
6724: The effective width $W_{\mathrm{eff}}$ of the parabolic potential is the separation of the
6725: equipotentials at the guiding center energy $E_{\mathrm{G}} \equiv E_{\mathrm{F}}-(n-\frac{1}{2})\hbar\omega_{\mathrm{c}}$.
6726:
6727: The two-terminal conductance of the periodically modulated channel
6728: drops by $e^{2}/h$ whenever $E_{\mathrm{F}}$ lies in a minigap. If the magnetic field dependence
6729: of $W_{\mathrm{eff}}$ is small, then Eq.\ (\ref{eq22.10}) shows that the magnetoconductance
6730: oscillations have approximately the periodicity $\Delta B\sim h/eaW_{\mathrm{eff}}$ of the
6731: Aharonov-Bohm effect in a single unit cell, in agreement with the calculations of Kouwenhoven et al.\cite{ref250} (Note that in their experiment the Fermi
6732: energy is tuned through the minigap by varying the gate voltage rather than
6733: the magnetic field.) The foregoing analysis is for a channel of infinite length.
6734: The interference of reflections at the entrance and exit of a finite superlattice
6735: of length $L$ leads to transmission resonances\cite{ref249,ref387} whenever $k=p\pi/L$, as
6736: described by Eqs.\ (\ref{eq22.9}) and (\ref{eq22.10}) after substituting $L$ for $a$. These
6737: transmission resonances are observed by Kouwenhoven et al.\ as rapid
6738: oscillations in the conductance. The number of conductance maxima between
6739: two deep minima from the minigap equals approximately the number $L/a$ of
6740: unit cells in the superlattice. The number of maxima may become somewhat
6741: larger than $L/a$ if one takes into account reflections at the transition from a
6742: narrow channel to a wide 2DEG. This might explain the observation in Ref.\
6743: \onlinecite{ref250} of 16, rather than 15, conductance maxima between two minigaps in one
6744: particular experiment on a 15-period superlattice.
6745:
6746: {\bf (b) Bloch oscillations.} In zero magnetic fields, an oscillatory current has
6747: been predicted to occur on application of a dc electric field to an electron gas
6748: in a periodic potential.\cite{ref509} This {\it Bloch oscillation\/} would result from Bragg
6749: reflection of electrons that, accelerated by the electric field, approach the
6750: band gap. A necessary condition is that the field be sufficiently weak that
6751: tunneling across the gap does not occur.\cite{ref510,ref511,ref512,ref513} The wave number increases
6752: in time according to $\dot{k}=eE/\hbar$ in an electric field $E$. The time interval between
6753: two Bragg reflections is $2\pi/a\dot{k}=h/eaE$. The oscillatory current thus would
6754: have a frequency $\Delta Ve/h$, with $\Delta V=aE$ the electrostatic potential drop over
6755: one unit cell. Bloch oscillations have so far eluded experimental observation.
6756:
6757: The successful demonstration\cite{ref250} of miniband formation in strong magnetic fields naturally leads to the question of whether Bloch oscillations might
6758: be observable in such a system. This question would appear to us to have a
6759: negative answer. The reason is simple, and it illustrates another interesting
6760: difference of magnetically induced band structure. In the quantum Hall effect
6761: regime the electric field is perpendicular to the current, so no acceleration of
6762: the electrons occurs. Since $\dot{k}=0$, no Bloch oscillations should be expected.
6763:
6764: \begin{thebibliography}{999}
6765: \bibitem{ref1} Y. Imry, in {\em Directions in Condensed Matter Physics}, Vol.\ 1 (G. Grinstein and G. Mazenko, eds.). World Scientific, Singapore, 1986.
6766: \bibitem{ref2} N. G. van Kampen, {\em Stochastic Processes in Physics and Chemistry}. North-Holland, Amsterdam, 1981.
6767: \bibitem{ref3} H. van Houten and C. W. J. Beenakker, in {\em Analogies in Optics and Microelectronics\/} (W.
6768: van Haeringen and D. Lenstra, eds.). Kluwer Academic, Dordrecht, 1990.
6769: \bibitem{ref4} R. Landauer, IBM J. Res. Dev. \textbf{1}, 223 (1957); \textbf{32}, 306 (1988).
6770: \bibitem{ref5} M. B\"{u}ttiker, Phys. Rev. Lett. \textbf{57}, 1761 (1986).
6771: \bibitem{ref6} B. J. van Wees, H. van Houten, C. W. J. Beenakker, J. G. Williamson, J. P. Kouwenhoven, D.
6772: van der Marel, and C. T. Foxon, Phys. Rev. Lett. 60, 848 (1988).
6773: \bibitem{ref7} D. A. Wharam, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, J. E. F. Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, J. Phys. C \textbf{21}, L209 (1988).
6774: \bibitem{ref8} K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. \textbf{45}, 494 (1980).
6775: \bibitem{ref9} Semiconductors and Semimetals, Vol.\ 35 (M. A. Reed, ed.). Academic, New York, 1992.
6776: \bibitem{ref10} P. A. Lee, R. A. Webb and B. L. Al'tshuler, eds., {\em Mesoscopic Phenomena in Solids.} Elsevier, Amsterdam, to be published.
6777: \bibitem{ref11} B. L. Al'tshuler, R. A. Webb, and R. B. Laibowitz, eds., IBM J. Res. Dev. \textbf{32}, 304--437, 439--579 (1988).
6778: \bibitem{ref12} {\em Proceedings of the International Conference on Electronic Properties of Two-Dimensional
6779: Systems,} IV-VIII, Surf Sci. \textbf{113} (1982); \textbf{142} (1984); \textbf{170} (1986); \textbf{196} (1988); \textbf{229} (1990).
6780: \bibitem{ref13} M. J. Kelly and C. Weisbuch, eds., {\em The Physics and Fabrication of Microstructures and
6781: Microdevices.} Proc. Winter School Les Houches, 1986, Springer, Berlin, 1986.
6782: \bibitem{ref14} H. Heinrich, G. Bauer, and F. Kuchar, eds., {\em Physics and Technology of Submicron
6783: Structures.} Springer, Berlin, 1988.
6784: \bibitem{ref15} M. Reed and W. P. Kirk, eds., {\em Nanostructure Physics and Fabrication.} Academic, New York, 1989.
6785: \bibitem{ref16} S. P. Beaumont and C. M. Sotomayor-Torres, eds., {\em Science and Engineering of 1- and 0-Dimensional Semiconductors.} Plenum, London, 1990.
6786: \bibitem{ref17} J. M. Chamberlain, L. Eaves, and J. C. Portal, eds., {\em Electronic Properties of Multilayers and
6787: Low-Dimensional Semiconductor Structures.} Plenum, London, to be published.
6788: \bibitem{ref18} R. Landauer, Phys. Today \textbf{42}, 119 (1989).
6789: \bibitem{ref19} R. T. Bate, Sci. Am. \textbf{258}, 78 (1988).
6790: \bibitem{ref20} T. Ando, A. B. Fowler, and F. Stern, Rev. Mod. Phys. \textbf{54}, 437 (1982).
6791: \bibitem{ref21} S. M. Sze, {\em Physics of Semiconductor Devices.} Wiley, New York, 1981.
6792: \bibitem{ref22} E. H. Nicollian and J. R. Brew, {\em Metal Oxide Semiconductor Technology.} Wiley, New York, 1982.
6793: \bibitem{ref23} B. J. F. Lin, M. A. Paalanen, A. C. Gossard, and D. C. Tsui, Phys. Rev. B \textbf{29}, 927 (1984).
6794: \bibitem{ref24} H. Z. Zheng, H. P. Wei, D. C. Tsui, and G. Weimann, Phys. Rev. B \textbf{34}, 5635 (1986).
6795: \bibitem{ref25} K. K. Choi, D. C. Tsui, and K. Alavi, Phys. Rev. B \textbf{36}, 7751 (1987); Appl. Phys. Lett. \textbf{50}, 110 (1987).
6796: \bibitem{ref26} H. van Houten, C. W. J. Beenakker, B. J. van Wees, and J. E. Mooij, Surf. Sci. \textbf{196}, 144 (1988).
6797: \bibitem{ref27} H. van Houten, C. W. J. Beenakker, M. E. I. Broekaart, M. G. J. Heijman, B. J. van Wees, J. E.
6798: Mooij, and J. P. Andr\'{e}, Acta Electronica, \textbf{28}, 27 (1988).
6799: \bibitem{ref28} D. J. Bishop, R. C. Dynes, and D. C. Tsui, Phys. Rev. B \textbf{26}, 773 (1982).
6800: \bibitem{ref29} W. J. Skocpol, L. D. Jackel, E. L. Hu, R. E. Howard, and L. A. Fetter, Phys. Rev. Lett. \textbf{49}, 951 (1982).
6801: \bibitem{ref30} K. K. Choi, Phys. Rev. B \textbf{28}, 5774 (1983).
6802: \bibitem{ref31} H. van Houten, B. J. van Wees, and C. W. J. Beenakker, in Ref.\ \protect\onlinecite{ref14}.
6803: \bibitem{ref32} A. B. Fowler, A. Hartstein, and R. A. Webb, Phys. Rev. Lett. \textbf{48}, 196 (1982).
6804: \bibitem{ref33} M. Pepper and M. J. Uren, J. Phys. C \textbf{15}, L617 (1982).
6805: \bibitem{ref34} C. C. Dean and M. Pepper, J. Phys. C \textbf{15}, L1287 (1982).
6806: \bibitem{ref35} A. B. Fowler, J. J. Wainer, and R. A. Webb, IBM J. Res. Dev. \textbf{32}, 372 (1988).
6807: \bibitem{ref36} S. B. Kaplan and A. C. Warren, Phys. Rev. B \textbf{34}, 1346 (1986).
6808: \bibitem{ref37} S. B. Kaplan and A. Hartstein, IBM J. Res. Dev. \textbf{32}, 347 (1988); Phys. Rev. Lett. \textbf{56}, 2403 (1986).
6809: \bibitem{ref38} R. G. Wheeler, K. K. Choi, A. Goel, R. Wisnieff, and D. E. Prober, Phys. Rev. Lett. \textbf{49}, 1674 (1982).
6810: \bibitem{ref39} R. F. Kwasnick, M. A. Kastner, J. Melngailis, and P. A. Lee, Phys. Rev. Lett. \textbf{52}, 224 (1984).
6811: \bibitem{ref40} J. C. Licini. D. J. Bishop, M. A. Kastner, and J. Melngailis, Phys. Rev. Lett. \textbf{55}, 2987 (1985).
6812: \bibitem{ref41} P. H. Woerlee, G. A. M. Hurkx, W. J. M. J. Josquin, and J. F. C. M. Verhoeven, Appl. Phys. Lett. \textbf{47}, 700 (1985); see also H. van Houten and P. H. Woerlee, {\em Proc. ICPS 18}, p.\ 1515 (O. Engstr\"{o}m, ed.). World Scientific, Singapore, 1987.
6813: \bibitem{ref42} S. E. Laux and F. Stern, Appl. Phys. Lett. \textbf{49}, 91 (1986).
6814: \bibitem{ref43} A. C. Warren, D. A. Antoniadis, and H. I. Smith, Phys. Rev. Lett. \textbf{56}, 1858 (1986).
6815: \bibitem{ref44} A. C. Warren, D. A. Antoniadis, and H. I. Smith, IEEE Electron Device Lett., \textbf{EDL-7}, 413 (1986).
6816: \bibitem{ref45} W. J. Skocpol, P. M. Mankiewich. R. E. Howard, L. D. Jackel, D. M. Tennant, and A. D.
6817: Stone, Phys. Rev. Lett. \textbf{56}, 2865 (1986).
6818: \bibitem{ref46} W. J. Skocpol, Physica Scripta \textbf{T19}, 95 (1987).
6819: \bibitem{ref47} K. S. Ralls, W. J. Skocpol, L. D. Jackel, R. E. Howard, L. A. Fetter, R. W. Epworth, and D. M. Tennant, Phys. Rev. Lett. \textbf{52}, 228 (1984).
6820: \bibitem{ref48} R. E. Howard, W. J. Skocpol, L. D. Jackel, P. M. Mankiewich, L. A. Fetter, D. M. Tennant, R. Epworth, and K. S. Ralls, IEEE Trans. \textbf{ED-32}, 1669 (1985).
6821: \bibitem{ref49} H. L. St\"ormer, R. Dingle, A. C. Gossard, and W. Wiegman, {\em Proc. 14th ICPS}, p. 6 (B. L. H. Wilson, ed.). Institute of Physics, London, 1978; R. Dingle, H. L. St\"ormer, A. C. Gossard, and W. Wiegman, Appl. Phys. Lett. \textbf{7}, 665 (1978).
6822: \bibitem{ref50} J. J. Harris, J. A. Pals, and R. Woltjer, Rep. Prog. Phys. \textbf{52}, 1217 (1989).
6823: \bibitem{ref51} S. Adachi, J. Appl. Phys. \textbf{58}, R1 (1985).
6824: \bibitem{ref52} D. Delagebeaudeuf and N. T. Linh, IEEE Trans. \textbf{ED-28}, 790 (1981).
6825: \bibitem{ref53} J. P. Kirtley, Z. Schlesinger, T. N. Theis. F. P. Milliken. S. L. Wright. and L. F. Palmateer, Phys. Rev. B \textbf{34}, 5414 (1986).
6826: \bibitem{ref54} L. Bliek. E. Braun. G. Hein. V. Kose, J. Niemeyer, G. Weimann, and W. Schlapp, Semicond.
6827: Sci. Technol. \textbf{1}, 110 (1986).
6828: \bibitem{ref55} K. K. Choi, D. C. Tsui, and S. C. Palmateer, Phys. Rev. B \textbf{33}, 8216 (1986).
6829: \bibitem{ref56} A. D. C. Grassie, K. M. Hutchings, M. Lakrimi, C. T. Foxon, and J. J. Harris, Phys. Rev. B \textbf{36}, 4551 (1987).
6830: \bibitem{ref57} T. Demel, D. Heitmann, P. Grambow. and K. Ploog, Appl. Phys. Lett. \textbf{53}, 2176 (1988).
6831: \bibitem{ref58} T. J. Thornton, M. Pepper, H. Ahmed, D. Andrews, and G. J. Davies, Phys. Rev. Lett. \textbf{56}, 1198 (1986).
6832: \bibitem{ref59} H. van Houten, B. J. van Wees, J. E. Mooij, C. W. J. Beenakker, J. G. Williamson, and C. T.
6833: Foxon, Europhys. Lett. \textbf{5}, 721 (1988).
6834: \bibitem{ref60} S. E. Laux, D. J. Frank, and F. Stern, Surf Sci. \textbf{196}, 101 (1988).
6835: \bibitem{ref61} A. Kumar, S. E. Laux, and F. Stern, Appl. Phys. Lett. \textbf{54}, 1270 (1989).
6836: \bibitem{ref62} K. Ismail, W. Chu, D. A. Antoniadis, and H. I. Smith, Appl. Phys. Lett. \textbf{52}, 1071 (1988).
6837: \bibitem{ref63} H. van Houten, B. J. van Wees, M. G. J. Heijman, and J. P. Andr\'e, Appl. Phys. Lett. \textbf{49}, 1781 (1986).
6838: \bibitem{ref64} R. E. Behringer, P. M. Mankiewich, and R. E. Howard, J. Vac. Sci. Technol. \textbf{B5}, 326 (1987).
6839: \bibitem{ref65} A. Scherer, M. L. Roukes, H. G. Craighead, R. M. Ruthen, E. D. Beebe, and J. P. Harbison,
6840: Appl. Phys. Lett. \textbf{51}, 2133 (1987).
6841: \bibitem{ref66} A. Scherer and M. L. Roukes, Appl. Phys. Lett. \textbf{55}, 377 (1989).
6842: \bibitem{ref67} M. L. Roukes, A. Scherer, S. J. Allen, Jr., H. G. Craighead, R. M. Ruthen, E. D. Beebe, and J. P. Harbison, Phys. Rev. Lett. \textbf{59}, 3011 (1987).
6843: \bibitem{ref68} G. Timp, A. M. Chang, P. Mankiewich, R. Behringer, J. E. Cunningham, T. Y. Chang, and R. E. Howard, Phys. Rev. Lett. \textbf{59}, 732 (1987).
6844: \bibitem{ref69} G. Timp, A. M. Chang, J. E. Cunningham, T. Y. Chang, P. Mankiewich, R. Behringer, and R. E. Howard, Phys. Rev. Lett. \textbf{58}, 2814 (1987).
6845: \bibitem{ref70} A. M. Chang, G. Timp, T. Y. Chang, J. E. Cunningham, P. M. Mankiewich, R. E. Behringer, and R. E. Howard, Solid State Comm. \textbf{67}, 769 (1988).
6846: \bibitem{ref71} K. Y. Lee, T. P. Smith, III, C. J. B. Ford, W. Hansen, C. M. Knoedler, J. M. Hong, and D. P.
6847: Kern, Appl. Phys. Lett. \textbf{55}, 625 (1989).
6848: \bibitem{ref72} J. H. Davies and J. A. Nixon, Phys. Rev. B \textbf{39}, 3423 (1989); J. H. Davies, in Ref.\ \protect\onlinecite{ref15}.
6849: \bibitem{ref73} C. J. B. Ford, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, G. J. Davies, and D. Andrews, Superlattices and Microstructures \textbf{4}, 541 (1988).
6850: \bibitem{ref74} C. J. B. Ford, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, C. T. Foxon, J. J. Harris, and C. Roberts, J. Phys. C. \textbf{21}, L325 (1988).
6851: \bibitem{ref75} T. P. Smith, III, H. Arnot, J. M. Hong, C. M. Knoedler, S. E. Laux, and H. Schmid, Phys. Rev. Lett. \textbf{59}, 2802 (1987).
6852: \bibitem{ref76} T. P. Smith, III, J. A. Brum, J. M. Hong, C. M. Knoedler, H. Arnot, and L. Esaki, Phys. Rev. Lett. \textbf{61}, 585 (1988).
6853: \bibitem{ref77} C. J. B. Ford, S. Washburn, M. B\"uttiker, C. M. Knoedler, and J. M. Hong, Phys. Rev. Lett. \textbf{62}, 2724 (1989).
6854: \bibitem{ref78} W. Hansen, M. Horst, J. P. Kotthaus, U. Merkt, Ch. Sikorski, and K. Ploog, Phys. Rev. Lett. \textbf{58}, 2586 (1987).
6855: \bibitem{ref79} F. Brinkop, W. Hansen, J. P. Kotthaus, and K. Ploog, Phys. Rev. B \textbf{37}, 6547 (1988).
6856: \bibitem{ref80} H. van Houten, C. W. J. Beenakker, J. G. Williamson, M. E. I. Broekaart, P. H. M. van Loosdrecht, B. J. van Wees, J. E. Mooij, C. T. Foxon, and J. J. Harris, Phys. Rev. B \textbf{39}, 8556 (1989).
6857: \bibitem{ref81} T. Hiramoto, K. Hirakawa, Y. Iye, and T. Ikoma, Appl. Phys. Lett. \textbf{54}, 2103 (1989).
6858: \bibitem{ref82} T. L. Cheeks, M. L. Roukes, A. Scherer, and H. G. Craighead, Appl. Phys. Lett. \textbf{53}, 1964 (1988).
6859: \bibitem{ref83} K. Kash, J. M. Worlock, M. D. Sturge, P. Grabbe, J. P. Harbison, A. Scherer, and P. S. D. Lin, Appl. Phys. Lett. \textbf{53}, 782 (1988).
6860: \bibitem{ref84} U. Meirav, M. Heiblum, and F. Stern, Appl. Phys. Lett. \textbf{52}, 1268 (1988).
6861: \bibitem{ref85} U. Meirav, M. A. Kastner, M. Heiblum, and S. J. Wind, Phys. Rev. B \textbf{40}, 5871 (1989).
6862: \bibitem{ref86} A. D. Wieck and K. Ploog, Surf Sci. \textbf{229}, 252 (1990); Appl. Phys. Lett. \textbf{56}, 928 (1990).
6863: \bibitem{ref87} P. M. Petroff, A. C. Gossard, and W. Wiegmann, Appl. Phys. Lett. \textbf{45}, 620 (1984).
6864: \bibitem{ref88} T. Fukui and H. Saito, Appl. Phys. Lett. \textbf{50}, 824 (1987).
6865: \bibitem{ref89} H. Asai, S. Yamada, and T. Fukui, Appl. Phys. Lett. \textbf{51}, 1518 (1987).
6866: \bibitem{ref90} T. Fukui, and H. Saito, J. Vac. Sci. Technol. \textbf{B6}, 1373 (1988).
6867: \bibitem{ref91} J. Motohisa, M. Tanaka, and H. Sakaki, Appl. Phys. Lett. \textbf{55}, 1214 (1989).
6868: \bibitem{ref92} H. Yamaguchi and Y. Horikoshi, Jap. J. Appl. Phys. \textbf{28}, L1456 (1989).
6869: \bibitem{ref93} L. D. Landau and E. M. Lifshitz, \emph{Quantum Mechanics.} Pergamon, Oxford, 1977.
6870: \bibitem{ref94} N. W. Ashcroft and N. D. Mermin, {\em Solid State Physics.} Holt, Rinehart and Winston, New York, 1976.
6871: \bibitem{ref95} R. Kubo, M. Toda, and N. Hashitsume, {\em Statistical Physics II.} Springer, Berlin, 1985.
6872: \bibitem{ref96} A. A. Abrikosov, {\em Fundamentals of the Theory of Metals.} North-Holland, Amsterdam, 1988.
6873: \bibitem{ref97} R. E. Prange and S. M. Girvin, eds., {\em The Quantum Hall Effect.} Springer, New York, 1987.
6874: \bibitem{ref98} I. Stone. Phys. Rev. \textbf{6}, 1 (1898).
6875: \bibitem{ref99} R. G. Chambers, in {\em The Physics of Metals,} Vol. 1 (J. M. Ziman, ed.). Cambridge University Press, Cambridge, 1969.
6876: \bibitem{ref100} G. Br\"{a}ndli and J. L. Olsen, Mater. Sci. Eng. \textbf{4}, 61 (1969).
6877: \bibitem{ref101} E. H. Sondheimer, Adv. Phys. \textbf{1}, 1 (1952).
6878: \bibitem{ref102} A. B. Pippard, {\em Magnetoresistance in Metals.} Cambridge University Press, Cambridge, 1989.
6879: \bibitem{ref103} K. Fuchs, Proc. Cambridge Philos. Soc. \textbf{34}, 100 (1938).
6880: \bibitem{ref104} D. K. C. MacDonald, Nature \textbf{163}, 637 (1949); D. K. C. MacDonald and K. Sarginson, Proc. Roy. Soc. A \textbf{203}, 223 (1950).
6881: \bibitem{ref105} E. H. Sondheimer, Nature \textbf{164}, 920 (1949); Phys. Rev. \textbf{80}, 401 (1950).
6882: \bibitem{ref106} S. B. Soffer, J. Appl. Phys. \textbf{38}, 1710 (1967).
6883: \bibitem{ref107} T. J. Thornton, M. L. Roukes, A. Scherer, and B. P. van der Gaag, Phys. Rev. Lett. \textbf{63}, 2128 (1989).
6884: \bibitem{ref108} K. Nakamura, D. C. Tsui, F. Nihey, H. Toyoshima, and T. Itoh, Appl. Phys. Lett. \textbf{56}, 385 (1990).
6885: \bibitem{ref109} C. W. J. Beenakker and H. van Houten, Phys. Rev. B \textbf{38}, 3232 (1988).
6886: \bibitem{ref110} Z. Tesanovic, M. V. Jaric, and S. Maekawa, Phys. Rev. Lett. \textbf{57}, 2760 (1986).
6887: \bibitem{ref111} C. W. J. Beenakker, Phys. Rev. Lett. \textbf{62}, 2020 (1989).
6888: \bibitem{ref112} M. B\"{u}ttiker, Phys. Rev. B \textbf{38}, 9375 (1988).
6889: \bibitem{ref113} H. van Houten, C. W. J. Beenakker, P. H. M. van Loosdrecht, T. J. Thornton, H. Ahmed, M. Pepper, C. T. Foxon, and J. J. Harris, Phys. Rev. B \textbf{37}, 8534 (1988); and unpublished.
6890: \bibitem{ref114} E. Ditlefsen and J. Lothe. Phil. Mag. \textbf{14}, 759 (1966).
6891: \bibitem{ref115} E. Abrahams, P. W. Anderson, D. C. Licciardello, and T. V. Ramakrishnan, Phys. Rev. Lett. \textbf{42}, 673 (1979).
6892: \bibitem{ref116} P. W. Anderson, E. Abrahams, and T. V. Ramakrishnan, Phys. Rev. Lett. \textbf{43}, 718 (1979).
6893: \bibitem{ref117} L. P. Gorkov, A. I. Larkin, and D. E. Khmel'nitskii, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{30}, 248 (1979) [JETP Lett. \textbf{30}, 228 (1979)].
6894: \bibitem{ref118} B. L. Al'tshuler, D. Khmel'nitskii, A. I. Larkin, and P. A. Lee, Phys. Rev. B \textbf{22}, 5142 (1980).
6895: \bibitem{ref119} A. Kawabata, J. Phys. Soc. Japan \textbf{49}, 628 (1980).
6896: \bibitem{ref120} V. K. Dugaev and D. E. Khmel'nitskii, Zh. Eksp. Teor. Fiz. \textbf{86}, 1784 (1984) [Sov. Phys. JETP \textbf{59}, 1038 (1984)].
6897: \bibitem{ref121} B. L. Al'tshuler and A. G. Aronov, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{33}, 515 (1981) [JETP Lett. \textbf{33}, 499 (1981)].
6898: \bibitem{ref122} P. G. De Gennes and M. Tinkham, Phys. (N.Y.) \textbf{1}, 107 (1964); see also P. G. De Gennes, {\em Superconductivity of Metals and Alloys,} Chapter 8. Benjamin, New York, 1966.
6899: \bibitem{ref123} G. Bergmann, Phys. Rep. \textbf{107}, 1 (1984); Phys. Rev. B \textbf{28}, 2914 (1983).
6900: \bibitem{ref124} A. I. Larkin and D. E. Khmel'nitskii, Usp. Fiz. Nauk \textbf{136}, 536 (1982) [Sov. Phys. Usp. \textbf{25}, 185 (1982)].
6901: \bibitem{ref125} D. E. Khmel'nitskii, Physica B \textbf{126}, 235 (1984).
6902: \bibitem{ref126} S. Chakravarty and A. Schmid, Phys. Rep. \textbf{140}, 193 (1986).
6903: \bibitem{ref127} P. A. Lee and T. V. Ramakrishnan, Rev. Mod. Phys. \textbf{57}, 287 (1985).
6904: \bibitem{ref128} B. L. Al'tshuler, A. G. Aronov, D. E. Khmel'nitskii, and A. I. Larkin, in {\em Quantum Theory of Solids,} p. 130. (I. M. Lifshitz, ed.) Advances in Science and Technology in the USSR, Physics Series, MIR, Moskow.
6905: \bibitem{ref129}R. P. Feynman and A. R. Hibbs, {\em Quantum Mechanics and Path Integrals.} McGraw-Hill, New York, 1965.
6906: \bibitem{ref130} H. P. Wittman and A. Schmid, J. Low Temp. Phys. \textbf{69}, 131 (1987).
6907: \bibitem{ref131} S. Hikami, A. I. Larkin, and Y. Nagaoka, Prog. Theor. Phys. \textbf{63}, 707 (1980).
6908: \bibitem{ref132} A. I. Larkin, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{31}, 239 (1980) [JETP Lett. \textbf{31}, 219 (1980)].
6909: \bibitem{ref133} Y. Kawaguchi and S. Kawaji. J. Phys. Soc. Japan \textbf{48}. 699 (1980).
6910: \bibitem{ref134} R. G. Wheeler, Phys. Rev. B \textbf{24}, 4645 (1981).
6911: \bibitem{ref135} M. J. Uren, R. A. Davis, M. Kaveh, and M. Pepper, J. Phys. C \textbf{14}, L395 (1981).
6912: \bibitem{ref136} D. A. Poole, M. Pepper, and R. W. Glew, J. Phys. C \textbf{14}, L995 (1981).
6913: \bibitem{ref137} M. A. Paalanen, D. C. Tsui, and J. C. M. Hwang, Phys. Rev. Lett. \textbf{51}, 2226 (1983).
6914: \bibitem{ref138} D. M. Pooke, R. Mottahedeh, M. Pepper, and A. Grundlach, Surf Sci. \textbf{196}, 59 (1988); D. M. Pooke, N. Paquin, M. Pepper, and A. Grundlach, J. Phys. Condens. Matter \textbf{1}, 3289 (1989).
6915: \bibitem{ref139} A. M. Chang, G. Timp, R. E. Howard, R. E. Behringer, P. M. Mankiewich, J. E. Cunningham, T. Y. Chang, and B. Chelluri, Superlattices and Microstructures \textbf{4}, 515 (1988).
6916: \bibitem{ref140} B. L. Al'tshuler, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{41}, 530 (1985) [JETP Lett. \textbf{41}, 648 (1985)].
6917: \bibitem{ref141} P. A. Lee and A. D. Stone, Phys. Rev. Lett. \textbf{55}, 1622 (1985).
6918: \bibitem{ref142} P. A. Lee, Physica A \textbf{140}, 169 (1986).
6919: \bibitem{ref143} D. S. Fisher and P. A. Lee, Phys. Rev. B \textbf{23}, 6851 (1981).
6920: \bibitem{ref144} A. D. Stone, in Ref. \protect\onlinecite{ref14}.
6921: \bibitem{ref145} P. A. Lee, A. D. Stone, and H. Fukuyama, Phys. Rev. B \textbf{35}, 1039 (1987).
6922: \bibitem{ref146} B. L. Al'tshuler and D. E. Khmel'nitskii, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{42}, 291 (1985) [JETP Lett. \textbf{42}, 359 (1985)].
6923: \bibitem{ref147} C. W. J. Beenakker and H. van Houten, Phys. Rev. B \textbf{37}, 6544 (1988).
6924: \bibitem{ref148} B. L. Al'tshuler, V. E. Kravtsov, and I. V. Lerner, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{43}, 342 (1986) [JETP Lett. \textbf{43}, 441 (1986)].
6925: \bibitem{ref149} M. B\"{u}ttiker, Phys. Rev. B \textbf{35}, 4123 (1987).
6926: \bibitem{ref150} S. Maekawa, Y. Isawa, and H. Ebisawa, J. Phys. Soc. Japan \textbf{56}, 25 (1987).
6927: \bibitem{ref151} H. U. Baranger: A. D. Stone, and D. P. DiVincenzo, Phys. Rev. B \textbf{37}, 6521 (1988).
6928: \bibitem{ref152} S. Hershfield and V. Ambegaokar, Phys. Rev. B \textbf{38}, 7909 (1988).
6929: \bibitem{ref153} C. L. Kane, P. A. Lee, and D. P. DiVincenzo, Phys. Rev. B \textbf{38}, 2995 (1988).
6930: \bibitem{ref154} D. P. DiVincenzo and C. L. Kane, Phys. Rev. B \textbf{38}, 3006 (1988).
6931: \bibitem{ref155} A. D. Benoit, C. P. Umbach, R. B. Laibowitz, and R. A. Webb, Phys. Rev. Lett. \textbf{58}, 2343 (1987).
6932: \bibitem{ref156} W. J. Skocpol, P. M. Mankiewich, R. E. Howard, L. D. Jackel, D. M. Tennant, and A. D. Stone, Phys. Rev. Lett. \textbf{58}, 2347 (1987).
6933: \bibitem{ref157} W. E. Howard and F. F. Fang, Solid State Electronics \textbf{8}, 82 (1965).
6934: \bibitem{ref158} A. Hartstein, R. A. Webb, A. B. Fowler, and J. J. Wainer, Surf Sci. \textbf{142}, 1 (1984).
6935: \bibitem{ref159} M. Ya. Azbel, Phys. Rev. B \textbf{28}, 4106 (1983).
6936: \bibitem{ref160} P. A. Lee, Phys. Rev. Lett. \textbf{53}, 2042 (1984).
6937: \bibitem{ref161} R. G. Wheeler, K. K. Choi, and R. Wisnieff, Surf. Sci. \textbf{142}, 19 (1984).
6938: \bibitem{ref162} W. J. Skocpol, L. D. Jackel, R. E. Howard, H. G. Craighead, L. A. Fetter, P. M. Mankiewich, P. Grabbe, and D. M. Tennant, Surf. Sci. \textbf{142}, 14 (1984).
6939: \bibitem{ref163} A. D. Stone, Phys. Rev. Lett. \textbf{54}, 2692 (1985).
6940: \bibitem{ref164} R. A. Webb, S. Washburn, H. J. Haucke, A. D. Benoit, C. P. Umbach, and F. P. Milliken, in Ref. \protect\onlinecite{ref14}.
6941: \bibitem{ref165} R. A. Webb, S. Washburn, C. P. Umbach, and R. B. Laibowitz, in {\em Localization, Interaction, and Transport Phenomena,} p. 121. (B. Kramer, G. Bergmann, and Y. Bruynseraede, eds.). Springer, New York, 1984.
6942: \bibitem{ref166} T. J. Thornton, M. Pepper, H. Ahmed, G. J. Davies, and D. Andrews, Phys. Rev. B \textbf{36}, 4514 (1987).
6943: \bibitem{ref167} H. van Houten, B. J. van Wees, J. E. Mooij, G. Roos, and K.-F. Berggren, Superlattices and Microstructures \textbf{3}, 497 (1987).
6944: \bibitem{ref168} R. P. Taylor, M. L. Leadbeater, G. P. Wittington, P. C. Main, L. Eaves, S. P. Beaumont, I. McIntyre, S. Thoms, and C. D. W. Wilkinson, Surf. Sci. \textbf{196}, 52 (1988).
6945: \bibitem{ref169} T. Fukui and H. Saito, Jap. J. Appl. Phys. \textbf{27}, L1320 (1988).
6946: \bibitem{ref170} A. D. Stone, Phys. Rev. B \textbf{39}, 10736 (1989).
6947: \bibitem{ref171} P. Debray, J.-L. Pichard, J. Vicente, and P. N. Tung, Phys. Rev. Lett. \textbf{63}, 2264 (1989).
6948: \bibitem{ref172} N. O. Birge, B. Golding, and W. H. Haemmerle, Phys. Rev. Lett. \textbf{62}, 195 (1989).
6949: \bibitem{ref173} D. Mailly, M. Sanquer, J.-L. Pichard, and P. Pari, Europhys. Lett. \textbf{8}, 471 (1989).
6950: \bibitem{ref174} S. Feng, P. A. Lee, and A. D. Stone, Phys. Rev. Lett. \textbf{56}, 1960 (1986); erratum \textbf{56}, 2772 (1986).
6951: \bibitem{ref175} B. L. Al'tshuler and B. Z. Spivak, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{42}, 363 (1985) [JETP Lett. \textbf{42}, 447 (1985)].
6952: \bibitem{ref176} A. M. Chang, K. Owusu-Sekyere, and T. Y. Chang, Solid State Comm. \textbf{67}, 1027 (1988).
6953: \bibitem{ref177} A. M. Chang, G. Timp, J. E. Cunningham, P. M. Mankiewich, R. E. Behringer, R. E. Howard, and H. U. Baranger, Phys. Rev. B \textbf{37}, 2745 (1988).
6954: \bibitem{ref178} J. A. Simmons, D. C. Tsui, and G. Weimann, Surf Sci. \textbf{196}, 81 (1988).
6955: \bibitem{ref179} S. Yamada, H. Asai, Y. Fukai, and T. Fukui, Phys. Rev. B. (to be published).
6956: \bibitem{ref180} Y. Aharonov and D. Bohm, Phys. Rev. \textbf{115}, 485 (1959).
6957: \bibitem{ref181} D. Yu. Sharvin and Yu. V. Sharvin, Pis'ma Zh. Teor. Fiz. \textbf{34}, 285 (1981) [JETP Lett. \textbf{34}, 272 (1981).
6958: \bibitem{ref182} B. L. Al'tshuler, A. G. Aronov, and B. Z. Spivak, Pis'ma Zh. Teor. Fiz. \textbf{33}, 101 (1981) [JETP Lett. \textbf{33}, 94 (1981)].
6959: \bibitem{ref183} B. L. Al'tshuler, A. G. Aronov, B. Z. Spivak, D. Yu Sharvin, and Yu. V. Sharvin, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{35}, 476 (1982) [JETP Lett. \textbf{35}, 588 (1982)].
6960: \bibitem{ref184} Yu. V. Sharvin, Physica B \textbf{126}, 288 (1984).
6961: \bibitem{ref185} M. Gijs, C. van Haesendonck, and Y. Bruynseraede, Phys. Rev. Lett. \textbf{52}, 2069 (1984); Phys. Rev. B \textbf{30}, 2964 (1984).
6962: \bibitem{ref186} R. A. Webb, S. Washburn, C. P. Umbach, and R. B. Laibowitz, Phys. Rev. Lett. \textbf{54}, 2696 (1985).
6963: \bibitem{ref187} Y. Gefen, Y. Imry, and M. Ya. Azbel, Surf Sci. \textbf{142}, 203 (1984); Phys. Rev. Lett. \textbf{52}, 129 (1984).
6964: \bibitem{ref188} M. B\"{u}ttiker, Y. Imry, R. Landauer, and S. Pinhas, Phys. Rev. B \textbf{31}, 6207 (1985).
6965: \bibitem{ref189} C. P. Umbach, C. Van Haesendonck, R. B. Laibowitz, S. Washburn, and R. A. Webb, Phys. Rev. Lett. \textbf{56}, 386 (1986).
6966: \bibitem{ref190} S. Washburn and R. A. Webb, Adv. Phys. \textbf{35}, 375 (1986).
6967: \bibitem{ref191} A. G. Aronov and Yu. V. Sharvin, Rev. Mod. Phys. \textbf{59}, 755 (1987).
6968: \bibitem{ref192} R. A. Webb, A. Hartstein, J. J. Wainer, and A. B. Fowler, Phys. Rev. Lett. \textbf{54}, 1577 (1985).
6969: \bibitem{ref193} K. Ishibashi, Y. Takagaki, K. Garno, S. Namba, S. Ishida, K. Murase, Y. Aoyagi, and M. Kawabe, Solid State Comm. \textbf{64}, 573 (1987).
6970: \bibitem{ref194} A. M. Chang, G. Timp, T. Y. Chang, J. E. Cunningham, B. Chelluri, P. M. Mankiewich, R. E. Behringer, and R. E. Howard, Surf Sci. \textbf{196}, 46 (1988).
6971: \bibitem{ref195} C. J. B. Ford, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, D. C. Peacock, D. A. Ritchie, J. E. F. Frost, and G. A. C. Jones, Appl. Phys. Lett. \textbf{54}, 21 (1989).
6972: \bibitem{ref196} C. P. Umbach, S. Washburn, R. B. Laibowitz, and R. A. Webb, Phys. Rev. B \textbf{30}, 4048 (1984).
6973: \bibitem{ref197} V. Chandrasekhar, M. J. Rooks, S. Wind, and D. E. Prober, Phys. Rev. Lett. \textbf{55}, 1610 (1985).
6974: \bibitem{ref198} S. Datta and S. Bandyopadhyay, Phys. Rev. Lett. \textbf{58}, 717 (1987).
6975: \bibitem{ref199} S. Datta, M. R. Melloch, S. Bandyopadhyah, R. Noren, M. Vaziri, M. Miller, and R. Reifenberger, Phys. Rev. Lett. \textbf{55}, 2344 (1985).
6976: \bibitem{ref200} J. R. Barker, in Ref. \protect\onlinecite{ref15}.
6977: \bibitem{ref201} G. Timp. A. M. Chang, P. DeVegvar, R. E. Howard, R. Behringer, J. E. Cunningham, and P. Mankiewich, Surf Sci. \textbf{196}, 68 (1988).
6978: \bibitem{ref202} S. Washburn, H. Schmid, D. Kern, and R. A. Webb, Phys. Rev. Lett. \textbf{59}, 1791 (1987).
6979: \bibitem{ref203} P. G. N. de Vegvar, G. Timp, P. M. Mankiewich, R. Behringer, and J. Cunningham, Phys. Rev. B \textbf{40}, 3491 (1989).
6980: \bibitem{ref204} B. L. Al'lshuler, A. G. Aronov, and P. A. Lee, Phys. Rev. Lett. \textbf{44}, 1288 (1980).
6981: \bibitem{ref205} H. Fukuyama, J. Phys. Soc. Japan, \textbf{48}, 2169 (1980).
6982: \bibitem{ref206} B. L. Al'tshuler, A. G. Aronov, A. I. Larkin, and D. E. Khmel'nitskii, Zh. Eksp. Teor. Fiz. \textbf{81}, 768 (1981) [Sov. Phys. JETP \textbf{54}, 411 (1981)].
6983: \bibitem{ref207} B. L. Al'tshuler, and A. G. Aronov, Solid State Comm. \textbf{46}, 429 (1983).
6984: \bibitem{ref208} E. Abrahams, P. W. Anderson, P. A. Lee, and T. V. Ramakrishnan, Phys. Rev. B \textbf{24}, 6783 (1981).
6985: \bibitem{ref209} H. Fukuyama, J. Phys. Soc. Japan, \textbf{50}, 3407, 3562 (1981); \textbf{51}, 1105 (1982).
6986: \bibitem{ref210} P. A. Lee and T. V. Ramakrishnan, Phys. Rev. B \textbf{26}, 4009 (1982).
6987: \bibitem{ref211} B. L. Al'tshuler and A. G. Aronov, in {\em Electron-Electron Interactions in Disordered Systems,} p. 1 (A. L. Efros and M. Pollak, eds.) North-Holland, Amsterdam, 1985.
6988: \bibitem{ref212} H. Fukuyama, in {\em Electron-Electron Interactions in Disordered Systems,} p. 155 (A. L. Efros and M. Pollak, eds.). North-Holland, Amsterdam, 1985.
6989: \bibitem{ref213} A. Isihara, {\em Solid State Physics,} Vol. 42, p. 271 (H. Ehrenreich and D. Turnbull, eds.). Academic, New York, 1989.
6990: \bibitem{ref214} G. Bergmann, Phys. Rev. B \textbf{35}, 4205 (1987).
6991: \bibitem{ref215} A. Houghton, J. R. Senna, and S. C. Ying, Phys. Rev. B \textbf{25}, 2196 (1982).
6992: \bibitem{ref216} S. M. Girvin, M. Jonson, and P. A. Lee, Phys. Rev. B \textbf{26}, 1651 (1982).
6993: \bibitem{ref217} K.-F. Berggren, T. J. Thornton, D, J. Newson, and M. Pepper, Phys, Rev. Lett. \textbf{57}, 1769 (1986).
6994: \bibitem{ref218} K.-F. Berggren, G. Roos, and H. van Houten, Phys. Rev. B \textbf{37}, 10118 (1988).
6995: \bibitem{ref219} R. P. Taylor, P. C. Main, L. Eaves, S. P. Beaumont, I. McIntyre, S. Thoms, and C. D. W. Wilkinson, J. Phys. Condens. Matter \textbf{1}, 10413 (1989).
6996: \bibitem{ref220} P. M. Mensz, R. G. Wheeler, C. T. Foxon, and J. J. Harris, Appl. Phys. Lett. \textbf{50}, 603 (1987); P. M. Mensz and R. G. Wheeler, Phys. Rev. B \textbf{35}, 2844 (1987).
6997: \bibitem{ref221} V. Fal'ko, J. Phys. Condens. Matter \textbf{2}, 3797 (1990).
6998: \bibitem{ref222} L. Smrcka, H. Havlova, and A. Ishara, J. Phys. C \textbf{19}, L457 (1986).
6999: \bibitem{ref223} K.-F. Berggren, and D. J. Newson, Semicond. Sci. Technol. \textbf{1}, 327 (1986).
7000: \bibitem{ref224} J. A. Brum and G. Bastard, Superlattices and Microstructures \textbf{4}, 443 (1988).
7001: \bibitem{ref225} M. J. Kearney and P. N. Butcher, J. Phys. C \textbf{20}, 47 (1987).
7002: \bibitem{ref226} S. Das Sarma and X. C. Xie, Phys. Rev. B \textbf{35}, 9875 (1987).
7003: \bibitem{ref227} P. Vasilopoulos and F. M. Peeters, Phys. Rev. B \textbf{40}, 10079 (1989).
7004: \bibitem{ref228} H. Sakaki, Jap. J. Appl. Phys. \textbf{19}, L735 (1980).
7005: \bibitem{ref229} G. Fishman, Phys. Rev. B \textbf{36}, 7448 (1987).
7006: \bibitem{ref230} J. Lee and M. O. Vassell, J. Phys. C \textbf{17}, 2525 (1984); J. Lee and H. N. Spector, J. Appl. Phys. \textbf{57}, 366 (1985).
7007: \bibitem{ref231} K. Ismail, D. A. Antoniadis, and H. I. Smith, Appl. Phys. Lett. \textbf{54}, 1130 (1989).
7008: \bibitem{ref232} M. Lakrimi, A. D. C. Grassie, K. M. Hutchings, J. J. Harris, and C. T. Foxon, Semicond. Sci. Technol. \textbf{4}, 313 (1989).
7009: \bibitem{ref233}J. J. Alsmeier, Ch. Sikorski, and U. Merkt, Phys. Rev. B \textbf{37}, 4314 (1988).
7010: \bibitem{ref234} Z. Tesanovic, J. Phys. C \textbf{20}, L829 (1987).
7011: \bibitem{ref235} L. Esaki and R. Tsu, IBM J. Res. Dev. \textbf{14}, 61 (1970).
7012: \bibitem{ref236} L. Esaki and L. L. Chang, Phys. Rev. Lett. \textbf{33}, 495 (1974).
7013: \bibitem{ref237} L. Esaki, Rev. Mod. Phys. \textbf{46}, 237 (1974).
7014: \bibitem{ref238} H. Sakaki, K. Wagatsuma, J. Hamasaki, and S. Saito, Thin Solid Films \textbf{36}, 497 (1976).
7015: \bibitem{ref239} R. T. Bate, Bull. Am. Phys. Soc. \textbf{22}, 407 (1977).
7016: \bibitem{ref240} M. J. Kelly, J. Phys. C \textbf{18}, 6341 (1985); Surf Sci. \textbf{170}, 49 (1986).
7017: \bibitem{ref241} P. F. Bagwell and T. P. Orlando, Phys. Rev. B \textbf{40}, 3735 (1989).
7018: \bibitem{ref242} A. C. Warren, D. A. Antoniadis, H. I. Smith, and J. Melngailis, IEEE Electron Device Letts, \textbf{EDL-6}, 294 (1985).
7019: \bibitem{ref243} G. Bernstein and D. K. Ferry, J. Vac. Sci. Technol. B \textbf{5}, 964 (1987).
7020: \bibitem{ref244} C. G. Smith, M. Pepper, R. Newbury, H. Ahmed, D. G. Hasko, D. C. Peacock, J. E. F. Frost, D. A. Ritchie, and G. A. C. Jones, J. Phys. Condens. Matter \textbf{2}, 3405 (1990).
7021: \bibitem{ref245} J. M. Gaines, P. M. Petroff, H. Kroemer, R. J. Simes, R. S. Geels, and J. H. English, J. Vac. Sci. Technol. B \textbf{6}, 1378 (1988).
7022: \bibitem{ref246} H. Sakaki, Jap. J. Appl. Phys. \textbf{28}, L314 (1989).
7023: \bibitem{ref247} T. Cole, A. A. Lakhani, and P. J. Stiles, Phys. Rev. Lett. \textbf{38}, 722 (1977).
7024: \bibitem{ref248} G. E. W. Bauer and A. A. van Gorkum, in Ref. \protect\onlinecite{ref16}.
7025: \bibitem{ref249} S. E. Ulloa, E. Casta\~{n}o, and G. Kirczenow, Phys. Rev. B. \textbf{41}, 12350 (1990).
7026: \bibitem{ref250} L. P. Kouwenhoven, F. W. J. Hekking, B. J. van Wees, C. J. P. M. Harmans, C. E. Timmering, and C. T. Foxon, Phys. Rev. Lett. \textbf{65}, 361 (1990).
7027: \bibitem{ref251} D. K. Ferry, in Ref. \protect\onlinecite{ref14}.
7028: \bibitem{ref252} P. A. Puechner, J. Ma, R. Mezenner, W.-P. Liu, A. M. Kriman, G. N. Maracas, G. Bernstein,
7029: and D. K. Ferry, Surf Sci. \textbf{27}, 137 (1987).
7030: \bibitem{ref253} J. C. Maan, Festk\"{o}rperprobleme \textbf{27}, 137 (1987).
7031: \bibitem{ref254} D. R. Hofstadter, Phys. Rev. B \textbf{14}, 2239 (1976).
7032: \bibitem{ref255} D. Weiss, K. von Klitzing, K. Ploog, and G. Weimann, Europhys. Lett. \textbf{8}, 179 (1989); also in {\em High Magnetic Fields in Semiconductor Physics II} (G. Landwehr, ed.). Springer, Berlin, 1989.
7033: \bibitem{ref256} D. Weiss, C. Zhang, R. R. Gerhardts, K. von Klitzing, and G. Weimann, Phys. Rev. B \textbf{39}, 13020 (1989).
7034: \bibitem{ref257} R. R. Gerhardts, D. Weiss, and K. von Klitzing, Phys. Rev. Lett. \textbf{62}, 1173 (1989).
7035: \bibitem{ref258} R. W. Winkler, J. P. Kotthaus, and K. Ploog, Phys. Rev. Lett. \textbf{62}, 1177 (1989).
7036: \bibitem{ref259} P. Vasilopoulos and F. M. Peeters, Phys. Rev. Lett. \textbf{63}, 2120 (1989); R. R. Gerhardts and C. Zhang, Phys. Rev. Lett. \textbf{64}, 1473 (1990).
7037: \bibitem{ref260} G. Knorr, F. R. Hansen, J. P. Lynov, H. L. Pecseli, and J. J. Rasmussen, Physica Scripta \textbf{38}, 829 (1988).
7038: \bibitem{ref261} A. V. Chaplik, Solid State Comm. \textbf{53}, 539 (1985).
7039: \bibitem{ref262} E. S. Alves, P. H. Beton, M. Henini, L. Eaves, P. C. Main, O. H. Hughes, G. A. Toombs, S. P. Beaumont, and C. D. W. Wilkinson, J. Phys. Condens. Matter \textbf{1}, 8257 (1989).
7040: \bibitem{ref263} D. Weiss, K. von Klitzing, G. Ploog, and G. Weimann, Surf Sci. (to be published).
7041: \bibitem{ref264} K. Ismail, T. P. Smith, III, W. T. Masselink, and H. I. Smith, Appl. Phys. Lett. \textbf{55}, 2766 (1989).
7042: \bibitem{ref265} W. Hansen, T. P. Smith, III, K. Y. Lee, J. A. Brum, C. M. Knoedler, J. M. Hong, and D. P. Kern, Phys. Rev. Lett. \textbf{62}, 2168 (1989).
7043: \bibitem{ref266} C. W. J. Beenakker, H. van Houten, and B. J. van Wees, Superlattices and Microstructures \textbf{5}, 127 (1989).
7044: \bibitem{ref267} R. E. Prange and T.-W. Nee, Phys. Rev. \textbf{168}, 779 (1968).
7045: \bibitem{ref268} C. W. J. Beenakker and H. van Houten, Phys. Rev. Lett. \textbf{60}, 2406 (1988).
7046: \bibitem{ref269} A. M. Kosevich and I. M. Lifshitz, Zh. Eksp. Teor. Fiz. \textbf{29}, 743 (1955) [Sov. Phys. JETP \textbf{2}, 646 (1956)]; M. S. Khaikin, Adv. Phys. \textbf{18}, 1 (1969).
7047: \bibitem{ref270} R. Vawter, Phys. Rev. \textbf{174}, 749 (1968); A. Isihara and K. Ebina, J. Phys. C \textbf{21}, L1079 (1988).
7048: \bibitem{ref271} I. B. Levinson, Zh. Eksp. Teor. Fiz. \textbf{95}, 2175 (1989) [Sov. Phys. JETP \textbf{68}, 1257 (1989)].
7049: \bibitem{ref272} M. C. Payne, J. Phys. Condens. Matter \textbf{1}, 4931 (1989); 4939 (1989).
7050: \bibitem{ref273} E. N. Economou and C. M. Soukoulis, Phys. Rev. Lett. \textbf{46}, 618 (1981).
7051: \bibitem{ref274} A. D. Stone and A. Szafer, IBM J. Res. Dev. \textbf{32}, 384 (1988).
7052: \bibitem{ref275} J. Kucera and P. Streda. J. Phys. C \textbf{21}, 4357 (1988).
7053: \bibitem{ref276} H. U. Baranger and A. D. Stone, Phys. Rev. B \textbf{40}, 8169 (1989).
7054: \bibitem{ref277} D. J. Thouless, Phys. Rev. Lett. \textbf{47}, 972 (1981).
7055: \bibitem{ref278} R. Landauer, Phys. Lett. A \textbf{85}, 91 (1981).
7056: \bibitem{ref279} E. N. Engquist and P. W. Anderson, Phys. Rev. B \textbf{24}, 1151 (1981).
7057: \bibitem{ref280} P. W. Anderson, D. J. Thouless, E. Abrahams, and D. S. Fisher, Phys. Rev. B \textbf{22}, 3519 (1980).
7058: \bibitem{ref281} D. C. Langreth and E. Abrahams, Phys. Rev. B \textbf{24}, 2978 (1981).
7059: \bibitem{ref282} M. Ya. Azbel, J. Phys. C \textbf{14}, L225 (1981).
7060: \bibitem{ref283} R. Landauer, J. Phys. Condens. Matter \textbf{1}, 8099 (1989); also in Ref. \protect\onlinecite{ref15}.
7061: \bibitem{ref284} M. B\"{u}ttiker, IBM J. Res. Dev. \textbf{32}, 317 (1988).
7062: \bibitem{ref285} H. B. G. Casimir, Rev. Mod. Phys. \textbf{17}, 343 (1945); Philips Res. Rep. \textbf{1}, 185 (1946); L. Onsager, Phys. Rev. \textbf{38}, 2265 (1931; see also S. R. de Groot and P. Mazur, {\em Non-Equilibrium Thermodynamics.} Dover, New York, 1984.
7063: \bibitem{ref286} A. D. Benoit, S. Washburn, C. P. Umbach, R. B. Laibowitz, and R. A. Webb, Phys. Rev. Lett. \textbf{57}, 1765 (1986).
7064: \bibitem{ref287} H. H. Sample, W. J. Bruno, S. B. Sample, and E. K. Sichel, J. Appl. Phys. \textbf{61}, 1079 (1987).
7065: \bibitem{ref288} L. L. Soethout, H. van Kempen, J. T. P. W. van Maarseveen, P. A. Schroeder, and P. Wyder, J. Phys. F \textbf{17}, L129 (1987).
7066: \bibitem{ref289} G. Timp, H. U. Baranger, P. deVegvar, J. E. Cunningham, R. E. Howard, R. Behringer, and P. M. Mankiewich, Phys. Rev. Lett. \textbf{60}, 2081 (1988).
7067: \bibitem{ref290} P. G. N. de Vegvar, G. Timp, P. M. Mankiewich, J. E. Cunningham, R. Behringer, and R. E. Howard, Phys. Rev. B \textbf{38}, 4326 (1988).
7068: \bibitem{ref291} A. I. Larkin and D. E. Khmel'nitskii, Zh. Eksp. Teor. Fiz. \textbf{91}, 1815 (1986) [Sov. Phys. JETP \textbf{64}, 1075 (1986)].
7069: \bibitem{ref292} P. H. M. van Loosdrecht, C. W. J. Beenakker, H. van Houten, J. G. Williamson, B. J. van
7070: Wees, J. E. Mooij, C. T. Foxon, and J. J. Harris, Phys. Rev. B \textbf{38}, 10162 (1988).
7071: \bibitem{ref293} S. Datta, Phys. Rev. B \textbf{40}, 5830 (1989).
7072: \bibitem{ref294} S. Feng. Phys. Lett. A \textbf{143}, 400 (1990).
7073: \bibitem{ref295} J. C. Maxwell, {\em A Treatise on Electricity and Magnetism.} Clarendon, Oxford, 1904.
7074: \bibitem{ref296} Yu. V. Sharvin, Zh. Eksp. Teor. Fiz. \textbf{48}, 984 (1965) [Sov. Phys. JETP 21, 655 (1965)].
7075: \bibitem{ref297} Yu. V. Sharvin and N. I. Bogatina, Zh. Eksp. Teor. Phys. \textbf{56}, 772 (1969) [Sov. Phys. JETP \textbf{29}, 419 (1969)].
7076: \bibitem{ref298} J. K. Gimzewski and R. M\"{o}ller, Phys. Rev. B \textbf{36}, 1284 (1987).
7077: \bibitem{ref299} N. D. Lang, Phys. Rev. B \textbf{36}, 8173 (1987).
7078: \bibitem{ref300} J. Ferrer, A. Martin-Rodero, and F. Flores, Phys. Rev. B \textbf{38}, 10113 (1988).
7079: \bibitem{ref301}N. D. Lang, Comm. Cond. Matt. Phys. \textbf{14}, 253 (1989).
7080: \bibitem{ref302} N. D. Lang, A. Yacoby, and Y. Imry, Phys. Rev. Lett. \textbf{63}, 1499 (1989).
7081: \bibitem{ref303} N. Garcia and H. Rohrer, J. Phys. Condens. Matter \textbf{1}, 3737 (1989).
7082: \bibitem{ref304} R. Trzcinski, E. Gmelin, and H. J. Queisser, Phys. Rev. B \textbf{35}, 6373 (1987).
7083: \bibitem{ref305} A. G. M. Jansen, A. P. van Gelder, and P. Wyder, J. Phys. C \textbf{13}, 6073 (1980).
7084: \bibitem{ref306} G. Timp, R. Behringer, S. Sampere, J. E. Cunningham, and R. E. Howard, in Ref. \protect\onlinecite{ref15}; see also G. Timp in Ref. \protect\onlinecite{ref9}.
7085: \bibitem{ref307} H. van Houten, C. W. J. Beenakker, and B. J. van Wees, in Ref. \protect\onlinecite{ref9}.
7086: \bibitem{ref308} B. J. van Wees, L. P. Kouwenhoven, E. M. M. Willems, C. J. P. M. Harmans, J. E. Mooij, H. van Houten, C. W. J. Beenakker, J. G. Williamson, and C. T. Foxon, submitted to Phys. Rev. B.
7087: \bibitem{ref309} P. F. Bagwell and T. P. Orlando, Phys. Rev. B \textbf{40}, 1456 (1989).
7088: \bibitem{ref310} G. Timp, in Ref. \protect\onlinecite{ref10}.
7089: \bibitem{ref311} H. van Houten and C. W. J. Beenakker, in Ref. \protect\onlinecite{ref15}.
7090: \bibitem{ref312} L. Escapa and N. Garcia, J. Phys. Condens. Matter \textbf{1}, 2125 (1989).
7091: \bibitem{ref313} E. G. Haanappel and D. van der Marel, Phys. Rev. B \textbf{39}, 5484 (1989); D. van der Marel and E. G. Haanappel, Phys. Rev. B \textbf{39}, 7811 (1989).
7092: \bibitem{ref314} G. Kirczenow, Solid State Comm. \textbf{68}, 715 (1988); J. Phys. Condens. Matter \textbf{1}, 305 (1989).
7093: \bibitem{ref315} A. Szafer and A. D. Stone, Phys. Rev. Lett. \textbf{62}, 300 (1989).
7094: \bibitem{ref316} E. Tekman and S. Ciraci, Phys. Rev. B \textbf{39}, 8772 (1989); Phys. Rev. B \textbf{40}, 8559 (1989).
7095: \bibitem{ref317} Song He and S. Das Sarma, Phys. Rev. B \textbf{40}, 3379 (1989).
7096: \bibitem{ref318} D. van der Marel, in Ref. \protect\onlinecite{ref15}.
7097: \bibitem{ref319} N. Garcia and L. Escapa, Appl. Phys. Lett. \textbf{54}, 1418 (1989).
7098: \bibitem{ref320} E. Casta\~{n}o and G. Kirczenow, Solid State Comm. \textbf{70}, 801 (1989).
7099: \bibitem{ref321} Y. Avishai and Y. B. Band, Phys. Rev. B \textbf{40}, 12535 (1989).
7100: \bibitem{ref322} A. Kawabata, J. Phys. Soc. Japan \textbf{58}, 372 (1989).
7101: \bibitem{ref323} I. B. Levinson, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{48}, 273 (1988) [JETP Lett. \textbf{48}, 301 (1988)].
7102: \bibitem{ref324} A. Matulis and D. Segzda, J. Phys. Condens. Matter \textbf{1}, 2289 (1989).
7103: \bibitem{ref325} L. I. Glazman, G. B. Lesovick, D. E. Khmel'nitskii, R. I. Shekhter, Pis'ma Zh. Teor. Fiz. \textbf{48}, 218 (1988) [JETP Lett. \textbf{48}, 238 (1988)].
7104: \bibitem{ref326} A. Yacoby and Y. Imry, Phys. Rev. B \textbf{41}, 5341 (1990).
7105: \bibitem{ref327} L. W. Molenkamp, A. A. M. Staring, C. W. J. Beenakker, R. Eppenga, C. E. Timmering, J. G.
7106: Williamson, C. J. P. M. Harmans, and C. T. Foxon, Phys. Rev. B \textbf{41}, 1274 (1990).
7107: \bibitem{ref328} R. Landauer, Z. Phys. B \textbf{68}, 217 (1987).
7108: \bibitem{ref329} C. W. J. Beenakker and H. van Houten, Phys. Rev. B \textbf{39}, 10445 (1989).
7109: \bibitem{ref330} Y. Hirayama, T. Saku, and Y. Horikoshi, Phys. Rev. B \textbf{39}, 5535 (1989).
7110: \bibitem{ref331} Y. Hirayama, T. Saku, and Y. Horikoshi, Jap. J. Appl. Phys. \textbf{28}, L701 (1989).
7111: \bibitem{ref332} R. J. Brown, M. J. Kelly, R. Newbury, M. Pepper, B. Miller, H. Ahmed, D. G. Hasko, D. C. Peacock, D. A. Ritchie, J. E. F. Frost, and G. A. C. Jones, Solid State Electron. \textbf{32}, 1179 (1989).
7112: \bibitem{ref333} L. Martin-Moreno and C. G. Smith, J. Phys. Condens. Matter \textbf{1}, 5421 (1989).
7113: \bibitem{ref334} B. J. van Wees, L. P. Kouwenhoven, H. van Houten, C. W. J. Beenakker, J. E. Mooij, C. T. Foxon, and J. J. Harris, Phys. Rev. B \textbf{38}, 3625 (1988).
7114: \bibitem{ref335} M. B\"{u}ttiker, Phys. Rev. B \textbf{41}, 7906 (1990); L. I. Glazman and A. V. Khaetskii, J. Phys. Condens. Matter \textbf{1}, 5005 (1989); Y. Avishai and Y. B. Band, Phys. Rev. B \textbf{40}, 3429 (1989); K. B. Efetov, J. Phys. Condens. Matter \textbf{1}, 5535 (1989).
7115: \bibitem{ref336} J. F. Weisz and K.-F. Berggren, Phys. Rev. B \textbf{40}, 1325 (1989).
7116: \bibitem{ref337} D. A. Wharam, U. Ekenberg, M. Pepper, D. G. Hasko, H. Ahmed, J. E. F. Frost, D. A. Ritchie, D. C. Peacock, and G. A. C. Jones, Phys. Rev. B \textbf{39}, 6283 (1989).
7117: \bibitem{ref338} H. Hirai, S. Komiyama, S. Sasa and T. Fujii, Solid State Comm. \textbf{72}, 1033 (1989).
7118: \bibitem{ref339} S. Washburn, A. B. Fowler, H. Schmid, and D. Kern, Phys. Rev. Lett. \textbf{61}, 2801 (1988).
7119: \bibitem{ref340} R. J. Haug, A. H. MacDonald, P. Streda, and K. von Klitzing, Phys. Rev. Lett. \textbf{61}, 2797 (1988).
7120: \bibitem{ref341} R. J. Haug, J. Kucera, P. Streda, and K. von Klitzing, Phys. Rev. B \textbf{39}, 10892 (1989).
7121: \bibitem{ref342} B. R. Snell, P. H. Beton, P. C. Main, A. Neves, J. R. Owers-Bradley, L. Eaves, M. Henini, O. H.
7122: Hughes, S. P. Beaumont, and C. D. W. Wilkinson, J. Phys. Condens. Matter \textbf{1}, 7499 (1989).
7123: \bibitem{ref343} Y. S. Tsoi, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{19}, 114 (1974) [JETP Lett. \textbf{19}, 70 (1974)]; Zh. Eksp. Teor. Fiz. \textbf{68}, 1849 (1975) [Sov. Phys. JETP \textbf{41}, 927 (1975)].
7124: \bibitem{ref344} P. C. van Son, H. van Kempen, and P. Wyder, Phys. Rev. Lett. \textbf{58}, 1567 (1987).
7125: \bibitem{ref345} I. K. Yanson, Zh. Eksp. Teor. Fiz. \textbf{66}, 1035 (1974) [Sov. Phys. JETP \textbf{39}, 506 (1974)].
7126: \bibitem{ref346} A. M. Duif, A. G. M. Jansen, and P. Wyder, J. Phys. Condens. Matter \textbf{1}, 3157 (1989).
7127: \bibitem{ref347} C. W. J. Beenakker, H. van Houten, and B. 1. van Wees, Europhys. Lett. \textbf{7}, 359 (1988).
7128: \bibitem{ref348} C. W. J. Beenakker, H. van Houten, and B. J. van Wees, Festk\"{o}rperprobleme \textbf{29}, 299 (1989).
7129: \bibitem{ref349} J. Spector, H. L. Stormer, K. W. Baldwin, L. N. Pfeiffer, and K. W. West, Surf Sci. \textbf{228}, 283 (1990).
7130: \bibitem{ref350} U. Sivan, M. Heiblum, and C. P. Umbach, and H. Shtrikman, Phys. Rev. B \textbf{41}, 7937 (1990); J. Spector, H. L. Stormer, K. W. Baldwin, L. N. Pfeiffer and K. W. West, Appl. Phys. Lett. \textbf{56}, 1290 (1990).
7131: \bibitem{ref351} P. A. M. Benistant, Ph.D. thesis, University of Nijmegen, The Netherlands, 1984; P. A. M. Benistant, A. P. van Gelder, H. van Kempen, and P. Wyder, Phys. Rev. B \textbf{32}, 3351 (1985).
7132: \bibitem{ref352} P. A. M. Benistant, G. F. A. van de Walle, H. van Kempen, and P. Wyder, Phys. Rev. B \textbf{33}, 690 (1986).
7133: \bibitem{ref353} L. Pfeiffer, K. W. West, H. L. Stormer, and K. W. Baldwin, Appl. Phys. Lett. \textbf{55}, 1888 (1989).
7134: \bibitem{ref354} C. T. Foxon, J. J. Harris, D. Hilton, J. Hewett, and C. Roberts, Semicond. Sci. Technol. \textbf{4}, 582 (1989); C. T. Foxon and J. J. Harris, Philips J. Res. \textbf{41}, 313 (1986).
7135: \bibitem{ref355} J. G. Williamson, H. van Houten, C. W. J. Beenakker, M. E. J. Broekaart, L. J. A. Spendeler, B. J. van Wees, and C. T. Foxon, Phys. Rev. B \textbf{41}, 1207 (1990); Surf Sci. \textbf{229}, 303 (1990).
7136: \bibitem{ref356} R. Landauer, in {\em Analogies in Optics and Microelectronics} (W. van Haeringen and D. Lenstra, eds.). Kluwer Academic, Dordrecht, 1990.
7137: \bibitem{ref357} D. A. Wharam, M. Pepper, H. Ahmed, J. E. F. Frost, D. G. Hasko, D. C. Peacock, D. A.
7138: Ritchie, and G. A. C. Jones, J. Phys. C \textbf{21}, L887 (1988).
7139: \bibitem{ref358} H. U. Baranger and A. D. Stone, Phys. Rev. Lett. \textbf{63}, 414 (1989); also in Ref. \protect\onlinecite{ref16}.
7140: \bibitem{ref359} C. W. J. Beenakker and H. van Houten, Phys. Rev. Lett. \textbf{63}, 1857 (1989).
7141: \bibitem{ref360} C. W. J. Beenakker and H. van Houten, in Ref. \protect\onlinecite{ref17}.
7142: \bibitem{ref361} L. D. Landau and E. M. Lifshitz, {\em Mechanics.} Pergamon, Oxford, 1976.
7143: \bibitem{ref362} N. S. Kapany, in {\em Concepts of Classical Optics} (J. Strong, ed.). Freeman, San Francisco, 1958.
7144: \bibitem{ref363} H. de Raedt, N. Garcia, and J. J. Saenz, Phys. Rev. Lett. \textbf{63}, 2260 (1989); N. Garcia, J. J. Saenz, and H. de Raedt, J. Phys. Condens. Matter \textbf{1}, 9931 (1989).
7145: \bibitem{ref364} Y. Takagaki, K. Garno, S. Namba, S. Ishida, S. Takaoka, K. Murase, K. Ishibashi, and Y. Aoyagi, Solid State Comm. \textbf{68}, 1051 (1988); \textbf{71}, 809 (1989).
7146: \bibitem{ref365} Y. Hirayama and T. Saku, Solid State Comm. \textbf{73}, 113 (1990); Phys. Rev. B \textbf{41}, 2927 (1990).
7147: \bibitem{ref366} P. H. Beton, B. R. Snell, P. C. Main, A. Neves, J. R. Owers-Bradley, L. Eaves, M. Henini, O. H. Hughes, S. P. Beaumont, and C. D. W. Wilkinson, J. Phys. Condens. Matter \textbf{1}, 7505 (1989).
7148: \bibitem{ref367} E. Casta\~{n}o and G. Kirczenow, Phys. Rev. B. \textbf{41}, 5055 (1990). Y. Avishai, M. Kaveh, S. Shatz, and Y. B. Band, J. Phys. Condens. Matter \textbf{1}, 6907 (1989).
7149: \bibitem{ref368} C. G. Smith, M. Pepper, R. Newbury, H. Ahmed, D. G. Hasko, D. C. Peacock, J. E. F. Frost, D. A. Ritchie, G. A. C. Jones, and G. Hill, J. Phys. Condens. Matter \textbf{1}, 6763 (1989).
7150: \bibitem{ref369} Y. Hirayama and T. Saku, Jap. J. Appl. Phys. (to be published).
7151: \bibitem{ref370} M. B\"{u}ttiker, Phys. Rev. B \textbf{33}, 3020 (1986).
7152: \bibitem{ref371} M. B\"{u}ttiker, IBM J. Res. Dev. \textbf{32}, 63 (1988).
7153: \bibitem{ref372} A. A. M. Staring, L. W. Molenkamp, C. W. J. Beenakker, L. P. Kouwenhoven, and C. T. Foxon, Phys. Rev. B. \textbf{41}, 8461 (1990).
7154: \bibitem{ref373} L. P. Kouwenhoven, B. J, van Wees, W, Kool, C. J. P, M. Harmans, A, A. M. Staring, and C. T. Foxon, Phys. Rev, B \textbf{40}, 8083 (1989).
7155: \bibitem{ref374} P. C. Main, P, H. Beton, B. R. Snell, A. J. M. Neves, J. R. Owers-Bradley, L. Eaves, S. P. Beaumont, and C. D. W. Wilkinson, Phys. Rev. B \textbf{40}, 10033 (1989).
7156: \bibitem{ref375} A. M. Chang, T. Y. Chang, and H. U. Baranger, Phys. Rev. Lett. 63, 996 (1989).
7157: \bibitem{ref376} D. G. Ravenhall, H. W. Wyld, and R. L. Schult, Phys. Rev. Lett. \textbf{62}, 1780 (1989); R. L. Schult,
7158: H. W. Wyld, and D. G. Ravenhall, Phys. Rev. B. \textbf{41}, 12760 (1990).
7159: \bibitem{ref377} G. Kirczenow, Phys. Rev. Lett. \textbf{62}, 2993 (1989); Phys. Rev. B \textbf{42}, 5375 (1990).
7160: \bibitem{ref378} H. Akera and T. Ando, Surf Sci. \textbf{229}, 268 (1990).
7161: \bibitem{ref379} C. J. B. Ford, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, D. C. Peacock, D. A.
7162: Ritchie, J. E. F. Frost, and G. A. C. Jones, Phys. Rev. B \textbf{38}, 8518 (1988).
7163: \bibitem{ref380} Y. Avishai and Y. B. Band, Phys. Rev. Lett. \textbf{62}, 2527 (1989).
7164: \bibitem{ref381} G. Kirczenow, Solid State Comm. \textbf{71}, 469 (1989).
7165: \bibitem{ref382} F. M. Peeters, in Ref. \protect\onlinecite{ref16}; Phys. Rev. Lett. \textbf{61}, 589 (1988); Superlattices and Microstructures \textbf{6}, 217 (1989).
7166: \bibitem{ref383} M. L. Roukes, T. J. Thornton, A. Scherer, J. A. Simmons, B. P. van der Gaag, and E. D. Beebe, in Ref. \protect\onlinecite{ref16}.
7167: \bibitem{ref384} M. L. Roukes, A. Scherer, and B. P. van der Gaag, Phys. Rev. Lett. \textbf{64}, 1154 (1990).
7168: \bibitem{ref385} C. J. B. Ford. S. Washburn, M. B\"{u}ttiker, C. M. Knoedler, and J. M. Hong, Surf Sci. \textbf{229}, 298 (1990).
7169: \bibitem{ref386} M. B\"{u}ttiker, in Ref. \protect\onlinecite{ref9}.
7170: \bibitem{ref387} R. Tsu and L. Esaki, Appl. Phys. Lett. \textbf{22}, 562 (1973).
7171: \bibitem{ref388} L. L. Chang, L. Esaki, and R. Tsu, Appl. Phys. Lett. \textbf{24}, 593 (1974).
7172: \bibitem{ref389} E. S. Alves, L. Eaves, M. Henini, O. H. Hughes, M. L. Leadbeater, F. W. Sheard, and G. A. Toombs, Electron. Lett. \textbf{24}, 1190 (1988).
7173: \bibitem{ref390} A. Zaslavsky, V. J. Goldman, D. C. Tsui, and J. E. Cunningham, Appl. Phys. Lett. \textbf{53}, 1408
7174: (1988).
7175: \bibitem{ref391} K. K. Likharev, IBM J. Res. Dev. \textbf{32}, 144 (1988).
7176: \bibitem{ref392} K. Ng and P. A. Lee, Phys. Rev. Lett. \textbf{61}, 1768 (1988).
7177: \bibitem{ref393} L. I. Glazman and K. A. Matveev, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{48}, 403 (1988) [JETP Lett. \textbf{48}, 445 (1988)].
7178: \bibitem{ref394} S. Washburn, A. B. Fowler, H. Schmid, and D. Kern, Phys. Rev. B \textbf{38}, 1554 (1988).
7179: \bibitem{ref395} J. H. Davies, Semicond. Sci. Technol. \textbf{3}, 995 (1988). See also in Ref. \protect\onlinecite{ref72}.
7180: \bibitem{ref396} S. Y. Chou, D. R. Allee, R. F. W. Pease, and J. S. Harris, Jr., Appl. Phys. Lett. \textbf{55}, 176 (1989).
7181: \bibitem{ref397} A. Palevski, M. Heiblum, C. P. Umbach, C. M. Knoedler, A. N. Broers, and R. H. Koch, Phys.
7182: Rev. Lett. \textbf{62}, 1776 (1989).
7183: \bibitem{ref398} Y. Avishai and Y. B. Band, Phys. Rev. B \textbf{41}, 3253 (1990).
7184: \bibitem{ref399} C. G. Smith, M. Pepper, H. Ahmed, J. E. Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, Superlattices and Microstructures \textbf{5}, 599 (1989).
7185: \bibitem{ref400} C. G. Smith, M. Pepper, H. Ahmed, J. E. F. Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, J. Phys. C \textbf{21}, L893 (1988).
7186: \bibitem{ref401} C. G. Smith, M. Pepper, H. Ahmed, J. E. F. Frost, D. G. Hasko, D. A. Ritchie, and G. A. C. Jones, Surf Sci. \textbf{228}, 387 (1990).
7187: \bibitem{ref402} S. J. Bending and M. R. Beasley, Phys. Rev. Lett. \textbf{55}, 324 (1985).
7188: \bibitem{ref403} A. B. Fowler, G. L. Timp, J. J. Wainer, and R. A. Webb, Phys. Rev. Lett. \textbf{57}, 138 (1986).
7189: \bibitem{ref404} T. E. Kopley, P. L. McEuen and R. G. Wheeler, Phys. Rev. Lett. 61, 1654 (1988); see also T. E.
7190: Kopley, Ph.D. thesis, Yale University, 1989.
7191: \bibitem{ref405} W. Xue and P. A. Lee, Phys. Rev. B \textbf{38}, 3913 (1988).
7192: \bibitem{ref406} V. Kalmeyer and R. B. Laughlin, Phys. Rev. B \textbf{35}, 9805 (1987).
7193: \bibitem{ref407} C. S. Chu and R. S. Sorbello, Phys. Rev. B \textbf{40}, 5941 (1989).
7194: \bibitem{ref408} J. Masek, P. Lipavsky, and B. Kramer, J. Phys. Condens. Matter \textbf{1}, 6395 (1989).
7195: \bibitem{ref409} P. L. McEuen, B. W. Alphenaar, R. G. Wheeler, and R. N. Sacks, Surf Sci. \textbf{229}, 312 (1990).
7196: \bibitem{ref410} G. J. Schulz, Rev. Mod. Phys. \textbf{45}, 378 (1973).
7197: \bibitem{ref411} C. G. Smith, M. Pepper, J. E. Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, J. Phys. Condens. Matter \textbf{1}, 9035 (1989).
7198: \bibitem{ref412} Y. Avishai, M. Kaveh, and Y. B. Band, preprint.
7199: \bibitem{ref413} F. Sols, M. Macucci, U. Ravioli, and K. Hess, Appl. Phys. Lett. \textbf{54}, 350 (1989).
7200: \bibitem{ref414} S. Datta, Superlattices and Microstructures \textbf{6}, 83 (1989).
7201: \bibitem{ref415} D. S. Miller, R. K. Lake, S. Datta, M. S. Lundstrom, and R. Reifenberger, in Ref. \protect\onlinecite{ref15}.
7202: \bibitem{ref416} J. H. F. Scott-Thomas, S. B. Field, M. A. Kastner, H. I. Smith, and D. A. Antoniadis, Phys. Rev. Lett. \textbf{62}, 583 (1989).
7203: \bibitem{ref417} A. I. Larkin and P. A. Lee, Phys. Rev. B \textbf{17}, 1596 (1978).
7204: \bibitem{ref418} P. A. Lee and T. M. Rice, Phys. Rev. B \textbf{19}, 3970 (1979).
7205: \bibitem{ref419} H. van Houten and C. W. J. Beenakker, Phys. Rev. Lett. \textbf{63}, 1893 (1989).
7206: \bibitem{ref420} K. Mullen, E. Ben-Jacob, R. C. Jaclevic, and Z. Schuss, Phys. Rev. B \textbf{37}, 98 (1988); M. Amman, K. Mullen, and E. Ben-Jacob, J. Appl. Phys. \textbf{65}, 339 (1989).
7207: \bibitem{ref421} M. A. Kastner, S. B. Field, U. Meirav, J. H. F. Scott-Thomas, D. A. Antoniadis, and M. I.
7208: Smith, Phys. Rev. Lett. \textbf{63}, 1894 (1989).
7209: \bibitem{ref422} L. I. Glazman and R. I. Shekhter, J. Phys. Condens. Matter \textbf{1}, 5811 (1989).
7210: \bibitem{ref423} R. J. Brown, M. Pepper, H. Ahmed, D. G. Hasko, R. A. Ritchie, J. E. F. Frost, D. C. Peacock,
7211: and G. A. C. Jones, J. Phys. Condens. Matter, \textbf{2}, 2105 (1990).
7212: \bibitem{ref424} L. P. Kouwenhoven, private communication; R. Haug, private communication.
7213: \bibitem{ref425} U. Meirav, M. A. Kastner, and S. J. Wind, Phys. Rev. Lett. \textbf{65}, 771 (1990).
7214: \bibitem{ref426} B. J. van Wees, E. M. M. Willems, C. J. P. M. Harmans, C. W. J. Beenakker, H. van Houten, J.
7215: G. Williamson, C. T. Foxon, and J. J. Harris, Phys. Rev. Lett. \textbf{62}, 1181 (1989).
7216: \bibitem{ref427} S. Komiyama, H. Hirai, S. Sasa, and S. Hiyamizu, Phys. Rev. B \textbf{40}, 12566 (1989).
7217: \bibitem{ref428} B. J. van Wees, E. M. M. Willems, L. P. Kouwenhoven, C. J. P. M. Harmans, J. G. Williamson, C. T. Foxon, and J. 1. Harris, Phys. Rev. B \textbf{39}, 8066 (1989).
7218: \bibitem{ref429} B. W. Alphenaar, P. L. McEuen, R. G. Wheeler, and R. N. Sacks, Phys. Rev. Lett. \textbf{64}, 677 (1990).
7219: \bibitem{ref430} R. J. Haug and K. von Klitzing, Europhys. Lett. \textbf{10}, 489 (1989).
7220: \bibitem{ref431} R. Kubo, S. J. Miyake, and N. Hashitsume, {\em Solid State Physics,} Vol. 17 (F. Seitz and D. Turnbull, eds.). Academic, New York, 1965. M. Tsukada, J. Phys. Soc. Jap., \textbf{41}, 1466 (1976).
7221: \bibitem{ref432} R. F. Kazarinov and S. Luryi, Phys. Rev. B \textbf{25}, 7626 (1982); S. Luryi and R. F. Kazarinov, Phys. Rev. B \textbf{27}, 1386 (1983); S. Luryi, in {\em High Magnetic Fields in Semiconductor Physics} (G. Landwehr, ed.). Springer, Berlin, 1987.
7222: \bibitem{ref433} S. V. Iordansky, Solid State Comm. \textbf{43}, 1 (1982).
7223: \bibitem{ref434} S. A. Trugman, Phys. Rev. B \textbf{27}, 7539 (1983).
7224: \bibitem{ref435} R. Joynt and R. E. Prange, Phys. Rev. B \textbf{29}, 3303 (1984).
7225: \bibitem{ref436} R. E. Prange, in Ref. \protect\onlinecite{ref97}.
7226: \bibitem{ref437} L. I. Glazman and M. Jonson, J. Phys. Condens. Matter \textbf{1}, 5547 (1989).
7227: \bibitem{ref438} L. I. Glazman and M. Jonson, Phys. Rev. B \textbf{41}, 10686 (1990).
7228: \bibitem{ref439} T. Martin and S. Feng, Phys. Rev. Lett. \textbf{64}, 1971 (1990).
7229: \bibitem{ref440} B. I. Halperin, Phys. Rev. B \textbf{25}, 2185 (1982).
7230: \bibitem{ref441} A. H. MacDonald and P. Streda, Phys. Rev. B \textbf{29}, 1616 (1984).
7231: \bibitem{ref442} S. M. Apenko and Yu. E. Lozovik, J. Phys. C \textbf{18}, 1197 (1985).
7232: \bibitem{ref443} P. Streda, J. Kucera, and A. H. MacDonald, Phys. Rev. Lett. \textbf{59}, 1973 (1987).
7233: \bibitem{ref444} J. K. Jain and S. A. Kivelson, Phys. Rev. B \textbf{37}, 4276 (1988).
7234: \bibitem{ref445} M. E. Cage, in Ref. \protect\onlinecite{ref97}.
7235: \bibitem{ref446} A. H. MacDonald, T. M. Rice, and W. F. Brinkman, Phys. Rev. B \textbf{28}, 3648 (1983).
7236: \bibitem{ref447} O. Heinonen and P. L. Taylor, Phys. Rev. B \textbf{32}, 633 (1985).
7237: \bibitem{ref448} D. J. Thouless, J. Phys. C \textbf{18}, 6211 (1985).
7238: \bibitem{ref449} Y. M. Pudalov and S. G. Semenchinskii, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{42}, 188 (1985) [JETP
7239: Lett. \textbf{42}, 232 (1985)].
7240: \bibitem{ref450} W. Maass, Europhys. Lett. \textbf{2}, 39 (1986).
7241: \bibitem{ref451} Y. Ono and T. Ohtsuki, Z. Phys. B \textbf{68}, 445 (1987); T. Ohtsuki and Y. Ono, J. Phys. Soc. Jap.
7242: \textbf{58}, 2482 (1989).
7243: \bibitem{ref452} R. Johnston and L. Schweitzer, Z. Phys. B \textbf{70}, 25 (1988).
7244: \bibitem{ref453} Y. Gudmundsson, R. R. Gerhardts, R. Johnston, and L. Schweitzer, Z. Phys. B \textbf{70}, 453 (1988).
7245: \bibitem{ref454} T. Ando, J. Phys. Soc. Jap. \textbf{58}, 3711 (1989).
7246: \bibitem{ref455} P. C. van Son, G. H. Kruithof, and T. M. Klapwijk, Surf Sci. \textbf{229}, 57 (1990); P. C. van Son and
7247: T. M. Klapwijk, Europhys. Lett. \textbf{12}, 429 (1990).
7248: \bibitem{ref456} G. Ebert, K. von Klitzing, and G. Weimann, J. Phys. C \textbf{18}, L257 (1985).
7249: \bibitem{ref457} H. Z. Zheng, D. C. Tsui, and A. M. Chang, Phys. Rev. B \textbf{32}, 5506 (1985).
7250: \bibitem{ref458} E. K. Sichel, H. H. Sample, and J. P. Salerno, Phys. Rev. B \textbf{32}, 6975 (1987); E. K. Sichel, M. L. Knowles, and H. H. Sample, J. Phys. C \textbf{19}, 5695 (1986).
7251: \bibitem{ref459} R. Woltjer, R. Eppenga, J. Mooren, C. E. Timmering, and J. P. Andr\'{e}, Europhys. Lett. \textbf{2}, 149 (1986).
7252: \bibitem{ref460} B. E. Kane, D. C. Tsui, and G. Weimann, Phys. Rev. Lett. \textbf{59}, 1353 (1987).
7253: \bibitem{ref461} M. B\"{u}ttiker, in Ref. \protect\onlinecite{ref15}.
7254: \bibitem{ref462} F. Kuchar, Festk\"{o}rperprobleme \textbf{28}, 45 (1988).
7255: \bibitem{ref463} P. F. Fontein, J. A. Kleinen, P. Hendriks, F. A. P. Blom, J. H. Wolter, H. G. M. Locks, F. A. J. M. Driessen, L. J. Giling, and C. W. J. Beenakker, submitted to Phys. Rev. B.
7256: \bibitem{ref464} C. W. J. Beenakker, unpublished.
7257: \bibitem{ref465} L. P. Kouwenhoven, Master's thesis, Delft University of Technology, 1988.
7258: \bibitem{ref466} K. yon Klitzing, G. Ebert, N. Kleinmichel, H. Obloh, G. Dorda, and G. Weimann, {\em Proc. ICPS 17} (J. D. Chadi and W. A. Harrison, eds.). Springer, New York, 1985.
7259: \bibitem{ref467} D. A. Syphers, F. F. Fang, and P. J. Stiles, Surf Sci. \textbf{142}, 208 (1984); F. F. Fang and P. J. Stiles, Phys. Rev. B \textbf{27}, 6487 (1983); F. F. Fang and P. J. Stiles, Phys. Rev. B \textbf{29}, 3749 (1984). A. B. Berkut, Yu. V. Dubrovskii, M. S. Nunuparov, M. I. Reznikov, and V. I. Tal'yanski, Pis'ma Zh. Teor. Fiz. \textbf{44}, 252 (1986) [JETP Lett. \textbf{44}, 324 (1986)].
7260: \bibitem{ref468} D. A. Syphers and P. J. Stiles, Phys. Rev. B \textbf{32}, 6620 (1985).
7261: \bibitem{ref469} Y. Zhu, J. Shi, and S. Feng, Phys. Rev. B \textbf{41}, 8509 (1990).
7262: \bibitem{ref470} H. Hirai, S. Komiyama, S. Hiyamizu, and S. Sasa, in {\em Proc. ICPS 19,} p. 55 (W. Zawadaski, ed.). Institute of Physics, Polish Academy of Sciences, 1988.
7263: \bibitem{ref471} S. Komiyama, H. Hirai, S. Sasa, and T. Fuji, Solid State Comm. \textbf{73}, 91 (1990); H. Hirai, S. Komiyama, S. Sasa, and T. Fujii, J. Phys. Soc. Japan \textbf{58}, 4086 (1989).
7264: \bibitem{ref472} S. Komiyama and H. Hirai, Phys. Rev. B \textbf{40}, 7767 (1989).
7265: \bibitem{ref473} U. Sivan, Y. Imry, and C. Hartzstein, Phys. Rev. B \textbf{39}, 1242 (1989).
7266: \bibitem{ref474} U. Sivan and Y. Imry, Phys. Rev. Lett. \textbf{61}, 1001 (1988).
7267: \bibitem{ref475} M. B\"{u}ttiker, Phys. Rev. B \textbf{38}, 12724 (1988).
7268: \bibitem{ref476} J. K. Jain, unpublished.
7269: \bibitem{ref477} P. L. McEuen, A. Szafer, C. A. Richter, B. W. Alphenaar, J. K. Jain, and R. N. Sacks, Phys. Rev. Lett. \textbf{64}, 2062 (1990).
7270: \bibitem{ref478} D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys. Rev. Lett. \textbf{48}, 1559 (1982).
7271: \bibitem{ref479} R. B. Laughlin, Phys. Rev. Lett. \textbf{50}, 1395 (1983).
7272: \bibitem{ref480} A. M. Chang and J. E. Cunningham, Solid State Comm. \textbf{72}, 651 (1989); Surf Sci. \textbf{229}, 216 (1990).
7273: \bibitem{ref481} L. P. Kouwenhoven, B. J. van Wees, N. C. van der Vaart, C. J. P. M. Harmans, C. E.
7274: Timmering, and C. T. Foxon, Phys. Rev. Lett. \textbf{64}, 685 (1990); and unpublished.
7275: \bibitem{ref482} C. W. J. Beenakker, Phys. Rev. Lett. \textbf{64}, 216 (1990).
7276: \bibitem{ref483} A. H. MacDonald, Phys. Rev. Lett. \textbf{64}, 220 (1990).
7277: \bibitem{ref484} T. Chakraborty and P. Pietil\"{a}inen, {\em The Fractional Quantum Hall Effect.} Springer, Berlin,
7278: 1988.
7279: \bibitem{ref485} R. B. Laughlin, Phys. Rev. B \textbf{27}, 3383 (1983).
7280: \bibitem{ref486} F. D. M. Haldane, Phys. Rev. Lett. \textbf{51}, 605 (1983).
7281: \bibitem{ref487} B. I. Halperin, Phys. Rev. Lett. \textbf{52}, 1583 (1984).
7282: \bibitem{ref488} R. B. Laughlin, Phys. Rev. B \textbf{23}, 5632 (1981).
7283: \bibitem{ref489} B. I. Halperin, Helv. Phys. Acta \textbf{56}, 75 (1983).
7284: \bibitem{ref490} A. M. Chang, Solid State Comm. \textbf{74}, 871 (1990).
7285: \bibitem{ref491} G. Timp, R. E. Behringer, J. E. Cunningham, and R. E. Howard, Phys. Rev. Lett. \textbf{63}, 2268
7286: (1989); G. Timp, in Ref. \protect\onlinecite{ref9}.
7287: \bibitem{ref492} S. T. Chui, Phys. Rev. Lett. \textbf{56}, 2395 (1986); Phys. Rev. B \textbf{36}, 2806 (1987).
7288: \bibitem{ref493} H. W. Jiang, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer, and K. W. West, Phys. Rev. B \textbf{40}, 12013 (1989).
7289: \bibitem{ref494} R. G. Clark, J. R. Mallett, S. R. Haynes, J. J. Harris, and C. T. Foxon, Phys. Rev. Lett. \textbf{60}, 1747 (1988).
7290: \bibitem{ref495} S. A. Kivelson and V. L. Pokrovsky, Phys. Rev. B \textbf{40}, 1373 (1989).
7291: \bibitem{ref496} J. A. Simmons, H. P. Wei, L. W. Engel, D. C. Tsui, and M. Shayegan, Phys. Rev. Lett. \textbf{63}, 1731 (1989).
7292: \bibitem{ref497} G. Timp, P. M. Mankiewich, P. DeVegvar, R. Behringer, J. E. Cunningham, R. E. Howard, H. U. Baranger, and J. K. Jain, Phys. Rev. B \textbf{39}, 6227 (1989).
7293: \bibitem{ref498} J. K. Jain, Phys. Rev. Lett. \textbf{60}, 2074 (1988).
7294: \bibitem{ref499} J. K. Jain and S. Kivelson, Phys. Rev. B \textbf{37}, 4111 (1988).
7295: \bibitem{ref500} B. J. van Wees, L. P. Kouwenhoven, C. J. P. M. Harmans, J. G. Williamson, C. E. T. Timmering, M. E. I. Broekaart, C. T. Foxon, and J. J. Harris, Phys. Rev. Lett. \textbf{62}, 2523 (1989).
7296: \bibitem{ref501} E. N. Bogachek and G. A. Gogadze, Zh. Eksp. Teor. Fiz. \textbf{63}, 1839 (1972) [Sov. Phys. JETP \textbf{36}, 973 (1973)].
7297: \bibitem{ref502} N. B. Brandt, D. V. Gitsu, A. A. Nikolaevna, and Va. G. Ponomarev, Zh. Eksp. Teor. Fiz. \textbf{72}, 2332 (1977) [Sov. Phys. JETP \textbf{45}, 1226 (1977)]; N. B. Brandt, D. B. Gitsu, V. A. Dolma, and Va. G. Ponomarev, Zh. Eksp. Teor. Fiz. \textbf{92}, 913 (1987) [Sov. Phys. JETP \textbf{65}, 515 (1987)].
7298: \bibitem{ref503} Y. Isawa, Surf Sci. \textbf{170}, 38 (1986).
7299: \bibitem{ref504} D. A. Wharam, M. Pepper, R. Newbury, H. Ahmed, D. G. Hasko, D. C. Peacock, J. E. F. Frost, D. A. Ritchie, and G. A. C. Jones, J. Phys. Condens. Matter \textbf{1}, 3369 (1989).
7300: \bibitem{ref505} R. J. Brown, C. G. Smith, M. Pepper, M. J. Kelly, R. Newbury, H. Ahmed, D. G. Hasko, J. E. F. Frost, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, J. Phys. Condens. Matter \textbf{1}, 6291 (1989).
7301: \bibitem{ref506} Y. Yosephin and M. Kaveh, J. Phys. Condens. Matter \textbf{1}, 10207 (1989).
7302: \bibitem{ref507} R. Mottahedeh, M. Pepper, R. Newbury, J. A. A. J. Perenboom, and K.-F. Berggren, Solid State Comm. \textbf{72}, 1065 (1989).
7303: \bibitem{ref508} L. D. Landau and E. M. Lifshitz, {\em Statistical Physics,} Part 2, Section 55. Pergamon, Oxford, 1980.
7304: \bibitem{ref509} F. Bloch, Z. Phys. \textbf{52}, 555 (1928).
7305: \bibitem{ref510} J. N. Churchill and F. E. Holmstrom, Phys. Lett. A \textbf{85}, 453 (1981).
7306: \bibitem{ref511} J. N. Churchill and F. E. Holmstrom, Am. J. Phys. \textbf{50}, 848 (1982).
7307: \bibitem{ref512} J. Zak, Phys. Rev. B \textbf{38}, 6322 (1988).
7308: \bibitem{ref513} J. B. Krieger and G. J. Iafrate, Phys. Rev. B \textbf{38}, 6324 (1988).
7309: \end{thebibliography}
7310:
7311: \end{document}