1: %\documentclass[prb,twocolumn,showpacs,floatfix,amsmath,eqsecnum,galley]{revtex4}
2: \documentclass[prb,twocolumn,showpacs,floatfix,amsmath,eqsecnum]{revtex4}
3: %\documentclass[prb,preprint,showpacs,floatfix,amsmath,eqsecnum]{revtex4}
4: \usepackage{epsfig}
5:
6: \begin{document}
7:
8: \title{Conductance anomalies and the
9: extended Anderson model\\ for nearly perfect quantum wires}
10:
11: \author{T. Rejec$^1$ and A. Ram\v sak$^{1,2}$}
12: \affiliation{$^1$J. Stefan Institute, 1001 Ljubljana, Slovenia\\
13: $^{2}$ Faculty of Mathematics and Physics, University of
14: Ljubljana, 1001 Ljubljana, Slovenia }
15: \author{J.H. Jefferson}
16: \affiliation{QinetiQ, Sensors and Electronic Division,
17: St. Andrews Road, Great Malvern,\\
18: Worcestershire WR14 3PS, England}
19:
20:
21: \date{\today}
22:
23: \begin{abstract}
24: Anomalies near the conductance threshold of nearly perfect
25: semiconductor quantum wires are explained in terms of singlet and
26: triplet resonances of conduction electrons with a single weakly-bound
27: electron in the wire. This is shown to be a universal effect for a
28: wide range of situations in which the effective single-electron
29: confinement is weak. The robustness of this generic behavior is
30: investigated numerically for a wide range of shapes and sizes of
31: cylindrical wires with a bulge. The dependence on gate voltage,
32: source-drain voltage and magnetic field is discussed within the
33: framework of an extended Hubbard model. This model is mapped onto an
34: extended Anderson model, which in the limit of low temperatures is
35: expected to lead to Kondo resonance physics and pronounced many-body
36: effects.
37: \end{abstract}
38:
39: \pacs{
40: %1996 PACS:
41: 73.23.-b, %Mesoscopic systems
42: 85.30.Vw, %Low-dimensional quantum devices(quantum dots,quantum wires,etc.)
43: 73.23.Ad, %Ballistic transport
44: 72.10.-d %Theory of electronic transport; scattering mechanisms
45: }
46: %]
47:
48: \maketitle
49:
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51:
52: \section{Introduction}
53: Semiconductor quantum wires can be fabricated with effective wire
54: widths down to a few nanometers; for example, by heteroepitaxial
55: growth on `v'-groove surfaces \cite{walther92} and ridges
56: \cite{ramvall97}, cleaved edge over-growth \cite{yacoby96}, etched
57: wires with gating \cite{kristensen98}, and gated two-dimensional
58: electron gas (2DEG) structures \cite{wees88,wharam88}. More recently,
59: there has been considerable interest in carbon nanotubes for which the
60: quantum wire cross section can approach atomic dimensions. Such
61: structures have potential for opto-electronic applications, such as
62: light-emitting diodes, low-threshold lasers, single-electron devices
63: and quantum information processing.
64:
65: Conductance steps in various types of quantum point contacts and
66: quantum wires were found more than a decade ago
67: \cite{wees88,wharam88}. These first experiments are broadly consistent
68: with a simple non-interacting picture\cite{houten92}. However, there
69: are certain anomalies, some of which are believed to be related to
70: electron-electron interactions and appear to be spin-dependent. In
71: particular, a structure is seen in the rising edge of the conductance
72: curve, starting at around $0.7(2e^{2}/h)$ and merging with the first
73: conductance plateau with increasing energy \cite{thomas96}. This
74: structure, already visible in the early experiments \cite{wees88}, can
75: survive to temperatures of a few degrees and also persists under
76: increasing source-drain bias, even when the conductance plateau has
77: disappeared. Under increasing in-plane magnetic field, the structure
78: moves down, eventually merging with the $e^{2}/h$ conductance plateau
79: at very high fields and is not a transmission effect through a
80: ballistic channel \cite{liang}. A structure is seen also in high
81: quality quantum wires \cite{pyshkin}. In some experiments, an anomaly
82: is seen at lower energy with conductance around $0.3(2e^{2}/h)$
83: \cite{kaufman99,ramvall97}. This can also survive to a few degrees,
84: though is less robust than the 0.7 anomaly and is more readily
85: suppressed by a magnetic field \cite{ramvall97}. Recently the anomaly
86: was confirmed also in back-gated \cite{nuttinck}, in shallow-etched
87: \cite{kristensen00} point contacts and in a ballistic quantum
88: wire\cite{liang00}. At low temperatures the anomaly exhibits a
89: puzzling similarity with Kondo resonance behavior \cite{kondo2002},
90: as do thermopower measurements\cite{appleyard00}.
91:
92: Theoretical work has attempted to explain these observations in various
93: ways, including conductance suppression in a Luttinger liquid with
94: repulsive interaction and disorder \cite{maslov95}, local
95: spin-polarized density-functional theory \cite{wang98} and
96: spin-polarized sub-bands \cite{fasol94}. Near the conduction
97: threshold, there is a 'Coulomb blockade' and we have shown that this
98: gives rise to spin-dependent resonances, for wires of both rectangular
99: \cite{rejec002d} and cylindrical \cite{rejec003d} cross-section, with
100: related anomalies in thermoelectric transport coefficients
101: \cite{rejec02}. A similar singlet-triplet scenario was presented in
102: Ref.~\onlinecite{flambaum00} and a phenomenological approach is presented in
103: Ref.~\onlinecite{landau}. Recent studies have investigated the
104: 0.7 anomaly in quantum point
105: contacts within the Hartree-Fock approximation \cite{sushkov1},
106: spin-fluctuation backscattering \cite{tokura} and in the framework of the
107: Anderson model with related Kondo resonance behavior \cite{meir2002}.
108:
109: In Refs.~\onlinecite{rejec002d,rejec003d,rejec02} we suggested that these
110: anomalies are related to weakly bound
111: states and resonant bound states within the wire. These would arise, for
112: example, from a small fluctuation in thickness of the wire in some
113: region giving rise to a weak bulge. If this bulge is very weak then only a
114: single electron will be bound. We may thus regard this system as an `open'
115: quantum dot in which the bound electron inhibits the transport of conduction
116: electrons via the Coulomb interaction. Near the conduction threshold, there
117: will be a Coulomb blockade and we show below that this also gives rise to a
118: resonance, analogous to that which occurs in the single-electron transistor
119: \cite{meirav91}. This is a generic effect arising from an electron bound in
120: some region of the wire and such binding may arise from a number of sources,
121: which we do not consider explicitly. For example, in addition to a weak
122: thickness fluctuation, a smooth variation in confining potential due to
123: remote gates, contacts and depletion regions could contribute to electron
124: confinement along the wire or gated 2DEG. A significant contribution to the
125: single-electron confinement could also arise from its electronic polarization
126: of the lattice or image charge.
127:
128: In this paper we extend our previous study of a particular geometry of
129: the quantum wire with a comprehensive analysis of a wide range of
130: shapes and sizes of wire in order to demonstrate the generic and wide
131: applicability of the phenomena. We study in particular the threshold
132: of the conductivity of nearly perfect wires for which a single
133: electron is bound. We express the conductance in terms of the
134: two-electron scattering matrix. In order to extend the exact
135: two-electron analysis into the true many-electron domain, we construct
136: an extended Anderson model and analyze the influence of the
137: corresponding momentum dependent coupling matrix elements.
138:
139: The model is introduced in the next section and the special case of a
140: cylindrical GaAs wire is derived in Appendix A. In Section III a
141: detailed analysis of the two-electron problem is presented in which
142: one electron is weakly bound in the wire, giving rise to
143: spin-dependent scattering of the other. Exact singlet and triplet
144: scattering states are computed near the conductance threshold. In
145: Section IV we then show how the solutions of the scattering problem
146: may be used to determine conductance by an extension of the
147: Landauer-B\"{u}ttiker formula. This gives excellent agreement with a
148: number of experiments on different kinds of quantum wire. The effect
149: of finite magnetic field on the anomalies is presented and it is shown
150: how they are related to the spin-split steps in perfect quantum wires. In
151: the last Section we also examine the dependence of the anomalies on
152: asymmetry introduced by finite source-drain voltage and
153: summarize. Additional appendices are devoted to technical details
154: on the solution of the two-electron wave function in an external
155: potential, and the Hartree-Fock analysis.
156:
157: \section{Basic model}
158:
159: In previous work\cite{rejec002d,rejec003d,rejec02}, we have considered
160: a straight quantum wire with a small fluctuation in
161: thickness giving rise to a weak `bulge'.
162: The precise details of the bulge are largely
163: unimportant for what follows, the main requirement being that the
164: change in the width of the wire is sufficiently gradual that
165: inter-channel mixing of the transverse modes is negligible and that
166: only one electron may be bound in the bulge region. The latter is
167: always the case for a weak symmetric bulge, which has at least one
168: bound state which can only sustain one electron due to Coulomb
169: repulsion. The problem reduces to electrons moving in an effective
170: weak potential well if we confine ourselves (by choice of gate
171: voltage) to the Fermi energies for which no more than one transverse
172: mode is occupied, i.e. the conductance threshold and the first
173: conductance step. A typical effective potential well for such a bulge
174: is shown in
175: Fig.~\ref{land1}. Such a potential well may arise in other ways, such as
176: an actual potential fluctuation due to a nearby unscreened charged
177: impurity, or even some self-consistent effect due
178: to the electrons themselves through electronic polarization and image
179: charge in a remote gate. We shall not consider
180: the possible cause of this weak potential further but emphasize that
181: because it may arise in many ways, the weak potential well model is
182: very general with widespread applicability.
183:
184:
185: \begin{figure}[htb]
186: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
187: {\includegraphics{Fig1.eps}}} \par}
188: \caption{\label{land1} Effective one-dimensional well caused by
189: thickness fluctuation, impurity charge, gate image charge, self-polarization
190: due to single electron or some combination of these.}
191: \end{figure}
192:
193:
194: Consider now the motion of electrons in the wire near the conductance
195: threshold. A single electron will be bound in the potential well
196: region and the remaining electrons will undergo scattering from the
197: localized electron via the Coulomb interaction as they propagate from
198: source to drain. At sufficiently low Fermi energy, the electrons in
199: the source contact will be totally reflected by the bound electron due
200: to Coulomb repulsion and there will be no current from source to drain
201: at $T=0$. As the Fermi energy is raised, the energy of the electrons
202: in the source contact will be sufficiently high for them to overcome
203: the Coulomb repulsion of the bound electron and a current will
204: flow. In calculating this current we will make the approximation that
205: the electrons flowing from source to drain only interact with the
206: bound electron via a screened Coulomb interaction. This is a
207: reasonable approximation provided that the electron density is not too
208: low in the region of interest, i.e. the rising edge to the first
209: conductance plateau. More precisely, the mean density of electrons in
210: the wire (number per unit length) should be at least of order the
211: inverse effective Bohr radius of the material. We return to this point
212: again in the final section. Within this approximation, the
213: many-electron problem is reduced to an effective two-electron problem
214: in which one electron is bound and the other is a representative
215: electron at the Fermi energy in the leads. We show below that by
216: solving this two-electron problem exactly and summing over all
217: electrons near the Fermi energy we may compute the conductance.
218:
219: \subsection{Extended Hubbard and Anderson model}
220:
221: The Hamiltonian corresponding to interacting electrons in the wire
222: with a small geometric or potential inhomogeneity and close to the
223: threshold of conduction is, within the effective-mass
224: approximation an extended Hubbard Hamiltonian on a
225: finite-difference lattice \cite{rejecand},
226: \begin{equation}
227: \label{hamdis1d}
228: H=\sum _{\sigma }H_{1\sigma }+\frac{1}{2}\sum _{i\neq
229: j}U_{ij}n_{i}n_{j}+
230: \sum _{i}U_{ii}n_{i\uparrow }n_{i\downarrow }.
231: \end{equation}
232: Here \( H_{1\sigma } \) is the single-particle
233: Hamiltonian:
234: \begin{equation}
235: \label{ham1dis1d}
236: H_{1\sigma }=-t\sum _{i}\left( c_{i+1\sigma }^{\dagger }c_{i\sigma
237: }+c_{i\sigma }^{\dagger }
238: c_{i+1\sigma }\right) +\sum _{i}\epsilon _{i}n_{i\sigma ,}
239: \end{equation}
240: where \( c_{i\sigma }^{\dagger }, c_{i\sigma } \) are electron creation,
241: annihilation operators,
242: \( n_{i\sigma }=c_{i\sigma }^{\dagger }c_{i\sigma } \), and
243: \( n_{i}=\sum _{\sigma }n_{i\sigma } \). Model parameters are
244: hopping $t$, local potential at site \( i \),
245: $\epsilon _{i}$,
246: and screened electron-electron interaction at sites \( i \) and \( j \),
247: $U_{ij}$.
248: This Hamiltonian is derived and justified in Appendix A.
249:
250: In order to study the many-electron problem, it is also convenient to
251: express the Hubbard Hamiltonian, Eq.~(\ref{hamdis1d}), in a basis
252: which distinguishes bound and unbound states
253: explicitly. Single-electron solutions corresponding to the
254: tight-binding Hamiltonian Eq.~(\ref{ham1dis1d}), follow from the
255: single-particle Schr\"odinger equation
256: %
257: \begin{equation}
258: H_{1}\left| \varphi \right\rangle =E_{1}\left| \varphi \right\rangle, \label{sin}
259: \end{equation}
260: and (with omitted spin index $\sigma$),
261: $\left| \varphi \right\rangle =\sum _{j}\varphi _{j}c_{j}^{\dagger
262: }\left| 0\right\rangle$. For large \( \left| j\right| \) the potential
263: \( \epsilon _{j} \) is
264: constant, therefore the solutions are asymptotically plane waves.
265: We thus diagonalize this
266: single-electron part of the Hamiltonian using the transformation
267: $c_{q\sigma }^\dagger=\sum_{j}c_{j\sigma }^\dagger \phi _{j}^{q}$,
268: where $\phi _{j}^{q}=\langle j| q \rangle \sim \exp ({i q j})$ asymptotically
269: for unbound states, with eigenenergies
270: $\varepsilon _{q}$. In this basis the Hamiltonian becomes,
271: \begin{equation}
272: H=\sum_{q}\epsilon _{q}n_{q}+\frac{1}{2}\!\!\!\!\!\sum\limits_{\scriptstyle q_1
273: q_2 q_3 q_4 \scriptstyle \sigma \sigma^\prime}\!\!\!\!{{\cal
274: U}(q_{1}q_{2}q_{3}q_{4})c_{q_{1}\sigma }^{\dagger }c_{q_{3}\sigma
275: ^{\prime }}^{\dagger }c_{q_{4}\sigma ^{\prime }}c_{q_{2}\sigma }},
276: \label{Rejec_h2q}
277: \end{equation}
278: where
279: \begin{equation}
280: {\cal U}(q_{1}q_{2}q_{3}q_{4})=\sum_{ij}U_{ij}(\phi
281: _{i}^{q_{1}})^{\ast }\phi _{i}^{q_{2}}(\phi _{j}^{q_{3}})^{\ast }\phi
282: _{j}^{q_{4}}.\label{uqq}
283: \end{equation}
284: We further denote the lowest bound state with
285: energy $\epsilon _{q}<0$ by $d_{\sigma }\equiv c_{q\sigma
286: },$ with $n_{d}=\sum_{\sigma }d_{\sigma }^{\dagger }d_{\sigma }$
287: and, similarly, the scattering states with positive $\epsilon _{q}$ are
288: distinguished by $q\rightarrow k$. There are two independent unbound states
289: corresponding to each $k$ and these are chosen to be plane waves
290: asymptotically, i.e. $\phi _{j}^{k}\rightarrow e^{ikj}$ as $j\rightarrow \pm
291: \infty $ and $\epsilon _{k}=\frac{\hbar ^{2}k^{2}}{2m^{\ast }}.$ Retaining
292: only those Coulomb matrix elements which involve both localized and
293: scattered electrons, omitting all terms which would give rise to states in
294: which the localized state is unoccupied, we arrive at an Anderson-type
295: Hamiltonian\cite{anderson61,rejecand},
296: \begin{eqnarray}
297: H &=&\sum_{k}\epsilon _{k}n_{k}+\epsilon _{d}n_{d}+ \sum_{k\sigma
298: }(V_{k}n_{d\bar{\sigma}}c_{k\sigma }^{\dagger }d_{\sigma }+
299: \mathrm{h.c.})+\\\nonumber
300: &+&Un_{d\uparrow }n_{d\downarrow } +\sum_{kk^{\prime
301: }\sigma }M_{kk^{\prime }}n_{d}c_{k\sigma }^{\dagger }c_{k^{\prime
302: }\sigma }+\sum_{kk^{\prime } }J_{kk^{\prime }}{\bf S }_{\bf d}\cdot
303: {\bf s}_{\bf kk^{\prime }}.
304: \label{Rejec_k}
305: \end{eqnarray}
306: Here $U={\cal U}(dddd)$ is the Hubbard repulsion,
307: $V_{k}={\cal U} (dddk) $ is mixing term, $M_{kk^{\prime
308: }}={\cal U}(ddkk^{\prime })-\frac{
309: 1}{2}{\cal U}(dkk^{\prime }d)$ corresponds to scattering of
310: electrons and the direct exchange coupling is $J_{kk^{\prime
311: }}=2\,{\cal U}(dkk^{\prime }d)$. Spin operators in
312: Eq.~(\ref{Rejec_k}) are defined as ${\bf S}_{\bf d}=\frac{1
313: }{2}\sum_{\sigma \sigma ^{\prime }}d_{\sigma }^{\dagger }{
314: \hbox{\boldmath$\sigma$}}_{\sigma \sigma ^{\prime }}d_{\sigma
315: ^{\prime }}$ and ${\bf s}_{\bf kk^{\prime
316: }}=\frac{1}{2}\sum_{\sigma \sigma ^{\prime }}c_{k\sigma
317: }^{\dagger }{\hbox{\boldmath$\sigma$}}_{\sigma \sigma ^{\prime
318: }}c_{k^{\prime }\sigma ^{\prime }}$, where the components of
319: ${\hbox{\boldmath$\sigma$}}$ are the
320: usual Pauli matrices. A similar model has been proposed recently in
321: Ref.~\onlinecite{meir2002}.
322: Although the Hamiltonian, Eq.~(\ref{Rejec_k}), is
323: similar to the usual Anderson Hamiltonian~\cite{anderson61}, we stress
324: the important difference that the $kd$-hybridization term above
325: arises solely from the Coulomb interaction, whereas in the usual
326: Anderson case it comes primarily from one-electron interactions.
327: These have been completely eliminated above by solving the
328: one-electron problem exactly. The resulting hybridization term
329: contains the factor $n_{d\bar{\sigma}}$, and hence disappears when the
330: localized orbital is unoccupied. This reflects the fact that an
331: effective double-barrier structure and resonant bound state
332: occurs via Coulomb repulsion only because of the presence of a
333: localized electron.
334:
335: To be specific, we consider in this paper a cylindrically symmetric
336: quantum wire with symmetry axis $z$ and lateral coordinates $r$
337: and $\varphi$
338: \cite{rejec003d}. Such a geometry corresponds to narrow 'v'-groove
339: $z$-dependent quantum
340: wires investigated recently, e.g. in Ref.~\onlinecite{kaufman99}. The diameter of the
341: wire is \( a\left( z\right) \), with zero potential within the wire and
342: constant \(V_{0}>0 \) outside, i.e.,
343: %
344: \begin{equation}
345: \label{pot3d}
346: V\left( r,z\right) =\left\{ \begin{array}{ll}
347: 0, & r<\frac{1}{2}a\left( z\right) \\
348: V_{0}, & r>\frac{1}{2}a\left( z\right)
349: \end{array}.\right.
350: \end{equation}
351: %
352: For the wire width, two generic shapes are taken, shown in
353: Fig.~\ref{oblika} with
354:
355: \begin{subequations}
356: \begin{eqnarray}
357: a\left( z\right) & = & \left\{ \begin{array}{ll}
358: a_{0}\left( 1-\xi \sin ^{2}\pi \frac{z}{a_{1}}\right) , & \left| z\right| <a_{1}\\
359: a_{0}, & \left| z\right| >a_{1}
360: \end{array}\right. ,\label{oblika2} \\
361: a\left( z\right) & = & \left\{ \begin{array}{ll}
362: a_{0}\left( 1+\xi \cos ^{2}\frac{\pi }{2}\frac{z}{a_{1}}\right) , &
363: \left| z\right| <a_{1}\\
364: a_{0}, & \left| z\right| >a_{1}
365: \end{array}\right. .\label{oblika1}
366: \end{eqnarray}
367: \end{subequations}
368:
369: \begin{figure}[htb]
370: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
371: {\includegraphics{Fig2.eps}}} \par}
372: \caption{\label{oblika} The geometry of the 'open quantum dot' for
373: the parameterization (a) Eq.~(\ref{oblika2}) and (b) Eq.~(\ref{oblika1}).}
374: \end{figure}
375: The region of interest is around \( z=0 \) and for large \( \left|
376: z\right| > a_1 \) the diameter is constant $a_0$. Single particle
377: solutions corresponding to this geometry as well as the derivation and
378: calculation of parameters of the corresponding Hubbard Hamiltonian are
379: presented in Appendix A.
380:
381: \section{Two-electron solutions}
382:
383: \subsection{Bound states}
384:
385: In order to calculate conductance though the system we first solve
386: the two interacting electron problem for the present geometry using the
387: extended Hubbard Hamiltonian Eq.~(\ref{hamdis1d}).
388: Solutions for bound states are determined
389: by numerical diagonalization of the system of
390: equations presented in Appendix B,
391: Eq.~(\ref{sistem2el}). In Fig.~\ref{2dens} is shown the result
392: of the two body electron density as
393: a function of $z/a_0$ for various shapes of the bulge [Fig. \ref{oblika}(b)].
394: A general tendency is that long/narrow bulges correspond to stronger
395: interaction resulting in formation of a double peak in density, as
396: known from other studies of one-dimensional quantum dots \cite{jauregui}. As
397: long as the two peaks are not well separated, the
398: approximate methods mentioned
399: below are excellent, becoming
400: gradually less reliable with increasing separation between
401: the peaks.
402:
403: %
404: \begin{figure}[htb]
405: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
406: {\includegraphics{Fig3.eps}}} \par}
407: \caption{\label{2dens}Two-electron density for various bulge parameters.}
408: \end{figure}
409:
410: In Fig.~\ref{1_24} we present typical examples of the energy of two
411: bound (singlet) electrons ($E<0$) where $\gamma$ is the electron-electron
412: coupling
413: strength, defined by replacement $U\to
414: \gamma U$. Exact results are represented by the solid line,
415: with other lines
416: representing results obtained with the
417: Hartree-Fock approximation, derived in Appendix C. At $E>0$ the lines
418: correspond to the position of the singlet resonance, calculated with
419: different methods and discussed below. In Fig.~\ref{1_06} the bulge is
420: longer and narrower, therefore both singlet and triplet bound states
421: exist for small $\gamma$, while for stronger coupling the triplet is
422: first pushed into continuum and finally, for $\gamma \sim 0.7$, both
423: states become resonances. Here approximate solutions are less
424: accurate, because the bulge is much larger than in the previous case
425: and therefore the problem is closer to the strong interaction limit,
426: as is seen also in Fig.~\ref{2dens}, dashed-dotted line, where the
427: dip in the electron density signals the strong interaction
428: regime. The Hartree-Fock approximation gives too large energies here,
429: which are, however, qualitatively correct.
430:
431: \begin{figure}[htb]
432: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
433: {\includegraphics{Fig4.eps}}} \par}
434: \caption{\label{1_24} Position of bound states ($E<0$) and resonances
435: ($E>0$) vs. coupling $\gamma$, calculated exactly and within the
436: Hartree-Fock approximation. Wire
437: shape corresponds to Eq.~(\ref{oblika1}) with parameters:
438: \protect\( a_{0}=a_{1}=10\textrm{ nm}\protect \), \protect\( \xi =0.24\protect \),
439: \protect\( V_{0}=0.4\textrm{ eV}\protect \) and \protect\( \kappa =50\textrm{ nm}
440: \protect \).}
441: \end{figure}
442:
443: \begin{figure}[htb]
444: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
445: {\includegraphics{Fig5.eps}}} \par}
446: \caption{\label{1_06} As in Fig.~\ref{1_24} but with parameters:
447: \protect\( a_{0}=10\textrm{ nm}\protect \), \protect\( a_{1}=4a_{0}\protect \),
448: \protect\( \xi =0.06\protect \), \protect\( V_{0}=0.4\textrm{ eV}\protect \)
449: and \protect\( \kappa =50\textrm{ nm}\protect \).}
450: \end{figure}
451:
452:
453: \subsection{Scattering states}
454:
455: Here we consider the scattering of an asymptotically free electron on
456: a bound electron within the bulge. Such a system may be regarded as
457: an ``open quantum dot'' in which one electron is bound and inhibits
458: the transport of conduction electrons via Coulomb repulsion. The
459: problem is analogous to treating the collision of an electron with a
460: hydrogen atom as, e.g., described in Ref.~\onlinecite{landauQM} and
461: studied by J.R. Oppenheimer and N.F. Mott\cite{oppenheimer28}. We
462: only consider here cases in which the energy of the scattered electron
463: is smaller than the binding energy of the bound electron. This ensures
464: that only elastic scattering is possible. Asymptotically the two-body
465: wave function is a properly symmetrized product of a single particle
466: bound state, \(\left|\varphi\right>\), and scattered state, \( \left|
467: \chi \left( E\right) \right\rangle \).
468:
469: For two electrons, the antisymmetrized wavefunction can be written as a
470: product of a spin-part and an orbital part. We write the orbital part as
471: \begin{equation}
472: \widetilde{\psi}_{ij}=\frac{\psi_{ij}+\left(-1\right)^{S}\psi_{ji}}{\sqrt{2}},
473: \end{equation}
474: ensuring that this is symmetric for singlets $(S=0)$ and antisymmetric for
475: triplets $(S=1)$ (see Appendix B).
476: For some large $N\gg1$, $\psi_{ij}$ takes the form
477: \begin{equation}
478: \label{nastavek}
479: \psi_{ij}=
480: \left\{\begin{array}{ll}\chi_{i}\left(E\right)\varphi_{j}
481: +r^{\left(S\right)}\chi_{i}^{\ast}\left(E\right)\varphi_{j},
482: & i<-N\\ t^{\left( S\right)}\chi _{i}\left(E\right)\varphi_{j}, & i>N
483: \end{array}\right. .
484: \end{equation}
485:
486: Asymptotic solutions for the unbound electron are obtained from the
487: single-electron Hamiltonian Eq.~(\ref{ham1dis1d}) with the potential
488: \begin{equation}
489: \label{asimppot}
490: \tilde{\epsilon }_{j}=\epsilon _{j}+\sum _{k}U_{jk}\left| \varphi _{k}\right| ^{2}
491: \end{equation}
492: for large $|j|$. Here \( \left| \varphi \right\rangle \) is the
493: single-particle bound state in the potential \( \epsilon _{j} \).
494: Solutions with forward and backward currents have the following
495: asymptotic form $(j \to \infty)$
496: \begin{equation}
497: \label{asimpinf}
498: \chi _{j}=\left\{ \begin{array}{ll}
499: e^{\pm ikj} & \kappa <\infty \\
500: e^{\pm i\left( kj-\eta(k) \ln 2kj\right) } & \kappa =\infty
501: \end{array}\right. ,
502: \end{equation}
503: for finite and infinite screening length $\kappa$, respectively (see
504: Appendix A). With no screening $(\kappa =\infty)$ $\chi _{j}$ are the
505: Coulomb functions \cite{abramovitz}.
506:
507: Numerically exact solutions are obtained by solving a set of linear
508: equations for the \( \left( 2N+1\right) ^{2} \) variables \(
509: \widetilde{\psi }_{ij} \) and transmission and reflection amplitudes.
510:
511: \section{Conductance}
512:
513: \subsection{Single-electron solutions}
514:
515: From the solution of the scattering problem, the conductance at zero
516: temperature is calculated using the usual Landauer-B\"{u}ttiker
517: formalism \cite{landauer57},
518: %
519: \begin{equation}
520: G=G_0 {\cal T}(E), \label{landauer}
521: \end{equation}
522: where $G_{0}=2e^{2}/h$, $E$ is the Fermi energy (in this case $E=E_1$)
523: and ${\cal T}(E)$ is the
524: total transmission probability.
525:
526: For an open bulge of shape Eq.~(\ref{oblika2}), Fig.~\ref{oblika}(a),
527: there are no bound states and only single-electron
528: solutions\cite{ramsak98} are relevant.
529: In Fig.~\ref{exact} we present $G$ as a function of
530: electron energy for wires with shape Eq.~(\ref{oblika2}) and three different
531: widths. The main effect is a change of energy scale, according to
532: scaling rule Eq.~(\ref{transform}), and the magnitude of the conductance at
533: the resonance energy, $G_0$.
534: In Fig.~\ref{exactb} the conductance through the bulge of the
535: shape from Eq.~(\ref{oblika1}) is presented. In contrast to the previous figure,
536: a bound state can exist here, indicated by the dashed vertical line.
537: Further lines ($n=1,2$) indicate bound
538: states of {\it individual channels} for the special case when channel
539: mixing terms in Eq.~(\ref{schrodkvazi1d}) are set to zero. Dips in the
540: conductivity in the second plateau ($G\sim 3 G_0$) correspond to Fano
541: resonances caused by interchannel mixing terms.
542: %
543: \begin{figure}[htb]
544: {\par\centering\resizebox*{0.9\columnwidth}{!}{\includegraphics{Fig6.eps}} \par}
545: \caption{\label{exact} $G$ for a wire with shape Eq.~(\ref{oblika2}) and
546: \protect\( V_{0}=0.4\textrm{ eV}\protect \), \protect\( \xi =0.8\protect \),
547: \protect\( a_{1}=a_{0}\protect \) and in particular (a) \protect\(
548: a_{0}=7\textrm{ nm}\protect \), (b) \protect\( a_{0}=10\textrm{
549: nm}\protect \), and (c) \protect\( a_{0}=15\protect \)
550: nm.}
551: \end{figure}
552:
553: \begin{figure}[htb]
554: {\par\centering \resizebox*{0.9\columnwidth}{!}{\includegraphics{Fig7.eps}} \par}
555: \caption{\label{exactb}
556: $G$ for wire shape Eq.~(\ref{oblika1}) and
557: \protect\( V_{0}=0.4\protect \) eV, \protect\( \xi =0.5\protect \),
558: \protect\( a_{0}=a_{1}=10\textrm{ nm}\protect \). Vertical lines
559: indicate positions of bound states for the lowest channels.}
560: \end{figure}
561:
562: In Fig.~\ref{2ch}(a) we again show the result of Fig.~\ref{exactb}
563: comparing it with the one-channel approximation. In this paper we
564: are interested in the rising edge of the conductance at the
565: threshold, but with Coulomb interactions between bound and
566: scattered electron included. Near threshold the
567: one-channel approximation is
568: excellent and therefore in the following we neglect higher channels.
569: In Fig.~\ref{2ch}(b) is presented the influence of discretization
570: parameter \protect\( \Delta \protect \), as introduced in Appendix A,
571: on the conductivity. The position of the bound state is not strongly
572: dependent on \protect\( \Delta \protect \) (inset in enlarged energy
573: scale) and for $\Delta < a_1/5$ the results obtained on the lattice
574: agree with the continuum calculation within a percent, which
575: justifies the use of the discretized Hamiltonian, Eq.~(\ref{hamdis1d}).
576:
577: \begin{figure}[htb]
578: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
579: {\includegraphics{Fig8.eps}}} \par}
580: \caption{\label{2ch}(a) Conductance from Fig.~\ref{exactb} in
581: comparison with the result obtained in the one channel
582: approximation. (b) Conductance calculated with different \protect\(
583: \Delta \protect \) for wire parameters as in (a). Bound states are
584: presented in the inset with enlarged energy scale.}
585: \end{figure}
586:
587:
588: \subsection{\label{nepel}Interacting electrons}
589:
590: We may extend the formula Eq.~(\ref{landauer}) to the case described in the
591: previous section in which one electron is bound in the wire and the remaining
592: electrons are transmitted with energy-dependent probability.
593: Let $P_{\sigma }$ be the probability that the bound electron has
594: spin $\sigma $. It follows directly that the conductance due to all spin-up
595: electrons in the leads is given by the extended Landauer B\"{u}ttiker
596: formula:
597: \begin{equation}
598: G_{\uparrow }=\frac{e^{2}}{h} \left[ P_{\uparrow }{\cal T}_{\uparrow
599: \uparrow }(E )+P_{\downarrow }{\cal T}_{\uparrow \downarrow
600: }(E )\right], \label{aa}
601: \end{equation}
602: where ${\cal T}_{\uparrow \uparrow }$ is the transmission probability when
603: the bound electron is spin up, ${\cal T}_{\uparrow \downarrow }$\ is the
604: transmission probability when the bound electron is spin down and $E$ is the
605: Fermi energy. We have a
606: similar expression for spin-down electrons in the leads and hence the total
607: conductance is
608: \begin{equation}
609: G(E)=\frac{e^{2}}{h} \left[ P_{\uparrow }{\cal T}_{\uparrow \uparrow
610: }(E )+P_{\downarrow }{\cal T}_{\uparrow \downarrow }(E
611: )+P_{\uparrow }{\cal T}_{\downarrow \uparrow }(E )+P_{\downarrow }%
612: {\cal T}_{\downarrow \downarrow }(E )\right] .
613: \label{bb}
614: \end{equation}
615: The transition probabilities ${\cal T}_{\uparrow \uparrow }$ and ${\cal T}%
616: _{\uparrow \downarrow }$ are different since in the former case the
617: conduction and bound electrons both have the same spin (up) before and after
618: scattering whereas in the latter case there are two possible final states,
619: with or without spin-flip, i.e.
620: \begin{equation}
621: {\cal T}_{\uparrow \downarrow }(E)=
622: \left| t_{\uparrow\downarrow\rightarrow\uparrow\downarrow}\right|
623: ^{2}+\left| t_{\uparrow\downarrow\rightarrow\downarrow\uparrow}\right|
624: ^{2}, \label{tud}
625: \end{equation}
626: where the scattering amplitudes are defined by
627: \begin{eqnarray}
628: \langle i \uparrow j \downarrow | \psi_{\uparrow \downarrow}\rangle &\rightarrow&
629: t_{\uparrow\downarrow\rightarrow\uparrow\downarrow} \chi_i \varphi_j \\ \nonumber
630: \langle i \downarrow j \uparrow | \psi_{\uparrow \downarrow}\rangle &\rightarrow&
631: t_{\uparrow\downarrow\rightarrow\downarrow\uparrow} \chi_i \varphi_j
632: \label{ijud}
633: \end{eqnarray}
634: as $i\to \infty$. $| \psi_{\uparrow \downarrow}\rangle$ is the exact
635: scattering wavefunction and $|i \sigma j \sigma'\rangle=c^\dagger_{i
636: \sigma} c^\dagger_{j \sigma'}
637: |0 \rangle $. $\varphi_j=\langle j | \varphi \rangle$ is the bound-state one-electron
638: wavefunction and $\chi_j=\langle j | \chi \rangle$ is a forward propagating one-electron
639: wavefunction at large $j$, as discussed in Section III.
640:
641: In zero magnetic field it is clear that $P_\uparrow=P_\downarrow=\frac{1}{2}$ in
642: Eq.~(\ref{bb}). We can express $G(E)$ in a simpler form since the
643: ${\cal T}_{\sigma \sigma'}(E)$ are not all independent. Transforming to singlet
644: and triplet base states (with $S_z=0$),
645: \begin{eqnarray}
646: |s,i,j\rangle &=&\frac{|i\uparrow ,j\downarrow \rangle -|i\downarrow ,j\uparrow
647: \rangle }{\sqrt{2}},\\ \nonumber
648: |t,i,j\rangle &=&\frac{|i\uparrow ,j\downarrow \rangle
649: +|i\downarrow ,j\uparrow \rangle }{\sqrt{2}},
650: \label{stij}
651: \end{eqnarray}
652: we get
653: \begin{eqnarray}
654: \langle s,i,j| \psi_{\uparrow \downarrow}\rangle &\rightarrow&
655: \frac{t^{(0)}}{\sqrt{2}} \chi_i \varphi_j \\ \nonumber
656: \langle t,i,j | \psi_{\uparrow \downarrow}\rangle &\rightarrow&
657: \frac{t^{(1)}}{\sqrt{2}} \chi_i \varphi_j,
658: \label{stud}
659: \end{eqnarray}
660: where
661: \begin{equation}
662: \label{tupdown}
663: \begin{array}{rcl}
664: t^{(0) } &=&
665: t_{\uparrow\downarrow\rightarrow\uparrow\downarrow
666: }+t_{\uparrow\downarrow\rightarrow\downarrow\uparrow } \\
667: t^{(1) }&=&
668: t_{\uparrow\downarrow\rightarrow\uparrow\downarrow
669: }-t_{\uparrow\downarrow\rightarrow\downarrow\uparrow }
670: \end{array}.
671: \end{equation}
672: Hence, Eq.~(\ref{aa}) becomes\cite{landauQM}
673: \begin{equation}
674: G=G_0\left(\frac{1}{4}\left| t^{\left( 0\right) }\right|
675: ^{2}+\frac{3}{4} \left| t^{\left( 1\right) }\right| ^{2}\right).
676: \end{equation}
677:
678:
679: \subsection{Results of numerical analysis}
680:
681: In Fig.~\ref{final1} we present the result of a comprehensive study
682: of conductance for a variety of shapes of bulge. In
683: Fig.~\ref{final1}(a) the bulge is wide and so short that only a singlet
684: resonance is developed. The conductance therefore exhibits a structure
685: similar to the 0.3 anomaly found in experiment \cite{patel91}. In
686: Figs.~\ref{final1}(b,c) both singlet and triplet resonances are visible
687: with a tendency for the resonances to sharpen as the bulge becomes weaker
688: ($\xi \to 0$).
689:
690: \begin{figure}[htb] {\par\centering
691: \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}{\includegraphics{Fig9.eps}}}
692: \par}
693: \caption{\label{final1} Conductance for different shapes of the
694: bulge. Each line is labeled with the parameter
695: \protect\(a_{1}/a_{0}\protect \). Other wire parameters:
696: \protect\( a_{0}=10\textrm{ nm}\protect \),
697: \protect\( V_{0}=0.4\textrm{ eV}\protect \) and \protect\( \kappa
698: =50\textrm{ nm}\protect \).}
699: \end{figure}
700: In Fig.~\ref{final2} the wire width is fixed at $10$nm of the wire and
701: positions of singlet (full lines) and triplet (dashed lines)
702: resonances (or the corresponding bound states for $E<0$) are plotted
703: for various lengths and widths of bulge, represented by $a_1/a_0$ and
704: $\xi$. We see that the resonances survive for a wide range of
705: parameters. In Fig.~\ref{final3} is shown the position of singlet and
706: triplet resonance energies vs. the width of the wire, $a_0$, with
707: fixed shape of the bulge. The insets show the energy dependence of
708: singlet and triplet transmission probabilities for selected special
709: cases. Note that we have scaled the energy by a factor $a_0^2
710: E$. This would produce identical curves for non-interacting electrons
711: [Eq.~(\ref{transform})].
712:
713: After performing calculations for a wide range of parameters, we
714: conclude that a singlet resonance is always lower in energy than
715: its corresponding
716: triplet, in accordance with Lieb-Mattis theorem, which however, is
717: strictly valid only for ground states \cite{lieb62}.
718:
719: \begin{figure}[htb]
720: {\par\centering \resizebox*{0.9\columnwidth}{!}
721: {\rotatebox{0}{\includegraphics{Fig10.eps}}} \par}
722: \caption{\label{final2} Energies of singlet (full lines) and triplet
723: (dashed lines) resonances and bound states for wire Eq.~(\ref{oblika1}) as
724: a function of \protect\( a_{1}/a_0\protect \) for different \protect\( \xi \protect \).
725: Other parameters are as in Fig.~\ref{final1}.
726: Full circles represent the energy, where
727: the resonance energy is above the 'ionization' energy and the underlying
728: wave function form is not valid anymore.}
729: \end{figure}
730:
731: \begin{figure}[htb]
732: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
733: {\includegraphics{Fig11.eps}}} \par}
734: \caption{\label{final3}The position of singlet (full line) and triplet
735: (dashed line) resonances as a function of the width of the wire,
736: $a_0$. Note that the energy of the resonances is presented in a scaled
737: form. $a_1=50\mathrm{~nm}$, $\xi=0.11$ and other parameters are as in
738: Fig.~\ref{final1}.}
739: \end{figure}
740:
741: \subsection{Magnetic field along the symmetry axis}
742:
743: The nature of conductance anomalies studied here can be further
744: illuminated with experiments done in a strong magnetic
745: field\cite{patel91}. The effect of the magnetic field
746: is, in our treatment, taken into account via the usual Zeeman
747: splitting of channel energies. The incoming electron with Fermi
748: energy \( E \) and spin component \( S_{z}=\pm\frac{1}{2} \) then has
749: kinetic energy
750: \begin{equation}
751: E_{\textrm{k}}\equiv E_\mp =E\mp E_{\textrm{B}},
752: \end{equation}
753: where $E_{\rm B}=\frac{1}{2}g^\ast\mu_{\rm B} B$. $g^\ast$ is the effective gyromagnetic ratio and \( \mu
754: _{\textrm{B}} \) is Bohr magneton.
755:
756: Near the conductance threshold we assume that the current is sufficiently low
757: that the localized electron is in its ground-state with spin $\downarrow$ before
758: each scattering event with a conduction electron. Hence $P_\downarrow=1$ and
759: $P_\uparrow=0$ in Eq.~(\ref{aa}) which becomes
760: \begin{equation}
761: G_\uparrow(E,B)=\frac{e^{2}}{h} {\cal T}_{\uparrow \downarrow}(E,B).
762: \end{equation}
763: In this case $G_\downarrow \neq G_\uparrow$ but, rather,
764: \begin{equation}
765: G_\downarrow(E,B)=\frac{e^{2}}{h} {\cal T}_{\downarrow \downarrow}(E,B).
766: \end{equation}
767: Since ${\cal T}_{\uparrow \downarrow}(E,B)={\cal T}_{\uparrow \downarrow}(E_-,0)$ and
768: ${\cal T}_{\downarrow \downarrow}(E,B)={\cal T}_{\downarrow \downarrow}(E_+,0)$,
769: then the conductance is
770: \begin{eqnarray}
771: G&=&G_{\uparrow}+G_{\downarrow}\\ \nonumber
772: &=&\frac{e^{2}}{h}[
773: {\cal T}_{\uparrow \downarrow}(E_-,0)+
774: {\cal T}_{\downarrow \downarrow}(E_+,0)] \\ \nonumber
775: &=& \frac{e^2}{h}[
776: {\cal T}_{t}(E_+,0)+
777: \frac{1}{2}{\cal T}_{t}(E_-,0)+
778: \frac{1}{2}{\cal T}_{s}(E_-,0)],\label{gb}
779: \end{eqnarray}
780: where ${\cal T}_{s}$ and ${\cal T}_{t}$ are the same functions
781: as in zero magnetic field.
782:
783: In Fig.~\ref{magnet}(a) are plotted individual transmission probabilities for
784: different spin configurations. Note that the spin-flip term
785: is in general dominant at higher energies. In Fig.~\ref{magnet}(b) the
786: corresponding results for conductance in the presence of a magnetic
787: field is shown. The full line corresponds to $B=0$, and other curves to $B$ in
788: increments $\Delta B=10$T.
789: %
790: \begin{figure}[htb]
791: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{-0}
792: {\includegraphics{Fig12.eps}}} \par}
793: \caption{\label{magnet} (a) Transmission probabilities for relevant
794: spin configurations. (b) Conductance for $B=0$ (full line) and other
795: lines for $B$ in increments $\Delta B=10$~T. Parameters of the wire:
796: \protect\( a_{0}=10 \textrm{ nm}\protect \), \protect\(
797: a_{1}=2.5a_{0}\protect \), \protect\( \xi =1.11\protect \), \protect\(
798: V_{0}=0.4 \textrm{ eV}\protect \) and \protect\( \kappa
799: =50\textrm{ nm}\protect \).}
800: \end{figure}
801:
802: \subsection{Results for the Anderson model}
803:
804: As shown in Section II, the Hubbard model studied above can be
805: mapped onto an extended Anderson model, Eq.~(\ref{Rejec_k}).
806: Conductance through a quantum dot described by a standard Anderson model
807: is basically described by a peak or several peaks and at higher energies the
808: conductance approaches zero \cite{anddot}. In the case of an
809: open quantum dot, studied here,
810: at higher energies the conductance tends toward unity, as a consequence of
811: additional coupling parameters in the extended
812: model. Here we analyze these terms individually
813: and show their relative importance.
814:
815: The coupling parameters are momentum dependent and in
816: Fig.~\ref{Rejec_Fig1} the couplings $V_{k}$, $M_{kk^\prime}$ and
817: $J_{kk^\prime}$ are shown. Note that $M_{kk^\prime}$ at higher
818: energies tends to a constant, while other parameters approach zero,
819: ensuring the correct behavior at high-energy with unit
820: transmission.
821: \begin{figure}[tbh]
822: {\par\centering \resizebox*{0.7\columnwidth}{!}{\rotatebox{0}
823: {\includegraphics{Fig13.eps}}} \par}
824: \caption{$k$-dependence of matrix elements of the extended Anderson
825: model. The wire is parameterized with $ a_{0}=10$~nm, $\protect\xi
826: =0.24$, $a_{1}/a_{0}=2$, $V_{0}=0.4$~eV, $\protect \kappa=50$~nm and
827: $\protect\gamma =1$. (a) Mixing coupling $V_{k}$. The energy
828: $\epsilon_d + U$ is indicated with an arrow. (b, c) Scattering
829: couplings $M_{kk'}$ and $J_{kk'}$. $L$ is the length of the wire, where the wave
830: functions are normalized.}
831: \label{Rejec_Fig1}
832: \end{figure}
833: The scattering solutions of the Hamiltonian Eq. (\ref{Rejec_k}) are then
834: obtained exactly for two electrons with the boundary condition
835: that for $z\rightarrow \infty $, one electron occupies the lowest
836: bound state, whilst the other is in a forward propagating plane
837: wave state, $\phi _{k}(z)\sim e^{ikz}.$ From these solutions we
838: compute the conductance again using the Landauer-B\"{u}ttiker formula.
839:
840: In Figs.~\ref{Rejec_Fig2}(a,b,c) we compare the results of
841: ${\cal T}_{ \mathrm{s}}$,
842: ${\cal T}_{\mathrm{t}}$ and conductance $G$ for a wire with the bulge
843: as in Fig.~\ref{1_24}.
844: The thin lines are the exact scattering result for two electrons. The
845: solid lines show the exact scattering solutions for the Anderson-type
846: Hamiltonian, for which the matrix elements, and their energy
847: dependence are calculated explicitly. The solution of this
848: Anderson-type model for two electrons, in which the localized level
849: always contains at least one electron, reproduce the main features of
850: the exact scattering solutions of the original model. The energy
851: dependence of the matrix elements is essential to get this good
852: agreement. Figs.~\ref{Rejec_Fig2}(d,e,f) show the corresponding results for
853: a longer bulge from Fig.~\ref{1_06}. Also shown in Fig.~\ref{Rejec_Fig2}, dashed
854: lines, are results with the direct exchange term omitted in
855: Eq.~(\ref{Rejec_k}). This term can have a significant quantitative effect,
856: but does not qualitatively change the conductance curves.
857:
858: We have also solved a similar model in which plane waves,
859: rather than exact scattering states of the non-interacting problem,
860: were used. However, this gave poor agreement with the exact results.
861: We conclude that an Anderson-type model is adequate for a near-perfect
862: quantum wire provided that a suitable basis set is used and the
863: energy-dependence of the matrix elements is accurately
864: determined. Future work will focus on the many-electron properties of
865: this effective Hamiltonian, including `Kondo' and `mixed valence'
866: regimes.
867:
868: \begin{figure}[tbh]
869: {\par\centering \resizebox*{0.8\columnwidth}{!}{\rotatebox{0}
870: {\includegraphics{Fig14.eps}}} \par}
871: \caption{Singlet (a, d) and triplet (b, e) transmission probabilities and
872: corresponding conductances (c, f). Parameters for the left set are as in
873: Fig.~\ref{Rejec_Fig1}, for the right set: $a_{0}=10$~nm, $\protect\xi =0.15$,
874: $a_{1}/a_{0}=4$, $V_{0}=0.4$~eV, $\protect\kappa=100$~nm and
875: $\protect\gamma =0.9$. Thin lines represent exact results from
876: Eq.~(\ref{hamdis1d}), thick lines are results from
877: Eq.~(\ref{Rejec_k}). Dashed lines show results where the exchange
878: term in Eq.~(\ref{Rejec_k}) is neglected.} \label{Rejec_Fig2}
879: \end{figure}
880:
881:
882: \subsection{Approximate methods}
883:
884: It is not easy to get sufficiently accurate numerical solutions
885: for the case of more
886: than two electrons. Therefore it would be extremely useful if an
887: accurate
888: approximative method could be applied. The simplest approximation
889: (presented here for the case of two electrons) is be the first iteration in
890: solving the Hartree-Fock equations, as we presented in Section III for
891: the case of bound
892: states.
893:
894: We assume the two-body wave functions consist of a single particle
895: state \( \left| \varphi ^{\left( 1\right) }\right\rangle \) and
896: scattering state $ | \chi \rangle$ with energy \( E \). The
897: two-electron wavefunction has then the form (see Appendix C),
898: \begin{equation}
899: \label{psi_approx_2}
900: \left| \psi \right\rangle =\sum _{ij}\varphi _{i}\chi _{j}c_{ij}^{
901: \left( S,S_{z}\right) \dagger }\left| 0\right\rangle .
902: \end{equation}
903: The coefficients \( \varphi _{i} \) are known, therefore
904: only coefficients \( \chi _{i} \) must be
905: determined. For the singlet state the simplest approximation is
906: obtained if we perform the first iteration of the Hartree-Fock method
907: subject to additional condition that the electron has energy $E$
908: (``restricted Hartree-Fock'' approximation):
909: %
910: \begin{equation}
911: \label{rest_sing_tr}
912: \left\langle 0\left| c_{i}H_{1}\right| \chi \right\rangle +\sum
913: _{j}U_{ij}
914: \left| \varphi _{j}\right| ^{2}\chi _{i}=E\chi _{i},
915: \end{equation}
916: where, using Eq.~(\ref{ham1dis1d}),
917: $\left\langle 0\left| c_{i}H_{1}\right| \chi \right\rangle=
918: -t(\chi_{i-1}+\chi_{i+1})+\epsilon_i \chi_i$. This is just the tight binding
919: results for a single electron moving in an effective potential
920: $\epsilon _{i}+\sum _{j}U_{ij}\left|
921: \varphi _{j}\right| ^{2}$.
922:
923: For the triplet state, the result is:
924: \begin{equation}
925: \label{trip_tr}
926: \left\langle 0\left| c_{i}H_{1}\right| \chi \right\rangle +
927: \sum _{j}U_{ij}\left| \varphi _{j}\right| ^{2}\chi _{i}-\sum _{j}
928: U_{ij}\varphi _{j}^{\ast }\varphi _{i}\chi _{j}=E\chi _{j}.
929: \end{equation}
930: %
931: A better approximation for the singlet case starts from the unrestricted
932: Hartree-Fock approximation, where the energy is
933: %
934: \begin{eqnarray}
935: &&\left\langle \psi \left|H\right| \psi \right\rangle =
936: \left\langle \varphi \left| H_{1}\right| \varphi
937: \right\rangle \left\langle \chi |\chi \right\rangle
938: +
939: \left\langle \chi \left| H_{1}\right| \chi \right\rangle
940: \left\langle \varphi |\varphi \right\rangle+ \nonumber \\
941: &&\quad\quad+\left( -1\right) ^{S}
942: \left( \left\langle \varphi \left| H_{1}\right| \chi \right\rangle
943: \left\langle \chi |\varphi \right\rangle +\left\langle \chi \left|
944: H_{1}\right| \varphi \right\rangle \left\langle \varphi |
945: \chi \right\rangle \right)+ \nonumber \\
946: &&\quad\quad+\frac{1}{2}\sum _{ij}U_{ij}\left| \varphi _{i}\chi _{j}+
947: \chi _{i}\varphi _{j}\right| ^{2},
948: \end{eqnarray}
949: and the norm is given with
950: \begin{equation}
951: \left\langle \psi |\psi \right\rangle =\sum _{i}\left| \chi _{i}
952: \right| ^{2}+\left| \sum _{i}\varphi _{i}^{\left( 1\right) \ast }\chi _{i}\right| ^{2}.
953: \end{equation}
954: The
955: coefficients $\chi_i$ are calculated from the Hartree-Fock
956: equations based on the variation principle (``unrestricted
957: Hartree-Fock'' approximation),
958: \begin{eqnarray}
959: \label{unrest_sing_tr}
960: &&\left\langle 0\left| c_{i}H_{1}\right| \chi \right\rangle +
961: \sum _{j}U_{ij}\left| \varphi _{j}\right| ^{2}\chi _{i}+
962: \sum _{j}U_{ij}\varphi _{j}^{\ast }\varphi _{i}\chi _{j}+ \nonumber \\
963: &&\quad\quad+
964: \left( E-E^{\left( 1\right) }_{1}\right) \sum _{j}\varphi _{j}^{\ast }
965: \varphi _{i}\chi _{j}=E\chi _{i}.
966: \end{eqnarray}
967:
968: In Fig.~\ref{efpot} the effective one
969: dimensional potential in Eq.~(\ref{rest_sing_tr}) is plotted for
970: $\gamma=0, 0.5$ and $1$. The shaded region
971: represents the position and the width of single particle resonance in
972: this effective potential. This resonance corresponds to the singlet
973: resonance presented in Fig.~\ref{approx}(a) (dashed line), for wire
974: parameters given in Fig.~\ref{1_24} together
975: with exact result, full line, and calculated from
976: Eq.~(\ref{rest_sing_tr}).
977:
978: We also show in Fig.~\ref{approx}(a) the exact result and the corresponding
979: result for an unrestricted Hartree-Fock scheme for which the wavefunctions
980: of up and down spin electrons are different. As expected, the unrestricted method
981: gives a more accurate result though both methods reproduce the main features,
982: the main discrepancy being an overall energy shift. Similarly in
983: Fig.~\ref{approx}(b) we
984: present the corresponding triplet resonance curve for parameters
985: from Fig.~\ref{1_06}. Again, the overall agreement with the exact result
986: is good apart from an overall energy shift.
987:
988: \begin{figure}[htb]
989: {\par\centering \resizebox*{0.6\columnwidth}{!}{\rotatebox{0}
990: {\includegraphics{Fig15.eps}}} \par}
991: \caption{\label{efpot} Effective potential from
992: Eq.~(\ref{rest_sing_tr}) and for the wire with parameters as in Fig.~\ref{1_24}.}
993: \end{figure}
994:
995: \begin{figure}[htb]
996: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
997: {\includegraphics{Fig16.eps}}} \par}
998: \caption{\label{approx}(a) Singlet resonance (parameters from
999: Fig.~\ref{1_24}). Exact result and approximations of
1000: Eqns.~(\ref{rest_sing_tr}) and (\ref{unrest_sing_tr}) are shown. (b)
1001: Triplet resonance (parameters from Fig.~\ref{1_06}). Exact result and
1002: approximation of Eqn.~(\ref{trip_tr}) are shown.}
1003: \end{figure}
1004:
1005:
1006: \section{Summary and conclusions}
1007:
1008: We have shown that quantum wires with weak longitudinal
1009: confinement, or open quantum dots, can give rise to spin-dependent, Coulomb
1010: blockade resonances when a single electron is bound in the confined region.
1011: This is a universal effect in one-dimensional systems with very weak
1012: longitudinal confinement. The emergence of a specific structure at $G(E)\sim
1013: \frac{1}{4}\frac{2e^{2}}{h}$ and $G\sim \frac{3}{4}\frac{2e^{2}}{h}$ is a
1014: consequence of the singlet and triplet nature of the resonances and the
1015: probability ratio 1:3 for singlet and triplet scattering and as such is a
1016: universal effect. A comprehensive numerical investigation of open quantum
1017: dots using a wide range of parameters shows that singlet resonances are
1018: always at lower energies than the triplets, in accordance with the
1019: corresponding theorem for bound states \cite{lieb62}. With increasing
1020: in-plane magnetic field, the resonances shift their position and eventually
1021: merge in the conductance plateau at $G\sim e^{2}/h$. With increasing
1022: source-drain bias we have shown why the higher triplet resonance weakens at
1023: the expense of the singlet, with the latter surviving to the point where the
1024: conductance steps themselves disappear.
1025:
1026: The existence of the conductance anomalies is a direct
1027: consequence of an effective double-barrier potential seen by the conduction
1028: electrons propagating from source to drain contacts under the influence of a
1029: bound electron. For a symmetric one-electron confining potential, the
1030: existence of a bound state is guaranteed but this is not necessarily the
1031: case when the confinement is asymmetric. Such asymmetry in the confining
1032: potential may be easily achieved under a finite source drain bias and
1033: indeed, this was reported in some of the experiments on gated quantum wires
1034: \cite{thomas96,patel91}. These experiments show that as the source-drain
1035: bias is increased from zero, an anomaly appears at\ $G\sim
1036: 0.25(2e^{2}/h)$, coexisting with the $0.7(2e^{2}/h)$ anomaly.
1037: Eventually, at larger bias, the remaining
1038: anomaly also disappears but only when the conductance steps themselves are
1039: on the point of disappearing, showing that the singlet anomaly is extremely
1040: robust. This behaviour is consistent with our model since under bias the
1041: triplet resonant bound-state will eventually disappear because the confining
1042: potential in the $x$-direction will only accommodate a single one-electron
1043: bound state, giving rise to a singlet resonance only. This is shown
1044: schematically in Fig.~\ref{land6} where we also indicate the surviving singlet
1045: becoming broader with increasing bias resulting in a more pronounced step, as
1046: observed.
1047: \begin{figure}[htb]
1048: {\par\centering \resizebox*{0.8\columnwidth}{!}{\rotatebox{0}
1049: {\includegraphics{Fig17.eps}}} \par}
1050: \caption{\label{land6}Effective double barrier showing singlet and triplet
1051: resonance with very small source-drain bias (a) and large source-drain
1052: bias (b).}
1053: \end{figure}
1054:
1055: Finally, we speculate on the exciting possibility that these anomalies
1056: in conduction are themselves a signature for a new kind of conducting
1057: state in ultra clean wires close to the conduction threshold. Indeed,
1058: there is some experimental evidence for this in that the anomalies
1059: discussed above merge into a conductance step at $e^{2}/h$ under quite
1060: moderate magnetic fields and in the cleanest samples this behaviour is
1061: sometimes even seen in zero magnetic field. This suggests that there
1062: may be an underlying spin polarized state associated with the
1063: propagating electrons in the quasi 1D region. Such a spin-polarized
1064: state would appear to violate the Lieb-Mattis theorem
1065: \cite{lieb62} and
1066: would also need to be made consistent with our above explanation in
1067: terms of singlet and triplet resonances. In this respect we emphasize
1068: that the above theory must break down at very low electron density in
1069: the wire such that the mean separation between electrons in the wire
1070: is somewhat greater than the effective Bohr radius, the so-called
1071: strong correlation regime. In practical situations it is very
1072: difficult to avoid some kind of weak potential fluctuation which traps
1073: one electron. Indeed this may ultimately be impossible since even in a
1074: nominally perfect wire, the presence of a single electron will
1075: polarise its environment leading to a potential well which will bind
1076: the electron giving rise to a Coulomb blockade for the remaining
1077: electrons, though the energy scale (temperature) for this may be very low making
1078: it susceptible to masking the other effects.
1079:
1080: The main question is whether or not this confinement is
1081: sufficiently large for the electron density to exceed to the inverse
1082: Bohr radius when the wire begins to conduct. If the density remains
1083: low at this conductance threshold then we cannot ignore the mutual
1084: interaction between all electrons in the wire region, or even treat
1085: them self-consistently. In this situation, a more appropriate picture
1086: would be one in which the Coulomb repulsion dominates and maintains
1087: roughly equal separation between the electrons as in a Wigner
1088: chain. On the other hand, if the mean electron density is of order,
1089: or greater,than the inverse Bohr radius, then an open quantum dot
1090: picture with effective resonance levels for the propagating electron is
1091: more appropriate, as discussed in this paper.
1092:
1093: At low temperatures strong many-body effects are indicated from the
1094: activation-like behaviour of the conductance \cite{thomas96,
1095: kondo2002} and the thermopower coeficient \cite{appleyard00}. As
1096: discused in our recend thermopower analysis, Ref.~\onlinecite{rejec02}, the
1097: anomaly at low-temperatures may well be a many-body Kondo-like effect
1098: contained within our extended Anderson model,
1099: Eq.~(\ref{Rejec_k}), and studied recently in Ref.~\onlinecite{meir2002}, but
1100: not within the two-electron approximation we have used here and in
1101: some our earlier papers. It may well be that the two-electron
1102: approximation breaks down at low temperatures. The model presented
1103: here differs from the standard Anderson model in that the
1104: hybridisation term contains the factor $n_{-\sigma }$, and hence
1105: disappears when the localized orbital is unoccupied. This reflects the
1106: fact that an effective double-barrier structure and resonant bound
1107: state occurs via Coulomb repulsion only because of the presence of a
1108: localized electron. The standard results for the single impurity
1109: problem \cite{hewsonbook}
1110: thus cannot be applied directly to this effective model,
1111: and are a subject of current reaserch \cite{rejecram}. However, a
1112: Kondo-like resonance is expected \cite{boese01,meir2002}, for which
1113: many-body effects would dominate with a breakdown of our two electron
1114: approximation.
1115:
1116: \appendix
1117:
1118: \section{Cylindrical wire}
1119:
1120: \subsection{Single electron basis}
1121:
1122: A single electron in the wire considered here is described with the wave function
1123: \( \Psi (r,\varphi ,z) \), which is a solution of
1124: the Schr\"{o}dinger equation
1125:
1126: \begin{equation}
1127: \label{schrod3d}
1128: -\frac{\hbar ^{2}}{2m^{\ast }}\nabla ^{2}\Psi \left( r,\varphi
1129: ,z\right)
1130: +V\left( r,z\right) \Psi \left( r,\varphi ,z\right) =E\Psi \left(
1131: r,\varphi
1132: ,z\right) ,
1133: \end{equation}
1134: where the effects of nonparabolicity are neglected and the effective
1135: mass is taken constant, $m^{\ast } = 0.067m_{\rm elec}$
1136: with dielectric constant 12.5, appropriate
1137: for GaAs \cite{ramsak98}.
1138:
1139: At fixed \( z \) the wave function \( \Psi (r,\varphi ,z) \) is
1140: expanded in a two-dimensional basis \( \Phi _{mn}\left( r,\varphi
1141: ;z\right) \) for the corresponding potential \( V\left( r;z\right) \)
1142: , Eq.~(\ref{pot3d}). The coefficients in such an expansion over {\it
1143: channels} are \( \psi _{mn}\left( z\right) \)
1144: \begin{equation}
1145: \Psi (r,\varphi ,z)=\sum _{n=0}^{\infty }\sum ^{n}_{m=-n}\psi
1146: _{mn}\left( z\right)
1147: \Phi _{mn}\left( r,\varphi ;z\right) .\label{a2}
1148: \end{equation}
1149: The transverse wavefunftions, $\Phi _{mn}\left( r,\varphi ;z\right)$, depend
1150: only parametrically on $z$ and take the form:
1151:
1152: \begin{subequations}
1153: \begin{equation}
1154: \Phi _{mn}\left( r,\varphi ;z\right) = \left\{ \begin{array}{ll}
1155: A_{mn}\left( z\right) J_{m}\left( k_{mn}\left( z\right) r\right) e^{im\varphi } &
1156: r<\frac{a(z)}{2} \\
1157: \begin{array}{l}
1158: \left[B_{mn}\left( z\right) B_{m}^{(1)}\left( \kappa _{mn}\left(
1159: z\right) r\right)+\right.\\
1160: \left.+C_{mn}\left( z\right) B_{m}^{(2)}\left( \kappa
1161: _{mn}\left( z\right) r\right)\right]e^{im\varphi }
1162: \end{array}& r>\frac{a(z)}{2}
1163: \end{array}\right.,
1164: \end{equation}
1165: \begin{equation}
1166: k_{mn}\left( z\right) = \sqrt{\frac{2m^{\ast }\epsilon _{mn}\left( z\right) }
1167: {\hbar ^{2}}},
1168: \end{equation}
1169: \begin{equation}
1170: \kappa _{mn}\left( z\right) = \sqrt{\frac{2m^{\ast }\left|
1171: \epsilon _{mn}\left( z\right)
1172: -V_{0}\right| }{\hbar ^{2}}}.
1173: \end{equation}
1174: \end{subequations}
1175: Here
1176: \( B_{m}^{(1)}=I_{m} \)
1177: and \( B_{m}^{(2)}= K_{m} \) are appropriate Bessel eigenfunctions \cite{abramovitz}
1178: for $\epsilon_{mn} \leq V_0$ with
1179: \( B_{m}^{(1)}=J_{m} \)
1180: and \( B_{m}^{(2)}= Y_{m} \), for $\epsilon_{mn} > V_0$ . The
1181: coefficients \( A_{mn} \), \( B_{mn} \),
1182: \( C_{mn} \) and energies \( \epsilon _{mn} \) are determined from the
1183: boundary conditions and the normalization of wave functions.
1184:
1185: Substituting Eq.~(\ref{a2}) into Eq.~(\ref{schrod3d}) and integrating over
1186: $r$ and $\varphi$ leads to following coupled ordinary differential
1187: equations for $\psi_{mn}$,
1188: \begin{eqnarray}
1189: \label{schrodkvazi1d}
1190: \psi ^{\prime \prime }_{mn}&+&\left[ k^{2}-k_{mn}^{2}\left( z\right)
1191: +a_{mnn}\left( z\right) \right]
1192: \psi _{mn}+ \\ \nonumber
1193: &+&\sum _{n\neq n^{\prime }}b_{mnn^{\prime }}(z)\psi ^{\prime
1194: }_{mn^{\prime }}+
1195: \sum _{n\neq n^{\prime }}a_{mnn^{\prime }}\left( z\right) \psi _{mn^{\prime }}=0,
1196: \end{eqnarray}
1197: where the coupling coefficients are
1198:
1199: \begin{subequations}
1200: \begin{equation}
1201: a_{mnn^{\prime }}\left( z\right) = 2\pi \int _{0}^{R}\Phi
1202: _{mn}^{\ast }\left( r,\varphi ;z\right)
1203: \frac{\partial ^{2}}{\partial z^{2}}\Phi _{mn^{\prime }}\left(
1204: r,\varphi ;z\right)
1205: r\textrm{d}r,\label{amnn}
1206: \end{equation}
1207: \begin{equation}
1208: b_{mnn^{\prime }}\left( z\right) = 4\pi \int _{0}^{R}\Phi
1209: _{mn}^{\ast }\left( r,\varphi ;z\right)
1210: \frac{\partial }{\partial z}\Phi _{mn^{\prime }}\left( r,\varphi ;z\right) r
1211: \textrm{d}r.\label{bmnn}
1212: \end{equation}
1213: \end{subequations}
1214: The coefficients coupling channels with different \( m \) are
1215: zero due to the orthogonality of \( e^{im\varphi } \) for different
1216: \( m \).
1217:
1218: Note that the Schr\"{o}dinger equation, Eq.~(\ref{schrod3d}), is invariant under the
1219: transformation
1220:
1221: \begin{eqnarray}
1222: \bf {r} & \rightarrow & \Lambda\bf {r},\label{invar} \\
1223: E,V & \rightarrow & \Lambda^{-2}E,\Lambda^{-2}V.\label{transform}
1224: \end{eqnarray}
1225:
1226: \subsection{Extended Hubbard Hamiltonian}
1227:
1228: We consider here the case when
1229: the variation in wire width is small, resulting in small derivatives of the
1230: coefficients in Eqs.~\ref{amnn},
1231: \ref{bmnn}. We consider only electrons with energy below the
1232: second channel and hence Eq.~(\ref{schrodkvazi1d}) reduces to a single equation for
1233: motion in $z$ direction, with the potential
1234:
1235: \begin{equation}
1236: \label{pot1d}
1237: \epsilon \left( z\right) =\epsilon _{00}\left( a\left(z\right)\right) +a^{\prime
1238: 2}\left( z\right)
1239: \tilde{\epsilon }_{00}\left( a\left(z\right)\right) .
1240: \end{equation}
1241:
1242: The first term is the energy of the first channel and the second is
1243: related to \( a_{000} ( z ) \) from
1244: Eq.~(\ref{schrodkvazi1d}). $a^{\prime}(z)$ is the derivative of the
1245: wire diameter with respect to $z$. The second term in Eq.~(\ref{pot1d})
1246: is always positive since
1247:
1248: \begin{equation}
1249: \tilde{\epsilon }_{00}\left( a\right) =\frac{\hbar ^{2}\pi }{m^{\ast
1250: }}\int _{0}^{R}\left( \frac{\partial \Phi _{00}\left( r,\varphi
1251: \right) }{\partial a}\right) ^{2}r\textrm{d}r .
1252: \end{equation}
1253: The potential Eq.~(\ref{pot1d}) is constant for large \( \left| z\right| \)
1254: and set to zero for convenience, i.e.
1255:
1256: \begin{equation}
1257: \epsilon \left( z\right) \rightarrow \epsilon \left( z\right) -\epsilon
1258: \left( \infty \right).
1259: \end{equation}
1260:
1261: \noindent In Fig.~\ref{epsil}(a) $\epsilon_{00}(a)$ and
1262: $\tilde{\epsilon}_{00}(a)$ are presented as a function of wire
1263: diameter. Figs.~\ref{epsil}(b) and (c) show the variation of
1264: one-dimensional potential $\epsilon(z)$ along the wire.
1265:
1266: \begin{figure}[htb]
1267: {\par\centering \resizebox*{0.7\columnwidth}{!}{\rotatebox{0}
1268: {\includegraphics{Fig18.eps}}} \par}
1269:
1270: \caption{\label{epsil}(a) Dependence of \protect\( \epsilon _{00}\protect \) and
1271: \protect\( \tilde{\epsilon }_{00}\protect \)
1272: in Eq.~(\ref{pot1d}) on wire diameter for
1273: \protect\( V_{0}=0.4\textrm{ eV}\protect \).
1274: (b) One dimensional potential Eq.~(\ref{pot1d}) for the wire shape
1275: Eq.~(\ref{oblika2}) and
1276: various values of \protect\( \xi \protect \). Dashed lines
1277: correspond to the contribution \protect\( \epsilon _{00}\protect
1278: \). (c) The same as in (b) but for wire shape Eq.~(\ref{oblika1}).}
1279: \end{figure}
1280: The single-electron Hamiltonian in the single channel approximation then becomes
1281: \begin{equation}
1282: H_1=-\frac{\hbar^2}{2m^\ast}\frac{\textrm{d}^{2}}{\textrm{d}z^{2}} +\epsilon(z).
1283: \label{h1kont}
1284: \end{equation}
1285: This is readily generalized to many electrons,
1286: \begin{eqnarray}
1287: &&H = \sum _{\sigma }\int \psi _{\sigma }^{\dagger }\left( z\right)
1288: \left[
1289: -\frac{\hbar ^{2}}{2m^\ast}\frac{\textrm{d}^{2}}{\textrm{d}z^{2}} \right] \psi _{\sigma
1290: }\left( z\right)
1291: \textrm{d}z+\\ \nonumber
1292: && +\frac{1}{2}\sum _{\sigma ,\sigma ^{\prime }}\int\!\!\! \int \psi
1293: _{\sigma }^{\dagger }\left( z\right)
1294: \psi _{\sigma ^{\prime }}^{\dagger }\left( z^{\prime }\right) U\left(
1295: z,z^{\prime }\right)
1296: \psi _{\sigma ^{\prime }}\left( z^{\prime }\right) \psi _{\sigma
1297: }\left( z\right)
1298: \textrm{d}z\textrm{d}z^{\prime },\label{hamcont1d}
1299: \end{eqnarray}
1300: where
1301: \( \psi _{\sigma }^{\dagger }\left( z\right) \) creates an electron with
1302: spin \( \sigma \) at coordinate \( z \) and
1303: %
1304: \begin{equation}
1305: \label{uab}
1306: U(z_i,z_j)=\frac{e^{2}}{4\pi \epsilon \epsilon _{0}d\left( z_{i},z_{j}\right) }
1307: \end{equation}
1308: with
1309: %
1310: \begin{equation}
1311: \label{averdist}
1312: \frac{1}{d\left( z_{i},z_{j}\right) }=\int \textrm{d}{\bf r}
1313: _{i}\textrm{d}
1314: {\bf r} _{j}\frac{\left| \Phi _{00}\left( {\bf r} _{i};z_{i}\right)
1315: \right| ^{2}\left| \Phi _{00}\left( {\bf r} _{j};z_{j}\right)
1316: \right| ^{2}}{\sqrt[]{\left( z_{i}-z_{j}\right) ^{2}+
1317: \left| {\bf r} _{i}-{\bf r} _{j}\right| ^{2}}}.
1318: \end{equation}
1319: %
1320: The Hamiltonian is further
1321: discretized at points \( z_{j}=j\Delta \), new creation operators
1322: are defined as
1323:
1324: \begin{equation}
1325: c_{j\sigma }^{\dagger }=\sqrt{\Delta }\psi _{\sigma }^{\dagger }\left( z_{j}\right) .
1326: \end{equation}
1327: For sufficiently small
1328: \( \Delta \) the difference formula is justified,
1329:
1330: \begin{equation}
1331: \left[ \frac{\textrm{d}^{2}}{\textrm{d}z^{2}}\psi _{\sigma }\left(
1332: z\right)
1333: \right] _{z=z_{i}}\approx \frac{\psi _{\sigma }\left( z_{i-1}\right) -
1334: 2\psi _{\sigma }\left( z_{i}\right) +\psi _{\sigma }\left( z_{i+1}\right) }
1335: {2\Delta ^{2}},
1336: \end{equation}
1337: and Eq.~(\ref{hamcont1d}) becomes the
1338: discretized extended Hubbard Hamiltonian,
1339: \begin{equation}
1340: \label{hamdis1d1}
1341: H=\sum _{\sigma }H_{1\sigma }+\frac{1}{2}\sum _{i\neq
1342: j}U_{ij}n_{i}n_{j}+
1343: \sum _{i}U_{ij}n_{i\uparrow }n_{j\downarrow },
1344: \end{equation}
1345: where \( H_{1\sigma } \) is single-particle contribution for spin $\sigma$,
1346: \begin{equation}
1347: \label{ham1dis1d1}
1348: H_{1\sigma }=-t\sum _{i}\left( c_{i+1\sigma }^{\dagger }c_{i\sigma
1349: }+c_{i\sigma }^{\dagger }
1350: c_{i+1\sigma }\right) +\sum _{i}\epsilon _{i}n_{i\sigma ,}
1351: \end{equation}
1352: with \( n_{i\sigma }=c_{i\sigma }^{\dagger }c_{i\sigma } \),
1353: \( n_{i}=\sum _{\sigma }n_{i\sigma } \),
1354: hoping parameter,
1355: \begin{equation}
1356: \label{diskr_t}
1357: t=\frac{\hbar ^{2}}{2m^{\ast }\Delta ^{2}},
1358: \end{equation}
1359: and $\epsilon _{i}=2t+\epsilon
1360: \left( z_{i}\right) $.
1361: The effective distance between electrons at $z_i$ and $z_j$ is after
1362: integrating Eq.~(\ref{uab}) over angular variables,
1363: %
1364: \begin{eqnarray}
1365: &&\frac{1}{d\left( z_{i},z_{j}\right) }=8\pi \int _{0}^{R}r_{i}
1366: \textrm{d}r_{i}\int _{0}^{R}r_{j}\textrm{d}r_{j}\left| \Phi _{00}
1367: \left( r_{i};z_{i}\right) \right| ^{2}\times \nonumber\\
1368: &&\quad\quad\times
1369: \left| \Phi \left( r_{j};z_{j}
1370: \right) \right| ^{2}\frac{\textrm{K}\left( -\frac{4r_{i}r_{j}}
1371: {\left( z_{i}-z_{j}\right) ^{2}+\left( r_{i}-r_{j}\right) ^{2}}\right)}
1372: {\sqrt[]{\left( z_{i}-z_{j}\right) ^{2}+\left( r_{i}-r_{j}\right) ^{2}}},
1373: \end{eqnarray}
1374: where $\rm K$ is the complete elliptic integral of the first kind. \( d\left(
1375: z_{i},z_{j}\right) \) can be decomposed into distance
1376: along the wire and effective distance in the lateral direction, \(
1377: \lambda \left( z_{i},z_{j}\right) a_{0} \), i.e.
1378:
1379: \begin{equation}
1380: \frac{1}{d\left( z_{i},z_{j}\right) }=\frac{1}{\sqrt[]{\left(
1381: z_{i}-z_{j}\right)^{2}+
1382: \left[ \lambda \left( z_{i},z_{j}\right) a_{0}\right] ^{2}}}.
1383: \end{equation}
1384: The distance \( \lambda \left( z_{i},z_{j}\right) \) is invariant under
1385: the transformation Eq.~(\ref{invar}), and hence the potential,
1386: Eq.~(\ref{uab}), transforms as
1387: %
1388: \begin{equation}
1389: \label{invar_u}
1390: U\rightarrow \Lambda^{-1}U.
1391: \end{equation}
1392: For convenience we also take into account possible screening with
1393: screening length \( \kappa \), i.e.
1394: %
1395: \begin{equation}
1396: U_{ij}\rightarrow U_{ij}e^{-\frac{\left| z_{i}-z_{j}\right| }{\kappa }}.
1397: \end{equation}
1398: Under the transformation Eq.~(\ref{invar_u}) the screening length should be
1399: multiplied by \( \Lambda \). In Fig.~\ref{lambde} the
1400: parameter $\lambda$ is plotted for some typical cases, showing its dependence
1401: on wire width.
1402:
1403: \begin{figure}[htb]
1404: {\par\centering \resizebox*{0.7\columnwidth}{!}{\rotatebox{0}
1405: {\includegraphics{Fig19.eps}}} \par}
1406:
1407:
1408: \caption{\label{lambde}
1409: (a) Lateral distance vs. separation along the wire. (b)
1410: Electron - electron interaction as a function of longitudinal
1411: separation. In both cases is \protect\( a_{0}=10\textrm{ nm}\protect
1412: \) and \protect\( V_{0}=0.4\textrm{ eV}\protect \) and $a=$const. (c)
1413: Lateral distance at fixed \protect\( z\protect \) vs. wire diameter
1414: at \protect\( V_{0}=0.4\textrm{ eV}\protect \).}
1415: \end{figure}
1416:
1417:
1418:
1419: \section{Two-electron wave functions}
1420:
1421:
1422: Wave function for the case of two electrons
1423: are expressed in terms of a set of operators
1424: \( c_{ij}^{\left( S,S_{z}\right) \dagger } \) creating an electron pair
1425: at sites \( i \) and \( j \) with spin
1426: \( S \) and $z$-component \( S_{z} \), i.e.
1427: \begin{equation}
1428: \label{psi2el}
1429: \left| \psi \right\rangle =\sum _{ij}\psi _{ij}c_{ij}^{\left(
1430: S,S_{z}\right)
1431: \dagger }\left| 0\right\rangle .
1432: \end{equation}
1433: The base states
1434: \begin{subequations}
1435:
1436: \begin{eqnarray}
1437: c_{ij}^{\left( 0,0\right) \dagger }\left| 0\right\rangle & = &
1438: \frac{c_{i\uparrow }^{\dagger }c_{j\downarrow }^{\dagger
1439: }-c_{i\downarrow }^{\dagger }
1440: c_{j\uparrow }^{\dagger }}{\sqrt{2}}\left| 0\right\rangle ,\\
1441: c_{ij}^{\left( 1,1\right) \dagger }\left| 0\right\rangle & =
1442: & c_{i\uparrow }^{\dagger }c_{j\uparrow }^{\dagger }\left| 0\right\rangle ,\\
1443: c_{ij}^{\left( 1,0\right) \dagger }\left| 0\right\rangle & = &
1444: \frac{c_{i\uparrow }^{\dagger }c_{j\downarrow }^{\dagger }+
1445: c_{i\downarrow }^{\dagger }c_{j\uparrow }^{\dagger }}{\sqrt{2}}\left| 0\right\rangle ,\\
1446: c_{ij}^{\left( 1,-1\right) \dagger }\left| 0\right\rangle & =
1447: & c_{i\downarrow }^{\dagger }c_{j\downarrow }^{\dagger }\left| 0\right\rangle .
1448: \end{eqnarray}
1449: \end{subequations}
1450: form a complete set.
1451:
1452: If $\left| \psi \right\rangle$ is a solution of Schr\"odinger equation
1453: \begin{equation}
1454: H\left| \psi \right\rangle =E\left| \psi \right\rangle ,
1455: \end{equation}
1456: then the coefficients \( \widetilde{\psi} _{ij} \) solve the system of
1457: linear equations
1458: \begin{eqnarray}
1459: \label{sistem2el}
1460: &&t\left( \widetilde{\psi }_{i-1j}+\widetilde{\psi
1461: }_{i+1j}+\widetilde{\psi }_{ij-1}+
1462: \widetilde{\psi }_{ij+1}\right)
1463: =\nonumber\\&&\quad\quad=\left( \epsilon _{i}+\epsilon
1464: _{j}+U_{ij}-E\right)
1465: \widetilde{\psi }_{ij},
1466: \end{eqnarray}
1467: where we use compact notation
1468: \begin{equation}
1469: \widetilde{\psi }_{ij}=\frac{1}{\sqrt{2}}\left( \psi _{ij}+(-1)^{S}\psi _{ji}\right) .
1470: \end{equation}
1471: In the basis Eq.~(\ref{psi2el}) the number of electrons on site $i$ is
1472: \begin{equation}
1473: \label{density}
1474: \left\langle \psi \left| n_{i}\right| \psi \right\rangle =2\sum
1475: _{j}\left|
1476: \widetilde{\psi }_{ij}\right| ^{2},
1477: \end{equation}
1478: the current for sites \( i \) and \( i+1 \) is
1479: \begin{equation}
1480: \left\langle \psi \left| j^{i,i+1}\right| \psi \right\rangle =
1481: -\frac{4t\Delta }{\hbar }\textrm{Im}\sum _{j}\widetilde{\psi }_{ij}^{\ast }
1482: \widetilde{\psi }_{i+1j},
1483: \end{equation}
1484: the energy is
1485: \begin{eqnarray}
1486: &&\left\langle \psi \left| H\right| \psi \right\rangle =\nonumber\\
1487: &&\quad\quad-t\sum_{ij} \widetilde{\psi }_{ij}^{\ast }
1488: \left( \widetilde{\psi }_{i+1j}+\widetilde{\psi }_
1489: {i-1j}
1490: +
1491: \widetilde{\psi }_{ij+1}+\widetilde{\psi }_{ij-1}
1492: \right)+\nonumber\\
1493: &&\quad\quad+\sum _{ij}\left( \epsilon _{i}+\epsilon _{j}+
1494: U_{ij}\right) \left| \widetilde{\psi }_{ij}\right| ^{2},
1495: \end{eqnarray}
1496: and the norm of the wave function Eq.~(\ref{psi2el}) is given with
1497: \begin{equation}
1498: \left\langle \psi |\psi \right\rangle =\sum _{ij}\left|
1499: \widetilde{\psi }_{ij}\right| ^{2}.
1500: \end{equation}
1501:
1502: We consider quantum wires which are almost perfect but for which there
1503: is a very weak effective potential, giving rise to bound states.
1504: The cross-sections of these wires are sufficiently small that
1505: the lowest transverse channel approximation is adequate for the energy
1506: range of interest. The smooth variation in
1507: cross-section also guarantees that inter-channel mixing is
1508: negligible. We study here only wires with one weak bulge around \( z=0
1509: \). There should exist single particle bound states of the system and \(
1510: E_{1}^{\left( \alpha \right) } \) is the energy of the state \( \alpha
1511: \). The energy of two electron states is shifted and
1512: defined to be zero, if one electron is
1513: bound and the other at the bottom of the single electron band, i.e.
1514: \begin{equation}
1515: \label{e2dogovor}
1516: E\rightarrow E-E_{1}^{\left( 1\right) }.
1517: \end{equation}
1518: With this definition, the energy of two bound electrons is negative
1519: whereas it is positive when only one electron is bound.
1520:
1521: \section{Hartree-Fock approximation}
1522:
1523: Here we neglect the Coulomb interaction between electrons
1524: ($\gamma=0$). In the ground state both electrons are in
1525: the same state \( \left| \varphi \right\rangle \) and
1526: the singlet wavefunction is
1527: %
1528: \begin{equation}
1529: \left| \psi \right\rangle =\frac{1}{\sqrt{2}}\sum _{ij}
1530: \varphi _{i}\varphi _{j}c_{ij}^{\left( 0,0\right) \dagger }\left| 0\right\rangle
1531: .\label{HFwv}
1532: \end{equation}
1533: For finite $\gamma$ the best one-electron wavefunctions, \( \varphi _{i} \), are
1534: determined by minimizing the energy,
1535: \begin{equation}
1536: \frac{\partial }{\partial \varphi ^{\ast }_{i}}\frac{\left\langle \psi
1537: \left| H\right| \psi \right\rangle }{\left\langle \psi |\psi
1538: \right\rangle }=0.\label{HFmin}
1539: \end{equation}
1540: Which leads to the equation
1541: %
1542: \begin{equation}
1543: \label{varprinc}
1544: \frac{\partial }{\partial \varphi ^{\ast }_{i}}\left\langle
1545: \psi \left| H\right| \psi \right\rangle -\left( E+E_{1}^{\left(
1546: 1\right) }\right) \frac{\partial }{\partial \varphi ^{\ast }_{i}}
1547: \left\langle \psi |\psi \right\rangle =0,
1548: \end{equation}
1549: where we have set
1550: \begin{equation}
1551: \frac{\left\langle \psi
1552: \left| H\right| \psi \right\rangle }{\left\langle \psi |\psi
1553: \right\rangle }=E+E^{(1)}_1,\label{php}
1554: \end{equation}
1555: taking into account the energy shift Eq.~(\ref{e2dogovor}).
1556: The expectation value of energy and the norm is
1557: \begin{equation}
1558: \left\langle \psi \left| H\right| \psi \right\rangle =
1559: 2\left\langle \varphi \left| H_{1}\right| \varphi
1560: \right\rangle \left\langle \varphi |\varphi \right
1561: \rangle +\sum _{ij}U_{ij}\left| \varphi _{i}\right| ^{2}\left| \varphi _{j}\right| ^{2},
1562: \end{equation}
1563: and
1564: \begin{equation}
1565: \left\langle \psi |\psi \right\rangle =\left\langle \varphi |\varphi \right\rangle ^{2}.
1566: \end{equation}
1567: From Eq.~(\ref{varprinc}) follows a system of equations for coefficients \( \varphi _{i} \)
1568: \begin{equation}
1569: \left\langle 0\left| c_{i}H_{1}\right| \varphi \right\rangle +\sum
1570: _{j}U_{ij}\left|
1571: \varphi _{j}\right| ^{2}\varphi _{i}=E_{\textrm{hf}}\varphi _{i},
1572: \end{equation}
1573: where
1574: $\left\langle 0\left| c_{i}H_{1}\right| \varphi \right\rangle=
1575: -t(\varphi_{i-1}+\varphi_{i+i})+\epsilon_i \varphi$ is a one-electron tight-binding
1576: Hamiltonian, $\sum _{j}U_{ij}\left|
1577: \varphi _{j}\right| ^{2}$ is a Hartree potential and
1578: \begin{equation}
1579: E_{\textrm{hf}}=E+E_{1}^{\left( 1\right) }-\frac{\left\langle \varphi
1580: \left| H_{1}\right| \varphi \right\rangle }{\left\langle \varphi |\varphi \right\rangle }
1581: \end{equation}
1582: is the so-called Hartree-Fock energy.
1583: The energy of a bound state is then given by
1584: \begin{equation}
1585: E=2E_{\textrm{hf}}-\sum _{ij}U_{ij}\left| \varphi _{i}\right|
1586: ^{2}\left|
1587: \varphi _{j}\right| ^{2}-E_{1}^{\left( 1\right) },
1588: \end{equation}
1589: where, due to double-counting of the interactions in single electron energies,
1590: \( E_{\textrm{hf}} \) is subtracted.
1591: In Fig.~\ref{dens1} electron density for singlet states with different \protect\(
1592: \gamma \protect \) is presented for a shorter bulge with parameters as in Fig.~\ref{1_24}.
1593:
1594: \begin{figure}[htb]
1595: {\par\centering \resizebox*{0.4\columnwidth}{!}{\includegraphics{Fig20.eps}} \par}
1596: \caption{\label{dens1} Electron density for singlet states with different \protect\(
1597: \gamma \protect \). Full line corresponds to exact results and dashed line to
1598: the Hartree-Fock
1599: approximation. Parameters as in Fig.~\ref{1_24}.}
1600: \end{figure}
1601:
1602: In the case of triplet two-electron states,
1603: the single electron states are different, \( \left|
1604: \varphi \right\rangle \) and \( \left| \bar{\varphi }\right\rangle
1605: \). Choosing these to be orthogonal, we get
1606:
1607: \begin{equation}
1608: \left| \psi \right\rangle =\sum _{ij}\varphi _{i}
1609: \bar{\varphi }_{j}c_{ij}^{\left( 1,1 \right) \dagger }\left| 0\right\rangle .
1610: \end{equation}
1611: The energy is now
1612:
1613: \begin{eqnarray}
1614: \left\langle \psi \left| H\right| \psi \right\rangle &=&
1615: \left\langle \varphi \left| H_{1}\right| \varphi
1616: \right\rangle \left\langle \bar{\varphi }|\bar{\varphi }\right\rangle
1617: \\ \nonumber
1618: &+&
1619: \left\langle \bar{\varphi }\left| H_{1}\right| \bar{\varphi
1620: }\right\rangle
1621: \left\langle \varphi |\varphi \right\rangle
1622: +\frac{1}{2}\sum
1623: _{ij}U_{ij}
1624: \left| \varphi _{i}\bar{\varphi }_{j}-\bar{\varphi }_{i}\varphi _{j}\right| ^{2},
1625: \end{eqnarray}
1626: and the norm is
1627: \begin{equation}
1628: \left\langle \psi |\psi \right\rangle =\left\langle \varphi
1629: |\varphi \right\rangle \left\langle \bar{\varphi }|\bar{\varphi }\right\rangle ,
1630: \end{equation}
1631: The system of equations for the coefficients \( \varphi _{i} \) (and
1632: equivalent for \( \bar{\varphi }_{i} \)) is
1633: \begin{equation}
1634: \left\langle 0\left| c_{i}H_{1}\right| \varphi _{i}\right\rangle +
1635: \sum _{j}U_{ij}\left| \bar{\varphi }_{j}\right| ^{2}\varphi _{i}-
1636: \sum _{j}U_{ij}\bar{\varphi }_{j}^{\ast }\bar{\varphi }_{i}\varphi _{j}=
1637: E_{\textrm{hf}}\varphi _{i},\label{d13}
1638: \end{equation}
1639: where
1640: \begin{equation}
1641: E_{\textrm{hf}}=E+E_{1}^{\left( 1\right) }-\frac{\left\langle
1642: \bar{\varphi }\left| H_{1}\right| \bar{\varphi }\right\rangle }
1643: {\left\langle \bar{\varphi }|\bar{\varphi }\right\rangle },
1644: \end{equation}
1645: Eq.~(\ref{d13})
1646: is a single-particle tight-binding Sch\" odinger equation with Hamiltonian
1647:
1648: \begin{equation}
1649: \tilde{H}_{1}=-\sum _{i}\tilde{t}_{ij}\left(
1650: c_{i}^{\dagger }c_{j}+c_{j}^{\dagger }c_{i}\right) +\sum _{i}\tilde{\epsilon }_
1651: {i}c_{i}^{\dagger }c_{i}
1652: \end{equation}
1653: with potential
1654: \begin{equation}
1655: \tilde{\epsilon }_{i}=\epsilon _{i}+\sum _{j}U_{ij}\left| \bar{\varphi }_{j}\right| ^{2}
1656: \end{equation}
1657: and renormalized hoping parameters
1658: \begin{equation}
1659: \tilde{t}_{ij}=t\delta _{j,i\pm 1}+\sum _{j}U_{ij}\bar{\varphi }_{j}^{\ast }
1660: \bar{\varphi }_{i}.
1661: \end{equation}
1662: The energy of the triplet bound state is then
1663: \begin{equation}
1664: E=E_{\textrm{hf}}+\bar{E}_{\textrm{hf}}-\frac{1}{2}\sum
1665: _{ij}U_{ij}\left|
1666: \varphi _{i}\bar{\varphi }_{j}-\bar{\varphi }_{i}\varphi _{j}\right| ^{2}-E_{1}^{
1667: \left( 1\right) }.
1668: \end{equation}
1669:
1670: In Fig.~\ref{dens2} the electron density for singlet (a) and
1671: for triplet (b) states are shown for different
1672: \protect\(\gamma \protect \). Other parameters are taken as in the case
1673: of the longer bulge, Fig.~\ref{1_06}. As discussed in the text relating to
1674: Fig.~\ref{1_06}, Hartree-Fock approximation for the singlet is less reliable since the
1675: Coulomb repulsion is stronger due to both electrons being in the same state. Indeed,
1676: for $\gamma ~ \sim 0.7$ the Hartree-Fock approximation does not yield the bound state
1677: found in the exact result.
1678:
1679:
1680: \begin{figure}[htb]
1681: {\par\centering \resizebox*{0.9\columnwidth}{!}{\rotatebox{0}
1682: {\includegraphics{Fig21.eps}}} \par}
1683: \caption{\label{dens2}
1684: Electron singlet (a) and triplet (b) state density for various \protect\( \gamma
1685: \protect \). Parameters are as in Fig.~\ref{1_06}. }
1686: \end{figure}
1687:
1688:
1689: \acknowledgments
1690: The authors wish to acknowledge N.J. Appleyard, A.V. Khaetskii, C.J.
1691: Lambert, M. Pepper and K.J. Thomas for helpful discussions. This work was
1692: supported by the EU and the MoD.
1693:
1694: \begin{references}
1695:
1696: % 1
1697: \bibitem{walther92} {M. Walther, E. Kapon, D.M. Hwang, E. Colas, and
1698: L. Nunes, Phys. Rev. B {\bf 45}, 6333 (1992); M. Grundmann {\em et
1699: al.}, Semicond. Sci. Tech. {\bf 9}, 1939 (1994); R. Rinaldi {\em et
1700: al.}, Phys. Rev. Lett. {\bf 73}, 2899 (1994).}
1701:
1702: % 2
1703: \bibitem{ramvall97} {P. Ramvall {\em et al.}, Appl. Phys. Lett. {\bf
1704: 71}, 918 (1997).}
1705:
1706: % 3
1707: \bibitem{yacoby96} {A. Yacoby {\em et al.}, Phys. Rev. Lett. {\bf 77},
1708: 4612 (1996).}
1709:
1710: % 4
1711: \bibitem{kristensen98} {A. Kristensen {\em et al.}, Contributed paper
1712: for ICPS24, Jerusalem, August 2-7, 1998.}
1713:
1714: % 5
1715: \bibitem{wees88} {B.J. van Wees {\em et al.}, Phys. Rev. Lett. {\bf
1716: 60}, 848 (1998).}
1717:
1718: % 6
1719: \bibitem{wharam88} {D.A. Wharam {\em et al.}, J. Phys. C: Solid state
1720: Phys. {\bf 21}, L209 (1988).}
1721:
1722: % 7
1723: \bibitem{houten92} {H. van Houten, C.W.J. Beenakker, and B.J. van
1724: Wees, in {\it Semiconductors and Semimetals}, Vol. {\bf 35}, edited by
1725: M.A. Reed (Academic Press, New York, 1992).}
1726:
1727: % 8
1728: \bibitem{thomas96} {K.J. Thomas {\em et al.}, Phys. Rev. Lett. {\bf
1729: 77}, 135 (1996); Phys. Rev. B {\bf 58}, 4846 (1998); Phys. Rev. B {\bf
1730: 59}, 12252 (1999).}
1731:
1732: % 9
1733: \bibitem{liang} {C.-T. Liang {\em et al.}, Phys. Rev. B {\bf 60}, 10687
1734: (1999).}
1735:
1736: % 10
1737: \bibitem{pyshkin} {K.S. Pyshkin {\em et al.}, Phys. Rev. B {\bf 62}, 15842
1738: (2000).}
1739:
1740: % 11
1741: \bibitem{kaufman99} {D. Kaufman {\em et al.}, Phys. Rev. B {\bf 59},
1742: R10433 (1999).}
1743:
1744: % 13
1745: \bibitem{nuttinck} {S. Nuttinck {\em et al.}, Jpn. J. Appl. Phys. {\bf 39},
1746: 655 (2000); K. Hashimoto {\em at al.}, Jpn. J. Appl. Phys. {\bf 40},
1747: 3000 (2001).}
1748:
1749: % 14
1750: \bibitem{kristensen00} {A. Kristensen {\em et al.}, Phys. Rev. B {\bf 62},
1751: 10950 (2000).}
1752:
1753: % 15
1754: \bibitem{liang00} {C.-T. Liang {\em et al.}, Phys. Rev. B {\bf 61}, 9952
1755: (2000).}
1756:
1757: % 12
1758: \bibitem{kondo2002} {C.M. Cronenwett {\em et al.},
1759: Phys. Rev. Lett. {\bf 88}, 226805 (2002)}
1760:
1761: % 16
1762: \bibitem{appleyard00} {N.J. Appleyard {\em et al.}, Phys. Rev. B {\bf
1763: 62}, R16275 (2000).}
1764:
1765: % 17
1766: \bibitem{maslov95} {D.L. Maslov, Phys. Rev. B {\bf 52}, R14368,
1767: 1995}.
1768:
1769: % 18
1770: \bibitem{wang98} {Chuan-Kui Wang and K.-F. Berggren, Phys. Rev. B {\bf
1771: 57}, 4552 (1998).}
1772:
1773: % 19
1774: \bibitem{fasol94} {G. Fasol and H. Sakaki, Jpn. J. Appl. Phys. {\bf
1775: 33}, 879 (1994).}
1776:
1777: % 20
1778: \bibitem{rejec002d} {T. Rejec, A. Ram\v sak, and J.H. Jefferson,
1779: J. Phys., Condens. Matter {\bf 12}, L233 (2000).}
1780:
1781: % 21
1782: \bibitem{rejec003d} {T. Rejec, A. Ram\v sak, and J.H. Jefferson,
1783: Phys. Rev. B {\bf 62}, 12985 (2000).}
1784:
1785: % 22
1786: \bibitem{rejec02} {T. Rejec, A. Ram\v sak, and J.H. Jefferson,
1787: Phys. Rev. B {\bf 65}, 235301 (2002).}
1788:
1789: % 23
1790: \bibitem{flambaum00} {V.V. Flambaum and M.Yu. Kuchiev, Phys. Rev. B
1791: {\bf 61}, R7869 (2000).}
1792:
1793: % 24
1794: \bibitem{landau} {H. Bruus, V. Cheianov, and K. Flensberg, Physica E
1795: {\bf 10}, 97 (2001).}
1796:
1797: % 25
1798: \bibitem{sushkov1} {O.P. Sushkov, Phys. Rev. B {\bf 64}, 155319
1799: (2001)}
1800:
1801: % 26
1802: \bibitem{tokura} {Y. Tokura and A. Khaetskii, Physica E {\bf 12}, 711 (2002).}
1803:
1804: % 27
1805: \bibitem{meir2002} {Y. Meir, K. Hirose, and N.S. Wingreen, \\cond-mat/0207044.}
1806:
1807: % 28
1808: \bibitem{meirav91} {U. Meirav {\em et al.}, Z. Phys. B {\bf 85}, 357
1809: (1991).}
1810:
1811: % 29
1812: \bibitem{rejecand} {T. Rejec, A. Ram\v sak, and J.H. Jefferson, in
1813: {\it Kondo Effect and Dephasing in Low-Dimensional Metallic Systems},
1814: edited by V. Chandrasekhar, C. Van Haesendonck, and A. Zawadowski,
1815: NATO ARW, Ser. II, Vol. {\bf 50} (Kluwer, Dordrecht, 2001).}
1816:
1817: % 30
1818: \bibitem{anderson61} {P.W. Anderson, Phys. Rev. {\bf 124}, 41 (1961);
1819: G.D. Mahan, {\it Many-Particle Physics}, Plenum
1820: Press, New York (1990).}
1821:
1822: % 31
1823: \bibitem{jauregui} {K. Jauregui, W. H\" ausler, and B. Kramer, Euriphys.
1824: Lett. {\bf 24} 581 (1993); D.L.J. Tipton, PhD Thesis, King's College London (2001).}
1825:
1826: % 32
1827: \bibitem{landauQM} {L.D. Landau and E.M. Lifshitz, {\it Quantum
1828: Mechanics} (Pergamon Press, Oxford, 1977).}
1829:
1830: % 33
1831: \bibitem{oppenheimer28} {J.R. Oppenheimer, Phys. Rev. {\bf 32}, 361
1832: (1928); N.F. Mott, Proc. Roy. Soc. A {\bf 126}, 259 (1930).}
1833:
1834: % 34
1835: \bibitem{abramovitz} {M. Abramowitz and I.A. Stegun, Handbook of
1836: mathematical functions, (Dover publications, Inc., New York).}
1837:
1838: % 35
1839: \bibitem{landauer57} {R. Landauer, IBM J. Res. Dev. {\bf 1}, 223
1840: (1957); {\bf 32}, 306 (1988); M. B\"{u}ttiker, Phys. Rev. Lett. {\bf
1841: 57}, 1761 (1986).}
1842:
1843: % 36
1844: \bibitem{ramsak98} {A. Ram\v sak, T. Rejec, and J.H. Jefferson,
1845: Phys. Rev. B {\bf 58}, 4014 (1998).}
1846:
1847: % 37
1848: \bibitem{patel91} {N.K. Patel {\em et al.}, Phys. Rev. B {\bf 44},
1849: 13549 (1991); K.J. Thomas {\em et al.}, Phil. Mag. B {\bf 77}, 1213
1850: (1998).}
1851:
1852: % 38
1853: \bibitem{lieb62} {E. Lieb and D. Mattis, Phys. Rev. {\bf 125}, 164
1854: (1962).}
1855:
1856: % 39
1857: \bibitem{anddot} {L.I. Glazman and M.E. Raikh, JETP Lett. {\bf 47},
1858: 452 (1988); T.K. Ng and P.A. Lee, Phys. Rev. Lett {\bf 61}, 1768 (1988).}
1859:
1860: % 40
1861: \bibitem{jeff} {C.E. Creffield, J.H. Jefferson, Sarben Sarkar, and
1862: D.L.J. Tipton, Phys. Rev. B {\bf 62}, 7249 (2000).}
1863:
1864: % 41
1865: \bibitem{hewsonbook} {see, e.g., A.C. Hewson, {\it The Kondo problem
1866: to heavy fermions}, (Cambridge university press, Cambridge, 1997).}
1867:
1868: % 42
1869: \bibitem{rejecram} {T. Rejec and A. Ram\v sak, unpublished.}
1870:
1871: % 43
1872: \bibitem{boese01} {D. Boese and R. Fazio, Europhys. Lett. {\bf 56},
1873: 576 (2001).}
1874:
1875: \end{references}
1876:
1877: \end{document}
1878: