1: % submitted to PRB on 21/4/01
2: %\documentstyle[12pt,epsfig]{article}
3: \documentclass[12pt]{article}
4: \usepackage{epsfig,amssymb}
5:
6: \setlength{\topmargin}{0.2cm}
7: \raggedbottom
8: \abovedisplayskip=3mm
9: \belowdisplayskip=3mm
10: \abovedisplayshortskip=0mm
11: \belowdisplayshortskip=2mm
12: \setlength{\baselineskip}{16pt}
13: \setlength{\evensidemargin}{0pt}
14: \setlength{\oddsidemargin}{0in}
15: \setlength{\parskip}{0.13cm}
16: \setlength{\textwidth}{16truecm}
17: \setlength{\textheight}{20cm}
18: \baselineskip=20pt
19:
20: \newcommand\beq{\begin{equation}}
21: \newcommand\eeq{\end{equation}}
22: \newcommand\bea{\begin{eqnarray}}
23: \newcommand\eea{\end{eqnarray}}
24: \newcommand{\nonum}{\nonumber}
25: \newcommand\sqpi{\sqrt{\pi}}
26: \newcommand\px{\partial_x}
27: \newcommand\pt{\partial_\tau}
28: \newcommand\bo{{\bar \omega}}
29: \newcommand\on{{\bar \omega}_n}
30: \newcommand\tphi{\tilde\phi}
31: \newcommand\tchi{\tilde\chi}
32: \newcommand\ua{\uparrow}
33: \newcommand\da{\downarrow}
34: \newcommand\phiou{\phi_{1\uparrow}}
35: \newcommand\phiod{\phi_{1\downarrow}}
36: \newcommand\phitu{\phi_{2\uparrow}}
37: \newcommand\phitd{\phi_{2\downarrow}}
38: \newcommand\tphiou{{\tilde \phi}_{1\uparrow}}
39: \newcommand\tphiod{{\tilde \phi}_{1\downarrow}}
40: \newcommand\tphitu{{\tilde \phi}_{2\uparrow}}
41: \newcommand\tphitd{{\tilde \phi}_{2\downarrow}}
42: \newcommand\ttheta{\tilde\theta}
43: \newcommand\tk{\tilde k}
44:
45: \begin{document}
46:
47: \centerline{\Large Conductance through contact barriers of a finite length
48: quantum wire}
49:
50: \vskip .5 true cm
51: \centerline{\bf Siddhartha Lal$^1$, Sumathi Rao$^2$ and Diptiman Sen$^1$}
52: \vskip .5 true cm
53:
54: \centerline{\it $^1$ Centre for Theoretical Studies}
55: \centerline{\it Indian Institute of Science, Bangalore 560012, India}
56: \vskip .5 true cm
57:
58: \centerline{\it $^2$ Harish-Chandra Research Institute}
59: \centerline{\it Chhatnag Road, Jhusi, Allahabad 211019, India}
60: \vskip .5 true cm
61:
62: \begin{abstract}
63:
64: We use the technique of bosonization to understand a variety of recent
65: experimental results on the conductivity of a quantum wire. The quantum wire
66: is taken to be a finite-length Luttinger liquid connected on two sides to
67: semi-infinite Fermi liquids through contacts. The contacts are modeled as
68: (short) Luttinger liquids bounded by localized one-body potentials. We use
69: effective actions and the renormalization group to study the effects of
70: electronic interactions within the wire, the length of the wire,
71: finite temperature and a magnetic field on the conductivity. We explain the
72: deviations of the conductivity away from $2Ne^2 /h$ in wires which are not
73: too short as arising from renormalization effects caused by the repulsive
74: interactions. We also explain the universal conductance corrections observed
75: in different channels at higher temperatures. We study the effects of an
76: external magnetic field on electronic transport through this system and
77: explain why odd and even spin split bands show different renormalizations
78: from the universal conductance values. We discuss the case of resonant
79: transmission and of the possibility of producing a spin-valve which only
80: allows electrons of one value of the spin to go through. We compare our
81: results for the conductance corrections with experimental observations.
82: We also propose an experimental test of our model of the contact regions.
83:
84: \end{abstract}
85: \vskip .5 true cm
86:
87: ~PACS number: 72.10.-d, 85.30.Vw, 71.10.Pm, 73.40.Cg
88:
89: \newpage
90:
91: \section{\bf Introduction}
92:
93: With the rapid advances made in the fabrication of high mobility
94: semiconductor heterojunctions, these systems have provided the
95: setting for the discovery of several new phenomena in quantum
96: systems. Popular examples include mesoscopic systems like
97: quantum dots, quantum wires, and the two-dimensional electron gas samples
98: in which the quantum Hall effects are observed \cite{datta}.
99: In particular, quantum wires are created by the electrostatic gating of
100: two-dimensional electron gases (2DEG) (with typical densities of
101: $n_{2DEG} \sim 0.5 - 6 \times 10^{15} m^{-2}$) in the inversion layer of GaAs
102: heterostructures. These GaAs samples typically have a very high mobility
103: (typically, $\mu \sim 3 - 8 \times 10^{2} m^{2} V^{-1} s^{-1}$) because there
104: is very little disorder in them; the mean free path of an electron in the
105: 2DEG is of the order of $\lambda_{MF} \sim 5 - 20\mu m$.
106: This makes it possible to create ballistic channels a few microns in length
107: for studying electron transport in such wires, especially
108: at low temperatures when the
109: thermal de Broglie wavelength of the electron is comparable to the channel
110: length. Furthermore, since it is possible to maintain a low carrier
111: concentration in these wires, it becomes possible for transport to take place
112: through only a few channels or even a single channel.
113:
114: Thus, several observations [2-12]
115: of the quantization of the conductance in electron
116: transport through such channels have been made over the last two decades. More
117: recently, several new ways have been found to produce such channels and this
118: has led to even more precise experimental studies. This has brought into focus
119: novel aspects of electron transport in such channels, not all of which
120: are understood as yet.
121:
122: Let us first briefly review some of the recent
123: experimental findings in quantum wires.
124: The first striking observation is that of a number of flat plateaus in the dc
125: conductance which are separated by steps of roughly the same value
126: \cite{yacoby}
127: \beq
128: g = N{\tilde g}~\frac{2e^2}{h} ~,
129: \eeq
130: where $N$ denotes positive integers starting from one. The factor
131: ${\tilde g} < 1$ and is found to vary with the length of the quantum
132: wire and the temperature; it has been seen to be as low as $0.75$. In fact,
133: the plateaus tend towards $N(2e^2/h)$ as either the temperature is raised or
134: the length of the quantum wire is shortened \cite{tarucha,yacoby}. This seems
135: to imply a uniform renormalization of the plateau heights for each channel
136: in the quantum wire as a function of wire length and temperature
137: \cite{yacoby,facer,liang2,reilly}. Also,
138: the flatness of the plateaus appears to indicate an insensitivity to the
139: electron density in the channel. Furthermore, kinks are observed on the
140: rise of the conductivity to some of the lowest plateaus. One such kink
141: has been named the ``$0.7$ effect" \cite{thomas,liang,reilly}. These kinks are
142: seen to wash away
143: quickly with increasing temperature \cite{thomas,liang} and an external
144: magnetic field placed in-plane and parallel to the channels \cite{liang}. Also,
145: upon increasing such a magnetic field, a splitting of the conductance steps
146: is observed together with an {\it odd-even} effect of the renormalization of
147: the plateau heights, with the odd and even plateau heights being
148: renormalized by smaller and larger amounts respectively. Finally, at very
149: high magnetic field, another kink is seen to arise near the first
150: spin-split plateau.
151:
152: Several of these experimental observations have found no satisfactory
153: explanation till date. It is the purpose of this work to provide a consistent
154: framework within which most of these observations of transport in quantum
155: wires can be explained. We will rely upon several concepts and techniques that
156: have become popular in the study of interacting mesoscopic systems. These
157: include the concept of bosonization, effective actions and the renormalization
158: group (RG) \cite{haldane,kane,furusaki1}. To be precise, we will employ
159: these techniques and ideas to understand the low energy transport properties
160: of ballistic electrons in a finite quantum wire attached to two semi-infinite
161: Fermi leads \cite{safi,maslov,furusaki}, but with a difference; the contacts
162: of the quantum wire with its leads will themselves be modeled as short
163: quantum wires with junction barriers at either end. The properties of the
164: contacts are unaffected by the
165: external gate voltage which causes the formation of the discrete sub-bands in
166: the quantum wire. The junction barriers will be
167: modeled as localized $\delta$-functions to account for the back-scattering of
168: electrons due to the imperfect coupling between the quantum wire and the 2DEG
169: reservoirs; these barriers will renormalize the conductance as
170: observed in the experiments. We will also study the effect of external electric
171: and magnetic fields on this system. The properties of such a model for the
172: quantum wire will be seen to account for several of the experimental
173: observations mentioned above, as well as predict the possibility of some more
174: interesting observations in future experiments in these systems. It should
175: be stated here that a possible mechanism for some of the experimental
176: observations \cite{yacoby} has been proposed in
177: Ref. \cite{alekseev}; this is based on the
178: anomalously enhanced back-scattering of electrons entering the 2DEG reservoir
179: from the quantum wire due to the formation of Friedel oscillations of the
180: electron density near the edges of the reservoirs, and it neglects
181: interactions between the electrons in the quantum wire. Our model, however,
182: attempts to understand these observations keeping in mind the importance
183: of electron-electron interactions, barrier back-scattering, finite
184: temperature and magnetic field as well as all
185: length scales in the quantum wire system.
186:
187: The paper is organized as follows. In Sec. 2, we discuss the basics of the
188: model outlined above. We show that the model can be described by a
189: $K_L$-$K_C$-$K_W$ Luttinger model
190: \cite{sumathi} with three different interaction parameters
191: in the lead ($K_L$), the contacts ($K_C$) and the wire ($K_W$). By assuming
192: that the electron-electron interactions get screened out rapidly as one goes
193: from one region to another, we get in a natural way the existence of
194: localized barriers at the junctions. We then discuss how our model goes
195: beyond the concept of {\it ideal} contact resistances as studied in the
196: Landauer-Buttiker formalism. In Sec. 3, we study the effective action of
197: our model in the presence of external electric fields after integrating
198: out all bosonic fields except those at the boundaries between the various
199: regions - the leads, the contacts and the wire. Depending on the
200: relative sizes of the contacts and the wire, we define two regimes
201: - a) the quantum wire (QW) limit, where
202: the length of the wire is much greater than the contact, and b) the
203: quantum point contact (QPC) limit, where the length of the contacts is
204: much greater than the length of the wire. We then study the symmetries
205: of the effective action to determine when resonant transmission is possible
206: as a function of a tunable gate voltage. All of this is first done
207: for spinless fermions and subsequently, we give the modifications of
208: the results for spinful fermions. In Sec. 4, we explicitly compute
209: the corrections to the conductance due to the barriers at finite
210: temperature ($T$) and for a finite length of the wire ($l$) and for finite
211: contact lengths ($d$). We compute the frequency dependent Green's
212: functions of the model, in the different frequency regimes and use the Kubo
213: formula to compute the conductance.
214: We also show how these results could have been anticipated from the
215: renormalization group (RG) equations for the barrier strengths. In Sec. 5,
216: we study our model in the presence of an external in-plane
217: magnetic field. Using RG
218: methods, we show that the spin-up and spin-down electrons see different
219: barrier heights at the junctions, and use this idea to explain the
220: odd-even effect mentioned earlier. We also outline all the
221: possible resonances that can be seen under such conditions. We point out
222: the possibility of producing a spin-valve at moderate magnetic fields.
223: In addition, we compute the conductance of our model
224: and discuss its qualitative features as a function of the strength of
225: the magnetic field. In Sec. 6, we compare the features of
226: the conductance expressions obtained with the observations made in
227: various experiments for transport in quantum wires with and without an external
228: magnetic field. We find that our model is applicable to a large class of
229: experiments and gives a unified and qualitatively correct explanation of all
230: of them. In particular, our model gives a possible explanation
231: for the uniform renormalization of all the conductance steps seen in several
232: experiments. We also explain the odd-even effect seen in
233: experiments in the presence of magnetic field. In addition, we propose
234: more precise experimental tests of our model.
235: Finally, we end in Sec. 7, with a summary of all the new results in our paper,
236: and outline further investigations that are possible.
237:
238: \section{\bf The Model}
239:
240: In this section, we will study the Tomonaga-Luttinger liquid (TLL)
241: model \cite{haldane} of a quantum wire of finite length with no disorder,
242: which is connected to the two 2DEG reservoirs modeled as two semi-infinite
243: Fermi leads through two contact regions. The contacts are modeled
244: as short quantum wires with the junctions at either end modeled as
245: $\delta$-function barriers. The inter-electron interactions in the
246: system, and hence the parameter $K$ which characterizes the interactions, vary
247: abruptly at each of the junctions. Hence, we study a
248: $K_L$-$K_C$-$K_W$-$K_C$-$K_W$ model (see Fig. 1). The motivation for the
249: above model is as follows. The electrons in the 2DEG are basically
250: free, and hence, in the equivalent
251: 1D model, they are modeled as semi-infinite leads with
252: Luttinger parameter $K_L=1$. This can be understood as follows: if each end
253: of the quantum wire is approximated by a point, only those
254: electrons in the 2DEG which are in a zero angular momentum state (with respect
255: to the appropriate end) can enter (or leave) the wire. Thus, the wave function
256: of such a state has the radial coordinate as its only variable and we may,
257: therefore, model the 2DEG as noninteracting 1D systems lying
258: on either side of the quantum wire. The electron velocity in the leads
259: $v_L$ is given by the Fermi velocity of the 2DEG electrons in the reservoirs
260: $v_F = \sqrt{2E_{F2D}/m}$. On the other hand, the externally applied
261: gate voltage $V_G$ is applied over a small region and this leads to the
262: formation of several discrete sub-bands where the electrons feel the
263: transverse confinement potential produced by $V_G$. This region is the
264: one-dimensional quantum wire where the density of electrons is controlled by
265: the gate voltage. The lowest energies $E_s$ in each sub-band are given by the
266: discrete energy levels for the transverse confinement potential (and can
267: therefore be shifted by changing $V_G$) \cite{buttiker}. The Fermi energy in
268: the $s^{\rm th}$ channel is given by $E_{F1D} = E_{F2D} - E_s$. A channel is
269: open when $E_{F1D} > 0$; the electron velocity $v_{W}(e)$ in the channel
270: is then related to the 2DEG Fermi velocity $v_F$ by $v_W (e) = {\sqrt {v_F^2 -
271: 2 E_s /m}}$. In this gate voltage constricted region, the electrons will be
272: considered as interacting via a short range (Coulomb-like) repulsion.
273: Thus each discrete channel is modeled by a separate TLL. Let us, for the
274: moment, consider one such channel with an interaction parameter $K_W$ and
275: quasiparticle velocity $v_W$.
276:
277: The contacts represent the regions where the geometry changes from
278: two-dimensional (2D) to 1D. In these regions of
279: changing geometry, interactions between
280: the electrons are likely to be very important; thus we model
281: the contact region as a Luttinger liquid with $K=K_C$. However, the gate
282: voltage $V_G$ is unlikely to affect the properties of the electrons in these
283: regions as the discrete sub-bands form a little deeper inside the wire.
284: We choose different parameters $K_W$ for the wire and $K_C$ for the contact,
285: because it is not obvious that inter-electrons interactions
286: within the quantum wire will be the same as in the contact. The density of
287: electrons in the quantum wire is controlled by the gate voltage, whereas the
288: density of electrons in the contacts is controlled by the density of the 2DEG
289: at or near the Fermi energy. Hence, we expect $K_C$ to be independent of
290: $V_G$, but $K_W$ is dependent on $E_{F1D}$, which, in turn depends on $V_G$.
291: We will also show below that the change in the
292: inter-electron interactions in each of the lead, contact and quantum wire
293: regions gives rise to barrier-like back-scattering of the electrons.
294:
295: Simpler versions of this model (but without junction barriers and without
296: contacts) have been studied by several authors \cite{safi,maslov,furusaki}
297: who found perfect conductance through the TLL channel which is independent
298: of the inter-electron interactions. Perfect conductance is also seen in
299: several of the experiments \cite{wees,tarucha,yacoby,thomas,liang}.
300: In the opposite limit, the model of a finite quantum wire connected to the
301: two reservoirs by tunneling through very large barriers has also been
302: studied \cite{fabrizio}. The idea of modeling 2DEG reservoirs by 1D
303: noninteracting Fermi leads has also earlier been
304: employed in studies of the fractional quantum Hall effect edge states coupled
305: to Fermi liquids through a tunneling term in the Hamiltonian \cite{chamon}.
306: Some studies of disordered quantum wires in such a model (again with perfect
307: junctions) have also been conducted and the corrections to the conductance
308: because of back-scattering impurities found \cite{safilong,maslovlong}.
309: The continuity of
310: the results found in these studies (which have quantum wires of a finite
311: length) with those found earlier for infinite quantum wires \cite{kane} has
312: also been established \cite{furusaki,safilong}.
313:
314: The main difference between our model and the earlier studies of the
315: quantum wire is that here we explicitly model the contacts as short
316: TLL wires bounded by junction barriers on either end and whose properties
317: are unaffected by the gate voltage $V_G$. As we will discuss later, an
318: experiment performed recently \cite{picciotto}
319: has conclusively shown the existence of a
320: region (of an appreciable length of $2-6\mu m$) in between the quantum
321: wire and the 2DEG reservoirs which leads to the back-scattering of 2DEG
322: electrons entering the quantum wire. Furthermore, the idea that the
323: properties of a one-dimensional system are determined by the Fermi
324: energy of the 2DEG reservoirs has been used in Ref. \cite{matveev} to
325: study the quantum point contact. In addition, we assume that the changes
326: in the inter-electron interactions take place abruptly in going from the
327: contacts into the quantum wire and that all inter-electron interactions get
328: screened out very quickly in going from the contacts into the leads. It can,
329: however, be shown that a smoother variation of the interaction parameter $K$
330: upon going from the quantum wire into the contacts and in going from the
331: contacts into the noninteracting leads does not affect any of the transport
332: properties in the $\omega \to 0$ (dc)
333: limit as long as we have no barriers of any kind in the
334: system. We will now show that changes in the inter-electron
335: interactions at the lead-contact and contact-quantum wire junctions give rise
336: to barrier-like terms in the Hamiltonian of the system; the existence of these
337: terms is mentioned briefly in the work of Safi and Schulz \cite{safilong}.
338: This is, however, only one reason why the junctions between the
339: $1$D channel and its leads can cause the back-scattering of electrons; another
340: reason is clearly the change in geometry
341: in going from the 2DEG reservoirs into the 1D channel.
342: This cause for the drop in the conductance of the channel has earlier been
343: studied within the purview of the Landauer-Buttiker formalism; see
344: \cite{yacimry} and references therein.
345:
346: Let us begin by studying the simpler case of a quantum wire (in which electrons
347: are interacting with each other) connected directly to the noninteracting,
348: semi-infinite leads without any intermediate contact regions. Then there is
349: only a single change in inter-electron interactions from zero in the leads to
350: a finite value in the quantum wire. The kinetic part of the Hamiltonian for
351: this system of interacting spinless electrons in the quantum wire when
352: expressed in terms of the bosonic field $\phi(x)$ and its canonically
353: conjugate momentum $\Pi(x)= \partial_t \phi/v_F$ is given by
354: \beq
355: H_0 = \frac{1}{2} \int dx v_F [ \Pi(x)^2 + (\partial_x\phi(x))^2 ] ~,
356: \eeq
357: where $v_F$ is the Fermi velocity of the electrons in the channel. The part of
358: the Hamiltonian which characterizes the short-ranged density-density
359: interactions between the electrons in a 1D channel of length $l$ is given by
360: \beq
361: H_{int} = \int_0^l dx \int_0^l dy ~{\cal U}(x,y) \rho(x) \rho(y) ~,
362: \eeq
363: where ${\cal U}(x,y)$ characterizes the strength of the
364: density-density interactions between the electrons, and $\rho(x)$ is the
365: electronic density at the point $x$. Using a truncated form of the Haldane
366: representation for the electronic density in terms of the bosonic field
367: $\phi(x)$ \cite{haldane}, the density $\rho(x)$ is given by
368: \beq
369: \rho(x) = \frac{1}{\sqpi}\partial_x\bar\phi(x) ~[c_0 +
370: 2c_1\cos(2\sqpi \bar\phi(x))]~,
371: \eeq
372: where $\bar \phi(x) = \phi(x) + k_Fx/\sqpi$, $c_0 = 1$, $c_1 =
373: \Lambda /(2k_F)$, and $k_F$ is the Fermi wave vector. $\Lambda$ is the
374: ultraviolet cutoff ($\Lambda < O(E_{F1D})$);
375: it is the energy limit up to which the linearization of
376: the bands and hence bosonization is expected to be applicable.
377: If we now characterize the short range inter-electron interactions
378: by ${\cal U}(x,y) = {\cal U}_0 \delta(x-y)$,
379: then we can substitute the expressions for the density and the inter-electron
380: interaction into the interaction term in the Hamiltonian. This gives us
381: \bea
382: H_{int} &=& {\cal U}_0 \int_0^l dx ~[\{\partial_x(\frac{c_0}{\sqpi}
383: \bar\phi)\}^2 ~-~2\partial_x(\frac{c_0}{\sqpi}\bar\phi)\partial_x (
384: \frac{c_1}{\sqpi} \sin (2\sqpi \bar\phi)) \nonum \\
385: & & ~~~~~~~~~~~~~~+~ \{\partial_x (\frac{c_1}{\sqpi}
386: \sin (2\sqpi\bar\phi) )\}^2] \nonum \\
387: &=& {\cal U}_0 \int_0^l dx ~[\frac{c_0^2}{\pi}(\partial_x\bar\phi)^2
388: ~+~2\frac{c_0^2k_F}{\pi}\partial_x\phi~+~2\frac{c_0c_1k_F}{\sqpi}
389: \partial_x\sin (2\sqpi \bar\phi) \nonum \\
390: & & ~~~~~~~~~~~~+~2\frac{c_0c_1}{\sqpi}\partial_x\phi\partial_x
391: \sin (2\sqpi \bar\phi) ~+~2\frac{c_1^2}{\pi}(\partial_x\bar\phi)^2
392: (\cos(4\sqpi\bar\phi) +1)] \nonum \\
393: &=& {\cal U}_0 \int_0^l dx ~[(\frac{c_0^2+2c_1^2}{\pi})
394: (\partial_x\phi)^2
395: ~+~2k_F(\frac{c_0^2+2c_1^2}{\pi})\partial_x\phi~+~2\frac{c_0c_1k_F}{\sqpi}
396: \partial_x\sin (2\sqpi\bar\phi) \nonum \\
397: & & ~~~~~~~~~~~~+~2\frac{c_0c_1}{\sqpi}\partial_x\phi\partial_x
398: \sin (2\sqpi\bar\phi) ~+~2\frac{c_1^2}{\pi}(\partial_x\bar\phi)^2
399: \cos(4\sqpi\bar\phi) ].
400: \eea
401: We can now simplify this expression by noting that several of the terms above
402: contain rapidly oscillating factors of $\cos(k_Fx)$ or $\sin(k_Fx)$ which
403: make those terms vanish upon performing the
404: integration (unless we are at very specific fillings of the electron density).
405: Thus, we can ignore the fifth term straightaway. The first term can
406: be added to a similar term in $H_0$ where it renormalizes the velocity and
407: introduces an interaction parameter $K$.
408: The second term is a chemical potential term and that too can be accounted for
409: by shifting the field $\phi$ accordingly. The third term is clearly a boundary
410: term, and it gives us two barrier like terms at $x=0$ and $x=l$, with
411: \bea
412: H_{barrier} &=& 2{\cal U}_0\frac{c_0c_1k_F}{\sqpi} \int_0^l dx ~
413: \partial_x \sin (2\sqpi\bar\phi) \nonum \\
414: &=& 2{\cal U}_0\frac{c_0c_1k_F}{\sqpi}~
415: [\sin (2\sqpi\phi(l)+ 2k_Fl)~-~\sin (2\sqpi\phi(0)) ].
416: \eea
417: Finally, the fourth term can also be rewritten as
418: \bea
419: H_{int,4} &=& 4{\cal U}_0\frac{c_0c_1}{\sqpi} \int_0^l dx ~(\partial_x
420: \phi)^2 \cos (2\sqpi\bar\phi)~+~2{\cal U}_0\frac{c_0c_1k_F}{\sqpi} \int_0^l dx
421: ~\partial_x\sin (2\sqpi\bar\phi) \nonum \\
422: & & -~4{\cal U}_0\frac{c_0c_1k_F^2}{\sqpi} \int_0^l dx ~\cos (2\sqpi\bar\phi)~,
423: \eea
424: in which the first and third terms again vanish because they contain rapidly
425: oscillating factors within the integrals, and the second term adds on to
426: $H_{barrier}$ exactly. All this finally gives us two $\delta$-function
427: barriers at the junctions of the quantum wire with its Fermi liquid leads as
428: \beq
429: H_{barrier} = 4{\cal U}_0 \frac{c_0c_1k_F}{\sqpi}~[ \sin(2\sqpi\phi(l)+
430: 2k_Fl)~-~\sin (2\sqpi\phi(0))] ~.
431: \eeq
432: The extension of the derivation given above to our model with two
433: intermediate contact regions where the inter-electron interactions are
434: ${\cal U}(x,y) = {\cal U}_1 \delta(x-y)$
435: (i.e., different from that in the quantum wire) is straightforward,
436: and it yields four barrier terms: two at the junctions of the
437: contacts with the leads, and two at the junctions of the contacts with the
438: quantum wire. It is also very likely that the inner two barriers are much
439: weaker than the outer two since the change in inter-electron interactions in
440: going from the contacts to the quantum wire is likely to be much smaller than
441: that in going from the contacts to the leads; also the change in geometry at
442: the contact-quantum wire junction is likely to be much more adiabatic. Thus,
443: we will from now on consider the junctions between the wire and the contacts
444: and between the contacts and the leads as local barriers whose heights
445: are determined by several factors, such as the nature of the inter-electron
446: interaction and its screening, and the deviations from adiabaticity in the
447: change in geometry in going from the reservoirs into the contacts or from the
448: contacts into the quantum wire. To be general, we should take these four
449: barriers to have different heights but it is very likely that any asymmetry
450: between the left two and right two contacts will be small. Thus, we
451: can finally write the complete Hamiltonian for the quantum wire of
452: spinless electrons, its contacts and its leads as
453: \bea
454: H &=& (\int_{-\infty}^0+\int_{l+2d}^\infty ) ~
455: dx ~\frac{v_F}{2} [ \Pi(x)^2 + (\partial_x\phi(x))^2] \nonum \\
456: && +(\int_0^d+\int_{l+d}^{l+2d}) ~
457: dx ~\frac{v_C}{2K_C} [\Pi_C(x)^2 + (\partial_x\phi(x))^2] \nonum \\
458: && +\int_d^{l+d} ~dx ~\frac{v_W}{2K_W} [\Pi_W(x)^2 +
459: (\partial_x\phi(x))^2] +V_{LC}\sin (2\sqpi\phi(0)) \nonum \\
460: && +V_{CW}\sin(2\sqpi\phi(d)+ 2k_Fd)
461: +V_{WC}\sin(2\sqpi\phi(l+d)+ 2k_F(l+d)) \nonum \\
462: && +V_{CL}\sin(2\sqpi\phi(l+2d)+ 2k_F(l+2d))~,
463: \label{model}
464: \eea
465: where $\Pi_{C,W}(x) = (1/v_{C,W})\partial_t \phi(x)$.
466: Finally, it is worth commenting here that since our
467: model shows that a quantum wire with no disorder
468: already has back-scattering junctions built into it, the notion of {\it ideal
469: contact resistances} (which are
470: seen in a study of this system using the Landauer-Buttiker formalism and arise
471: from the {\it ideal} connection of the quantum wire to its reservoirs) which
472: are universal in value, $h/2e^2$ to be precise, does not seem
473: to hold true even for the so-called clean quantum wire with adiabatic
474: junctions in the presence of inter-electron interactions within the quantum
475: wire. We will show later that these junction barriers are likely to be weak
476: when the lengths of the quantum wires are quite short or temperatures are
477: not very
478: low, and that the junction barriers are likely to remain weak even after some
479: small renormalization that might take place due to the electron-electron
480: interactions in the quantum wire. Thus, the contact resistances between the
481: wire and the reservoirs due to the junction barriers will be very nearly the
482: universal value quoted above only for very short quantum wires (i.e., quantum
483: point contacts) or when the
484: temperatures are not very low. This is also observed in all the experiments
485: till date \cite{tarucha,yacoby,thomas,liang}.
486:
487: The generalization of the model to spinful fermions is straightforward. For
488: completeness, we give below the Hamiltonian for spinful electrons in a quantum
489: wire connected to external reservoirs through the contacts and junction
490: barriers,
491: \bea
492: H_{spin} &=& \sum_{i=\ua,\da}\Big[ (\int_{-\infty}^0+\int_{l+2d}^\infty) ~
493: dx ~\frac{v_{iF}}{2} [ \Pi_i(x)^2 + (\partial_x\phi_i(x))^2] \nonum \\
494: && +(\int_0^d+\int_{l+d}^{l+2d}) ~dx ~\frac{v_{iC}}{2K_{iC}}[\Pi_{iC}
495: (x)^2 + (\partial_x\phi_i(x))^2] \nonum \\
496: && + \int_d^{l+d} ~dx ~\frac{v_{iW}}{2K_{iW}} [\Pi_{iW}(x)^2 +
497: (\partial_x\phi_i(x))^2] +V_{iLC}\sin (2\sqpi\phi_i(0)) \nonum \\
498: && + V_{iCW}\sin(2\sqpi\phi_i(d)+ 2k_{iF}d)
499: +V_{iWC}\sin(2\sqpi\phi_i(l+d)+ 2k_{iF}(l+d))\nonum \\
500: && + V_{iCL}\sin(\sqpi\phi_i(l+2d)+ 2k_{iF}(l+2d)) \Big].
501: \label{modelspin}
502: \eea
503: Note that we have allowed for independent velocities and
504: interaction strengths for the $\ua$ and $\da$ electrons. This
505: generality will be required when we study the model in the presence of
506: a magnetic field. Finally, let us note the fact that we will be taking into
507: account only the outer two junction barriers (i.e., those at the junctions of
508: the contacts and the leads) in all our subsequent calculations as these are
509: likely to be the more significant junction barriers in the system as
510: long as transport through fully open quantum wires is considered.
511:
512: \section{Effective Actions}
513:
514: In this section, the aim is to obtain an effective action
515: in terms of the fields at the junction barriers for both spinless
516: and spinful electrons. We then analyze the symmetries of the
517: effective action and obtain the resonance conditions.
518:
519: \subsection{The case of spinless electrons}
520:
521: In Sec. 2, it was shown that the screening out of the
522: interactions in the 2DEG leads to a Hamiltonian with junction barriers given
523: in Eq. (\ref{model}). The effective action for this model of spinless
524: electrons can be written as
525: \beq
526: S = S_0 + S_{barrier} + S_{gate} ~,
527: \eeq
528: where we have defined each of the actions separately below.
529: \beq
530: S_0= \int d\tau ~[ \int_{-\infty}^0 dx {\cal L}_1 + \int_0^d dx {\cal
531: L}_2 + \int_d^{l+d} dx {\cal L}_3 + \int_{l+d}^{l+2d} dx {\cal L}_2
532: +\int_{l+2d}^{\infty} dx {\cal L}_1 ]~,
533: \label{s0}
534: \eeq
535: where
536: \beq
537: {\cal L}_1 ~=~ {\cal L} (\phi ; K_L , v_L ), ~~~
538: {\cal L}_2 ~=~ {\cal L} (\phi ; K_{C} , v_{C} ),
539: ~~~{\rm and} ~~~ {\cal L}_3 ~=~ {\cal L} (\phi ; K_{W} , v_{W}) ~,
540: \label{sol}
541: \eeq
542: and we have defined ${\cal L} (\phi ; K,v) = (1/2Kv) (\pt \phi)^2 +(v/2K)
543: (\px \phi)^2$, and used the imaginary time $\tau=it$ notation.
544: \bea
545: S_{barrier} &=& \int d\tau ~\Lambda [ V_1 \cos (2\sqpi\phi_1) + V_1
546: \cos (2\sqpi\phi_4 +2k_FL) \nonum \\
547: && ~~~~~~~~~ +V_2\cos(2\sqpi\phi_2 + 2k_Fd) + V_2\cos(2\sqpi\phi_3 +
548: 2k_F(l+d)) ]~,
549: \eea
550: where we have set $V_{LC} = V_{CL}=V_1 \Lambda$ and
551: $V_{CW}=V_{WC} =V_2 \Lambda$ assuming
552: left-right symmetric barriers ($V_1$ and $V_2$ are dimensionless), and have
553: used $\phi(0,\tau) = \phi_1(\tau)
554: \equiv \phi_1$, $\phi(d,\tau) = \phi_2(\tau) \equiv \phi_2$, $\phi(l+d,\tau)
555: = \phi_3(\tau) \equiv \phi_3$ and $\phi(L,\tau) = \phi_4(\tau) \equiv
556: \phi_4$. The total length of the wire is denoted by $L=l+2d$.
557: We shall henceforth assume that $V_2\ll V_1$ and can be
558: dropped; as we have explained earlier, the inner two
559: barriers are likely to be weaker than the outer two barriers.
560: We also include the coupling of the electrons in the wire
561: to an external gate voltage $V_G$ given by
562: \beq
563: S_{gate} ~=~ V_G \int d\tau \int_d^{l+d} dx \rho(x,\tau) ~
564: =~ V_G\int d\tau {[\phi_3 - \phi_2]\over \sqpi}.
565: \eeq
566: This coupling is necessary because it is the
567: gate voltage which controls the density of electrons in the wire,
568: which, in turn, controls the number of channels in the quantum
569: wire. Experimentally, an external voltage drop across the wire drives
570: the current through the wire, which is measured as a function of
571: the gate voltage or the density of electrons in the wire.
572:
573: Since the Luttinger liquid action is quadratic, the effective action
574: can be obtained in terms of the fields $\phi_i$, $i=1...4$, by integrating
575: out all degrees of freedom except those at the positions of the
576: four junction barriers, following Ref. \cite{kane}.
577: Using the (imaginary time) Fourier transform of the fields
578: \beq
579: \phi_1(\tau) = \sum_{\on} e^{-i\on \tau} \tphi_{1n} (\on),
580: \quad \phi_2(\tau) = \sum_{\on} e^{-i\on \tau} \tphi_{2n} (\on) ~,
581: \eeq
582: we explicitly obtain the $S_0$ part of the effective action; this is
583: presented in Appendix A. (The $\on$ are the Matsubara frequencies which are
584: quantized in multiples of the temperature as $\on = 2\pi n k_B T$).
585: In the high frequency limit, or, equivalently at high temperatures,
586: where $\on d/v_C, \on l/v_W \gg 1$, the effective action reduces to
587: \beq
588: S_{0,eff,high} (\tphi_1,\tphi_2,\tphi_3,\tphi_4)
589: ={K_L+K_C \over 2K_L K_C} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2)
590: +{K_W+K_C \over 2K_W K_C} \sum_{\on} |\on| (\tphi_2^2+\tphi_3^2).
591: \label{highfreq}
592: \eeq
593: In this limit, all the barriers are seen as the sum of individual
594: barriers with no interference. In fact, if we integrate out the two
595: inner fields $\phi_2$ and $\phi_3$, we are just left with
596: \beq
597: S_{0,eff,high}(\tphi_1,\tphi_4)
598: ={K_L+K_C \over 2K_L K_C} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2).
599: \eeq
600: The surprising point to note is that the effective interaction
601: strength $K_{eff} =K_L K_C /(K_L+K_C)$ depends only on the
602: interaction strengths in the contacts and in the leads (where there are
603: no interactions), and not on the interaction strength in the wire!
604: Furthermore, since the gate voltage $V_G$ couples only to the inner fields
605: $\phi_2$ and $\phi_3$ and these two fields are completely decoupled
606: from the outer fields $\phi_1$ and $\phi_4$ in ${\cal L}_{0,eff,high}
607: (\phi_1,\phi_2,\phi_3,\phi_4)$ above, integrating out $\phi_2$ and
608: $\phi_3$ does not lead to any gate voltage term in the final effective
609: action in this temperature regime.
610:
611: Depending on whether $d\gg l$ or $l\gg d$, we can have two possible
612: scenarios of intermediate regimes, each with two
613: crossovers. We can express all our lengths in terms of equivalent
614: temperatures by defining $v_C/ d = k_B T_d$ and $v_W/l = k_B T_l$.
615: So the high temperature limit defined above is just $T\gg T_d,T_l$.
616:
617: \begin{itemize}
618:
619: \item{}
620:
621: Let us first consider the quantum wire limit where $l\gg d$.
622:
623: In the intermediate frequency (or temperature) regime of
624: $T_l \ll T \ll T_d$, the action becomes
625: \bea
626: S_{0,eff,int} (\tphi_1,\tphi_2,\tphi_3,\tphi_4) &=&
627: {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) +
628: {1\over 2K_W} \sum_{\on} |\on| (\tphi_2^2+\tphi_3^2) \nonum \\
629: & & + {U_C \over 2} \sum_{\on} [ (\tphi_1- \tphi_2) ^2
630: +(\tphi_3- \tphi_4)^2] + S_{gate} ,
631: \eea
632: where $U_C = v_C/(K_Cd)$ is an energy whose significance will
633: become clear shortly. As the action is quadratic, we can integrate
634: out $\tphi_2$ and $\tphi_3$ to be left with an action dependent only
635: on $\tphi_1$ and $\tphi_4$ as given by
636: \bea
637: & & S_{0,eff,int} (\tphi_1,\tphi_4) \nonum \\
638: &=& {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) \nonum \\
639: & & + ~~{1\over 2K_W} \sum_{\on} |\on|
640: [\frac{U_C^2}{A^2}(\tphi_1^2+\tphi_4^2) +
641: \frac{2{\tilde V}_G U_C}{A^2}(\tphi_1-\tphi_4) ] \nonum \\
642: & & + ~~\frac{U_C}{2} \sum_{\on} \Bigl[ [ (\frac{U_C}{A} - 1)\tphi_1 +
643: \frac{{\tilde V}_G}{A}]^2 + [ (\frac{U_C}{A} - 1)\tphi_4 -
644: \frac{{\tilde V}_G}{A}]^2 \Bigr] \nonum \\
645: & & + \int d\tau \frac{{\tilde V}_G U_C}{A}(\phi_4 - \phi_1) ~,
646: \eea
647: where $A = U_C + |\on|/K_W$ and ${\tilde V}_G = V_G/\sqrt{\pi}$. We can
648: approximate $A$ by $U_C$ which is justified in the
649: intermediate regime as $T \ll T_d$ and $K_C \sim K_W$.
650: Then we are finally left with the expression
651: \beq
652: S_{0,eff,int} (\tphi_1,\tphi_4) = {{K_L+K_W}\over{2K_LK_W} } \sum_{\on} |\on|
653: (\tphi_1^2+\tphi_4^2) ~+~ \int d\tau {\tilde V}_G(\phi_4 - \phi_1) ~.
654: \label{intertemp}
655: \eeq
656:
657: \item{}
658:
659: Now, we consider the QPC limit where $d \gg l$.
660:
661: In the regime where $T_d \ll T \ll T_l$, the action becomes
662: \bea
663: S_{0,eff,int} (\tphi_1,\tphi_2,\tphi_3,\tphi_4) &=&
664: {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) \nonum \\
665: & & + {1\over 2K_C} \sum_{\on} |\on| (\tphi_1^2+\tphi_2^2+
666: \tphi_3^2+\tphi_4^2) \nonum \\
667: & & + {U_W \over 2}\sum_{\on} (\tphi_2- \tphi_3)^2 + S_{gate} ,
668: \eea
669: where $U_W = v_W/(K_Wl)$ is again a frequency independent
670: energy. As before, we can integrate out $\tphi_2$ and $\tphi_3$ to be left
671: with an action dependent only on $\tphi_1$ and $\tphi_4$ given by
672: \beq
673: S_{0,eff,int} (\tphi_1 ,\tphi_4 ) ~=~ S_{0,eff,high} (\tphi_1 ,\tphi_4 )~.
674: \eeq
675: Thus there is no difference between the intermediate and high energy scales in
676: the QPC limit because the gate voltage is applied over too short a length to
677: affect the conductance even at intermediate temperatures.
678:
679: \end{itemize}
680:
681: Finally, in the low frequency limit where $\on \ll v_W/l , v_c/d$
682: (i.e., $T \ll T_d$ and $T \ll T_l$), $S_0$ reduces to
683: \bea
684: S_{ 0,eff,low} (\tphi_1,\tphi_2,\tphi_3,\tphi_4)
685: &=& {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) + {U_C\over 2}
686: \sum_{\on} [ (\tphi_1- \tphi_2) ^2 +(\tphi_3- \tphi_4)^2] \nonum \\
687: && ~~~+{U_W\over 2} \sum_{\on} (\tphi_2- \tphi_3)^2 ~.
688: \eea
689: Since the action is still quadratic, it is possible to integrate out
690: the two inner fields $\tphi_2$ and $\tphi_3$ and get the effective action
691: wholly in terms of the $\tphi_1$ and $\tphi_4$ fields, remembering
692: however, to also include the gate voltage term which couples to the inner
693: fields. After doing out, we are left with the full effective action as
694: \bea
695: S_{eff,low} (\phi_1,\phi_4) &=& S_{0,eff,low} +
696: S_{gate}+ S_{barrier} \nonum \\
697: &=& {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) +
698: {U_C U_W \over 2(U_C+2U_W)}\sum_{\on} (\tphi_1-\tphi_4
699: -{{\tilde V}_G \over U_W})^2. \nonum \\
700: &&
701: \eea
702: In this limit, the full action can be rewritten in terms of a
703: ``current" field $\chi(\tau)$ and a ``charge" field $n(\tau)$
704: (and their Fourier transforms $\tilde \chi$ and $\tilde n$) defined as
705: \beq
706: \chi(\tau) = {\phi_1 +\phi_4 \over 2 }+ {k_F L\over 2\sqpi}, \quad {\rm
707: and} \quad n(\tau) = {\phi_4 - \phi_1 \over \sqpi} + {k_F L\over \pi} ~.
708: \eeq
709: The action is given by
710: \bea
711: S_{eff,low} &=& S_0 + S_{barrier} +S_{gate} \nonum \\
712: &=& {1\over 2K_L} \sum_{\on} |\on| [(\tchi - {k_F
713: L\over 2\sqpi})^2 + {\pi\over 4} (\tilde n -{k_F L\over \pi})^2] \nonum \\
714: && + \int d\tau [{U_{eff} \over 2} (n-n_0)^2 + V_1 \Lambda
715: \cos (2\sqpi \chi) \cos (\pi n) ] ~,
716: \label{lowfreq}
717: \eea
718: where $n_0 = (2k_C d + k_W l)/\pi - V_G/(\pi^{3/2}U_W)$ and
719: $U_{eff} =\pi U_C U_W / (U_C+2U_W)$. The derivation of the effective
720: action in this limit follows the method outlined
721: in Ref. \cite{kane}; however, their derivation was for a uniform wire with a
722: single interaction parameter $K$, whereas we have three interaction
723: parameters here. $K_W$ acts only within the quantum wire delimited by the two
724: contact regions, $K_C$ acts within the contact region, and $K_L =1$ outside
725: the contact and wire region.
726: The current field is interpreted as the number of particles transferred
727: across the two barriers, and the charge field is the number of particles
728: between the barriers. In the low frequency limit, the two barriers are
729: clearly being seen as one coherent object with charge and current degrees of
730: freedom. Since
731: in the limit of weak barriers, $V_1 \ll U_{eff}$, the action is minimized when
732: $n=n_0$, we can integrate out the quadratic fluctuations of $n-n_0$ to obtain
733: an effective action only in terms of the single variable $\chi$; we obtain
734: \beq
735: S_{low} = \int d\tau ~[~2V_1 \Lambda \cos (2\sqpi\chi) \cos (\pi n_0)
736: -{2(V_1 \Lambda)^2 \over U_{eff}} (\pi \cos (2\sqpi \chi) \sin (\pi n_0))^2
737: + \cdots ~]~.
738: \eeq
739: The first term in this effective action is precisely the
740: same term that is obtained for the impurity potential for
741: a single barrier in terms of the variable $\chi$.
742:
743: In the low frequency limit, from Eq. (\ref{lowfreq}), we see that the
744: effective action contains extra terms due to the interference between the
745: two barriers. It is easy to check that this effective action is invariant
746: under $\chi\rightarrow \chi+\sqrt{\pi}, n \rightarrow n$; this is
747: the same symmetry which exists for a single barrier \cite{kane}, and it
748: corresponds to the transfer of a single electron across the two barriers, and
749: hence in our model, from the left lead to the right lead. But when $n_0$ is
750: precisely equal to a half-odd-integer, the action is also invariant
751: under $\chi \rightarrow \chi+\sqpi/2, n\rightarrow 2n_0-n$. As explained in
752: Ref. \cite{kane}, this corresponds to the `transfer of half an electron
753: across the wire' accompanied by a change in the charge state of the
754: wire. In the language of scattering, this corresponds to resonant tunneling
755: through a virtual state. Within the TLL theory, this is the explanation
756: of the Coulomb blockade phenomenon, which leads to steps or plateaus
757: in the current versus gate voltage for quantum dots.
758:
759: \subsection{The case of spinful electrons}
760:
761: The spinless electron model is expected to be valid for real systems
762: in the presence of strong magnetic fields which completely
763: polarizes all the electrons in a given channel.
764: However, for a real system without magnetic
765: field, or in the presence of weak magnetic fields which do not
766: polarize all the electrons, one has to study a model of electrons
767: with spin. We shall study such a model here. Its modification due
768: to the presence of a magnetic field will be studied in Sec. 5.
769: The basic action of the model is a straightforward extension of
770: the model for spinless fermions given in Eqs. (\ref{s0}) and (\ref{sol}),
771: with $\phi_{\ua}$ denoting the spin up boson and $\phi_{\da}$ denoting
772: the spin down boson. However, since the Coulomb interaction couples the spin
773: up and spin down fermions (for instance, remember the Hubbard term which is
774: $U\sum_i n_{i\ua} n_{i\da}$), the Luttinger model is diagonal only in terms of
775: the charge and spin fields $\phi_\rho =(\phi_{\ua} + \phi_{\da})/{\sqrt 2}$
776: and $\phi_\sigma =(\phi_{\ua} - \phi_{\da})/{\sqrt 2}$. In terms of
777: these fields $S_0$ is given by
778: \beq
779: S_0 = \int dt ~[ \int_{-\infty}^0 dx {\cal L}_1
780: + \int_0^d dx {\cal L}_2 + \int_d^{l+d} dx {\cal L}_3
781: + \int_{l+d}^{l+2d} dx {\cal L}_2 + \int_{l+2d}^{\infty} dx {\cal L}_1 ]~,
782: \label{s0spin}
783: \eeq
784: where
785: \bea
786: {\cal L}_1 ~&=&~ {\cal L} (\phi_\rho ; K_{L\rho} , v_L ) ~+~
787: {\cal L} (\phi_\sigma ; K_{L\sigma} , v_L ) ~, \nonum \\
788: {\cal L}_2 ~&=&~ {\cal L} (\phi_\rho ; K_{C\rho} , v_{C\rho} ) ~+~
789: {\cal L} (\phi_\sigma ; K_{C\sigma} , v_{C\sigma} ) ~, \nonum \\
790: {\cal L}_3 ~&=&~ {\cal L} (\phi_\rho ; K_{W\rho} , v_{W\rho} ) ~
791: +~ {\cal L} (\phi_\sigma ; K_{W\sigma} , v_{W\sigma} ) ~,
792: \eea
793: with $\cal L$ defined as before, ${\cal L} (\phi ; K,v) = (1/2Kv)
794: (\pt \phi)^2 +(v/2K) (\px \phi)^2$,
795: $K_{L\rho} = K_{L\sigma}=1$ are the interaction parameters in
796: the two external leads, and $K_{C/W\rho}$ and $K_{C/W\sigma}$ are the
797: interaction parameters in the contacts and wire respectively.
798: As for the spinless case, we include junction barrier terms of the form
799: \beq
800: S_{barrier} = \int d\tau \sum_{i=\da,\ua} ~V_i \Lambda ~[\cos (2\sqpi\phi_i(0,
801: \tau) ) +\cos (2\sqpi\phi_i(L,\tau) +2k_F L)] ,
802: \eeq
803: at the junctions of the contacts and the leads, and we assume that the
804: barriers at the junctions of the wire and the contact are weak and can
805: be ignored. The barrier action can be re-expressed in terms of the
806: diagonal fields of the model (using $V_{\ua} = V_{\da} =V$) as
807: \beq
808: S_{barrier} = V \Lambda~\int d\tau [\cos (\sqrt{2\pi}\phi_{1\rho}) \cos (
809: \sqrt{2\pi} \phi_{1\sigma}) + \cos (\sqrt{2\pi} \phi_{4\rho} +2k_F L)\cos (
810: \sqrt{2\pi} \phi_{4\sigma})] ~,
811: \eeq
812: where, as before, we define $\phi_\rho(0) = \phi_{1\rho}$ and $\phi_\rho(L) =
813: \phi_{4\rho}$ and similarly for the $\phi_\sigma$ fields. The gate voltage
814: only couples to the charge degree of freedom within the wire region as
815: \beq
816: S_{gate} = ~\frac{V_G}{\sqpi} ~\int d\tau
817: (\phi_{3\rho} - \phi_{2\rho}) ~,
818: \eeq
819: where $\phi_\rho(d)=\phi_{2\rho}$ and $\phi_\rho(l+d)=\phi_{3\rho}$ as
820: before. So just as in the spinless case, we can integrate out all degrees
821: of freedom except at $x=0,d,l$ and $L$ and obtain the effective
822: action. The full details of the effective action are spelt out in
823: Appendix B. By taking its high, intermediate and
824: low frequency limits, we will be able to
825: obtain conductance corrections just as we did for the spinless fermions.
826:
827: In the high frequency limit where $\omega \gg v_{Ca} /d$ and $v_{Wa} /l$, the
828: two barriers are seen as decoupled barriers, with
829: \bea
830: && S_{0,eff,high} (\phi_{1,\rho/\sigma},\phi_{2,\rho/\sigma},
831: \phi_{3,\rho/\sigma},\phi_{4,\rho/\sigma}) \nonum \\
832: && \quad = \sum_{\on}
833: \vert \on \vert [ \frac{K_{L\rho} + K_{C\rho}}{2 K_{L\rho} K_{C\rho}}
834: (\tphi_{1\rho}^2 + \tphi_{4 \rho}^2 ) + \frac{K_{C\rho} + K_{W\rho}}{2
835: K_{C\rho} K_{W\rho}} (\tphi_{2\rho}^2 ~+~ \tphi_{3\rho}^2 ) \nonum \\
836: && ~~~~~~~~~~~~~ +{\rm ~~~similar ~terms ~with~}
837: \rho \rightarrow \sigma ].
838: \eea
839: The fields $\phi_{2a}$ and $\phi_{3a}$ are completely decoupled
840: from the fields at $x=0$ and $L$ and can be integrated out yielding
841: \bea
842: && S_{0,eff,high} (\phi_{1,\rho/\sigma},\phi_{4,\rho/\sigma})
843: \nonum \\
844: && \quad = {K_{L\rho}+K_{C\rho} \over 2K_{L\rho} K_{C\rho}} \sum_{\on} |\on|
845: (\tphi_{1\rho}^2+\tphi_{4\rho}^2 ) + {K_{L\sigma}+ K_{C\sigma} \over 2 K_{L
846: \sigma} K_{C\sigma}} \sum_{\on} |\on| (\tphi_{1\sigma}^2+\tphi_{4\sigma}^2).
847: \eea
848: Just as for the spinless case, we see that the parameters of the wire
849: do not enter $K_{eff,\rho/\sigma} = K_{L\rho/\sigma}
850: K_{C\rho/\sigma}/(K_{L\rho/\sigma}+K_{C\rho/\sigma})$. Nor does the
851: gate voltage affect the action.
852:
853: Just as in the spinless case, we have two possibilities for the intermediate
854: frequency regime, the QW limit or the QPC limit.
855:
856: For the QW limit, we have $v_W/l \ll \on \ll v_C/d$, and we obtain
857: \bea
858: &&S_{eff,int} (\tphi_{1,\rho/\sigma},\tphi_{2,\rho/\sigma},\tphi_{3,
859: \rho/\sigma}, \tphi_{4,\rho/\sigma}) \nonum \\
860: && \quad = {1\over 2K_{L\rho}} \sum_{\on} |\on| (\tphi_{1\rho}^2+
861: \tphi_{4\rho}^2) + {1\over 2K_{L\sigma}} \sum_{\on} |\on| (\tphi_{1\sigma}^2
862: + \tphi_{4\sigma}^2) \nonum \\
863: && \quad + {1\over 2K_{W\rho}} \sum_{\on} |\on| (\tphi_{2\rho}^2+
864: \tphi_{3\rho}^2) + {1\over 2K_{W\sigma}} \sum_{\on} |\on| (\tphi_{2\sigma}^2 +
865: \tphi_{3\sigma}^2) \nonum \\
866: && \quad + {U_{C\rho} \over 2} \sum_{\on} [ (\tphi_{1\rho}- \tphi_{2\rho}) ^2
867: + (\tphi_{3\rho}- \tphi_{4\rho})^2] \nonum \\
868: && \quad + {U_{C\sigma} \over 2} \sum_{\on} [ (\tphi_{1\sigma}-
869: \tphi_{2\sigma})^2 + (\tphi_{3\sigma}- \tphi_{4\sigma})^2] + S_{gate} ,
870: \eea
871: where $U_{C\rho,\sigma} = v_{C\rho,\sigma}/(K_{C\rho,\sigma}d)$ is the
872: charging energy
873: for the charge degrees of freedom. As the action is quadratic, we can
874: integrate out the $\tphi_{2,\rho/\sigma}$ and $\tphi_{3,\rho/\sigma}$ spin
875: and charge fields to be
876: left with an action dependent only on the $\tphi_{1,\rho/\sigma}$ and
877: $\tphi_{4,\rho/\sigma}$ spin and charge fields as given by
878: \bea
879: \hspace*{-1cm}S_{0,eff,int} (\phi_{1,\rho/\sigma}, \phi_{4,\rho/\sigma})
880: \hspace*{-0.3cm} &=& \frac{K_{L\rho}+K_{W\rho}}{2K_{L\rho}K_{W\rho}}
881: \sum_{\on} |\on| (\tphi_{1\rho}^2+\tphi_{4\rho}^2) \nonum \\
882: && \hspace*{-0.5cm}+ \frac{K_{L\sigma}+K_{W\sigma}}{2K_{L\sigma}K_{W\sigma}}
883: \sum_{\on} |\on| (\tphi_{1\sigma}^2+\tphi_{4\sigma}^2) + \int d\tau
884: {\tilde V}_G (\phi_{4\rho} - \phi_{1\rho} ) ,
885: \eea
886: where we have approximated
887: $U_C + \on/(K_{W,\rho/\sigma})$ by $U_C$; this is justified in the
888: intermediate regime as $T \ll v_{C,\rho/\sigma}/d$ and
889: $K_{C,\rho/\sigma} \sim K_{W,\rho/\sigma}$.
890:
891: In the QPC limit, we have
892: \bea
893: &&S_{eff,int} (\tphi_{1,\rho/\sigma},\tphi_{2,\rho/\sigma},\tphi_{3,
894: \rho/\sigma}, \tphi_{4,\rho/\sigma}) \nonum \\
895: && \quad = {1\over 2K_{L\rho}} \sum_{\on} |\on| (\tphi_{1\rho}^2+
896: \tphi_{4\rho}^2) + {1\over 2K_{L\sigma}} \sum_{\on} |\on| (\tphi_{1\sigma}^2
897: + \tphi_{4\sigma}^2) \nonum \\
898: && \quad + {1\over 2K_{C\rho}} \sum_{\on} |\on|
899: (\tphi_{1\rho}^2 + \tphi_{2\rho}^2+\tphi_{3\rho}^2 +\tphi_{4\rho}^2)
900: + {1\over 2K_{C\sigma}} \sum_{\on} |\on| (\tphi_{1\sigma}^2 +
901: \tphi_{2\sigma}^2 + \tphi_{3\sigma}^2 +\tphi_{4\sigma}^2) \nonum \\
902: && \quad + {U_{W\rho} \over 2}\sum_{\on} (\tphi_{2\rho}- \tphi_{3\rho}) ^2
903: + S_{gate} ,
904: \eea
905: where $U_{W\rho} = v_{W\rho}/(K_{W\rho}l)$ is the charging energy
906: for the charge degrees of freedom in the wire.
907: As before, we may integrate out the inner degrees of freedom to find that
908: \beq
909: S_{0,eff,int} (\tphi_{1,\rho/\sigma}, \tphi_{4,\rho/\sigma}) ~=~
910: S_{0,eff,high} (\tphi_{1,\rho/\sigma}, \tphi_{4,\rho/\sigma})
911: \eeq
912: as expected.
913:
914: Finally, in the low frequency limit $\on \ll v_C /d$ and $v_W /l$,
915: as in the spinless case, the terms multiplying $1/K_{C\rho/\sigma}$ and
916: $1/K_{W\rho/\sigma}$ in Eq. (\ref{spineff}) in Appendix B
917: become constant `mass' terms. We get the full effective action as
918: \bea
919: &&S_{eff,low}
920: (\phi_{1,\rho/\sigma},\phi_{2,\rho/\sigma},\phi_{3,\rho/\sigma},
921: \phi_{4,\rho/\sigma}) \nonum \\
922: && \quad = \frac{1}{2 K_{L\rho}} \sum_{\on} |\on|
923: (\tphi_{1\rho}^2 + \tphi_{4\rho}^2)
924: + \int d\tau [\frac{v_{C\rho}}{K_{C\rho}d} \{ (\phi_{1\rho} -\phi_{2
925: \rho})^2 + (\phi_{3\rho} - \phi_{4\rho})^2 \} \nonum \\
926: && \quad ~~ + \frac{v_{W\rho}}{K_{W\rho} } (\phi_{2\rho} - \phi_{3\rho})^2
927: + {\rm similar ~terms ~with ~} ~\rho \rightarrow \sigma ]
928: + \frac{eV_G}{\sqpi} \int d\tau [\phi_{3\rho} - \phi_{2\rho} ] \nonum \\
929: && \quad ~~ + V \Lambda \int d\tau [\cos (\sqrt{2\pi}\phi_{1\rho}) \cos (
930: \sqrt{2\pi} \phi_{1\sigma}) + \cos (\sqrt{2\pi}\phi_{4\rho} +2k_F L)\cos (
931: \sqrt{2\pi} \phi_{4\sigma})] ~. \nonum \\
932: &&
933: \label{lowfreqspin}
934: \eea
935: Just as we did in the spinless case, we now integrate out the fields
936: at $x=d$ and $l+d$, in terms of which the above action is quadratic, to get
937: \bea
938: &&S_{eff,low} (\phi_{1,\rho/\sigma},\phi_{4,\rho/\sigma})
939: \nonum \\
940: && \quad = {1\over 2K_{L\rho}}\sum_{\on} |\on| (\tphi_{1\rho}^2 +\tphi_{4
941: \rho}^2) + {U_{C\rho} U_{W\rho} \over 2(U_{C\rho}+2U_W{\rho})} \sum_{\on}
942: (\tphi_{1\rho}-\tphi_{4\rho} -{{\tilde V}_G\over U_{W\rho}})^2 \nonum \\
943: && \quad ~~ + {1\over 2K_{L\sigma}}\sum_{\on} |\on| (\tphi_{1\sigma}^2 +
944: \tphi_{4\sigma}^2) + {U_{C\sigma} U_{W\sigma} \over 2(U_{C\sigma}+2
945: U_W{\sigma})} \sum_{\on} (\tphi_{1\sigma}-\tphi_{4\sigma})^2.
946: \eea
947: Here, we see that the effective mass terms are given by $U_{eff,\rho}
948: =\pi U_{C\rho} U_{W\rho} / (U_{C\rho}+2U_W{\rho})$ and
949: $U_{eff,\sigma} =\pi U_{C\sigma} U_{W\sigma} / (U_{C\sigma}+2U_{W\sigma})$ for
950: the `charge' and `spin charge' fluctuations respectively.
951: We denote the `charge on the quantum wire'
952: fields as $n_\rho=\sqrt{2/\pi} (\phi_{1\rho} - \phi_{4\rho})$ and $n_\sigma=
953: \sqrt{2/\pi} (\phi_{1\sigma} - \phi_{4\sigma})$ respectively, and
954: the `current' fields as $\chi_\rho = (\phi_{1\rho} + \phi_{4\rho})/\sqrt{2},
955: \chi_\sigma= (\phi_{1\sigma} + \phi_{4\sigma})/\sqrt{2}$ along with their
956: appropriate Fourier transforms $\tchi_{\rho/\sigma}$ and
957: ${\tilde n}_{\rho/\sigma}$ just as we did in the spinless case. The action
958: then takes the following form,
959: \bea
960: S_{eff,low} &=& {1\over 2K_{L\rho}}
961: \sum_{\on} |\on| [|\tchi_{\rho} - {k_F d\over \sqpi}|^2 + {\pi\over 4}
962: |\tilde n_{\rho} -{2k_F d\over \pi}|^2] + \int d\tau {U_{eff,\rho} \over 2}
963: (n_{\rho} -n_{0\rho})^2 \nonum \\
964: && ~+ {1\over 2K_{L\sigma}} \sum_{\on}|\on| [|\tchi_{\sigma}|^2 + {\pi\over 4}
965: |\tilde n_{\sigma}|^2] + \int d\tau {U_{eff,\sigma} \over 2}
966: (n_{\sigma})^2 \nonum \\
967: && ~+ 2V \Lambda \int d\tau [\cos (\sqpi\chi_{\rho})
968: \cos ({\pi n_\rho\over 2}) \cos (
969: \sqpi\chi_{\sigma}) \cos({\pi n_\sigma\over 2}) \nonum \\
970: && ~~~~~~~~~~~~~~~~ +\sin (\sqpi\chi_{\rho}) \sin ({\pi n_\rho\over 2}) \sin (
971: \sqpi\chi_{\sigma}) \sin ({\pi n_\sigma\over 2}) ]~.
972: \label{reson}
973: \eea
974: We have used the fact that since it is only the $\rho$ field which couples
975: to the gate voltage and not the $\sigma$ fields, we only get
976: $n_{0\rho} = (2k_C d + k_W l)/\pi - V_G/(\pi^{3/2}U_{W\rho})$ and
977: $n_{0\sigma} =0$.
978:
979: We now study the symmetries of the effective action to find out the
980: possible resonances. As in the spinless fermion case, this effective action
981: is invariant under the transformation $\chi_{\rho} \rightarrow \chi_{\rho} +
982: \sqpi$ and $\chi_{\sigma} \rightarrow \chi_{\sigma} + \sqpi$, which
983: corresponds to the transfer of either an up electron or a down
984: electron through the two barriers. But besides this symmetry, there
985: are also some special gate voltages at which one can get resonance
986: symmetries. This can happen when we adjust the gate voltage so as to make
987: $n_{0\rho}$ an odd integer. In that case, $V_{eff}$ is invariant under
988: \bea
989: && n_\sigma \rightarrow -n_\sigma, ~~ n_\rho \rightarrow 2n_{0\rho} - n_\rho
990: {\rm ~~in ~conjunction ~with} \nonum \\
991: && {\rm either} ~~ (i)~\chi_\rho\rightarrow \chi_{\rho} +\sqpi, ~~
992: \chi_\sigma \rightarrow \chi_\sigma ~~ {\rm or} ~~ (ii)~\chi_\sigma
993: \rightarrow \chi_{\sigma} +\sqpi, ~~ \chi_\rho \rightarrow \chi_\rho ~.
994: \eea
995: As explained in Ref. \cite{kane}, this resonance which occurs when $n_{0\rho}$
996: is tuned to be an odd integer, is called a Kondo resonance because it happens
997: when two spin states in the island with $n_{\sigma}=\pm 1$ become degenerate.
998:
999: The kind of resonance which was seen for spinless fermions when two
1000: charge states on the island becomes degenerate is harder to see for
1001: spinful fermions. Two charge states become degenerate when $n_{0\rho}$
1002: is tuned to be a half-odd-integer. But, in that case, the effective
1003: action in Eq. (\ref{reson}) does not have any extra `resonance symmetry'
1004: unless $n_{0\sigma}$ (which we have set to be zero) is also tuned to be
1005: a half-odd-integer. But non-zero $n_{0\sigma}$ is only possible
1006: when there is an effective magnetic field or $SU(2)$ breaking field
1007: just over the quantum wire. This is because the Zeeman term is given by the
1008: Hamiltonian density
1009: \beq
1010: {\cal H}_{Zeeman} = -h(\partial_x\phi_\ua - \partial_x\phi_\da)
1011: = -\sqrt{2} h\partial_x\phi_\sigma ~,
1012: \eeq
1013: and it does not lead to any boundary terms as long as the magnetic field
1014: is felt through the full sample. However,
1015: although in current experiments it is not possible to tune the $SU(2)$
1016: breaking to occur only between the two barriers, it could
1017: be possible in future experiments. Hence it is of interest to look
1018: for possible resonances in this case as well. We see that if one could arrange
1019: to tune both the gate voltage and the magnetic field (adjusted to be
1020: just over the quantum wire) so that $n_{0\rho}$ and $n_{0\sigma}$ are
1021: both half-odd-integers, the effective action in Eq. (\ref{reson}) is symmetric
1022: under $n_\rho \rightarrow 2n_{0\rho}-n_\rho$, $n_\sigma \rightarrow
1023: 2n_{0\sigma}-n_\sigma$, $\chi_{\rho} \rightarrow \chi_{\rho} +\sqpi /2$, and
1024: $\chi_{\sigma} \rightarrow \chi_{\sigma} +\sqpi /2$.
1025: This resonance is exactly analogous to the resonance that existed for
1026: spinless fermions and corresponds to hopping an electron from either
1027: of the leads to the wire. But since this requires the tuning of two
1028: parameters, it is a `higher' order resonance and will be more
1029: difficult to achieve experimentally.
1030:
1031: In fact, if we allow for non-zero $n_{0\sigma}$, then the effective
1032: action also has the symmetry
1033: \bea
1034: && n_\rho \rightarrow -n_\rho, ~~ n_\sigma \rightarrow 2n_{0\sigma} -
1035: n_\sigma ~~ {\rm in ~conjunction ~with} \nonum \\
1036: && {\rm either} ~~ (i) ~\chi_\rho\rightarrow \chi_{\rho} +\sqpi, ~~
1037: \chi_\sigma \rightarrow \chi_\sigma ~{\rm or~} ~~ (ii) ~\chi_\sigma
1038: \rightarrow \chi_{\sigma} +\sqpi, ~~ \chi_\rho \rightarrow \chi_\rho ~,
1039: \eea
1040: when $n_{0\sigma}$ is an odd integer and $n_{0\rho}=0$. But this is hard
1041: to achieve, because one needs to tune the external gate voltage so as to
1042: cancel the field due to the presence of all the other electrons within the
1043: two barriers as well. Hence, this resonance will not be easy to see in
1044: experiments. Moreover, it will show up in the spin conductance and not the
1045: charge conductance.
1046:
1047: In conclusion, we have studied in this section the effective actions
1048: of our model for both spinless and spinful fermions, and used them to
1049: obtain conductance corrections away from resonances (where the conduction
1050: is perfect) as a function of finite temperature and finite length of
1051: the wire. The same technique will again be used in Sec. 5, where it
1052: will be used to study the symmetries and obtain the conductance
1053: corrections of the quantum wire in the presence of a magnetic field.
1054:
1055: \section{\bf Computation of the conductances}
1056:
1057: In this section, we compute the conductances of our TLL quantum wire with
1058: contacts, two semi-infinite Fermi liquid leads and two weak barriers at the
1059: junctions of the contacts and the leads, for both spinless
1060: and spinful electrons, perturbatively in the barrier strength. We explicitly
1061: derive an expression for the conductance to lowest order in
1062: barrier strength (quadratic) in terms of the Green's functions of
1063: the model. The RG flow of the barrier strengths has been incorporated
1064: through a function $\chi(x,y)$. Thus, the behavior of
1065: the Green's functions in the different frequency regimes
1066: determines the conductance corrections. The conductance corrections for a
1067: simpler version of the model of the quantum wire (i.e., one in which the
1068: quantum wire is directly connected to the Fermi leads through two weak
1069: junction barriers) has already been studied by Safi
1070: and Schulz \cite{safilong}, who used a real time formulation
1071: and computed time-dependent Green's functions. The perturbative corrections
1072: in the Kane-Fisher imaginary time formalism was also extended to the case of
1073: finite length wires by Maslov \cite{maslovlong} and Furusaki and Nagaosa
1074: \cite{furusaki}, who computed frequency dependent Green's functions. For our
1075: model, with five distinct spatial regions with their boundaries, the real
1076: time picture of TLL quasiparticle waves reflecting back and forth
1077: between the boundaries (as developed by Safi and Schulz \cite{safi}) is more
1078: cumbersome; hence, we use the imaginary
1079: time formulation and compute frequency dependent Green's functions.
1080:
1081: \subsection{\bf The formulation of the conductance expressions}
1082:
1083: The current through a clean quantum wire through which spinless
1084: electrons are traveling can be found using the Kubo formula
1085: \beq
1086: j(x) = \lim_{\omega\rightarrow 0}\int dy \sigma(x,y,\omega) E(y,\omega) ~,
1087: \eeq
1088: where $\sigma(x,y,\omega)$ is the non-local conductivity and is related to the
1089: two-point Green's function $G(x,y,\omega)$ at finite frequency $\omega$ as
1090: \beq
1091: \sigma(x,y,\omega) = ~-i\frac{2 \omega e^2}{h} G(x,y,\omega) ~.
1092: \label{cond}
1093: \eeq
1094: For our model of the quantum wire, $G_{\bo}(x,y)$ has been computed in
1095: Appendix D. Note that the real frequency $\omega$ is related to $\bar \omega$
1096: (used in the earlier sections) by the analytic continuation $\omega = i
1097: {\bar\omega} +\epsilon$. From Appendix D, we find that
1098: $G_\bo(x,y) = K_L/(2|\bo|)$ + non-singular terms in $\bo$ in the limit
1099: $\omega \rightarrow 0$ for our model. hence the dc conductance $g_0$ is given
1100: by
1101: \beq
1102: g_0 = \lim_{\omega\rightarrow 0}\sigma(x,y,\omega) = \frac{e^2}{h}~.
1103: \eeq
1104: This shows perfect dc conductance through the system
1105: as in the earlier models without contacts\cite{safi,maslov}.
1106: This result remains unchanged for the case of electrons with spin,
1107: except for a multiplication of the conductance by a factor of two.
1108:
1109: For a quantum wire in the presence of stationary impurities,
1110: an explicit expression for the conductance can be derived to lowest
1111: (quadratic) order in the impurity strength from the partition function,
1112: using perturbation theory \cite{kane,safilong,maslovlong}. The
1113: renormalization group (RG) equations for the barriers (discussed in detail in
1114: subsection 4.4) imply that the barrier strengths grow under renormalization.
1115: However, it is only for very low temperatures or very long wire
1116: lengths that there will be considerable renormalization.
1117: In real experimental setups, the length of the
1118: wire is in the range of micrometers and the temperatures in the
1119: range of a Kelvin; hence one does not expect much renormalization.
1120: Hence, it is expected that the barrier strengths remain small
1121: enough for perturbation theory to be applicable. We follow the methods
1122: of Safi and Schulz \cite{safilong} and Maslov \cite{maslovlong}, who
1123: derived explicitly a conductance expression for a non-translationally
1124: invariant system and obtained
1125: \beq
1126: g = g_0 K_L ( 1 - {\cal R} )~,
1127: \eeq
1128: where {\cal R} is the perturbative correction to second
1129: order in the impurity strength. $\cal R$ is given by
1130: \beq
1131: {\cal R} = g_0 K_L \sum_{m=1}^{\infty} m^2 c_m^2 {\cal R}^{(m)}~,
1132: \eeq
1133: where the $c_m$'s are the coefficients for the terms in the Haldane
1134: representation of the fermionic density, and ${\cal R}^{(m)}$ is the
1135: correction due to the back-scattering of {\it m} electrons given by
1136: \beq
1137: {\cal R}^{(m)} = \int\int dx dy V(x)V(y) \cos [ 2m(\xi (x)
1138: - \xi (y))] \chi_m(x,y)~.
1139: \eeq
1140: In the above expression, $V(x)$ is the bare potential of the
1141: impurities, $\xi$ is a phase factor which includes the
1142: $k_F x$ factor coming from the
1143: back-scattering process and other factors which arise due to the removal
1144: of the forward scattering terms from the Hamiltonian by shifts in the
1145: bosonic field $\phi$, and $\chi_m(x,y)$ is a factor which
1146: incorporates the renormalization group (RG) flows of the barrier strengths.
1147: In general, it is given by a two-point correlation function defined as
1148: \beq
1149: \chi_m(x,y) = \frac{1}{T^2} e^{-2m^2G_0(x,x,\tau_0) - 2m^2G_0(y,y,\tau_0)}
1150: \int_0^{\infty} dt e^{4m^2G_0(x,y,it+\pi/2)}~,
1151: \eeq
1152: $G_0 (x,y,it)$ is the two-point Green's function for a clean quantum wire.
1153: Here, the Green's functions are in terms of the imaginary time $\tau =it$.
1154: $\tau_0 \sim 1/\Lambda$ is the inverse of the high energy cutoff.
1155: In a later subsection, we show how the one-point function
1156: $\chi_m(x,x)$ can be obtained directly from the RG equation for the barriers.
1157:
1158: For our system,
1159: we shall instead compute the two-point Green's function in terms of $\bar
1160: \omega$, in terms of which, the correction to the conductance is given by
1161: (specializing to the case $m=1$)
1162: \bea
1163: R^{(1)} &=& {\rm lim}_{\omega \rightarrow 0}
1164: \frac{\omega}{(g_0 K_L)^2} \int dx' \int dy' G_0(x,x',\omega )
1165: G_0(y,y',\omega)\cos [2(\xi (x') - \xi (y'))] \times \nonum \\
1166: && \quad \quad \quad \quad ~~~~~~~~~~~~~~~~~~~~ V(x') V(y')
1167: Im(F(x',y',\omega)-F(x',y',0)) \nonum \\
1168: &=& \int dx' \int dy' \cos [2(\xi (x') - \xi (y'))] V(x') V(y')
1169: {\rm lim}_{\omega \rightarrow 0} {dF(x',y',\omega) \over d \omega} ,
1170: \eea
1171: where
1172: \bea
1173: && \hspace*{-1cm} F(x',y',\omega) \nonum \\
1174: && \hspace*{-1cm}= \int_{-\infty}^{\infty} dt e^{i\omega t} \exp
1175: ( - 2\pi \sum_{\on'} [G_0(x',x',\on') + G_0(y',y', \on') - 2 G_0(x',y',\on')
1176: \cos (\on' \tau)]).
1177: \eea
1178: Note that the Green's functions in the prefactors of $R^{(1)}$ are dependent
1179: on the external driving frequency, but the Green's functions in
1180: the exponential depend only on the Matsubara frequencies $\on'$, and not
1181: on the external driving frequency $\omega$ or its analytic continuation
1182: $\bo$. The sum over the Matsubara frequencies are cutoff at the low energy
1183: end by $ {\bar\omega}'_{n=1} \sim k_B T$ and at the upper end
1184: by the high energy cutoff $\Lambda$. In evaluating $R^{(1)}$, we will
1185: approximate $\sum_{\on'}$ by $\int d\bo' /(2\pi)$ which is reasonable
1186: since we always assume that the temperature $T$ is much smaller than the
1187: cutoff $\Lambda$.
1188:
1189: \subsection{Results for the Quantum Wire}
1190:
1191: We will concentrate here on calculating the conductance of a quantum wire
1192: system in which the length of the quantum wire $l$ is much greater than
1193: the length of the contact regions $d$. Also, we will finally be interested
1194: in studying the effects of the junction barriers placed at the two
1195: lead-contact junctions (as explained earlier). Hence, for our model
1196: \beq
1197: V(x) = V_1 \Lambda \delta(x) + V_2 \Lambda \delta(x-L).
1198: \eeq
1199: For this potential, we can obtain the expression for the conductance
1200: corrections
1201: as
1202: \bea
1203: R^{(1)} &=& \Lambda^{2} {\rm lim}_{\omega \rightarrow 0}
1204: [V_1^2 Im {dF(0,0,\omega)
1205: \over d\omega} + V_2^2 Im {dF(L,L,\omega)\over d\omega} \nonum \\
1206: && \quad \quad ~~~~~+ 2V_1 V_2 Im {dF(0,L,\omega)\over d\omega}
1207: \cos(\xi(0)-\xi(L))].
1208: \label{r1}
1209: \eea
1210:
1211: \begin{itemize}
1212:
1213: \item{} Spinless electrons
1214:
1215: Now, the expression for the one-point Green's function (in frequency space)
1216: for a barrier placed inside
1217: the contact region on the left of the QW and at a distance $a$ from the left
1218: lead-contact junction (which is taken to be the origin, giving the hierarchy
1219: of length scales $a \ll d \ll l$) can be easily obtained from Appendix D.
1220: It is given by
1221: \beq
1222: G_{\bar{\omega}} (x=a,y=a) \simeq \frac{K}{2\vert\bar{\omega}\vert} ~,
1223: \eeq
1224: where
1225: \begin{displaymath}
1226: K = \left\{ \begin{array}{ll}
1227: K_C & \textrm{for $\vert\bar{\omega}\vert \gg v_C/a$}\\ \\
1228: \frac{2K_LK_C}{K_L+K_C} & \textrm{for $v_C/d \ll \vert\bar{\omega}\vert
1229: \ll v_C/a$}\\ \\
1230: \frac{2K_LK_W}{K_L+K_W} & \textrm{for $v_W/l \ll \vert\bar{\omega}\vert
1231: \ll v_C/d$}\\ \\
1232: K_L & \textrm{for $\vert\bar{\omega}\vert \ll v_W/l$ ~.}
1233: \end{array} \right.
1234: \end{displaymath}
1235: For our model with a barrier at the left lead-contact junction, $a=0$ and
1236: the first frequency regime does not exist.
1237:
1238: The two-point Green's function $G(x,y)$
1239: for $y$ in the left contact region and $x$ anywhere is given in
1240: Appendix D. By setting $y=0$ (i.e., at the first barrier) and $x=L=l+2d$
1241: (i.e., at the second barrier) we
1242: obtain the conductances in the different frequency regimes given by
1243: \bea
1244: G &=& {2K_C(1+\gamma_1)^2 \over (2+K_C/K_W + K_W/K_C)} {\exp[-|\bo|
1245: ({2d\over v_C} +{l\over v_W})] \over |\bo|} \quad {\rm for}\quad
1246: |\bo| \gg v_W/l \gg v_C/d
1247: \nonum \\
1248: &=& {2K_C(1+\gamma_1)^2 \over (2+K_C/K_W + K_W/K_C)} {\exp[-|\bo|
1249: ({2d\over v_C})] \over |\bo|} ~~\quad \quad \quad {\rm for} \quad
1250: v_W/l \gg |\bo| \gg v_C/d \nonum \\
1251: &=& {K_L \over 2|\bo|} \quad \quad {\rm for} \quad v_C/d,v_W/l \gg |\bo| ,
1252: \label{twopoint}
1253: \eea
1254: where $\gamma_1 = {K_L-K_C\over K_L+K_C}$.
1255: Thus, we see that $G(x,y, \bo')$ decays
1256: exponentially to zero except at the lowest frequency
1257: regime where $G(x,y,\bo') = G(x,x,\bo') = G(y,y,\bo')$.
1258:
1259: To obtain the conductance corrections, we
1260: use the above Green's functions to compute $F(x',y',\bo')$ in each of these
1261: frequency regimes. For the high frequency regime, it is simply given by
1262: \bea
1263: F(x',y',\omega) &=& \int_{-\infty}^\infty dt
1264: e^{i\omega t} \exp [-2\pi \int_T^\Lambda {d\bo'\over 2\pi}
1265: ({K_{eff} \over |\bo'|} - 2 G(x',y',\bo') \cos \bo' \tau )] \nonum \\
1266: &=& \int_{-\infty}^\infty {1\over T} dz e^{i\omega z/T}
1267: ({T\over \Lambda})^{2K_{eff}} \exp [-2\pi \int_T^\Lambda
1268: {d\bo'\over 2\pi} 2 G(x',y',\bo') \cos \bo' \tau] , \nonum \\
1269: &&
1270: \eea
1271: where $K_{eff} = K_L K_C/(K_L + K_C)$, and
1272: in the second line, we have scaled $t$ by $T$, i.e., we
1273: have used $t=z/T$ to write the integral in terms of dimensionless
1274: variables so that the temperature power-laws can be made explicit.
1275: When $x'\ne y'$, $G(x',y', \bo') \rightarrow 0$, so that one can check that
1276: $lim_{\omega \rightarrow 0}Im {dF\over d\omega}$ also tends to zero.
1277: This means that the cross-term in Eq. (\ref{r1}) does not contribute.
1278: For each of the terms involving just one barrier, we find that
1279: \beq
1280: lim_{\omega \rightarrow 0}Im {dF\over d\omega} =
1281: {1\over T^2}({T\over \Lambda})^{2K_{eff}} \exp [- \int_T^\Lambda
1282: {d\bo'} {K_{eff}\over |\bo'|} \cos \bo' \tau].
1283: \eeq
1284: Hence, we obtain the following answer for the conductance correction
1285: for high temperatures $T\gg T_d \equiv v_C/(k_B d)$,
1286: \beq
1287: g = g_0 K_L [ 1 - c_1 ({T\over \Lambda})^{2(K_{eff} - 1)}(|V_1|^2+ |V_2|^2)],
1288: \label{hightemp}
1289: \eeq
1290: where $c_1$ is a dimensionful constant
1291: dependent on factors like the contact quasiparticle velocity $v_C$, but is
1292: independent of the gate voltage $V_G$.
1293:
1294: For intermediate temperatures
1295: where $T_l \equiv v_W/(k_B l) \ll T \ll T_d$ the calculation is very
1296: similar to that performed for the high frequency case, except that the
1297: integral over the Matsubara frequencies is now split into two regions
1298: \bea
1299: F(x',y',\omega) &=& \int_{-\infty}^\infty dt
1300: e^{i\omega t} \exp [-2\pi (\int_T^{T_d} + \int_{T_d}^\Lambda)
1301: {d\bo'\over 2\pi} {K \over |\bo'|} ]~ \nonum \\
1302: && \quad \quad \quad \exp
1303: [2\pi \int_T^\Lambda {d\bo'\over 2\pi} 2 G(x',y', \bo') \cos \bo' \tau] ~.
1304: \eea
1305: The rest of the calculations go through as above, and we find that
1306: the conductance expression is
1307: \beq
1308: g = g_0 K_L [ 1 - c_2 ({T_d\over \Lambda})^{2(K_{eff} - 1)}
1309: ({T\over T_d})^{2({\tilde K}_{eff}-1)} (|V_1|^2 + |V_2|^2) ],
1310: \eeq
1311: where ${\tilde K}_{eff} = K_L K_W/(K_L + K_W)$ and $c_2$ is a dimensionful
1312: constant which is dependent on $V_G$.
1313:
1314: Finally, for very low temperatures $T \ll T_l$, the sum over
1315: Matsubara frequencies split into three regions so that we have
1316: \bea
1317: F(x',y',\omega) &=& \int_{-\infty}^\infty dt
1318: e^{i\omega t} \exp [-2\pi (\int_T^{T_l} + \int_{T_l}^{T_d}
1319: +\int_{T_d}^\Lambda) {d\bo'\over 2\pi} {K \over |\bo'|} ] ~ \nonum \\
1320: && \quad \quad \quad \exp
1321: [2\pi \int_T^\Lambda {d\bo'\over 2\pi} 2 G(x',y', \bo') \cos \bo' \tau] ~.
1322: \eea
1323: Furthermore, in this regime, the cross-term does not vanish; in
1324: fact, for $x' \ne y'$, we have $G(x',y',\bo) = G(x',x',\bo)=G(y',y',\bo)$,
1325: so that $F(x',y',\omega) =F(x',x',\omega) =F(y',y',\omega)$.
1326: The contribution of the cross term is hence identical to that
1327: of the terms due to a single barrier. Hence, we obtain the
1328: corrections to the conductance as
1329: \beq
1330: g = g_0 K_L [ 1 - c_3 ({T\over T_l})^{2(K_L - 1)} ({T_d \over
1331: \Lambda})^{2(K_{eff} -1)} ({T_l\over T_d})^{2({\tilde K}_{eff} - 1)}
1332: |V_1 + V_2|^2 ] ~,
1333: \eeq
1334: where $c_3$ is a dimensionful constant similar in nature to $c_2$; thus the
1335: two barriers are seen coherently.
1336:
1337: Note that the power-laws come purely from the one-point Green's functions,
1338: whereas the phase coherence between the barriers is determined by the
1339: behavior of the two-point correlation function. At high or intermediate
1340: frequencies, $lim_{\omega \rightarrow 0} Im {dF (x',y',\omega) \over
1341: d\omega}$ tends to zero for $x'\ne y'$ leading to the lack of phase coherence
1342: between the two barriers. At very high temperatures, the
1343: interaction parameters of the contact region $K_C$ and the lead region $K_L$
1344: controls the renormalization of a barrier in the contact region.
1345: As the temperature is lowered, the phase coherence length of the electronic
1346: excitations increases, and the renormalization exponent makes a crossover to
1347: a combination of the interaction parameters of the contact and QW, and finally
1348: to that of the lead alone at the lowest temperature regime. The lowest
1349: temperature regime is also the one in which resonant transport through both
1350: the lead-contact junction barriers can take place as phase coherence over the
1351: entire system is achieved at these temperatures.
1352:
1353: \item{} Electrons with spin
1354:
1355: The above expressions were given for a
1356: model of the QW system but for spinless electrons. Let us now see what the
1357: conductance expressions are for electrons with spin. These expressions can
1358: be derived in the same way as for spinless electrons by using the appropriate
1359: Green's functions for spin and charge fields. This
1360: gives us for the high temperature regime of $T_d \ll T $
1361: \beq
1362: g = 2 g_0 K_L [ 1 - c_4 ({T \over \Lambda})^{2(K_{eff} - 1)}
1363: (|V(0)|^2 + |V(l+2d)|^2) ],
1364: \eeq
1365: where now $K_{eff} = K_L K_{C\rho}/(K_L + K_{C\rho}) + K_L K_{C\sigma}/(K_L
1366: + K_{C\sigma})$, and $c_4$ is a dimensionful constant much like $c_1$ for the
1367: spinless case (i.e., dependent on the contact charge velocity $v_{C,\rho}$ but
1368: independent of the gate voltage $V_G$). For the intermediate temperature range
1369: $T_l \ll T \ll T_d$, we find
1370: \beq
1371: g = 2 g_0 K_L [ 1 - c_5 ({T_d \over \Lambda})^{2(K_{eff} - 1)}
1372: ({T \over T_d})^{2({\tilde K}_{eff}-1)} (|V(0)|^2 + |V(l+2d)|^2) ] ~,
1373: \eeq
1374: where ${\tilde K}_{eff} = K_L K_{W\rho}/(K_L + K_{W\rho}) +
1375: K_L K_{W\sigma}/(K_L + K_{W\sigma})$, and $c_5$ is a constant similar to
1376: $c_2$ for the spinless case (i.e., dependent on $V_G$). Finally, for the low
1377: temperature regime $T \ll T_l$, we obtain
1378: \beq
1379: g = 2 g_0 K_L [ 1 - c_6 ({T \over T_l})^{2(K_L - 1)} ({T_d \over \Lambda})^{2
1380: (K_{eff} -1)} ({T_l \over T_d})^{2({\tilde K}_{eff} - 1)}
1381: |V(0) + V(l+2d)|^2 ] ~,
1382: \eeq
1383: where $c_6$ is similar in nature to $c_3$ for the spinless case.
1384:
1385: \end{itemize}
1386:
1387: \subsection{Results for the Quantum Point Contact}
1388:
1389: The Quantum Point Contact (QPC) is simply a quantum wire system in which the
1390: length of the quantum wire region $l \sim 0.2 - 0.5 \mu m$ (i.e., the region
1391: undergoing the constriction due to the application of the gate voltage)
1392: is much reduced in comparison to typical lengths for a quantum wire
1393: $l \sim 2 - 20 \mu m$. Thus, in our model of the quantum wire system, we can
1394: reach the QPC by studying the limit when the contact region length $d$ is much
1395: greater than the wire length $l$. Let us then study the effects of
1396: barriers/impurities placed in the contact and wire region of the QPC.
1397:
1398: \begin{itemize}
1399:
1400: \item{} Spinless electrons
1401:
1402: In order to study the effect of a weak barrier placed in the contact
1403: region such that its distance $a$ from the left lead-contact junction falls
1404: in the hierarchy of $a \ll l \ll d$, we again start by computing the one-point
1405: Green's function for such an impurity. We find that
1406: \beq
1407: G_{\bar{\omega}} (x=a,y=a) \simeq \frac{K}{2\vert\bar{\omega}\vert},
1408: \eeq
1409: where
1410: \begin{displaymath}
1411: K = \left\{ \begin{array}{ll}
1412: K_C & \textrm{for $\vert\bar{\omega}\vert \gg v_C/a$}\\ \\
1413: \frac{2K_LK_C}{K_L+K_C} & \textrm{for $v_W/l \ll \vert\bar{\omega}\vert
1414: \ll v_C/a$}\\ \\
1415: \frac{2K_LK_C}{K_L+K_C} & \textrm{for $v_C/d \ll \vert\bar{\omega}\vert
1416: \ll v_W/l$}\\ \\
1417: K_L & \textrm{for $\vert\bar{\omega}\vert \ll v_C/d$ ~.}
1418: \end{array} \right.
1419: \end{displaymath}
1420:
1421: As before, for the two-point function, we find that the high and low
1422: frequency limits are the same as that given in Eq. (\ref{twopoint}) for the
1423: QW, but for $T_d \ll T \ll T_l$, the answer turns out to be
1424: the same as in the high frequency limit. This is similar
1425: to what one sees for the one-point Green's functions above as well.
1426: So, without giving any further derivations,
1427: we directly quote the expressions for the conductance corrections.
1428: In the high and intermediate frequency regimes, the conductance is given by
1429: \beq
1430: g = g_0 K_L [ 1 - c_i
1431: ({T\over \Lambda})^{2(K_{eff} - 1)} (|V_1|^2+ |V_2|^2) ],
1432: \eeq
1433: where $i=4,5$ allows for the constant to be different in the high and
1434: intermediate frequency regimes. For the low frequency regime, we get
1435: \beq
1436: g = g_0 K_L [ 1 - c_6
1437: ({T\over T_d})^{2(K_L - 1)} ({T_d\over \Lambda})^{2(K_{eff} - 1)}
1438: (|V_1 + V_2|^2) ],
1439: \eeq
1440: It is clear from the above expressions
1441: that the contributions of barriers in the contacts of a QPC are always
1442: going to be independent of the gate voltage $V_G$ as the QPC interaction
1443: parameter $K_W$ does not enter anywhere. Thus, such an impurity would always
1444: lead to a flat and channel independent renormalized conductance.
1445: It should be noted that we have found from a similar calculation that even
1446: for an impurity placed deep inside the contact (i.e., with the hierarchy of
1447: $l \ll a \ll d$), the above conclusions still remain true; this is because
1448: the only change that takes place is that $K = K_C$ (rather than the
1449: combination of $K_L$ and $K_C$ found earlier) for the regime of $v_W/l \ll
1450: \vert\bar{\omega}\vert \ll v_C/a$.
1451:
1452: Finally, let us study the effect of an impurity placed inside the QPC itself.
1453: We find the one-point Green's function for such a case (with the hierarchy of
1454: $l \ll d < a$) to be
1455: \beq
1456: G_{\bar{\omega}} (x=a,y=a) \simeq \frac{K}{2\vert\bar{\omega}\vert},
1457: \eeq
1458: where
1459: \begin{displaymath}
1460: K = \left\{ \begin{array}{ll}
1461: K_W & \textrm{for $\vert\bar{\omega}\vert \gg v_W/l$}\\ \\
1462: K_C & \textrm{for $v_C/d \ll \vert\bar{\omega}\vert
1463: \ll v_W/l$}\\ \\
1464: K_L & \textrm{for $\vert\bar{\omega}\vert \ll v_C/a$ ~.}
1465: \end{array} \right.
1466: \end{displaymath}
1467: We have here only three frequency regimes as the regime of
1468: $v_C/d \ll \vert\bar{\omega}\vert \ll v_W/l$ cannot be taken sensibly within
1469: the given hierarchy of length scales. This shows again that the effect of an
1470: impurity placed within the QPC will always be dependent on the gate voltage
1471: $V_G$, and can never lead to flat and channel independent renormalizations of
1472: the conductance.
1473:
1474: \item{} Electrons with spin
1475:
1476: The generalization to spinful electrons can be obtained just
1477: as was done for quantum wires with the appropriate substitutions.
1478:
1479: \end{itemize}
1480:
1481: \subsection {Evaluation of the conductances from the effective
1482: actions using the RG equations}
1483:
1484: Here, we note that the above results for the conductances
1485: could have been anticipated by computing the RG equation for the impurity
1486: potentials using the effective actions calculated in Sec. 3.
1487:
1488: The conductance is governed by the renormalized barrier potentials at the two
1489: junctions. Since the interaction is repulsive, the barrier potentials
1490: are expected to grow as a function of the frequency cutoff. This is
1491: what leads to the result that any impurity potential,
1492: however small, eventually cuts the wire; in the zero temperature
1493: limit, there is no transmission at all \cite{kane}. However,
1494: at a finite temperature $T$, finite wire length $l$ or finite contact
1495: length $d$, the growth is cutoff by either $T$, $v_W/L$ or $v_C/d$.
1496: In fact, since the energy scales in the problem are the
1497: temperature $k_BT$, the high frequency cutoff $\Lambda$ and those related
1498: to the contact length $k_BT_d= v_C/d$ and
1499: the wire length $k_BT_l= v_W/l$, we can see that there will
1500: exist two energy scale crossovers in the system --- one from $T/\Lambda$
1501: to $T/T_d$ and the other from $T/T_d$ to $T/T_l$ for $d \ll l$ (the QW
1502: limit), or from $T/\Lambda$ to $T/T_l$ and then from $T/T_l$ to $T/T_d$ for
1503: $l \ll d$ (the QPC limit).
1504:
1505: In fact, an explicit RG calculation of either of the individual barrier
1506: strengths in the high, intermediate and low frequency regimes
1507: simply involves computing the dimension
1508: of $\cos (2\sqpi\phi_1)$ or $\cos (2\sqpi\phi_4)$ (which turn out to be
1509: the same) using those respective actions. For example, for the high
1510: frequency effective action given in Eq. (\ref{highfreq}), the RG
1511: equation for a single barrier is given by
1512: \beq
1513: {dV_1\over d\lambda} = (1 - {2K_C K_L\over K_C+K_L}) V_1 \equiv
1514: (1 - K_{eff})V_1 ~,
1515: \eeq
1516: where $\lambda={\rm ln} \frac{\Lambda(\lambda)}{\Lambda}$.
1517: Using this, we can get the renormalized barrier strength to be
1518: \beq
1519: V_1^{ren} = V_1 ({T\over \Lambda})^{K_{eff}-1}
1520: \eeq
1521: in the high frequency regime $T\gg T_d$, where we have used $T$ to
1522: cutoff the RG flow (which begins from $\Lambda$). From this, we infer that
1523: to quadratic order, the $T$ dependence of the conductance
1524: corrections is given by
1525: \beq
1526: g = {e^2\over h} [1 - {\tilde c}_1V_1^2 (\frac{T}{\Lambda})^{2(K_{eff}-1)}]
1527: \quad {\rm for} \quad
1528: T\gg T_d ~,
1529: \label{offres}
1530: \eeq
1531: where ${\tilde c}_1$ is a dimensionful constant like $c_1$ defined
1532: earlier containing factors like $v_C$ but
1533: is, most importantly, independent of the gate voltage $V_G$.
1534: Comparing with Eq. (\ref{hightemp}), we see that if we include the
1535: subtraction due to two barriers, the expressions are identical.
1536:
1537: In the intermediate regime of $T_l \ll T \ll T_d$, the RG equation for the
1538: same barrier now becomes
1539: \beq
1540: {dV_1\over d\lambda} = (1 - {2K_C K_W\over K_W+K_L}) V_1 \equiv
1541: (1 - {\tilde K}_{eff})V_1 ~.
1542: \eeq
1543: using the effective action in Eq. (\ref{intertemp}).
1544: At the same time, the appearance of the energy scale $v_C/d$ (through
1545: $U_C$) in the effective action in this temperature regime and the taking
1546: of the approximation $\on \ll U_C$ means that $v_C/d$
1547: has replaced $\Lambda$ as the high energy cutoff in the expression for
1548: the $T$ dependence of the conductance correction. The influence of
1549: those degrees of freedom whose energies lie between $v_C/d$ and $\Lambda$
1550: can be taken into account by noting that they will contribute a
1551: factor of $(T_d/\Lambda)^{2(K_{eff}-1)}$; this is because
1552: these degrees of freedom have been integrated away during the RG
1553: procedure, and there must be continuity between the
1554: conductance expressions for $T \gg T_d$ and $T_l \ll T \ll T_d$ at
1555: $T = T_d$. Thus, we get the conductance expression in this regime as
1556: \beq
1557: g = {e^2\over h} [1 - {\tilde c}_2V_1^2 (\frac{T_d}{\Lambda})^{2(K_{eff}-
1558: 1)}) (\frac{T}{T_d})^{2({\tilde K}_{eff}-1)}]
1559: \quad {\rm for} \quad T_l\ll T\ll T_d ~,
1560: \label{intcond}
1561: \eeq
1562: where ${\tilde c}_2$ is a constant similar in nature to ${\tilde c}_1$, but it
1563: can depend on $v_W$ and is hence dependent on the gate voltage $V_G$. Thus the
1564: conductance is no longer independent of $V_G$. This again is the same as the
1565: expression obtained by the explicit computation of the conductance.
1566:
1567: Finally, in the low temperature limit,
1568: we recognize the fact that there is phase coherence over the distance
1569: between the two barriers; this follows from
1570: the low frequency effective action which has cross terms between
1571: the fields at the two barriers. This is what leads to resonant transmission.
1572: To compute the conductance corrections away from resonance in this limit,
1573: we note the following. Since the resonance occurs precisely when the
1574: $2k_F$ component of the barrier term goes to zero, the relevant
1575: term away from this resonance is precisely the back-scattering potential
1576: $V\cos (2\sqpi \chi)$. Computing the dimension of this operator gives us the
1577: RG equation for our barriers in this temperature regime as
1578: \beq
1579: {dV_1\over d\lambda} = (1 - K_L) V_1 .
1580: \eeq
1581: This makes the $T$ dependence of the conductance correction clear. Again,
1582: the appearance of the energy scale $v_W/l$ (through $U_W$) and
1583: the approximation $\on \ll U_W$ indicate
1584: that $v_W/l$ has now replaced $v_C/d$ as the high energy cutoff in the
1585: expression for the $T$ dependence of the conductance correction.
1586: As before, the influence of those degrees of freedom whose energies lie in
1587: between $v_W/l$ and $v_C/d$ is shown by the appearance of the term
1588: $(T_l/T_d)^{2({\tilde K}_{eff}-1)}$. This is because these degrees
1589: of freedom have also been integrated away during the RG procedure, and
1590: there must be continuity in the conductance expressions at $T=T_l$ whether we
1591: come from the $T_d \gg T \gg T_l$ regime or the $T_l \ll T$ regime. Thus,
1592: we obtain the conductance in this regime as
1593: \beq
1594: G= {e^2\over h} [1 - {\tilde c}_3V_1^2 (\frac{T_d}{\Lambda})^{2(K_{eff}-1)})
1595: (\frac{T_l}{T_d})^{2({\tilde K}_{eff}-1)} (\frac{T}{T_l})^{2(K_L-1)}]
1596: \quad {\rm for} \quad T\gg T_L ~,
1597: \eeq
1598: where ${\tilde c}_3$ is a constant similar in nature to ${\tilde c}_2$. We can
1599: now see that, as $K_L=1$ (for 2DEG Fermi reservoirs), the conductance has no
1600: temperature dependence in the low temperature regime.
1601:
1602: A similar analysis can be done for the QPC limit, which reproduces the
1603: conductance expressions for the QPC limit that were obtained explicitly in
1604: the earlier subsection. We note, however, that the conductance
1605: corrections are small in this case as the RG flow for the barriers
1606: is restricted by the small length scales in the system.
1607:
1608: The conductance corrections for electrons with spin can also be
1609: obtained using the effective actions and the RG equations for the
1610: barriers, by proceeding in the same way as was done for spinless
1611: electrons. Since the conductance expressions have already been
1612: given in the previous section, we do not repeat them here.
1613:
1614: Thus, we emphasize that just by using the effective action and the RG
1615: equations for the barriers, we can actually obtain the conductance
1616: corrections. However, all that we actually do here is
1617: to compute the RG flows of the individual barriers, and then infer
1618: the temperature and length power-laws in the
1619: conductance corrections. Hence, even in principle, there is no way of
1620: obtaining the constants ${\tilde c}_1 ,\ldots ,{\tilde c}_6$ from this method,
1621: whereas the explicit computation of the conductance in the earlier subsection
1622: can give the explicit forms of the constants as well.
1623: In fact, the correlation functions computed there
1624: can be directly related to the coefficients which appear in the RG
1625: equations. In the various frequency regimes, the RG equation for a
1626: single barrier for spinless fermions can be written as
1627: \beq
1628: {dV\over d\lambda} = (1-2 |\bo| G_\bo (x,x))V.
1629: \eeq
1630: On Fourier transforming, this gives
1631: \beq
1632: \frac{dV (x)}{d\lambda} = ( 1 + 2\frac{dG}{d\lambda}(x,x,\tau_0 e^\lambda))
1633: V(x) ~,
1634: \eeq
1635: where $\tau_0$ is the high energy cutoff $1/\Lambda$. Integrating this
1636: gives the renormalized strength of the impurity $V_{ren}$ as
1637: \beq
1638: V_{ren}(x,\lambda) = V(x,\lambda=0) \exp [\lambda - U(x,x,\tau_0
1639: e^\lambda)] ~,
1640: \eeq
1641: where $U(x,x,\tau_0e^\lambda) = -2 (G_0(x,x,\tau_0 e^\lambda) -
1642: G_0(x,x,\tau_0))$.
1643: Thus, in this case,
1644: \beq
1645: \chi_1 (x,x) = \exp [l -U(x,x,\tau_0 e^\lambda )] ~.
1646: \eeq
1647: However, the non-local $\chi_1 (x,y)$ is not so easy to obtain just
1648: from the RG equations.
1649:
1650: Now, let us study the conclusions that can be drawn from the conductance
1651: expressions. To begin with, the expressions in the various frequency regimes
1652: reveal that as {\it either} the temperature $T$
1653: is raised {\it or} the total length $L$ of the contacts and QW is decreased,
1654: the conductance corrections become smaller and the conductance approaches
1655: integer multiples of $2g_0$ as expected \cite{tarucha,yacoby}. Furthermore,
1656: we can see from these expressions that in the high temperature limit
1657: i.e., when $T \gg T_d, ~T_l$, the conductance corrections are independent
1658: of the QW parameters. Hence, they are {\it independent} of the gate voltage
1659: $V_G$ and of all factors dependent on the channel index. Thus they yield
1660: renormalizations to the ideal values which are themselves plateau-like
1661: and uniform for all channels. Such corrections to
1662: the conductance explain some of the puzzling features observed in the
1663: experiments of Ref. \cite{yacoby}. A more detailed
1664: comparison of these results against experimental findings will be made
1665: in a later section; it is important to note here that our results are in
1666: qualitative agreement with most experimental observations on electronic
1667: transport through a variety of quantum wire systems.
1668:
1669: Let us now compare these observations with what we find as the
1670: perturbative renormalizations to the perfect conductance of a barrier/impurity
1671: placed anywhere within the quantum wire itself such that its distance from
1672: the left lead-contact junction (taken as the origin) is again denoted by $a$.
1673: An exactly similar computation of the one-point Green's function in this
1674: case reveals that (note that we are now working with the hierarchy of
1675: $d \ll a \ll l$)
1676: \beq
1677: G_{\bar{\omega}} (x=a,y=a) \simeq \frac{K}{2\vert\bar{\omega}\vert} ~,
1678: \eeq
1679: where
1680: \begin{displaymath}
1681: K = \left\{ \begin{array}{ll}
1682: K_W & \textrm{for $\vert\bar{\omega}\vert \gg v_C/d$}\\ \\
1683: K_W & \textrm{for $v_W/a \ll \vert\bar{\omega}\vert
1684: \ll v_C/d$}\\ \\
1685: \frac{2K_LK_W}{K_L+K_W} & \textrm{for $v_W/l \ll \vert\bar{\omega}\vert
1686: \ll v_C/a$}\\ \\
1687: K_L & \textrm{for $\vert\bar{\omega}\vert \ll v_W/l$ ~.}
1688: \end{array} \right.
1689: \end{displaymath}
1690: Now, for a barrier at the left contact-QW junction, $a \rightarrow d$,
1691: the second frequency regime of $v_W/a \ll \vert\bar{\omega}\vert
1692: \ll v_C/d$ does not exist. We also
1693: find that $K = K'_{eff} \equiv 2K_CK_W/(K_C+K_W)$ rather than $K_W$ for
1694: $\vert\bar{\omega}\vert \gg v_C/d$.
1695: Thus for a quantum wire system which has the two
1696: contact regions and only barriers at the two contact-QW junctions, the
1697: conductance in the highest temperature regime of $T \gg T_d$ is
1698: \beq
1699: g = g_0 K_L [ 1 - {\tilde c}_1 ({T \over \Lambda})^{2(K'_{eff} - 1)}
1700: (|V(d)|^2 + |V(l+d)|^2) ] ~.
1701: \eeq
1702: Here ${\tilde c}_1$ is a dimensionful
1703: constant which will depend on factors like $v_W$ and hence also the gate
1704: voltage $V_G$. For the intermediate temperature regime of $T_l \ll T \ll T_d$,
1705: we find the conductance to be
1706: \beq
1707: g = g_0 K_L [ 1 - {\tilde c}_2 ({T_d \over \Lambda})^{2(K'_{eff} - 1)}
1708: ({T \over T_d})^{2({\tilde K}_{eff}-1)} (|V(d)|^2 + |V(l+d)|^2) ] ~,
1709: \eeq
1710: where ${\tilde K}_{eff} = 2K_L K_W/(K_L + K_W)$ as before, and ${\tilde c}_2$
1711: is also a dimensionful constant dependent on $V_G$. Finally, for the lowest
1712: temperature regime of $T \ll T_l$, we obtain
1713: \beq
1714: g = g_0 K_L [ 1 - {\tilde c}_3 ({T \over T_l})^{2(K_L - 1)} ({T_d \over
1715: \Lambda})^{2(K'_{eff} - 1)} ({T_l \over T_d})^{2({\tilde K}_{eff} - 1)}
1716: |V(0) + V(l+2d)|^2 ] ~,
1717: \eeq
1718: where ${\tilde c}_3$ too is a dimensionful constant dependent on $V_G$. Thus,
1719: we can see that any barrier or impurity placed anywhere inside the QW will
1720: always give a perturbative renormalization to the conductance which will be
1721: dependent on the gate voltage and hence can {\it never} be flat or even
1722: channel independent. The conductance expressions for the case of spinful
1723: electrons can be found for this case in exactly the same way as before.
1724:
1725: \section{\bf Effects of a Magnetic Field}
1726:
1727: In this section, we will study the effects of an in-plane magnetic field on
1728: the conductivity of a quantum wire. In general, a magnetic field couples to
1729: both the spin (Zeeman coupling) and the orbital motion of an electron.
1730: However, orbital motion is not possible in an in-plane magnetic field
1731: because the electrons are constrained to move only in the plane.
1732: Thus we will only consider the
1733: effect of the Zeeman term. This term couples differently to spin up
1734: and spin down electrons; here up and down are defined with respect to the
1735: direction of the magnetic field which may or not be parallel to the quantum
1736: wire. Thus the $SU(2)$ symmetry of rotations is explicitly broken. We will now
1737: see that the spin and charge degrees of freedom do not decouple any longer.
1738: Our findings reveal that
1739:
1740: \noindent (a) for low magnetic fields (of about $0-3T$ for
1741: Ga-As systems), the Zeeman splitting of the Fermi energies of the
1742: two spin species of electrons in the QW is very small, and its effects
1743: can be ignored.
1744:
1745: \noindent (b) for intermediate magnetic fields (of about $3-8T$),
1746: the Zeeman splitting of the Fermi energies of up and down spins becomes
1747: appreciable; up and down spins see the two barrier
1748: strengths renormalize differently because of the Zeeman splitting, giving
1749: rise to an odd-even effect in the conductance of the two spin species.
1750:
1751: \noindent (c) at still higher magnetic fields (of about $8-16T$),
1752: when each of the earlier sub-bands is completely Zeeman split into two
1753: spin-split sub-bands, conductance steps will be seen in multiples of
1754: $g_0=e^2/h$; the odd-even effect will be most pronounced here with
1755: odd numbered spin-split sub-bands (containing only aligned
1756: moments) having a much less renormalized conductance and
1757: even numbered spin-split sub-bands (containing only
1758: anti-aligned moments) having a much more renormalized conductance, and
1759: we can treat each spin-split sub-band as an effectively spinless TLL system.
1760:
1761: \noindent (d) at magnetic fields much higher than this, all the spins in the
1762: system will be spin polarized.
1763:
1764: \subsection{The infinite TLL Quantum Wire and the Odd-Even Effect}
1765:
1766: Let us first consider an infinitely long wire containing noninteracting
1767: electrons. A magnetic field $h$ contributes the following term to the
1768: Hamiltonian
1769: \beq
1770: - ~g \mu_B h ~S_{z,total} ~=~ - ~\frac{g \mu_B h}{2} ~(~ \rho_{0\ua} ~
1771: +~ \frac{1}{\sqrt \pi} ~\partial_x \phi_\ua ~-~ \rho_{0\da} ~-~
1772: \frac{1}{\sqrt \pi} ~\partial_x \phi_\da ~)~,
1773: \label{magham}
1774: \eeq
1775: where $g$ is the gyromagnetic ratio (which is $2$ for free electrons but
1776: may be substantially smaller in quantum wire systems), $\rho_{0\ua} ,
1777: \rho_{0\da}$ respectively denote the mean density of the spin up
1778: and spin down electrons, and $\phi_\ua , \phi_\da$ denote the
1779: bosonic fields for the spin up and down electrons. The density terms
1780: $\rho_{0\sigma} = \rho_{0\ua} - \rho_{0\da}$
1781: have a bigger effect than the derivative terms $\partial_x
1782: \phi_\sigma$; by altering the chemical potentials for spin up and down
1783: electrons, these terms lead to different Fermi momenta
1784: and therefore to different Fermi velocities $v_{F\ua}$
1785: and $v_{F\ua}$ for the two kinds of electrons.
1786:
1787: We now add a density-density interaction of the form ${\cal U} \rho^2 /2$
1788: where $\rho = \rho_\ua + \rho_\da$. (For instance, this may describe a
1789: short-range Coulomb repulsion as in the Hubbard model; in that case $\cal U$
1790: is positive). The bosonized Lagrangian density takes the form
1791: \bea
1792: {\cal L} ~=~ & & \frac{1}{2v_{F\ua}} ~(\partial_t \phi_\ua )^2 ~-~
1793: \frac{v_{F\ua}}{2} ~(\partial_x \phi_\ua )^2 ~+~
1794: \frac{1}{2v_{F\da}} ~(\partial_t \phi_\da)^2 ~-~
1795: \frac{v_{F\da}}{2} ~(\partial_t \phi_\da)^2 \nonum \\
1796: & & -~ \frac{\cal U}{2\pi} ~(~\partial_x \phi_\ua ~+~ \partial_x
1797: \phi_\da ~)^2 ~,
1798: \label{hamlag1}
1799: \eea
1800: where we have dropped some additive constants, and have only kept terms
1801: which are quadratic in the fields. We can rediagonalize this Lagrangian by
1802: defining two new fields $\phi_i$, velocities $v_i$ and interaction parameters
1803: $K_i$ (where $i=+,-$), and a mixing angle $\gamma$, where
1804: \bea
1805: \phi_+ &=& {\sqrt {K_+ v_+}} ~(\frac{1}{\sqrt {v_{F\ua}}} \cos
1806: \gamma ~\phi_\ua + \frac{1}{\sqrt {v_{F\da}}} \sin \gamma ~\phi_\da )
1807: \equiv p \phi_\ua + q \phi_\da , \nonum \\
1808: \phi_- &=& {\sqrt {K_- v_-}} ~( - \frac{1}{\sqrt {v_{F\ua}}} \sin
1809: \gamma ~\phi_\ua + \frac{1}{\sqrt {v_{F\da}}} \cos \gamma ~\phi_\da )
1810: \equiv r \phi_\ua + s \phi_\da ,
1811: \label{transf1}
1812: \eea
1813: and
1814: \bea
1815: v_+^2 ~+~ v_-^2 ~&=&~ v_{F\ua}^2 ~+~ v_{F\da}^2 ~+~ \frac{\cal U}{\pi} ~
1816: (~ v_{F\ua} ~+~ v_{F\da} ~)~, \nonum \\
1817: (~ v_+^2 ~-~ v_-^2 ~)~ \cos ~(2\gamma) ~&=&~ v_{F\ua}^2 ~-~
1818: v_{F\da}^2 ~+~ \frac{\cal U}{\pi} ~ (~ v_{F\ua} ~-~ v_{F\da} ~)~, \nonum \\
1819: (~ v_+^2 ~-~ v_-^2 ~)~ \sin ~(2\gamma) ~&=&~ \frac{2{\cal U}}{\pi} ~{\sqrt {
1820: v_{F\ua} v_{F\da}}}~, \nonum \\
1821: K_+ v_+ ~(~ \frac{\cos^2 \gamma}{v_{F\ua}} ~+~ \frac{\sin^2
1822: \gamma}{v_{F\da}} ~) ~&=& ~1 ~, \nonum \\
1823: K_- v_- ~(~ \frac{\sin^2 \gamma}{v_{F\ua}} ~+~ \frac{\cos^2
1824: \gamma}{v_{F\da}} ~) ~&=& ~1 ~.
1825: \label{transf2}
1826: \eea
1827: The Lagrangian density in Eq. (\ref{hamlag1}) then takes the decoupled form
1828: \beq
1829: {\cal L} ~=~ \frac{1}{2K_+ v_+} ~(\partial_t \phi_+ )^2 ~-~ \frac{v_+}{2K_+} ~
1830: (\partial_x \phi_+ )^2 ~+~ \frac{1}{2K_- v_-} ~(\partial_t \phi_- )^2 ~-~
1831: \frac{v_-}{2K_-} ~(\partial_t \phi_- )^2 ~.
1832: \label{hamlag2}
1833: \eeq
1834: Thus the charge and spin degrees of freedom get mixed since the fields
1835: $\phi_+$ and $\phi_-$ which diagonalize the Lagrangian will generally be
1836: different from the fields $\phi_\rho = (\phi_\ua + \phi_\da)/\sqrt{2}$
1837: and $\phi_\sigma = (\phi_\ua - \phi_\da)/\sqrt{2}$. Note that if the
1838: magnetic field $h$
1839: is zero, then $v_{F\ua} = v_{F\da}$ and $\gamma = \pi /4$; $\phi_+$ and
1840: $\phi_-$ are then identical (up to a sign) to the charge and spin fields
1841: $\phi_\rho$ and $\phi_\sigma$.
1842:
1843: We will now present the RG equations for a weak $\delta$-function
1844: impurity placed at the origin. For this, we need to
1845: compute the scaling dimension of the impurity term
1846: in the Lagrangian. We first write the impurity term in terms of fermionic
1847: fields $\psi_{\ua}(0)$ and $\psi_{\da}(0)$, and then in terms of
1848: the bosonic fields $\phi_{\ua}(0)$ and $\phi_{\da}(0)$ at the origin as
1849: \bea
1850: L_{imp} &=& V(0) (\psi_{\ua}^{\dag}(0) \psi_{\da}(0)
1851: + \psi_{\da}^{\dag}(0) \psi_{\ua}(0)) \nonum \\
1852: &=& V_1 \Lambda ~[\cos(2\sqpi\phi_{\ua}(0)) +\cos(2\sqpi\phi_{\da}(0))] ~.
1853: \eea
1854: We then invert the relations between $\phi_{\pm}$ and
1855: $\phi_{\ua,\da}$ given above in order to rewrite the above
1856: expression for the impurity in terms of the diagonal fields $\phi_{\pm}$.
1857: The scaling dimensions of the impurity terms are then found to be
1858: \beq
1859: D_{V_{\ua}} = v_{F\ua} (\frac{\cos^{2}\gamma}{v_+}
1860: + \frac{\sin^{2}\gamma}{v_-}) ~~~,~~~ D_{V_{\da}} =
1861: v_{F\da} (\frac{\sin^{2}\gamma}{v_+} + \frac{\cos^{2}\gamma}{v_-}).
1862: \eeq
1863: Hence the RG equations for $V_{\ua}$ and $V_{\da}$ are given by
1864: \bea
1865: \frac{dV_{\ua}}{d\lambda} &=& ( 1 - D_{V_{\ua}}) V_{\ua} \nonum \\
1866: \frac{dV_{\da}}{d\lambda} &=& ( 1 - D_{V_{\da}}) V_{\da} ~,
1867: \eea
1868: where $V_\ua$ and $V_\da$ both start from the value $V_1$ at the microscopic
1869: length scale.
1870:
1871: We can now study what happens in the presence of strong and
1872: weak magnetic fields. But let us first remind ourselves of the following
1873: relations (which result from the Zeeman splitting),
1874: \bea
1875: v_{F\ua,\da} &=& v_F \sqrt{1 \pm \frac{g\mu_{B} h}{2E_{F1D}}} \nonum \\
1876: \tan(2\gamma) &=& \frac{2{\cal U} \sqrt{v_{F\ua}v_{F\da}}/\pi}
1877: {(v_{F\ua}-v_{F\da})(v_{F\ua}+v_{F\da} + {\cal U}/\pi)} ~,
1878: \eea
1879: where $v_F ={\sqrt {2E_{F1D}/m}}$ is the Fermi velocity in the absence of a
1880: magnetic field. Therefore, in the limit of a strong magnetic field where
1881: the Zeeman splitting of the two spin species is much larger
1882: than the short ranged interaction energy $\cal U$ (i.e., ${\cal U} \ll \vert
1883: v_{F\ua}-v_{F\da}\vert$ and $\gamma \ll \pi/2$), we can approximate
1884: the relations for the two velocities
1885: $v_{\pm}$ (to linear order in ${\cal U}/(v_{F\ua}-v_{F\da})$) as
1886: \bea
1887: v_+ &=& v_{F\ua} + \delta_+ ~~~,~~~ \delta_+\ll v_{F\ua} \nonum \\
1888: v_- &=& v_{F\da} + \delta_- ~~~,~~~ \delta_-\ll v_{F\da} ~.
1889: \eea
1890: By putting these relations for $v_{\pm}$ into the expressions given
1891: above relating $v_{\pm}$ and $v_{F\ua,\da}$, we get
1892: \beq
1893: \delta_+ ~=~ \delta_- ~=~ \frac{\cal U}{2\pi} ~.
1894: \eeq
1895: Now, using this (together with the fact that $\gamma \ll \frac{\pi}{2}$) in
1896: the two RG equations obtained above gives us
1897: \bea
1898: \frac{dV_{\ua}}{d\lambda} &=& (1 - \frac{v_{F\ua}}{v_+}) V_{\ua} \nonum \\
1899: &\simeq& \frac{\cal U}{2\pi v_{F\ua}} V_{\ua} ~,
1900: \eea
1901: and
1902: \bea
1903: \frac{dV_{\da}}{d\lambda} &=& ( 1 - \frac{v_{F\da}}{v_-}) V_{\da} \nonum \\
1904: &\simeq& \frac{\cal U}{2\pi v_{F\da}} V_{\da} ~.
1905: \eea
1906: Now, these two RG equations indicate that as $v_{F\ua}$ is larger than
1907: $v_{F\da}$, the renormalized impurity strength felt by those electrons
1908: which have their magnetic moments aligned with the external B field is less
1909: than the renormalized impurity strength felt by the electrons which have their
1910: magnetic moments anti-aligned with the B field. Furthermore, if the B field
1911: is further increased, we will reach a situation when alternate sub-bands
1912: in the QW will be populated by either only the aligned or only the
1913: anti-aligned electrons. In this regime, the difference in back-scattering
1914: felt by the two species of electrons will be very clear from the alternating
1915: weak and strong corrections to the conductance. To be more specific, all odd
1916: numbered sub-bands will show much less corrections to the perfect conductance
1917: (as they will be populated by electrons aligned with the magnetic field),
1918: while all even numbered
1919: sub-bands will show much greater corrections to the perfect conductance
1920: (as they will be populated by the anti-aligned electrons). This
1921: {\it odd-even} effect had, in fact, been predicted by a two-band TLL study of
1922: Kimura {\it et al} \cite{kimura}, but its explanation
1923: on the grounds of impurity
1924: renormalization is now made clear. Furthermore, this effect has been recently
1925: observed by Liang {\it et al} \cite{liang}, and we will discuss their
1926: observations in a later section. It should be noted here that though the
1927: odd-even effect is easy to show upon taking the limit of ${\cal U} \ll
1928: \vert v_{F\ua}
1929: - v_{F\da} \vert$, the existence of this phenomenon does not need
1930: this limit to be taken. Furthermore, we also find that upon taking the limit
1931: of $h \ll E_{F1D}$ (i.e., weak magnetic field with $v_{F\ua} = v_{F\da} =
1932: v_F$),
1933: \beq
1934: \frac{dV_{\ua}}{d\lambda} = \frac{dV_{\da}}{d\lambda} = \frac{\cal U}{2\pi
1935: v_F} ~,
1936: \eeq
1937: which tells us that the odd-even effect vanishes in the weak magnetic field
1938: limit. Finally, we comment on the fact that the odd-even effect discussed
1939: above gives rise to the possibility of the creation of a spin-valve (i.e.,
1940: a spin polarized current creating device) in these QW systems. Even though
1941: the odd-even effect needs a high magnetic field to be observed in current
1942: day experiments \cite{liang}, it may be possible to employ artificial barriers
1943: like negatively-biased finger gates to heighten the difference in
1944: renormalization of the up and down spin electrons at lower magnetic fields.
1945: At this point, however, quantitative predictions are difficult to make.
1946:
1947: \subsection{A study of our model for the QW with a magnetic field}
1948:
1949: Having discussed how to obtain a diagonal Lagrangian when both a magnetic
1950: field and interactions are present as well as shown the interesting odd-even
1951: effect that takes place because of an impurity in an infinite TLL in the
1952: presence of an external B field, we will now study what happens when the
1953: model in Sec. 2 is placed in a magnetic field. In the regions $x < 0$ and
1954: $x > l+2d$, we have a system of noninteracting electrons parametrized by
1955: velocities $v_{F\ua}$ and $v_{F\da}$. In the regions of the
1956: contacts $0<x<d$ and $l+d<x<l+2d$, we have an interacting system parametrized
1957: by two velocities $v_{C+}$, $v_{C-}$ and a mixing angle $\gamma_C$. In the
1958: quantum wire $d<x<l+d$, the system is parametrized by the velocities $v_{W+}$,
1959: $v_{W-}$ and a mixing angle $\gamma_W$. The last six parameters are functions
1960: of $v_{F\ua}$, $v_{F\da}$ and the strengths of the interaction in the contacts
1961: and quantum wire. The action for this model is given by
1962: \bea
1963: S_0 &=& \int d\tau ~[\int_{-\infty}^0 +\int_{l+2d}^\infty]~ dx ~\{ ~{1 \over
1964: 2K_L}[ {1\over v_{F\ua}} (\pt\phi_\ua)^2 + v_{F\ua} (\px\phi_\ua)^2 ]~
1965: +~ [\ua \rightarrow \da ] ~\} \nonum \\
1966: &+& \int d\tau ~[\int_{0}^d +\int_{l+d}^{l+2d}] ~dx ~\{ ~{1 \over 2 K_{C+}}
1967: [{1\over v_{C+}} (\pt\phi_{C+})^2 + v_{C+} (\px \phi_{C+})^2 ] ~+~
1968: [ C+ \rightarrow C- ] ~\} \nonum \\
1969: &+& \int d\tau ~\int_{d}^{l+d} ~dx ~\{ ~{1\over 2 K_{W+} }[{1\over v_{W+}}
1970: (\pt \phi_{W+})^2 + v_{W+} (\px\phi_{W+})^2 ] ~+~ [W+ \rightarrow W- ] ~\} ~.
1971: \nonum \\
1972: &&
1973: \label{actmag}
1974: \eea
1975: Note that we have
1976: ignored the junction barriers and the gate voltage for the moment. It is
1977: worth mentioning here that we have verified, by performing a calculation of
1978: the kind outlined in Ref. \cite{maslov}, that our model for the
1979: QW system when placed in an external magnetic field and in the absence of
1980: any barriers/impurities still gives perfect conductance in the dc limit for
1981: each sub-band.
1982:
1983: We begin by noting that we will present here the calculation for the
1984: case when the mixing angle $\gamma$ is the same in both the contacts as well
1985: as the QW, i.e., the short ranged electron-electron interaction $\cal U$ is
1986: equal in all the three TLLs. Though this is not necessarily the case in a real
1987: system, we will present it as it considerably simplifies the computations
1988: while providing us with an adequate discussion of all the important results
1989: for effective actions, conductance expressions, resonances, etc.
1990: Later, we will briefly discuss the case in which
1991: the mixing angle is different in the contact and QW regions. The explicit
1992: derivation of the effective action is given in Appendix C.
1993: The high frequency effective Lagrangian density ($\bo \gg v_{C\pm}/d ,
1994: v_{W\pm}/l$) simplifies to
1995: \bea
1996: {\cal L}_{eff, high} &=& {|\bo| \over 2}[(\frac{1}{K_L} +
1997: \frac{p^2}{K_{C+}} + \frac{r^2}{K_{C-}})(\tphi_{1\ua}^2 + \tphi_{4\ua}^2)
1998: + (\frac{1}{K_L} + \frac{q^2}{K_{C+}} + \frac{s^2}{K_{C-}})
1999: (\tphi_{1\da}^2 + \tphi_{4\da}^2) \nonum \\
2000: &&~~~~~ +~2(\frac{pq}{K_{C+}}+\frac{rs}{K_{C-}})(\tphi_{1\ua}
2001: \tphi_{1\da} + \tphi_{4\ua}\tphi_{4\da})] \nonum \\
2002: && +~~{|\bo| \over 2}[\{p^2(\frac{1}{K_{C+}}+\frac{1}{K_{W+}}) +r^2(
2003: \frac{1}{K_{C-}}+\frac{1}{K_{W-}})\}(\tphi_{2\ua}^2 +\tphi_{3\ua}^2) \nonum \\
2004: &&~~~~~~~~~~ +~ \{q^2(\frac{1}{K_{C+}}+\frac{1}{K_{W+}}) +s^2(
2005: \frac{1}{K_{C-}}+\frac{1}{K_{W-}})\}(\tphi_{2\da}^2 +\tphi_{3\da}^2) \nonum \\
2006: &&~~~~~~~~~~ +~ 2\{pq(\frac{1}{K_{C+}}+\frac{1}{K_{W+}})
2007: +rs(\frac{1}{K_{C-}}+\frac{1}{K_{W-}})\}(\tphi_{2\ua}\tphi_{2\da}
2008: + \tphi_{3\ua}\tphi_{3\da})] \nonum \\
2009: && +~~ {\cal L}_{gate}(\phi_{2\ua,\da},
2010: \phi_{3\ua,\da}) + {\cal L}_{imp}(\phi_{1\ua,\da}, \phi_{4\ua,\da}) .
2011: \eea
2012: We can clearly see the separation between the outer two and inner two fields.
2013: This means that we can integrate out the inner fields
2014: $\tphi_{2\ua,\da}$ and $\tphi_{3\ua,\da}$ without any
2015: further work and be left with a high frequency effective action dependent
2016: on $\tphi_{1\ua,\da}$ and $\tphi_{4\ua,\da}$ exactly as
2017: given above (and without any influence of the gate voltage $V_G$ either).
2018: We can also make the prediction that the conductance corrections due to
2019: barriers at the outer two junctions will have temperature power-laws which
2020: will be combinations of $K_L$ and $K_{C\pm}$ (much like those seen before) and
2021: also that it will be not be dependent on the gate voltage.
2022:
2023: In the intermediate frequency range of $v_{W\pm}/l \ll \bo \ll v_{C\pm}/d$, we
2024: get (after integrating out the inner fields $\tphi_{2\ua,\da}$ and
2025: $\tphi_{3\ua,\da}$),
2026: \bea
2027: {\cal L}_{eff,int} &=& {|\bo| \over 2}[(\frac{1}{K_L} + \frac{p^2}{K_{W+}}
2028: + \frac{r^2}{K_{W-}})(\tphi_{1\ua}^2 + \tphi_{4\ua}^2)
2029: + (\frac{1}{K_L} + \frac{q^2}{K_{W+}} + \frac{s^2}{K_{W-}})(\tphi_{1\da}^2
2030: + \tphi_{4\da}^2) \nonum \\
2031: &&~~~~ +~2(\frac{pq}{K_{W+}}+\frac{rs}{K_{W-}})(\tphi_{1\ua}
2032: \tphi_{1\da} + \tphi_{4\ua}\tphi_{4\da})] \nonum \\
2033: && + \frac{{\tilde V}_G (p-q)(s-r)}{(ps-qr)}(\phi_{4\ua} - \phi_{1\ua} +
2034: \phi_{4\da} - \phi_{1\da}) ~,
2035: \eea
2036: where $\alpha_{C\pm} = v_{C\pm}/(K_{C\pm}d)$ are the charging energies for
2037: the $\tphi_{\pm}$ fields, and they appear because of the growth of the
2038: coherence in the system over the contact regions. We can now predict that
2039: the conductance corrections due to barriers at the outer two junctions will
2040: have temperature power-laws which will be combinations of $K_L$ and
2041: $K_{W\pm}$ (again like those seen previously), and that this correction will
2042: definitely be dependent on the gate voltage.
2043:
2044: Finally, in the low frequency
2045: limit of $\bo \ll v_{W\pm}/l$, we get
2046: \bea
2047: {\cal L}_{eff,low} &=& {|\bo| \over 2K_L} (\tphi_{1\ua}^2 + \tphi_{1\da}^2 +
2048: \tphi_{4\ua}^2 +\tphi_{4\da}^2) \nonum \\
2049: && +\sum_{\pm} {v_{C\pm}\over 2d}[\phi_{2\pm}^2 +\phi_{3\pm}^2 +
2050: \phi_{1\pm}^2 + \phi_{4\pm}^2 -2\phi_{2\pm} \phi_{1\pm} -2\phi_{3\pm}
2051: \phi_{4\pm}] \nonum \\
2052: && +\sum_\pm {v_{W\pm}^2\over 2l v_{C\pm}}(\phi_{2\pm} -\phi_{3\pm})^2
2053: \nonum \\
2054: && +\sum_\pm \frac{{\tilde V}_G}{2(ps-qr)}[(s-r)(\phi_{3+}-\phi_{2+}) +
2055: (p-q)((\phi_{3-}-\phi_{2-})] .
2056: \eea
2057: Integrating out the inner fields, we are then left with an effective action in
2058: terms of the new current and charge variables respectively
2059: \bea
2060: \phi_{14\ua,\da} &=& \frac{\phi_{1\ua,\da} + \phi_{4\ua,\da}}{2} \nonum \\
2061: n_{14\ua,\da} &=& \frac{\phi_{4\ua,\da} - \phi_{1\ua,\da}}{\sqpi} ~.
2062: \eea
2063: that span the coherent TLL system
2064: between the two lead-contact junction barriers in this
2065: low frequency regime. Thus, we get the effective Lagrangian density in this
2066: regime as
2067: \bea
2068: {\cal L}_{eff,low} &=& {|\bo| \over K_L}[\tphi_{14\ua}^2 + \tphi_{14\da}^2 +
2069: \frac{\pi}{4} {\tilde n}_{14\ua}^2 + \frac{\pi}{4} {\tilde n}_{14\da}^2]
2070: \nonum \\
2071: && + \frac{U_1}{2}(p n_{14\ua} + q n_{14\da} + n_{01})^2 + \frac{U_2}{2}(r
2072: n_{14\ua} + s n_{14\da} + n_{02})^2 + {\cal L}_{imp},
2073: \eea
2074: where
2075: \bea U_1 &=& \frac{\pi v_{W+}\alpha_{+}\beta_{+}} {v_{C+}\alpha_+
2076: + 2v_{W+}\beta_+} ~~,~~ U_2 = \frac{\pi v_{W-}\alpha_{-}\beta_{-}}
2077: {v_{C-}\alpha_- + 2v_{W-}\beta_-} \nonum \\
2078: n_{01} &=& \frac{{\tilde V}_G v_{C+}(s-r)}{\pi v_{W+}\beta_+ (ps-qr)} ~~,~~
2079: n_{02} = \frac{{\tilde V}_G v_{C-}(p-q)}{\pi v_{W-}\beta_- (ps-qr)} \nonum \\
2080: {\rm and} \quad \alpha_{\pm} &=& \frac{v_{C\pm}}{K_{C\pm}d} ~~,~~
2081: \beta_{\pm} = \frac{v_{W\pm}}{K_{W\pm}d},
2082: \eea
2083: and the term coming from the two barriers can be written as
2084: \beq
2085: {\cal L}_{imp} = 2V \Bigl( \cos(2\sqpi\phi_{14\ua}) \cos(\pi n_{14\ua}) +
2086: \cos(2 \sqpi\phi_{14\da}) \cos(\pi n_{14\da}) \Bigr) ~.
2087: \eeq
2088: It becomes clear from the above expression that the temperature power-law
2089: will be dependent on the lead interaction parameter $K_L = 1$, and the
2090: conductance correction will, in this regime, be temperature independent (as
2091: seen previously). Furthermore, the length corrections will be gate
2092: voltage dependent.
2093: Let us now study the possible resonance symmetries of the low frequency
2094: effective action given above. Even though the structure of this expression
2095: is more complicated than those encountered previously, we can rewrite the
2096: two charging terms as follows:
2097: \beq
2098: {\cal L}_{charging} = \frac{(U_1 p^2 + U_2 r^2)}{2} [n_{14\ua}
2099: + a n_{14\da} -b]^2 + f(n_{14\da}) + {\rm const} ~,
2100: \eeq
2101: where
2102: \bea
2103: a &=& \frac{U_1 pq + U_2 rs}{U_1 p^2 + U_2 r^2} \nonum \\
2104: b &=& \frac{U_1 n_{01} + U_2 n_{02}}{U_1 p^2 + U_2 r^2} ~,
2105: \eea
2106: and $f(n_{14\da})$ is a quadratic function of the field
2107: $n_{14\da}$ only. Thus, we can see that if we set
2108: \beq
2109: a n_{14\da} - b = - Z - \frac{1}{2} ~,
2110: \eeq
2111: we get a resonance in the $n_{14\ua}$ parts of the charging and barrier
2112: terms whenever $n_{14\ua} = Z$ or $Z+1$ and one makes the transformation
2113: of $\phi_{14\ua}\rightarrow \phi_{14\ua}\pm \sqpi/2$. This means that
2114: only the transport of all up-spin electrons through the two barriers
2115: is at resonance, and this is clearly a one-parameter tuned resonance.
2116: A one-parameter tuned resonance for only the transport of all down-spin
2117: electrons through the two barriers can be found in exactly the same manner
2118: by rewriting the above charging expressions but for $n_{14\da}$
2119: instead of $n_{14\ua}$. We also find another resonance given by
2120: \bea
2121: \frac{q}{p}n_{14\da} + \frac{1}{p}n_{01} &=& - Z - \frac{1}{2} ~, \nonum \\
2122: {\rm and} \quad \frac{s}{r}n_{14\da} + \frac{1}{r}n_{02} &=& - Z -
2123: \frac{1}{2} ~,
2124: \eea
2125: where $Z$ is the same integer in both equations; then there exists a possible
2126: resonance whenever $n_{14\ua} = Z$ or $Z+1$, and one makes the
2127: transformation of $\phi_{14\ua,\da}\rightarrow
2128: \phi_{14\ua,\da}\pm \sqpi/2$. This resonance symmetry would
2129: lead to a vanishing of the barrier terms
2130: in the effective action, and would
2131: correspond to the transfer of an electron across the system. One
2132: can immediately see that the above two conditions on $n_{14\da}$
2133: mean that two parameters, here the gate voltage $V_G$ and the external
2134: magnetic field $h$, have to be manipulated to achieve this resonance
2135: condition. Such a resonance will, therefore, be much more difficult to
2136: observe experimentally. However, this resonance will lead to a complete
2137: vanishing of all back-scattering (and hence the conductance corrections)
2138: while the two one-parameter tuned resonances will give only partial
2139: lessening of the conductance corrections.
2140:
2141: The two one-parameter tuned resonances can prove useful in
2142: creating a spin-valve. Even at a low magnetic field, the odd-even effect
2143: can be enhanced by using stronger artificial barriers (e.g., by
2144: employing finger gates over the channel) or making the length of the channel
2145: longer or working at lower temperatures, together with tuning the
2146: transport of only one spin species of electrons through the two barriers at
2147: resonance. Thus one can create an enhanced spin polarized electron current
2148: output from the QW system.
2149:
2150: Before we go on to computing the conductance through the system for the above
2151: model in the presence of the magnetic field, let us make some remarks about
2152: the case when the mixing angle in the QW is taken to be different from that
2153: in the contacts. A long calculation does give expressions for the
2154: effective action in the three frequency regimes similar to those obtained
2155: above, but with two sets of the transformation coefficients
2156: relating the $\phi_{\pm}$ and $\phi_{\ua,\da}$ fields. However,
2157: the integrating out of the inner fields is a far more difficult task;
2158: furthermore, the analysis reveals that the only possible resonance symmetry
2159: of the low frequency effective action is one that needs at least four
2160: parameters to be manipulated. We will, therefore, not present these
2161: results as we do not find anything substantially
2162: new from the analysis compared to the simpler case of equal mixing angles.
2163:
2164: \subsection{Conductance of our model for the QW with a magnetic field}
2165:
2166: We will begin by a re-writing the RG equation, obtained by Safi and Schulz
2167: \cite{safilong} for an impurity placed within a QW of a finite size and
2168: connected to Fermi leads, in a way which will be convenient to use
2169: in computing the conductance expressions for our model of the QW with
2170: barriers even in the presence of the external magnetic field. We begin by
2171: quoting the expression for the RG flow found for an impurity in Ref.
2172: \cite{safilong},
2173: \beq
2174: \frac{dV_{m_{\rho},m_{\sigma}}(x)}{d\lambda} = [1 - \frac{1}{2}(m_{\rho}^2
2175: \frac{dU_{\rho} (x,x,\tau_0 e^\lambda)}{d\lambda} + m_{\sigma}^2 \frac{dU_{
2176: \sigma} (x,x,\tau_0 e^\lambda)}{d\lambda})] V_{m_{\rho},m_{\sigma}}(x),
2177: \eeq
2178: where
2179: \bea
2180: U_{\rho}(x,x,\tau_0 e^\lambda) &=& 2 [G_{\rho}(x,x,\tau_0) - G_{\rho}
2181: (x,x,\tau_0 e^\lambda)] \nonum \\
2182: U_{\sigma}(x,x,\tau_0 e^\lambda) &=& 2 [G_{\sigma}(x,x,\tau_0)
2183: - G_{\sigma}(x,x,\tau_0 e^\lambda)].
2184: \eea
2185: Now with $\phi_{\rho} = (\phi_{\ua} + \phi_{\da})/{\sqrt 2}$ and
2186: $\phi_{\sigma} = (\phi_{\ua} - \phi_{\da})/{\sqrt 2}$, we can write
2187: \bea
2188: G_{\rho}(x,x,\tau_0 e^\lambda) &=& <\phi_{\rho}(x,\tau_0 e^\lambda)
2189: \phi_{\rho}(x,0)> = \frac{1}{2}[<\phi_{\ua}\phi_{\ua}> + <\phi_{\da}
2190: \phi_{\da}> + 2<\phi_{\ua}\phi_{\da}>] \nonum \\
2191: G_{\sigma}(x,x,\tau_0 e^\lambda) &=& <\phi_{\sigma}(x,\tau_0 e^\lambda)
2192: \phi_{\sigma}(x,0)> = \frac{1}{2}[<\phi_{\ua}\phi_{\ua}> + <\phi_{\da}
2193: \phi_{\da}> - 2<\phi_{\ua}\phi_{\da}>], \nonum \\
2194: &&
2195: \eea
2196: where the space-time indices are implicit on the right hand sides.
2197: Substituting the expressions
2198: for $U_{\rho}$ and $U_{\sigma}$ given above in the RG equation, and
2199: working with the case for the back-scattering of one electron
2200: $m_{\rho} = m_{\sigma} = 1$, we write the RG equation as
2201: \beq
2202: \frac{dV_{1,1}(x)}{d\lambda} = [1 + \frac{d}{d\lambda}(G_{\ua}(x,x,\tau_0
2203: e^\lambda) + G_{\da}(x,x,\tau_0 e^\lambda))] V_{1,1}(x) ~,
2204: \eeq
2205: where $G_{\ua} = <\phi_{\ua}\phi_{\ua}>$ and
2206: $G_{\da} = <\phi_{\da}\phi_{\da}>$. We will now
2207: use the effective actions found in the various frequency regimes
2208: to obtain the two Green's functions $G_{\ua}$ and $G_{\da}$,
2209: put these in the RG equations and thereby infer the corrections
2210: to the conductance caused by the junction barriers.
2211:
2212: We start with the high and intermediate frequency/temperature effective
2213: actions given earlier for the model when the mixing angle is the same in the
2214: contact and QW regions. Here, we can
2215: see that the final effective action (in terms of only the fields at the outer
2216: two junctions) is the sum of two distinct parts, each of which is an
2217: expression of the kind $\frac{A}{2}\phi_{\ua}^2 +
2218: \frac{B}{2}\phi_{\da}^2 + C\phi_{\ua}\phi_{\da}$
2219: separately for fields $\phi_1$ and $\phi_4$. This tells us that we can simply
2220: take the sum of the contributions from each of the two incoherent barriers.
2221: Thus, the general expression
2222: \beq
2223: {\cal L}_{eff} = \frac{A}{2\vert\bo\vert}\tphi_{\ua}^2 + \frac{B}{2\vert\bo
2224: \vert}\tphi_{\da}^2 + \frac{C}{\vert\bo\vert}\tphi_{\ua}\tphi_{\da},
2225: \eeq
2226: can be diagonalized in terms of two new fields $\phi_a$ and
2227: $\phi_b$ i.e., written as
2228: \beq
2229: {\cal L}_{eff} = \frac{1}{2\vert\bo\vert} (\lambda_a {\tphi_a}^2
2230: + \lambda_b {\tphi_b}^2) ~,
2231: \eeq
2232: where $\lambda_a$ and $\lambda_b$ are the eigenvalues of the transformation
2233: given by
2234: \beq
2235: \lambda_{a,b} = \frac{A+B}{2} \pm \frac{1}{2}\sqrt{(A-B)^2+C^2} ~.
2236: \eeq
2237: Then,
2238: \bea
2239: G_{\bo a} &=& <\tphi_a \tphi_a> = \frac{1}{\lambda_a
2240: \vert\bo\vert} \nonum \\
2241: G_{\bo b} &=& <\tphi_b \tphi_b> = \frac{1}{\lambda_b \vert\bo\vert}.
2242: \eea
2243: Using the two eigenvectors corresponding to these two eigenvalues, we obtain
2244: $G_{\bo\ua}$ and $G_{\bo\da}$ as
2245: \beq
2246: G_{\bo\ua} = \frac{C^2}{\vert\bo\vert \{(\lambda_a-A)(\lambda_b-B)-C^2\}^2}
2247: [C^2 G_{\bo a} + (\lambda_b - B)^2 G_{\bo b}] ,
2248: \eeq
2249: and
2250: \beq
2251: G_{\bo\da} = \frac{C^2}{\vert\bo\vert \{(\lambda_a-A) (\lambda_b-B)-C^2\}^2}
2252: [(\lambda_a-A)^2 G_{\bo a} + C^2 G_{\bo b}].
2253: \eeq
2254: We finally obtain an expression for $G_{\bo\ua} + G_{\bo\da}$ as
2255: \bea
2256: G_{\bo\ua} + G_{\bo\da} &=& \frac{C^2}{\vert\bo\vert \{(\lambda_a-A)
2257: (\lambda_b-B)-C^2\}^2} [\frac{C^2 + (\lambda_a - A)^2}{\lambda_a}
2258: + \frac{C^2 + (\lambda_b - B)^2}{\lambda_b}] \nonum \\
2259: &\equiv& \frac{K_{mag}}{2\vert\bo\vert} .
2260: \eea
2261: We can now use the Fourier transform of the above expression to obtain
2262: the temperature and length power-laws for the conductance corrections in
2263: the high and
2264: intermediate frequency regimes. In the high temperature regime of
2265: $T \gg T_d ~(\sim v_{C\pm}/d)$, we get the conductance as
2266: \beq
2267: g = g_0 K_L [1 - c_1(|V(0)|^2 + |V(l+2d)|^2)(\frac{T}{\Lambda})^{(K_{eff,mag}
2268: - 2)}] ,
2269: \eeq
2270: where $c_1$ is a dimensionful constant independent of the gate voltage $V_G$,
2271: and $K_{eff,mag}$ is given by the expression for $K_{mag}$ where
2272: the coefficients $A$, $B$ and $C$ are given by
2273: \bea
2274: A &=& \frac{1}{K_L} + \frac{p^2}{K_{C+}} + \frac{r^2}{K_{C-}} \nonum \\
2275: B &=& \frac{1}{K_L} + \frac{q^2}{K_{C+}} + \frac{s^2}{K_{C-}} \nonum \\
2276: C &=& \frac{pq}{K_{C+}} + \frac{rs}{K_{C-}} ~.
2277: \eea
2278: Similarly, the conductance expression for the intermediate temperature defined
2279: by $T_l ~(\sim v_{W\pm}/l) \ll T \ll T_d$ is given by
2280: \beq
2281: g = g_0 K_L [1 - c_2(|V(0)|^2 + |V(l+2d)|^2)(\frac{{\tilde T}_d}{\Lambda})^{
2282: (K_{eff,mag} -2)}(\frac{T}{{\tilde T}_d}) ^{({\tilde K}_{eff,mag} - 2)}],
2283: \eeq
2284: where $c_2$ is another dimensionful constant which is dependent on the gate
2285: voltage, $T_d$ has replaced $\Lambda$ as the correct cutoff for the
2286: temperature, and ${\tilde K}_{eff,mag}$ is found in exactly the same way as
2287: $K_{eff,mag}$ but with coefficients $A_1$, $B_1$ and $C_1$ defined as
2288: \bea
2289: A_1 &=& \frac{1}{K_L} + \frac{p^2}{K_{W+}} + \frac{r^2}{K_{W-}} \nonum \\
2290: B_1 &=& \frac{1}{K_L} + \frac{q^2}{K_{W+}} + \frac{s^2}{K_{W-}} \nonum \\
2291: C_1 &=& \frac{pq}{K_{W+}} + \frac{rs}{K_{W-}}.
2292: \eea
2293: Finally, we obtain the low frequency conductance expression for the
2294: temperature regime of $T \ll T_l$ as
2295: \beq
2296: g = g_0 K_L [1 - c_3(|V(0) + V(l+2d)|^2)(\frac{{\tilde T}_d}{\Lambda})^{
2297: (K_{eff,mag} -2)}(\frac{T_l}{T_d})^{({\tilde K}_{eff,mag} - 2)}
2298: (\frac{T}{T_l})^{2(K_L - 1)}],
2299: \eeq
2300: where $c_3$ is a dimensionful constant similar in nature to $c_2$, i.e.,
2301: dependent on gate voltage. This expression is also independent of the
2302: temperature for Fermi leads with $K_L = 1$, and the coherence between the
2303: barriers means that this correction term could go to zero at resonance.
2304:
2305: We end by noting that we can again take the limit of ${\cal U} \ll
2306: \vert v_{C\ua} - v_{C\da} \vert$ in our equations to highlight the
2307: existence of the odd-even effect within our model of the QW as well. Upon
2308: taking this limit in the high temperature regime, we find that
2309: \beq
2310: \frac{dV_{\ua}}{d\lambda} \simeq \frac{\cal U}{4\pi v_{C\ua}},
2311: \eeq
2312: while
2313: \beq
2314: \frac{dV_{\da}}{d\lambda} \simeq \frac{\cal U}{4\pi v_{C\da}},
2315: \eeq
2316: where $\cal U$ is the inter-electron interaction term. This clearly shows
2317: that as $v_{C\ua}$ increases and $v_{C\da}$
2318: decreases with an increasing magnetic field, the renormalized barrier seen
2319: by the two spin species of electrons will be different. We also note that,
2320: just like the case of the infinite, homogeneous QW, a weak field of $h \ll
2321: E_{F1D}$ in our model of the QW does not give rise to the odd-even effect.
2322:
2323: In summary, we can see that by turning on an external magnetic field in the
2324: QW system, the up and down spin electrons see different renormalized strengths
2325: of any barriers (or impurities) --- this is the odd-even effect. We speculate
2326: on the possible use of this effect in creating a spin-valve using QW
2327: systems. The effective actions, their resonance symmetries as well as the
2328: temperature and length power-law corrections to the conductance in the
2329: various temperature regimes, however, still follow a pattern similar to
2330: that for a QW without a magnetic field.
2331:
2332: \section{\bf Comparison with the Experiments}
2333:
2334: We now discuss the relevance of this model to many of the experiments that
2335: have been performed so far on quantum wire systems fabricated using
2336: cleaved-edge overgrowth as well as split-gate techniques. But before
2337: doing that, let us reiterate some well-known
2338: observations about the experimental system that we are trying to
2339: model here. In this system, the electrons enter the wire from the
2340: 2DEG reservoirs lying outside the wire with
2341: a Fermi energy $E_F$ whose value (typically around $5 - 10meV$) is fixed by
2342: the parameters of the 2DEG.
2343: Within the quantum wire, the gate voltage produces a discrete set of sub-bands
2344: labeled by an integer $s$ (see Fig. 2); let $E_s$ denote the energies of the
2345: bottoms of these sub-bands. In a sufficiently long quantum wire, we expect
2346: $E_s$ to be constant along the length of the wire provided we are not too
2347: close to either of the junctions. Thus an electron which has energy $E_F$
2348: and enters the sub-band $s$ will have a wave number $k_{Fs}$ inside the
2349: wire given by $k_{Fs}^2 /2m = E_F - E_s$ and a velocity given by
2350: $v_{Fs} = k_{Fs} /m$ \cite{buttiker}.
2351: We know that if $N$ of the $1D$ sub-bands lie below the 2DEG
2352: Fermi energy $E_{F2D}$ (which itself at any finite temperature is
2353: surrounded by a small thermal spread), we will get
2354: $N$ quantized steps in the conductance when the quantum wire is completely
2355: free of any impurities; this statement is true irrespective of the electron
2356: velocities, densities or how they interact among themselves while in the
2357: various channels \cite{safi,maslov,buttiker}.
2358: Now, upon increasing the gate voltage $V_G$, one adds an
2359: energy $eV_G$ to every electron in each of the $1$D sub-bands in the quantum
2360: wire. This has the effect of pushing up each of the sub-bands
2361: by the same energy and can even de-populate the
2362: sub-bands by pushing them above $E_{F2D}$ (see Fig. 2). Thus, changing
2363: the gate voltage decreases the electron density in the quantum wire and
2364: allows the transport process to take place through only a few channels, and
2365: in the extreme limit, only one channel, before cutting off the wire altogether
2366: by pushing all the $1$D sub-bands above the $E_{F2D}$ (this is called
2367: {\it pinch-off}). The conductance measurement which
2368: shows step quantization in terms of rises and plateaus can then be
2369: explained in the following way. Whenever, by decreasing the
2370: gate voltage $V_G$, the bottom one of the 1D sub-bands (which is initially
2371: well above $E_{F2D}$) first touches the top of the thermal spread just
2372: above $E_{F2D}$, that band starts filling up and so
2373: we can see a rise. Once the bottom of this sub-band
2374: crosses the bottom of the thermal spread just below $E_{F2D}$, the rise
2375: is topped off by a plateau which signals that another channel is fully open
2376: to electron transport between the two reservoirs (see Fig. 2). Some of the
2377: earliest experiments with quantum wires free of impurities did indeed reveal
2378: quantization of the conductance in integer steps of $2g_0$ \cite{wees}.
2379:
2380: But later Tarucha {\it et al} \cite{tarucha} performed experiments
2381: with wires of lengths of $2\mu m$ to $10\mu m$ fabricated using split-gate
2382: methods at temperatures from $0.3K$ to $1.1K$,
2383: and found deviations from the perfect quantization of the steps. Attempts
2384: were then made to explain these deviations as due to electron-electron
2385: interactions. Although, a clean TLL wire between Fermi liquid
2386: leads would not lead to renormalization of the conductance
2387: quantization, several authors \cite{furusaki,safilong,maslovlong}, showed
2388: that the presence of impurities in a
2389: TLL connected to Fermi leads would cause renormalization.
2390: However, they expected the renormalizations to be gate voltage
2391: dependent; this was indeed seen by Tarucha {\it et al} \cite{tarucha}.
2392:
2393: However, Yacoby {\it et al} \cite{yacoby} made the following surprising
2394: observation for a quantum wire $2\mu m$ long fabricated in
2395: cleaved-edge overgrowth systems: the dc conductance
2396: showed several nearly flat plateaus whose
2397: heights are uniformly renormalized from the ideal
2398: values of integer multiples of $2g_0$ for measurements made over a
2399: temperature range of $0.3 - 25K$. Similar observations were subsequently
2400: made in several experiments on quantum wires made using the split-gate
2401: technique \cite{facer,liang,liang2,reilly}. In all these experiments
2402: the step heights were increasingly renormalized as either the temperature
2403: was lowered (for a fixed length of quantum wire) or the length of the
2404: quantum wire was increased at a fixed temperature.
2405: Such renormalizations would require back-scattering of electrons.
2406: If these back-scatterings were due to impurities within
2407: the quantum wires, the conductance corrections would be gate
2408: voltage dependent as shown in our calculations. This can
2409: certainly not lead to flat conductance plateaus as seen in the experiments.
2410:
2411: Our model, however, has contact regions independent of the gate voltage and
2412: has barriers at the contacts arising due to the changes in the
2413: nature of the electron-electron interactions and geometry.
2414: Thus, the back-scattering at these barriers is independent of gate
2415: voltage and the sub-band index (as can be seen in our results), and will
2416: lead to conductance plateaus which are flat as a function of
2417: gate voltage and uniform for all the sub-bands at the highest temperatures.
2418: We note that a recent experiment
2419: \cite{picciotto} on a quantum wire system similar to that used by Yacoby
2420: {\it et al} \cite{yacoby} revealed the existence of a
2421: region of length $2 - 6\mu m$ which lies in between the gated quantum wire
2422: region and the 2DEG reservoirs and gives rise to the back-scattering
2423: that causes the flat and uniform renormalization of the conductance of each
2424: sub-band. Such contact regions
2425: correspond to $T_d \sim 0.2 - 0.7 K$. This is much less than most
2426: of the temperature range shown in Fig. 3 of Ref. \cite{yacoby}.
2427: The similar flat and uniform conductance corrections
2428: seen in the experiments of Refs. \cite{facer,liang,liang2,reilly}
2429: seem to suggest that their QW systems also include contact regions and have
2430: $T \gg T_d$.
2431:
2432: Now, as explained earlier for a quantum wire
2433: system in which the contact length $d \ll l$, in the intermediate and
2434: low temperature regimes of $T_l \ll T \ll T_d$ and $T \ll T_l$,
2435: we know that the correct cutoffs for the RG procedure are
2436: $T_d$ and $T_l$ respectively; that is why the length power-laws of
2437: $d$ and $l$ appear in the conductance corrections in these two regimes
2438: besides the customary temperature power-law. We can clearly see that the
2439: inverse length scale $d^{-1}$ (for the contact region) and $l^{-1}$ (for
2440: the wire region) have similar power-laws to those obtained for the
2441: temperature. Thus, one
2442: can qualitatively understand the increase in the conductance with increasing
2443: temperature and its decrease as the length of the quantum wire is increased.
2444: This has been observed by several groups \cite{tarucha,yacoby,facer,liang2}.
2445: Furthermore, one recent experiment using a split-gate QW system
2446: \cite{reilly} shows that the conductance of a $2\mu m$ long QW
2447: at $T = 1K$ shows flat, renormalized plateaus which are replaced by uneven
2448: conductance fluctuations at $T = 50mK$. However a different experiment
2449: \cite{facer} reveals that a QPC created using similar
2450: split-gate methods shows plateaus which are hardly renormalized at higher
2451: temperatures, and no conductance fluctuations are seen at lower temperatures.
2452: This can also be understood from our model: the
2453: conductance corrections due the junction barriers for a quantum wire are
2454: gate voltage independent at higher temperatures, but are dependent on it at
2455: lower temperatures. For the experiment in Ref. \cite{reilly}, $T_l = 0.4K >>
2456: T=50mK$. Hence, resonance effects are expected at these temperatures.
2457: This is in contrast to the conductance corrections for a QPC which are
2458: gate voltage independent at all temperatures. In fact, if the quantum wire
2459: samples of Yacoby {\it et al} have contact regions as long
2460: as $2 - 6\mu m$ (as found by the authors of Ref. \cite{picciotto} on similar
2461: samples), this would suggest that their $2\mu m$ long wire is actually
2462: closer to a quantum point contact. This would help explain the flatness
2463: of the renormalizations seen over a wide temperature range of $0.3 - 25K$.
2464:
2465: We now discuss our attempt to quantitatively understand the variation of
2466: conductance against temperature as given in the inset of Fig. 3 of the work
2467: of Yacoby {\it et al} \cite{yacoby}. The conductance given there is
2468: measured at a fixed value
2469: of the gate voltage on the plateau of the first sub-band (i.e., close to
2470: $2g_0$). We find that the conductance correction versus temperature variation
2471: found by them (i.e., $\delta g \equiv 2g_0 - g$ vs. $T$) is best fitted by a
2472: function of the form
2473: \beq
2474: \delta g = - 0.3512 ~T^{-0.1058 - 0.0345 T}
2475: \eeq
2476: as shown in Fig. 3. We find that the goodness of this fit is given by
2477: the correlation coefficient $R^2 = 0.9955$.
2478: Clearly, this expression for the conductance corrections does not match the
2479: simple form $\delta g \sim T^{-\alpha}$ given in Sec. $4$ for the QW or QPC
2480: systems. The presence
2481: of the $T$ dependent piece in the exponent implies that our model is
2482: only qualitatively correct. Several factors could be important in determining
2483: this complicated temperature dependent power-law. Some of these are:
2484:
2485: \begin{itemize}
2486:
2487: \item a more extended transition region between the leads and the contacts
2488: in which the parameters $K$ and $v$ vary smoothly as a function of $x$,
2489:
2490: \item more extended junction barriers lying within the contacts rather than
2491: the local $\delta$-function barriers that we have studied, and
2492:
2493: \item the possibility of the electron-electron interactions having
2494: a finite range instead of the short-ranged interactions that
2495: we have used to study our TLL systems.
2496:
2497: \end{itemize}
2498:
2499: A detailed quantitative comparison of our model with the experiments would,
2500: therefore, need a more sophisticated treatment taking these factors into
2501: consideration. We should emphasize here that a temperature and length
2502: dependence of the conductance correction of the form that we have obtained
2503: (decreasing at high temperatures or short lengths)
2504: is a nontrivial effect of the electron interactions, and our simple model has
2505: already captured this qualitatively. A non-interacting theory does not have
2506: temperature or length dependences of this kind.
2507:
2508: We now discuss the important experimental finding of Liang {\it et al}
2509: \cite{liang} of the odd-even effect in the transport of electrons through
2510: a quantum wire in the presence of a magnetic field.
2511: Liang {\it et al} find that as they turn up the external magnetic field (kept
2512: in plane and aligned along the direction of the channel) from $0$ to $11T$,
2513: the increasing magnetic field expectedly lifts the spin degeneracy and splits
2514: each conductance step into two steps, with the heights of both being less than
2515: $g_0$. Furthermore, at a magnetic field strength of $11T$, they find that the
2516: difference between the conductance of successive pairs of spin-split sub-bands
2517: alternates. This shows that the conductance of the odd numbered
2518: spin-split sub-bands containing the moments aligned with the magnetic field
2519: undergoes little renormalization (i.e., is close to $g_0$ in their Fig. 4),
2520: while the conductance of the even numbered spin-split sub-bands containing
2521: the moments anti-aligned with the magnetic field undergoes a large
2522: renormalization correction; their Fig. 4 indicates a correction as large
2523: as $0.3 g_0$. As discussed earlier, this phenomenon can be simply
2524: understood as the aligned moments seeing a much weaker barrier and the
2525: anti-aligned moments seeing a much stronger barrier. This is due to the Zeeman
2526: splitting of the Fermi levels of the up and down spin electrons
2527: and their interactions with each other.
2528: Since the difference in renormalizations between the aligned and
2529: anti-aligned electrons occur for all magnetic fields (i.e., even when the
2530: up and down sub-bands are not spin-split), we suggest the following
2531: possibility. One can artificially enhance the barrier
2532: strengths so that the difference in renormalizations of the up
2533: and down spins can be made substantial at moderate magnetic fields. More
2534: importantly, we can vary the gate voltage so as to tune the
2535: spin polarization with greater transmission to
2536: resonance. This would mean that at these values of the magnetic field
2537: and gate voltage, transmission of one of the polarizations is completely
2538: suppressed and the other one greatly enhanced. This leads us to
2539: the possibility of creating a spin-valve at moderate magnetic fields.
2540:
2541: Finally, we comment on a new set of experiments \cite{topinka,crook} which
2542: have used scanning probe microscopy techniques to study transport through
2543: QPCs and propose a test for our model based on such a study. In these
2544: experiments, a negatively charged atomic force microscope tip is held
2545: at a distance of $100-150nm$ above the 2DEG gas on which the QPC is created
2546: via split-gate methods. A capacitive coupling between the 2DEG and the
2547: tip reduces the density of the 2DEG in a small spot directly beneath the tip,
2548: thereby creating a small depletion region (negatively charged ``bubble") which
2549: can back-scatter electrons approaching it. The tip then scans the surface
2550: of the 2DEG reservoir into which the electrons are entering after traveling
2551: through the QPC, and the two-probe conductance is measured. This allows one
2552: to ``image" the electron current flowing out from the QPC. Topinka {\it et al}
2553: \cite{topinka} have made such measurements at a temperature of $T = 1.7K$ and
2554: find that the electrons flow out into the 2DEG reservoir in streaks from each
2555: sub-band. The number and nature of the streaks is governed by the
2556: electron wave function in each sub-band caused by the quantization due to the
2557: confinement in the transverse direction. They find that the electron flow is
2558: coherent along
2559: these streaks quite far from the QPC mouth where they finally disperse into
2560: the 2DEG. Furthermore, they find that placing the depletion bubble in the
2561: path of a particular streak (at a distance of about $0.3 - 0.5\mu m$ from
2562: the mouth of the QPC) gives rise to a flat, renormalized plateau only for the
2563: particular sub-band from which it is emanating, while the other sub-bands give
2564: the universal conductance value of $2g_0$. This tells us that the effect
2565: of the gate voltage must vanish quickly since it is not felt beyond distances
2566: as short as $0.3 - 0.5\mu m$ from the mouth of the QPC. Crook {\it et al}
2567: \cite{crook} find a series of peaks and troughs upon measuring the
2568: differential conductance $dg/dV_G$ versus the gate voltage $V_G$ (which are
2569: caused by the step rises and plateaus for each sub-band respectively) while
2570: scanning the tip through the QPC. Their finding that the troughs do not fall
2571: to zero indicate that the conductance corrections caused by the depletion
2572: bubble (when placed within the QPC) is gate
2573: voltage dependent as would have been expected.
2574:
2575: Now, the availability of the tip generated depletion bubble as a controlled
2576: barrier to the flow of electrons through the QPC also
2577: suggests a possible use of scanning probe microscopy
2578: techniques to test the predictions of
2579: our model in a quantitative fashion. This would require the gate voltage to be
2580: first fixed such that only the lowest sub-band is fully open to the flow
2581: of electron current, and then the depletion bubble to be placed somewhere on a
2582: streak emanating from this lowest sub-band at a distance from the QPC mouth;
2583: the conductance can then be measured by changing the gate
2584: voltage but holding the temperature fixed. The nature of the
2585: conductance versus gate voltage curve will tell us whether the gate voltage
2586: does or does not have any effect on the electrons on the streak at that
2587: distance from the mouth of the QPC. Furthermore, the gate voltage can then be
2588: held fixed somewhere on a plateau and the conductance measured as the
2589: temperature is varied. The form of the conductance corrections versus
2590: temperature can then be obtained. This entire chain of measurements can then
2591: be repeated after taking the depletion bubble closer to the QPC mouth and
2592: into the QPC in a series of steps. Such a series of measurements would
2593: help answer questions about where the conductance corrections start becoming
2594: dependent on the gate voltage as well as
2595: how the conductance corrections vary with temperature when a barrier is
2596: placed within the QPC or away from the QPC. Such
2597: experiments could also be carried out with longer QWs to
2598: check the length dependences of the conductance corrections.
2599:
2600: \section{\bf Summary and Outlook}
2601:
2602: The main idea in this paper is to introduce a model which explicitly
2603: describes the regions in between the quantum wire and the 2DEG
2604: reservoirs as interacting 1D systems which
2605: are independent of the density of electrons in the quantum wire.
2606: We show that the difference in the strengths of the interactions in
2607: the different regions leads to local junction barriers between the regions;
2608: the barriers simulate the effects of the imperfect coupling between
2609: the 2DEG and the quantum wire. Our model leads to the following results for
2610: wires with no impurities, all of which are in agreement with a large body of
2611: experimental observations.
2612:
2613: \begin{itemize}
2614:
2615: \item{} Flat (independent of gate voltage) and uniform (for all the sub-bands)
2616: renormalizations of the quantized conductance plateaus.
2617:
2618: \item{} The renormalizations increase as the temperature is
2619: lowered or the length of the quantum wire is increased.
2620:
2621: \item{} At still lower temperatures, the flatness of the plateaus
2622: disappears and oscillatory features in the conductance can be observed
2623: which we interpret as resonant transmission through the quantum wire.
2624:
2625: \item{} In the presence of a magnetic field, an odd-even effect
2626: is found in the conductance of alternate spin-split sub-bands. This effect
2627: may be used to construct a spin-valve, which allows only electrons with one
2628: particular spin to transmit through the wire even if the magnetic field is
2629: not high enough to completely spin-split the sub-bands.
2630:
2631: \end{itemize}
2632:
2633: For quantum wires with impurities, which are either intrinsic or
2634: externally imposed as finger gates, the conductance corrections
2635: are always gate voltage dependent and therefore, are neither flat
2636: nor sub-band independent.
2637:
2638: Some interesting questions for future studies include the following.
2639: A quantitative fit to the conductance corrections as a function
2640: of the temperature and wire length still remains to be done. This
2641: would require an even more realistic modeling of the quantum wire
2642: system (including some of the features itemized in the
2643: previous section) as well as more experimental data.
2644: Theoretical studies at finite frequencies and finite external voltages
2645: across the quantum wire also need to be pursued.
2646: Finally, one needs to understand several features which are observed
2647: on the rise between two successive plateaus, such as the ``$0.7$ effect''
2648: mentioned in the introduction, the observation of continuous oscillations as
2649: a function of the gate voltage upon introducing finger gate barriers
2650: \cite{tkachenko}, and the fixed point that exists on the rise as the
2651: temperature is varied \cite{senz}. For all of these, one needs to study the
2652: model when some sub-band is partially opened.
2653:
2654: \vskip .5 true cm
2655: \leftline{\bf Acknowledgments}
2656: \vskip .5 true cm
2657:
2658: SL thanks I. Safi for useful correspondence. DS thanks the Council of
2659: Scientific and Industrial Research, India for financial support through
2660: grant No. 03(0911)/00/EMR-II.
2661:
2662: \vskip .5 true cm
2663:
2664: \appendix
2665:
2666: \section{\bf Effective action for spinless fermions}
2667:
2668: In this Appendix, we will obtain explicitly the $S_0$ part of the effective
2669: action in terms of the fields $\phi_i,i=1,..4$, at the junctions $x=0,d,l+d$
2670: and $x=l+2d=L$ for the $K_L$-$K_C$-$K_W$-$K_C$-$K_L$ model described by the
2671: action in Eq. (\ref{s0}) by integrating out all degrees of freedom except at
2672: the positions of the junctions. We will also give the effective action of the
2673: simpler model $K_L$-$K_W$-$K_L$ for comparison, since they have also
2674: not been explicitly given anywhere.
2675:
2676: We first start with the simpler $K_L$-$K_W$-$K_C$ model, which is
2677: defined as a length $L$ quantum wire with interaction parameter $K_W$
2678: between $x=0$ and $x=L$, and with leads defined by $K_L=1$ for
2679: $x<0$ and $x>L$, described by the Lagrangian
2680: \beq
2681: {\cal L}= \int_{-\infty}^0 dx {\cal L}_1 + \int_{0}^{L}
2682: dx {\cal L}_2 + \int_L^\infty dx {\cal L}_1 ,
2683: \eeq
2684: where
2685: \bea
2686: {\cal L}_1 ~&=&~ {1\over 2K_Lv_L} (\pt\phi)^2 +{v_L\over 2K_L}
2687: (\px\phi)^2, \nonum \\
2688: {\rm and } \quad
2689: {\cal L}_2 ~&=&~ {1\over 2K_Wv_W} (\pt\phi)^2 +{v_W\over 2K_W} (\px \phi)^2.
2690: \eea
2691: to set the notation. There
2692: are three ways to derive the effective action. We can (a) integrate out
2693: the fields at all points in space except at $x=0$ and $L$, or (b) find
2694: the solution of the equations of motion in terms of the above two fields
2695: and then compute the action from that solution, or (c) compute the
2696: Green's function $G_\bo (x,x^\prime )$, set $x,x^\prime$ equal
2697: to $0$ or $L$, and invert $G$ to get $S_{eff}$. All the methods
2698: produce the same result since the original action is purely quadratic.
2699: We will use the second method here because it is technically simpler.
2700:
2701: As in other sections, we will work with the
2702: Euclidean time action for convenience. If all the
2703: fields have a time dependence of the form $\exp (-i\on \tau)$, then
2704: normalizability of the solutions imply that they should decay exponentially
2705: at $x \rightarrow \pm \infty$. We assume that the solution of the equation
2706: of motion has the following forms in the three regions,
2707: \bea
2708: \tphi(x,\on) &=& \tphi(0,\on)e^{-i\on\tau +|\on|x/v_L}, \quad x<0 \nonum \\
2709: &=& e^{-i\on\tau} (\ttheta_1(\on) e^{|\on| x/v_W} +\ttheta_2(\on)
2710: e^{-\on x/v_W}), \quad 0<x<L \nonum \\
2711: &=& \tphi(L,\on)e^{-i\on\tau+|\on|(L-x)/v_L}, \quad x>L ~.
2712: \eea
2713: Matching solutions at the boundaries $x=0$ and $x=L$ to eliminate $\ttheta_i
2714: (\on)$, and using this solution in the effective action and carrying out the
2715: spatial integration, we obtain the action
2716: \bea
2717: S_0 &=& {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_2^2)
2718: + {1\over 2K_W} \sum_{\on} {|\on| \over e^{k_{nW}L} - e^{-k_{nW}L} }
2719: \times \nonum \\
2720: && \quad \quad \quad [(e^{k_{nW}L}+e^{-k_{nW}L} )(\tphi_1^2+\tphi_2^2) -
2721: 4\tphi_1\tphi_2] ,
2722: \eea
2723: where $\tphi_1 \equiv \tphi(0,\on)$ and $\tphi_2 \equiv \tphi(L,\on)$
2724: and $k_{nW}$ and $k_{nL}$ are defined as $|\on|/v_W$ and $|\on|/v_L$
2725: respectively.
2726: In the limit $\on \gg v_L/L, v_W/L$, we get the high frequency effective action
2727: \beq
2728: S_{high} = {K_L+K_W \over 2K_LK_W} \sum_{\on}|\on|
2729: (\tphi_1^2 +\tphi_2^2) ~,
2730: \eeq
2731: where the two junctions are decoupled as expected. In the low
2732: frequency limit $\on \ll v_L/L, v_W/L$, we get
2733: \beq
2734: S_{low} = {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+
2735: \tphi_2^2) + { U_W\over 2} \sum_{\on} |\on| (\tphi_1-\tphi_2)^2 ~,
2736: \eeq
2737: where $U_W = v_W/K_W L$.
2738:
2739: Using the same method as above, we can also obtain the full effective
2740: action for the $K_L$-$K_C$-$K_W$-$K_C$-$K_L$ model in terms of the fields at
2741: the four junctions $\phi_i, i=1,..4$, where $\tphi(0,\on) = \tphi_1(\on)
2742: \equiv \tphi_1$, $\tphi(d,\on) \equiv \tphi_2$, $\tphi(l+d,\on)
2743: \equiv \tphi_3$ and $\tphi(L=l+2d,\on) \equiv
2744: \tphi_4$. The solutions in the five regions can be written as
2745: \bea
2746: \tphi(x, \on) &=& \tphi_1 e^{k_{nL}x}, \quad x<0 \nonum \\
2747: &=& B e^{k_{nC}x} + Ce^{-k_{nC}x}, \quad 0<x<d \nonum \\
2748: &=& D e^{k_{nW}x} + Ee^{-k_{nW}x}, \quad d<x<l+d \nonum \\
2749: &=& F e^{k_{nC}x} + Ge^{-k_{nC}x}, \quad l+d<x<L \nonum \\
2750: &=& \tphi_4 e^{k_{nL}(L-x)}, \quad x<0 ~.
2751: \eea
2752: where by matching the solutions at $x=0,d,l+d$ and $L$, we can
2753: obtain the functions of $\on$, $B,C,D,E,F$ and $G$ in terms of the
2754: $\tphi_i, i=1...4$, and $k_{nC}$ is defined as $k_{nC} = |\on| /v_C$.
2755: Substituting this solution in the
2756: action and integrating over all space, we get the effective action
2757: \bea
2758: S_{eff} &=& \sum_{\on}\{{1\over 2K_L} |\on| (\tphi_1^2+\tphi_4^2) \nonum \\
2759: && +{1\over 2K_C} |\on| [(B^2+G^2e^{-2k_n L}) (e^{2k_n d} - 1) + (C^2+F^2
2760: e^{2k_nL})(1-e^{-2k_n d})] \nonum \\
2761: && +{1\over 2K_W} [|\on| [D^2(e^{2k_n (L-d)} - e^{2k_n d}) -
2762: E^2(e^{-2k_n (L-d)} - e^{-2k_n d})] \}~,
2763: \eea
2764: with
2765: \bea
2766: B={\tphi_2 -\tphi_1e^{-k_{nC}d}\over e^{k_{nC} d} -e^{-k_{nC}d}}, \quad && C=
2767: {\tphi_1e^{k_{nC}d}-\tphi_2\over e^{k_{nC} d} -e^{-k_{nC}d}}\nonum \\
2768: D={\tphi_3 -\tphi_2e^{-k_{nW}l}\over e^{k_{nC}d} (e^{k_{nW}l} -e^{-k_{nW}l})},
2769: \quad && E= {\tphi_2e^{k_{nW}l}-\tphi_2\over e^{-k_{nC}d}
2770: (e^{k_{nW} l} -e^{-k_{nW}l})}\nonum \\
2771: F={\tphi_4 -\tphi_3e^{-k_{nC}d}\over e^{k_{nC}d +k_{nW}l}
2772: (e^{k_{nC} d} -e^{-k_{nC}d})},
2773: \quad && G= {\tphi_3e^{k_{nC}d}-\tphi_4\over e^{-k_{nC}d -k_{nW}l}
2774: (e^{k_{nC} d} -e^{-k_{nC}d})} ~.
2775: \eea
2776: The action finally simplifies to
2777: \bea
2778: S_{eff} &=& {1\over 2K_L} \sum_{\on} |\on| (\tphi_1^2+\tphi_4^2) \nonum \\
2779: &&\hspace*{-0.2cm} +{1\over 2K_C} \sum_{\on} {|\on| \over e^{k_{nC}d}
2780: - e^{-k_{nC}d}} [(e^{k_{nC}d} +e^{-k_{nC}d}) (\tphi_1^2 +
2781: \tphi_2^2 + \tphi_3^2+\tphi_4^2) -4\tphi_1\tphi_2 -4\tphi_3\tphi_4] \nonum \\
2782: &&\hspace*{-0.2cm} +{1\over 2K_W} \sum_{\on} {|\on| \over e^{k_{nW}l}
2783: - e^{-k_{nW}l}} [(e^{k_{nW}l} +e^{-k_{nW}l}) (\tphi_2^2+\tphi_3^2)
2784: -4\tphi_2\tphi_3] ~,
2785: \eea
2786: where $\tphi_1 \equiv \tphi(0,\on)$ and $\tphi_2 \equiv \tphi(L,\on)$.
2787: The high and low frequency limits of this effective
2788: action have been used in Sec. 3, to compute the finite temperature and
2789: finite length corrections off-resonance.
2790:
2791: \section{\bf Effective action for spinful fermions}
2792:
2793: In this section, we will explicitly compute the effective action for
2794: spinful fermions in the $K_L$-$K_C$-$K_W$-$K_C$-$K_L$ model. Although,
2795: the method followed is exactly the same as that in the previous
2796: section for spinless fermions, we do it explicitly because there are
2797: a few points where the inclusion of spin makes a difference.
2798:
2799: The effective action for spinful fermions is normally computed in
2800: terms of the `charge' and `spin' field variables defined as $\phi_\rho=
2801: (\phi_\ua +\phi_\da)/\sqrt{2}$ and $\phi_\sigma =(\phi_\ua
2802: -\phi_\da)/\sqrt{2}$ because in the presence of interactions, the spin $\ua$
2803: and $\da$ fermions are mixed (remember the Hubbard term ${\cal U} \sum_i
2804: n_{i\ua} n_{i\da}$). Here, in our model with contacts, the interaction
2805: term $\cal U$ is different in the contact region and in the wire
2806: region. But since the linear combination that diagonalizes the interaction
2807: is independent of the value of $\cal U$, the action in terms of the
2808: $\phi_\rho$ and $\phi_\sigma$ fields are decoupled. In the
2809: presence of a magnetic
2810: field, in the next Appendix, we will see that the action continues to be
2811: diagonalizable; however, the diagonal fields are defined in terms of
2812: mixing angles which explicitly depend on $\cal U$ and the magnetic
2813: field and hence are different in the leads, the contacts and the wire.
2814:
2815: The starting action for the spinful fermions is given in
2816: Eqs. (\ref{s0}) and (\ref{sol}) in the text in terms of the charge and
2817: spin fields. As in the earlier Appendix, we will obtain the solution of the
2818: equations of motion in terms of the eight fields $\tphi_{ia}, i=1...4,
2819: a=\rho , \sigma$ defined to be at the positions $x=0,d,l+d$ and $L=l+2d$ and
2820: then compute the effective action from that solution. We
2821: assume that the solutions in the five regions can be written as
2822: \bea
2823: \tphi_{a}(x, \on) &=& \tphi_{1a} e^{k_{nLa}x}, \quad x<0 \nonum \\
2824: &=& B_a e^{k_{nCa}x} + C_ae^{-k_{nCa}x}, \quad 0<x<d \nonum \\
2825: &=& D_a e^{k_{nWa}x} + E_ae^{-k_{nWa}x}, \quad d<x<l+d \nonum \\
2826: &=& F_a e^{k_{nCa}x} + G_ae^{-k_{nCa}x}, \quad l+d<x<L \nonum \\
2827: &=& \tphi_{4a} e^{k_{nL}(L-x)}, \quad x<0 ~.
2828: \eea
2829: and as before, the coefficients, $B_a,C_a,D_a,E_a,F_a$ and $G_a$ can
2830: be found in terms of the $\tphi_{ia}, i=1...4, a=\rho , \sigma$
2831: by matching the solutions at $x=0,d,l+d$ and $L$. Note that
2832: $k_{nWa}$, $k_{nCa}$ and $k_{nL}$ are defined as $|\on|/v_{Wa}$,
2833: $|\on|/v_{Ca}$ and $|\on|/v_{La}$ respectively. Substituting this solution
2834: in the action and integrating over all space, we get the effective action as
2835: \bea
2836: S &=& \sum_{a=\rho , \sigma}\{{1\over 2K_{La}}
2837: \sum_{\on} |\on| (\tphi_{1a}^2+\tphi_{4a}^2) + {1\over 2K_{Ca}}
2838: \sum_{\on} {|\on| \over e^{k_{nCa}d} - e^{-k_{nCa}d}} \times\nonum \\
2839: && \quad \quad \quad [(e^{k_{nCa}d} +e^{-k_{nCa}d}) (\tphi_{1a}^2+
2840: \tphi_{2a}^2 + \tphi_{3a}^2+\tphi_{4a}^2) -4\tphi_{1a}\tphi_{2a}
2841: -4\tphi_{3a}\tphi_{4a}] \nonum \\
2842: &+& {1\over 2K_{Wa}} \sum_{\on} {|\on| \over e^{k_{nWa}l} - e^{-k_{nWa}l}}
2843: [(e^{k_{nWa}l} +e^{-k_{nWa}l}) (\tphi_{2a}^2+\tphi_{3a}^2)
2844: -4\tphi_{2a} \tphi_{3a}]\}.
2845: \label{spineff}
2846: \eea
2847: The high and low frequency limits of this effective
2848: action have been used in Sec. 3, to discuss the various resonances that
2849: are possible in the low temperature limit and to explicitly
2850: compute the off-resonance corrections to the conductances at
2851: finite temperatures and for finite length wires.
2852:
2853: \section{\bf Effective action for spinful electrons in the
2854: presence of a magnetic field}
2855:
2856: We present here the calculation for the effective action for
2857: spinful fermions in a magnetic field
2858: when the mixing angle $\gamma$ is the same in both the contacts as well as
2859: the QW, i.e., the short ranged electron-electron interaction $\cal U$ is equal
2860: in all the three TLLs. We start with the action given in Eq. (\ref{actmag})
2861: in Sec. 5 and will now integrate out the fields at all points except at the
2862: four junctions as these will be
2863: the sites for the two outer barriers while the two inner junctions are the
2864: ends of the region to which the gate voltage couples. Thus, we write down
2865: the equations of motion in each of the five regions and solve them. If all the
2866: fields have a time dependence of the form $\exp (-i\on \tau)$, then
2867: normalizability of the solutions imply that they should decay exponentially
2868: at $x \rightarrow \pm \infty$. The general solution is given by
2869: \bea
2870: \tphi_\ua(x) &=& A_\ua e^{k_\ua x}, \quad \tphi_\da(x) = A_\da e^{k_\da x},
2871: \quad x<0 \nonum \\
2872: \tphi_\pm(x) &=& B_{\pm}e^{k_\pm x}+C_{\pm}e^{-k_\pm x}, \quad 0<x<d \nonum \\
2873: \tphi_\pm(x) &=& D_{\pm}e^{\tk_\pm x}+E_{\pm}e^{-\tk_\pm x}, \quad d<x<l+d
2874: \nonum \\
2875: \tphi_\pm(x) &=& F_{\pm}e^{k_\pm x}+G_{\pm}e^{-k_\pm x}, \quad l+d<x<l+2d
2876: \nonum \\
2877: \tphi_\ua(x) &=& H_\ua e^{k_\ua x}, \quad \tphi_\da = H_\da e^{k_\da x}, \quad
2878: x<0 ~,
2879: \label{coeff}
2880: \eea
2881: where we have defined $\tk_\pm = |\on| /v_{W\pm}$, $k_\pm = |\on| /
2882: v_{C\pm}$, $k_\ua = |\on|/v_{F\ua}$ and $k_\da = |\on|/v_{F\da}$.
2883:
2884: We now solve for the coefficients $A,...,H$ in Eq. (\ref{coeff})
2885: by matching the fields
2886: $\tphi_\ua$ and the $\tphi_\da$ at the four junctions. At this point, we make
2887: the simplifying assumption that the mixing angles $\gamma_C$ and $\gamma_W$
2888: (defined as in Eqs. (\ref{transf1}) and (\ref{transf2})) in the contact and
2889: wire regions are equal to each other, $\gamma_C = \gamma_W = \gamma$.
2890: This implies that $\tphi_{W\pm} = \sqrt{{K_{W\pm} v_{W\pm}} \over {K_{C\pm}
2891: v_{C\pm}}} \tphi_{C\pm}$ at $x=d$ and $x=l+d$.
2892: We find that the coefficients are given by
2893: \bea
2894: A_\ua &=& \tphi_{1\ua}, \quad A_\da = \tphi_{1\da} \nonum \\
2895: B_{\pm} &=& {\tphi_{2\pm} - \tphi_{1\pm}
2896: e^{-k_\pm d}\over D_{C\pm}} \nonum \\
2897: C_{\pm} &=&- {\tphi_{2\pm} + \tphi_{1\pm}
2898: e^{k_\pm d}\over D_{C\pm}} \nonum \\
2899: D_{\pm} &=& \sqrt{{K_{W\pm} v_{W\pm}} \over {K_{C\pm} v_{C\pm}}}
2900: e^{-\tk_\pm d} ~{(-\tphi_{2\pm}
2901: e^{-\tk_\pm l} + \tphi_{3\pm}) \over D_{W\pm}}
2902: \nonum \\
2903: E_{\pm} &=& \sqrt{{K_{W\pm} v_{W\pm}} \over {K_{C\pm} v_{C\pm}}}
2904: e^{\tk_\pm d}~{(\tphi_{2\pm}
2905: e^{\tk_\pm l} - \tphi_{3\pm}) \over D_{W\pm}}
2906: \nonum \\
2907: F_{\pm} &=& -{\tphi_{3\pm}e^{-k_\pm(l+2d)} - \tphi_{4\pm}
2908: e^{-k_\pm d}\over D_{C\pm}} \nonum \\
2909: G_{\pm} &=& {\tphi_{3\pm}e^{k_\pm(l+2d)} - \tphi_{4\pm}
2910: e^{k_\pm d}\over D_{C\pm}} \nonum \\
2911: H_{\ua} &=& \tphi_{4\ua}e^{k_\ua(l+2d)}, \quad H_\da = \tphi_{4\da}
2912: e^{k_\da(l+2d)} ~,
2913: \eea
2914: where
2915: \bea
2916: D_{C\pm} &=& e^{k_\pm d} - e^{-k_\pm d} ~ \nonum \\
2917: {\rm and}\quad D_{W\pm} &=& e^{\tk_\pm l} - e^{-\tk_\pm l} ~.
2918: \eea
2919:
2920: Then using the relations written down in Eq. (\ref{transf1})
2921: connecting the $\phi_{\pm}$ and $\phi_{\ua,\da}$ fields,
2922: we find that the effective Lagrangian density is given by
2923: \bea
2924: {\cal L} = && \frac{|\on|}{2K_L}(\tphi_{1\ua}^2 + \tphi_{1\da}^2
2925: \tphi_{4\ua}^2 + \tphi_{4\da}^2) \nonum \\
2926: && + \frac{|\on|}{2}[(p^2\frac{N_{C+}} {K_{C+}D_{C+}} + r^2\frac{N_{C-}}
2927: {K_{C-}D_{C-}})(\tphi_{1\ua}^2 + \tphi_{2\ua}^2
2928: + \tphi_{3\ua}^2 + \tphi_{4\ua}^2) \nonum \\
2929: && + (q^2\frac{N_{C+}} {K_{C+}D_{C+}} + s^2\frac{N_{C-}}
2930: {K_{C-}D_{C-}})(\tphi_{1\da}^2
2931: + \tphi_{2\da}^2 + \tphi_{3\da}^2 + \tphi_{4\da}^2) \nonum \\
2932: && - 4(\frac{p^2}{K_{C+}D_{C+}} +
2933: \frac{r^2}{K_{C-}D_{C-}})(\tphi_{1\ua}\tphi_{2\ua}
2934: + \tphi_{3\ua}\tphi_{4\ua}) \nonum \\
2935: && - 4(\frac{q^2}{K_{C+}D_{C+}} + \frac{s^2}{K_{C-}D_{C-}})
2936: (\tphi_{1\da} \tphi_{2\da} + \tphi_{3\da}\tphi_{4\da})] \nonum \\
2937: && + |\on| [(\frac{pq N_{C+}}{K_{C+} D_{C+}}
2938: + \frac{rs N_{C-}}{K_{C-} D_{C-}})
2939: (\tphi_{1\ua}\tphi_{1\da} + \tphi_{2\ua}\tphi_{2\da}
2940: + \tphi_{3\ua}\tphi_{3\da} + \tphi_{4\ua}\tphi_{4\da}) \nonum \\
2941: && - 2(\frac{pq}{K_{C+}D_{C+}} + \frac{rs}
2942: {K_{C-}D_{C-}})(\tphi_{1\ua}\tphi_{2\da} +
2943: \tphi_{1\da}\tphi_{2\ua} + \tphi_{3\ua}\tphi_{4\da}
2944: + \tphi_{3\da}\tphi_{4\ua})] \nonum \\
2945: && + \frac{|\on|}{2}[(p^2\frac{N_{W+}} {K_{W+}D_{W+}} + r^2\frac{N_{W-}}
2946: {K_{W-}D_{W-}})(\tphi_{2\ua}^2 + \tphi_{3\ua}^2) \nonum \\
2947: && + (q^2\frac{N_{W+}} {K_{W+}D_{W+}} + s^2\frac{N_{W-}}
2948: {K_{W-}D_{W-}})(\tphi_{2\da}^2 + \tphi_{3\da}^2) \nonum \\
2949: && - 4(\frac{p^2}{K_{W+}D_{W+}} +
2950: \frac{r^2}{K_{W-}D_{W-}})\tphi_{2\ua}\tphi_{3\ua}
2951: \nonum \\
2952: && - 4(\frac{q^2}{K_{W+}D_{W+}} +
2953: \frac{s^2}{K_{W-}D_{W-}})\tphi_{2\da}\tphi_{3\da}] \nonum \\
2954: && + |\on| [(\frac{pq N_{W+}}{K_{W+} D_{W+}} + \frac{rs N_{W-}}{K_{W-}D_{W-}})
2955: (\tphi_{2\ua}\tphi_{2\da} + \tphi_{3\ua}\tphi_{3\da}) \nonum \\
2956: && - 2(\frac{pq}{K_{W+}D_{W+}} + \frac{rs}
2957: {K_{W-}D_{W-}})(\tphi_{2\ua}\tphi_{3\da} + \tphi_{2\da}\tphi_{3\ua})],
2958: \eea
2959: where
2960: \bea
2961: N_{C\pm} &=& e^{k_\pm d} + e^{-k_\pm d} ~, \nonum \\
2962: N_{W\pm} &=& e^{\tk_\pm l} + e^{-\tk_\pm l} ~,
2963: \eea
2964: and $D_{C\pm}, D_{W\pm}, k_{\pm}$ and $\tk_{\pm}$ have already been
2965: defined above.
2966:
2967: \section{\bf Calculation of the Green's function in our model for the
2968: Quantum Wire}
2969:
2970: Here, we will present a calculation of the Green's function for the bosonic
2971: excitations in the model that we have presented for the quantum wire system
2972: of spinless fermions. The method we follow is along the lines of the
2973: calculation presented by Maslov and Stone \cite{maslov}. We will study
2974: the case when there are no barriers present anywhere in the system. Then,
2975: we see that the Euclidean action $S_E$ in all the five distinct TLL regions
2976: in our model (Fermi lead, contact, QW, contact and Fermi lead) is given by
2977: \beq
2978: S_E = \frac{1}{2} \int_{0}^{\beta} d\tau \int_{-\infty}^{\infty}
2979: [\frac{1}{K(x)v(x)}(\partial_{\tau}\phi)^2 +
2980: \frac{v(x)}{K(x)}(\partial_{x}\phi)^2],
2981: \eeq
2982: with $K(x) = K_L, v(x) = v_L$ in the first and fifth (Fermi lead) regions,
2983: $K(x) = K_C, v(x) = v_C$ in the second and fourth (contact) regions and
2984: $K(x) = K_W, v(x) = v_W$ in the third (QW) region. Then, defining the
2985: two-point bosonic Green's function/propagator (in Euclidean time $\tau$) as
2986: \beq
2987: G(x,x',\tau) = <T_{\tau}\phi(x,\tau)\phi(x',0)>,
2988: \eeq
2989: it can be shown that the equation satisfied by the Fourier transform of
2990: the above Green's function $G_{\bar{\omega}}(x,x')$ is
2991: \beq
2992: \{-\partial_{x}(\frac{v(x)}{K(x)}\partial_{x}) +
2993: \frac{1}{K(x)v(x)}\bar{\omega}^2\} G_{\bar{\omega}}(x,x') = \delta(x-x').
2994: \eeq
2995: We now have to solve the above equation to obtain a functional form for
2996: $G_{\bar{\omega}}(x,x')$. We know that the interaction parameter $K$ and
2997: the velocity $v$ change abruptly at each of the junctions and that the
2998: two Fermi leads are semi-infinite in length (i.e. $G_{\bar{\omega}}(x,x')$
2999: must decay to zero as $x\rightarrow \pm\infty$). As we are interested in
3000: finding the one-point Green's function at a point in the left contact, we
3001: will choose $x'$ to lie between $0$ (the left lead-contact junction) and
3002: $d$ (the left contact-QW junction). Furthermore, we know that the
3003: Green's function $G_{\bar{\omega}}(x,x')$ must satisfy the following
3004: boundary conditions: (a) $G_{\bar{\omega}}(x,x')$ must be continuous at
3005: $x = 0, x', d, l+d$ and $l+2d$ (b) $(\frac{v(x)}{K(x)})\partial_{x}
3006: G_{\bar{\omega}}(x,x')$ must be continuous at $x = 0, d, l+d$ and $l+2d$
3007: and
3008: \beq
3009: - \frac{v(x)}{K(x)}\partial_{x}G_{\bar{\omega}}(x,x')
3010: \vert_{x=x'-0}^{x=x'+0} = 1,
3011: \eeq
3012: i.e., $(\frac{v(x)}{K(x)})\partial_{x}G_{\bar{\omega}}(x,x')$ undergoes a
3013: jump of unity at $x=x'$. It is then easily seen that the solution for
3014: $G_{\bar{\omega}}(x,x')$ is of the form
3015: \bea
3016: G_{\bar{\omega}}(x,x') &=& A e^{\vert\bar{\omega}\vert x/v_L}
3017: \quad \quad \quad \quad ~~~~~~~~~\textrm{for $x\leq 0$} \nonum \\
3018: &=& B e^{\vert\bar{\omega}\vert x/v_C} +
3019: C e^{-\vert\bar{\omega}\vert x/v_C} \quad ~\textrm{for $0<x\leq x'$} \nonum \\
3020: &=& D e^{\vert\bar{\omega}\vert x/v_C} +
3021: E e^{-\vert\bar{\omega}\vert x/v_C} \quad ~\textrm{for $x'<x\leq d$} \nonum \\
3022: &=& F e^{\vert\bar{\omega}\vert x/v_C} +
3023: G e^{-\vert\bar{\omega}\vert x/v_C} \quad ~\textrm{for $d<x\leq l+d$} \nonum \\
3024: &=& H e^{\vert\bar{\omega}\vert x/v_C} +
3025: I e^{-\vert\bar{\omega}\vert x/v_C} \quad ~~\textrm{for $l+d<x\leq l+2d$}
3026: \nonum \\
3027: &=& J e^{-\vert\bar{\omega}\vert x/v_C} \quad \quad \quad \quad
3028: ~~~~~~~~\textrm{for $l+2d<x$} ~.
3029: \eea
3030: The coefficients $A, B, \ldots , J$ are found by matching the boundary
3031: conditions. To begin with, it is worth noting that in the dc limit
3032: of $\bar{\omega}\rightarrow 0$, we find that
3033: \beq
3034: A = B + C = D + E = F + G = H + I = J = \frac{K_L}{2\vert\bar{\omega}\vert},
3035: \eeq
3036: which gives the dc conductance to be
3037: \beq
3038: g = \frac{2e^2}{h} K_L = 2g_0 K_L.
3039: \eeq
3040: This gives the perfect quantized conductance observed in several experiments
3041: on transport of electrons through a QPC when we take
3042: the leads to be Fermi liquids with $K_L = 1$.
3043:
3044: We now give the expressions for the Green's functions for the case when
3045: both $x$ and $x'$ are taken equal to $a$ at a point in the left contact:
3046: \bea
3047: &&\hspace{-1cm}G_{\bar{\omega}}(a,a)= \nonum \\
3048: &&\hspace{-1cm}\frac{K_C}{2\vert\bar{\omega}\vert}
3049: \frac{\{(r-p-q-s)e^{\vert\bar{\omega}\vert d/v_C}(1 +
3050: \gamma_1 e^{-\vert\bar{\omega}\vert 2a/v_C}) -
3051: (r+p+q-s)e^{-\vert\bar{\omega}\vert d/v_C}
3052: (e^{\vert\bar{\omega}\vert 2a/v_C} + \gamma_1)\}}
3053: {\{(r-p-q-s)e^{\vert\bar{\omega}\vert d/v_C} +
3054: \gamma_1 (r+p+q-s)e^{-\vert\bar{\omega}\vert d/v_C}\}} , \nonum \\
3055: &&
3056: \eea
3057: where
3058: \bea
3059: p &=& e^{-\vert\bar{\omega}\vert l(\frac{1}{v_C}+\frac{1}{v_W})}
3060: (\gamma_1 e^{-2\vert\bar{\omega}\vert d/v_C} (1+\frac{K_W}{K_C})
3061: + (1-\frac{K_W}{K_C})) \nonum \\
3062: q &=& e^{-\vert\bar{\omega}\vert l(\frac{1}{v_C}-\frac{1}{v_W})}
3063: (\gamma_1 e^{-2\vert\bar{\omega}\vert d/v_C} (1-\frac{K_W}{K_C})
3064: + (1+\frac{K_W}{K_C})) \nonum \\
3065: r &=& e^{-\vert\bar{\omega}\vert l(\frac{1}{v_C}+\frac{1}{v_W})}
3066: (\gamma_1 e^{-2\vert\bar{\omega}\vert d/v_C} (1+\frac{K_C}{K_W})
3067: + (\frac{K_C}{K_W}-1)) \nonum \\
3068: s &=& e^{-\vert\bar{\omega}\vert l(\frac{1}{v_C}-\frac{1}{v_W})}
3069: (\gamma_1 e^{-2\vert\bar{\omega}\vert d/v_C} (\frac{K_C}{K_W}-1)
3070: + (1+\frac{K_C}{K_W})) \nonum \\
3071: \gamma_1 &=& \frac{K_L - K_C}{K_L + K_C}.
3072: \eea
3073: The results upon taking the limits corresponding to the various
3074: frequency (or temperature) regimes are given in the section
3075: where the conductance is computed for quantum wires and quantum
3076: point contacts with a junction barrier in the left contact region and we
3077: will not repeat them here.
3078:
3079: We also give the general form of the two-point propagator
3080: $G_{\bar{\omega}}(x,y)$ for when $x$ is a point in the right lead
3081: and $y$ is a point in the left contact:
3082: \beq
3083: G_{\bar{\omega}}(x,y) = \frac{2K_2}{\vert\bar{\omega}\vert}
3084: \frac{e^{\vert\bar{\omega}\vert (l+2d)(\frac{1}{v_L}-\frac{1}{v_C})}
3085: e^{\vert\bar{\omega}\vert \frac{d}{v_C}}
3086: (1+\gamma_1)^{2} e^{\vert\bar{\omega}\vert (\frac{y}{v_C}- \frac{x}{v_L})}}
3087: {\{(p+q+s-r)e^{\vert\bar{\omega}\vert d/v_C} +
3088: \gamma_1 (s-r-p-q)e^{-\vert\bar{\omega}\vert d/v_C}\}} ~,
3089: \eeq
3090: where the expressions for $p, q, r, s$ and $\gamma_1$ have already been
3091: given earlier.
3092:
3093: Now, we give the expression for the one-point propagator at a
3094: point $a$ inside the quantum wire:
3095: \beq
3096: G_{\bar{\omega}}(a,a)= \frac{K_W}{2\vert\bar{\omega}\vert}
3097: \frac{(je^{\vert\bar{\omega}\vert a/v_W} +
3098: ke^{-\vert\bar{\omega}\vert a/v_W}) (me^{\vert\bar{\omega}\vert a/v_W} +
3099: ne^{-\vert\bar{\omega}\vert a/v_W})} {(jn - km )} ,
3100: \eeq
3101: where
3102: \bea
3103: j &=& (1+\frac{K_W}{K_C})e^{\vert\bar{\omega}\vert
3104: d(\frac{1}{v_C}-\frac{1}{v_W})} +
3105: \gamma_1 (1-\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3106: d(\frac{1}{v_C}+\frac{1}{v_W})} \nonum \\
3107: k &=& (1-\frac{K_W}{K_C})e^{\vert\bar{\omega}\vert
3108: d(\frac{1}{v_C}+\frac{1}{v_W})} +
3109: \gamma_1 (1+\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3110: d(\frac{1}{v_C}-\frac{1}{v_W})} \nonum \\
3111: m &=& (1-\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3112: (l+d)(\frac{1}{v_C}+\frac{1}{v_W})} +
3113: \gamma_1 (1+\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3114: (l+d)(\frac{1}{v_C}+\frac{1}{v_W})}e^{-2\vert\bar{\omega}\vert
3115: \frac{d}{v_C}} \nonum \\
3116: n &=& (1+\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3117: (l+d)(\frac{1}{v_C}-\frac{1}{v_W})} +
3118: \gamma_1 (1-\frac{K_W}{K_C})e^{-\vert\bar{\omega}\vert
3119: (l+d)(\frac{1}{v_C}-\frac{1}{v_W})}e^{-2\vert\bar{\omega}\vert \frac{d}{v_C}}.
3120: \eea
3121: Again, we will not give the results of taking the various limits
3122: corresponding to the different frequency (or temperature) regimes
3123: as these have already been quoted in the section on the conductance
3124: of a quantum wire and quantum point contact.
3125:
3126: \newpage
3127:
3128: \begin{thebibliography}{99}
3129:
3130: \bibitem{datta} S. Datta, {\it Electronic transport in mesoscopic systems}
3131: (Cambridge University Press, Cambridge, 1995); {\it Transport phenomenon in
3132: mesoscopic systems}, edited by H. Fukuyama and T. Ando (Springer Verlag,
3133: Berlin, 1992); Y. Imry, {\it Introduction to Mesoscopic Physics} (Oxford
3134: University Press, New York, 1997).
3135:
3136: \bibitem{wees} B. J. van Wees, H. van Houten, C. W. J. Beenakker, J. G.
3137: Williamson, L. P. Kouwenhoven, D. van der Marel and C. T. Foxon, Phys. Rev.
3138: Lett. {\bf 60}, 848 (1988); D. A. Wharam, T. J. Thornton, R. Newbury, M.
3139: Pepper, H. Ahmed, J. E. Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie and
3140: G. A. C. Jones, J. Phys. C {\bf 21}, L209 (1988).
3141:
3142: \bibitem{tarucha} S. Tarucha, T. Honda and T. Saku, Sol. St. Comm. {\bf 94},
3143: 413 (1995).
3144:
3145: \bibitem{yacoby} A. Yacoby, H. L. Stormer, N. S. Wingreen, L. N. Pfeiffer, K.
3146: W. Baldwin and K. W. West, Phys. Rev. Lett. {\bf 77}, 4612 (1996).
3147:
3148: \bibitem{thomas} K. J. Thomas, J. T. Nicholls, M. Y. Simmons, M. Pepper, D. R.
3149: Mace and D. A. Ritchie, Phys. Rev. Lett. {\bf 77}, 135 (1996).
3150:
3151: \bibitem{facer} B. E. Kane, G. R. Facer, A. S. Dzurak, N. E. Lumpkin, R. G.
3152: Clark, L. N. Pfeiffer and K. W. West, App. Phys. Lett. {\bf 72}, 3506 (1998).
3153:
3154: \bibitem{liang} C.-T. Liang, M. Pepper, M. Y. Simmons, C. G. Smith and D. A.
3155: Ritchie, Phys. Rev. B {\bf 61}, 9952 (2000).
3156:
3157: \bibitem{liang2} C.-T. Liang, M. Y. Simmons, C. G. Smith, D. A. Ritchie and
3158: M. Pepper, App. Phys. Lett. {\bf 75}, 2975 (1999).
3159:
3160: \bibitem{picciotto} R. de Picciotto, H. L. Stormer, A. Yacoby, L. N. Pfeiffer,
3161: K. W. Baldwin and K. W. West, Phys. Rev. Lett. {\bf 85}, 1730 (2000).
3162:
3163: \bibitem{topinka} M. A. Topinka, B. J. LeRoy, S. E. J. Shaw, E. J. Heller,
3164: R. M. Westervelt, K. D. Maranowski and A. C. Gossard, Science {bf 289}, 2323
3165: (2000).
3166:
3167: \bibitem{crook} R. Crook, C. G. Smith, M. Y. Simmons and D. A. Ritchie,
3168: cond-mat/9909017.
3169:
3170: \bibitem{reilly} D. J. Reilly, G. R. Facer, A. S. Dzurak, B. E. Kane, R. G.
3171: Clark, P. J. Stiles, J. L. O'Brien, N. E. Lumpkin, L. N. Pfeiffer and K. W.
3172: West, cond-mat/0001174.
3173:
3174: \bibitem{haldane} F. D. M. Haldane, Phys. Rev. Lett. {\bf 47}, 1840 (1981).
3175:
3176: \bibitem{kane} C. L. Kane and M. P. A. Fisher, Phys. Rev. B {\bf 46}, 15233
3177: (1992).
3178:
3179: \bibitem{furusaki1} A. Furusaki and N. Nagaosa, Phys. Rev. B {\bf 47}, 4631
3180: (1993).
3181:
3182: \bibitem{safi} I. Safi and H. J. Schulz, Phys. Rev. B {\bf 52}, 17040 (1995).
3183:
3184: \bibitem{maslov} D. L. Maslov and M. Stone, Phys. Rev. B {\bf 52}, 5539
3185: (1995); V. V. Ponomarenko, {\it ibid.} {\bf 52}, R8666 (1995).
3186:
3187: \bibitem{furusaki} A. Furusaki and N. Nagaosa, Phys. Rev. B {\bf 54}, 5239
3188: (1996).
3189:
3190: \bibitem{alekseev} A. Yu. Alekseev and V. V. Cheianov, Phys. Rev. B {\bf 57},
3191: 6834 (1998).
3192:
3193: \bibitem{sumathi} For a review, see S. Rao and D. Sen, cond-mat/005492.
3194:
3195: \bibitem{buttiker} M. B\"uttiker, Phys. Rev. B {\bf 41}, 7906 (1990).
3196:
3197: \bibitem{fabrizio} M. Fabrizio and A. O. Gogolin, Phys. Rev. B {\bf 51},
3198: 17827 (1995).
3199:
3200: \bibitem{chamon} C. de C. Chamon and E. Fradkin, Phys. Rev. B {\bf 55}, 4534
3201: (1997); N. P. Sandler, C. C. Chamon and E. Fradkin, Phys. Rev. B {\bf 56},
3202: 2012 (1997).
3203:
3204: \bibitem{safilong} I. Safi and H. J. Schulz, Phys. Rev. B {\bf 59}, 3040
3205: (1999); I. Safi, Ph.D. thesis, Laboratoire de Physique des Solides, Orsay
3206: (1996); I. Safi, Ann. Phys. (Paris) {\bf 22}, 463 (1997).
3207:
3208: \bibitem{matveev} K. A. Matveev, Phys. Rev. B {\bf 51}, 1743 (1995).
3209:
3210: \bibitem{yacimry} A. Yacoby and Y. Imry, Phys. Rev. B {\bf 41}, 5341 (1990).
3211:
3212: \bibitem{maslovlong} D. L. Maslov, Phys. Rev. B {\bf 52}, R14386 (1995).
3213:
3214: \bibitem{kimura} T. Kimura, K. Kuroki and H. Aoki, Phys. Rev. B {\bf 53},
3215: 9572 (1996).
3216:
3217: \bibitem{tkachenko} O. A. Tkachenko, V. A. Tkachenko, D. G. Baksheyev, C.-T.
3218: Liang, M. Y. Simmons, C. G. Smith, D. A. Ritchie, G.-H. Kim and M. Pepper,
3219: cond-mat/0003021.
3220:
3221: \bibitem{senz} V. Senz, T. Heinzel, T. Ihn, S. Lindemann, R. Held, K. Ensslin,
3222: W. Wegscheider and M. Bichler, cond-mat/0012205.
3223:
3224: \end{thebibliography}
3225:
3226: \newpage
3227:
3228: \noindent {\bf Figure Captions}
3229: \vskip 1 true cm
3230:
3231: \noindent {1.} Schematic diagram of the model showing the lead regions
3232: (marked FL for Fermi liquid), the contact regions (C) of length $d$, and the
3233: quantum wire (QW) of length $l$. The interaction parameters in these three
3234: regions are denoted by $K_L$, $K_C$ and $K_W$ respectively.
3235:
3236: \noindent {2.} Diagram showing the Fermi energy $E_F$ (with a thermal spread
3237: of $k_B T$) in relation to the sub-bands within the quantum wire. The
3238: conductance will lie on a plateau if the energy levels are as shown in (a),
3239: while the conductance will be at a step between one plateau and the next if
3240: the energy levels are as shown in (b).
3241:
3242: \noindent {3.} A plot showing the curve fitted to the conductance corrections
3243: $\delta g$ versus temperature $T$ data obtained from the inset of Fig. 3 of
3244: Yacoby {\it et al} \cite{yacoby}. The expression for the curve and the value
3245: of the correlation coefficient are given in the text.
3246:
3247: \newpage
3248:
3249: \begin{figure}
3250: \begin{center}
3251: \epsfig{figure=fig1.ps,bbllx=50,bblly=0,bburx=550,bbury=800,height=21cm}
3252: \end{center}
3253: \vspace*{-3 cm}
3254: \centerline{Fig. 1}
3255: \end{figure}
3256:
3257: \begin{figure}
3258: \begin{center}
3259: %\epsfig{figure=fig2.ps,angle=-90,bbllx=50,bblly=0,
3260: %bburx=500,bbury=800,height=15cm}
3261: \epsfig{figure=fig2.ps,angle=-90,width=16cm}
3262: \end{center}
3263: \vspace*{2 cm}
3264: \centerline{Fig. 2}
3265: \end{figure}
3266:
3267: \begin{figure}
3268: \begin{center}
3269: \epsfig{figure=fig3.ps,bbllx=50,bblly=-200,bburx=550,bbury=600,height=21cm}
3270: \end{center}
3271: \vspace*{-5 cm}
3272: \centerline{Fig. 3}
3273: \end{figure}
3274:
3275: \end{document}
3276:
3277: