cond-mat0507097/hf3.tex
1: % re-submitted in 11/05
2: \documentstyle[prl,aps,epsfig,floats]{revtex}
3: 
4: \newcommand{\bib}{\bibitem}
5: \newcommand\bea{\begin{eqnarray}}
6: \newcommand\eea{\end{eqnarray}}
7: \newcommand\beq{\begin{equation}}
8: \newcommand\eeq{\end{equation}}
9: \newcommand\non{\nonumber}
10: \newcommand{\ua}{\uparrow}
11: \newcommand{\da}{\downarrow}
12: 
13: \begin{document}
14: 
15: \draft
16: 
17: \textheight=24cm
18: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
19: 
20: \title{\Large \bf Conductance of quantum wires: a numerical study of the 
21: effects of an impurity and interactions}
22: \author{\bf Amit Agarwal and Diptiman Sen}
23: \address{\it Centre for High Energy Physics, Indian Institute of Science,
24: Bangalore 560012, India}
25: 
26: \date{\today}
27: \maketitle
28: 
29: \begin{abstract}
30: We use the non-equilibrium Green's function formalism and a 
31: self-consistent Hartree-Fock approximation to numerically study the effects 
32: of a single impurity and interactions between the electrons (with and without 
33: spin) on the conductance of a quantum wire. We study how the conductance varies
34: with the wire length, the temperature, and the strengths of the impurity and 
35: interactions. The numerical results for the dependence of the conductance on 
36: the wire length and temperature are compared with the results obtained from a
37: renormalization group analysis based on the Hartree-Fock approximation. For the
38: spin-1/2 model with a repulsive on-site interaction or the spinless model with
39: an attractive nearest neighbor interaction, we find that the conductance 
40: increases with increasing wire length or decreasing temperature. This can be 
41: explained using the Born approximation in scattering theory. For a strong 
42: impurity, the conductance is significantly different for a repulsive and an 
43: attractive impurity; this is due to the existence of a bound state in the 
44: latter case. In general, the large density deviations close to the impurity
45: have an appreciable effect on the conductance at short distances which is 
46: not captured by the renormalization group equations.
47: \end{abstract}
48: \vskip .5 true cm
49: 
50: \pacs{~~ PACS number: ~73.23.-b, ~73.63.Nm, ~71.10.Pm}
51: \vskip.5pc
52: ]
53: 
54: \section{Introduction}
55: 
56: The conductance of electrons in a quantum wire has been the subject of 
57: intensive study in recent years, both experimentally 
58: \cite{tarucha,liang,bkane,yacoby,aus,reilly} and theoretically
59: \cite{kane,furusaki,lal1}. For a wire in which only one channel is 
60: available to the electrons and the transport is 
61: ballistic (i.e., there are no impurities inside the wire, and there is no 
62: scattering from phonons or from the contacts between the wire and its two 
63: leads), the conductance is given by $G = 2e^2 /h$ for infinitesimal bias 
64: \cite{datta,imry}. This result is expected to hold even if one takes into 
65: account the interactions between the electrons since such two-body scatterings
66: conserve the momentum. However, if there is an impurity inside the wire which 
67: scatters the electrons, then the conductance is reduced 
68: because such a scattering does not conserve the momentum of the electron.
69: For a one-dimensional system containing a $\delta$-function impurity
70: with strength $V$, we obtain
71: \beq
72: G ~=~ \frac{2e^2}{h} ~ (1 ~-~ c ~V^2 ~) ~,
73: \eeq
74: to lowest order in $V$, where $c$ is a constant related to the Fermi velocity
75: of the electrons (see the discussion below Eq. (\ref{rt1})). (The situation 
76: is different if the wire is only quasi-one-dimensional, and the impurity 
77: potential has a finite range \cite{chu}). In the absence of interactions, 
78: $G$ does not depend on the wire length $L$ or the 
79: temperature $T$ (as long as $k_B T$ is much less than the Fermi energy). 
80: But in the presence of interactions, it turns out that $V$ effectively 
81: becomes a function of the length scale (which is related to either $L$ 
82: or $T$ as will be explained below), and $G$ therefore varies with $L$ and 
83: $T$. The variation of $V$ with length scale is governed by a renormalization
84: group (RG) equation. (In this paper, we will only consider a non-magnetic
85: impurity which scatters electrons in a spin independent way).
86: 
87: There are three length scales which are of interest in the problem. The 
88: smallest of them is $\lambda \equiv \pi /k_F$ (where $k_F$ is the Fermi 
89: wavenumber); this is the wavelength of the oscillations in the electronic 
90: density near impurities as we will see. The other two length scales are the 
91: length of the wire $L$, and the thermal coherence length $L_T$ which is equal 
92: to $\hbar v_F /(k_B T)$ (where $v_F$ is the Fermi velocity). $L_T$ gives an
93: idea of the distance beyond which an electron wave function loses its
94: phase coherence. At a temperature $T$, the conductance typically receives 
95: contributions from a number of states near the Fermi energy whose energies
96: have a spread of the order of $\Delta E = k_B T$. Hence the spread in 
97: momentum $\Delta p = \Delta E /v_F$ is of the order of $k_B T /v_F$. 
98: A superposition of waves with such a spread of momenta loses phase coherence 
99: in a distance of the order of $L_T =\hbar v_F /(k_B T)$. Electronic transport
100: is therefore thermally incoherent if the wire length $L >> L_T$, coherent if 
101: $L << L_T$, and partially coherent in the intermediate range. This will
102: become clearer when we discuss our numerical results.
103: 
104: The RG equations for the transmission coefficient $|t|^2$ has been derived 
105: from continuum theories in several ways \cite{kane,furusaki,yue,lal2,polyakov}.
106: We will discuss these equations for two models in Sec. II. (The conductance 
107: is related to the transmission as $G = (2e^2 /h) |t|^2$, where the factor of 
108: 2 is due to the electron spin). The RG equation is used to analytically
109: follow the evolution of $|t|^2$ starting from a short distance scale 
110: and going up to a length scale which is the smaller of 
111: the two quantities $L$ and $L_T$. If $L >> L_T$, the conductance is
112: governed by $L_T$ and not $L$, and vice versa if $L_T >> L$. The RG equation 
113: has two fixed points which lie at $|t|^2 =0$ and 1; the system
114: approaches one of these fixed points if $L$ and $L_T$ are both very large.
115: 
116: The above statements are only valid for length scales much longer than 
117: $\lambda$ because only then can one use a continuum description from which 
118: the RG equation is derived. It would therefore be useful to consider an 
119: alternative method for computing the conductance which works even at short 
120: length scales where the continuum description is not valid. An interesting 
121: thing which occurs at short distances is that if the impurity provides an 
122: attractive potential to the electrons, there is a bound state whose
123: wave function decays exponentially away from the impurity. We may wonder what 
124: effect a bound state has on the conductance; the RG equations mentioned above
125: do not take this state into account. One may think that a bound state (whose 
126: energy lies outside the bandwidth of the leads) cannot directly affect the 
127: conductance, because electrons coming in from or going out to the leads 
128: cannot enter or leave such states in the absence of any inelastic scattering.
129: However, such a state contributes to the electronic density near the impurity,
130: and that can affect the transport in the presence of interactions. 
131: 
132: In a series of papers, the technique of functional RG has been used to 
133: numerically study the conductance and other properties of interacting electron
134: systems in one dimension \cite{enss,ander,barnabe,meden}. The authors of those
135: papers have gone up to very large system sizes and have found excellent 
136: agreement at those length scales between their numerical results and the 
137: asymptotic scaling forms given by the RG equations.
138: 
139: In this paper, we will use the non-equilibrium Green's function (NEGF) 
140: formalism to numerically study the conductance of a quantum wire 
141: \cite{datta,meir,datta2,dhar,tsukada}. Since we will use lattice models 
142: and the Hartree-Fock (HF) approximation for dealing with interactions
143: in our numerical studies, we will first discuss those topics in Sec. III.
144: In that section, we will also show how the Born approximation for scattering
145: can qualitatively explain the dependence of the conductance on the wire
146: length which we obtain from the RG equations in Sec. II.
147: 
148: The NEGF formalism will be briefly described in Sec. IV. The advantage of 
149: this method is that it treats the infinitely extended leads (reservoirs) in 
150: an exact way, and it can be used for all values of the wire length and the 
151: impurity strength. However, it is accurate only for weak interactions between 
152: the electrons because we are forced to use a HF approximation for 
153: dealing with the interactions (for reasons which will be explained below). We 
154: will use a lattice model for both the wire and its leads. The 
155: Hamiltonian will have hopping terms in the wire and in the 
156: leads, and a density-density interaction (on-site for spin-1/2 electrons, and
157: between nearest neighbor sites for spinless electrons) only in the wire.
158: 
159: In Sec. V, we will describe our results for spin-1/2 and spinless electrons 
160: for different values of the impurity potential and the interaction strength, 
161: and we will compare our results with those obtained by the RG analysis. For 
162: the case of spinless electrons, we find that the agreement between the RG
163: and numerical results is excellent if we follow a certain procedure. In 
164: Sec. VI, we will make some concluding remarks.
165: 
166: \section{Renormalization group equation for scattering from a point}
167: 
168: There are several ways of studying the renormalization group (RG) evolution 
169: of the scattering from one or more points in a one-dimensional system of 
170: interacting electrons. One can use the technique of bosonization 
171: \cite{kane,furusaki,gogolin}, a fermionic RG method \cite{yue,lal2,polyakov},
172: and the functional RG method \cite{enss,ander,barnabe,meden}. Since our 
173: numerical calculations use a HF approximation, the method of Refs. 
174: \cite{yue,lal2} will be the most useful for us. 
175: Before considering the HF approach, however, we will briefly discuss the 
176: RG equations obtained by bosonization for spinless electrons.
177: 
178: \subsection{Spinless electrons}
179: 
180: Let us first consider the Hamiltonian for non-interacting electrons in
181: the presence of a $\delta$-function impurity placed at the origin,
182: \beq
183: H ~=~ \frac{{\vec p}^2}{2m} ~+~ V \delta (x) ~.
184: \label{ham}
185: \eeq
186: It is easy to check that for plane waves incident either from the left or
187: from the right with wavenumber $k$, the reflection and transmission 
188: amplitudes are given by \cite{merzbacher}
189: \bea
190: r(k) &=& - ~\frac{imV}{\hbar^2 k ~+~ imV} ~, \non \\
191: {\rm and} \quad t(k) &=& \frac{\hbar^2 k}{\hbar^2 k ~+~ imV} ~.
192: \label{rt1}
193: \eea
194: If the Fermi energy of the electrons is given by $E_F = \hbar^2 k_F^2 /(2m)$ 
195: (where $k_F$ is the Fermi wavenumber, and $v_F = \hbar k_F /m$ is the Fermi 
196: velocity), then the conductance $G$ for spinless electrons at zero temperature
197: is given by $e^2 /h$ times $|t (k_F)|^2$. For $|V| << \hbar v_F$, we see that
198: $|t (k_F)|^2 = 1 - (V/\hbar v_F)^2$ up to order $V^2$. (We will usually set 
199: Planck's constant $\hbar = 1$). Eq. (\ref{ham}) has a bound state if
200: $V < 0$, but this does not play any role in the RG analysis described below.
201: 
202: Now let us introduce interactions between the electrons. We assume a 
203: density-density interaction between spinless electrons of
204: the form 
205: \beq
206: H_{\rm int} ~=~ \frac{1}{2} ~\int \int ~dx dy ~\rho (x) ~U(x-y) ~\rho (y) ~,
207: \label{hint}
208: \eeq
209: where the density $\rho$ is given in terms of the second-quantized electron 
210: field $\Psi (x)$ as $\rho = \Psi^\dagger \Psi$. The electron field can be 
211: written in terms of the right and left moving fields $\Psi_R$ and $\Psi_L$ 
212: (whose variations in space are governed by wavenumbers much smaller than 
213: $k_F$) as
214: \beq
215: \Psi (x) ~=~ \Psi_R e^{ik_F x} ~+~ \Psi_L e^{-ik_F x} ~.
216: \label{psirl}
217: \eeq
218: If the range of the interaction $U(x)$ is short (of the order of $\lambda$),
219: such as that of a screened Coulomb repulsion, the Hamiltonian in (\ref{hint})
220: can be written in the form
221: \beq
222: H_{\rm int} ~=~ g_2 ~\int dx ~\Psi_R^\dagger \Psi_R \Psi_L^\dagger \Psi_L ~,
223: \label{hint2}
224: \eeq
225: where $g_2$ is related to the Fourier transform of $U(x)$ as $g_2 = {\tilde U}
226: (0)- {\tilde U} (2k_F)$. It is convenient to define the dimensionless constant
227: \beq
228: \alpha ~=~ \frac{g_2}{2\pi v_F} ~.
229: \label{alpha}
230: \eeq
231: 
232: A system of interacting electrons in one dimension such as the one introduced 
233: above is described by Tomonaga-Luttinger liquid (TLL) theory. The low energy 
234: excitations of a TLL are particle-hole pairs which are bosonic in nature and 
235: have a linear relation between energy and momentum. For spinless electrons, 
236: the low energy and long distance properties of the TLL are governed by three 
237: quantities, namely, the velocity $v$ of the low energy excitations, a 
238: dimensionless parameter $K$ which is related to the interaction strength, and 
239: the Fermi wavenumber $k_F$. For the model described above, we find that 
240: \cite{gogolin}
241: \bea
242: v ~&=&~ v_F ~(1 - \alpha^2 )^{1/2} ~, \non \\
243: {\rm and} \quad K ~&=&~ \bigl( \frac{1-\alpha}{1+\alpha} \bigr)^{1/2} ~.
244: \eea
245: Thus $K=1$ for non-interacting fermions.
246: For weak interactions, $v=v_F$ and $K = 1 -\alpha$
247: to first order in $\alpha$. In this paper, we will be interested in the
248: case in which the interaction is weak, i.e., $|\alpha| << 1$.
249: 
250: It turns out that in the presence of interactions, the impurity strength $V$ 
251: effectively becomes a function of a length scale $l$, and satisfies a RG
252: equation. On bosonizing the TLL theory \cite{kane,furusaki,gogolin}, we 
253: obtain the equation
254: \beq
255: \frac{dV}{d \ln l} ~=~(1 ~-~ K) ~ V ~, 
256: \label{rg1} 
257: \eeq
258: to first order in $V$, i.e., this is valid in the weak barrier limit.
259: In the strong barrier limit, which implies $|V| >> v_F$, bosonization leads
260: to a different RG equation
261: \beq
262: \frac{d (1/V)}{d \ln l} ~=~(1 ~-~ \frac{1}{K}) ~ (1/V) ~, 
263: \label{rg2} 
264: \eeq
265: which is valid to first order in $1/V$. We see that $V=0$ ($V = \infty$)
266: is a stable fixed point if $K > 1$ ($K < 1$ respectively). Since the above 
267: RG equations are not valid if $|V|/v_F$ is of order 1, we cannot 
268: conclude from them whether or not there is an intermediate fixed point. 
269: 
270: For a given value of the wire 
271: length $L$ and temperature $T$, the RG equations can be used to compute the 
272: conductance analytically as follows. We begin at a short length scale
273: (of order $\lambda$) with an initial value of $V$; we then use Eq. (\ref{rg1})
274: or (\ref{rg2}), depending on whether $V$ is small or large, to follow the 
275: evolution of $V$ with the length scale $l$. The RG flow stops when $l$ reaches
276: a distance of the order of $L$ or $L_T$, whichever is {\it smaller}. When the 
277: RG flow stops, we take the value of $V$ obtained at that point, and compute 
278: the conductance $G$ in terms of the transmission amplitude given in Eq. 
279: (\ref{rt1}). [We will see later that the RG flow of the conductance does not 
280: stop abruptly at one particular length scale. There is an intermediate
281: range of length scales where the conductance continues to evolve slowly.]
282: 
283: Let us now discuss the second method for obtaining the RG equations, namely,
284: using a HF approximation for the case of weak interactions
285: \cite{yue,lal2,polyakov}. This method directly gives an RG equation for
286: the scattering matrix which is produced by the impurity. The idea is 
287: that reflection from the impurity leads to an interference between the
288: incoming and outgoing electron waves; this leads to
289: Friedel oscillations in the density with an amplitude proportional to $r$
290: and a wavelength given by $\lambda = \pi /k_F$.
291: In the presence of interactions, these density oscillations cause the
292: electrons to scatter; these scatterings therefore renormalize the scattering 
293: caused by the impurity. The RG equations obtained by this method are given by
294: \bea
295: \frac{dt}{d \ln l} &=& - ~\alpha ~t ~|r|^2 ~, \non \\
296: {\rm and} \quad \frac{dr}{d \ln l} &=& \alpha ~r ~|t|^2 ~, \non \\
297: \label{rg3}
298: \eea
299: to first order in the interaction parameter $\alpha$. (Eq. (\ref{rg3}) is 
300: consistent with the unitarity of the scattering matrix, namely, $|r|^2 + |t|^2
301: = 1$ and $r^* t + t^* r = 0$). The solution of Eq. (\ref{rg3}) is given by
302: \beq
303: |t (l)|^2 ~=~ \frac{|t (d)|^2 ~(d/l)^{2\alpha}}{1 ~-~ |t(d)|^2 ~+~ |t(d)|^2 ~
304: (d/l)^{2\alpha}} ~,
305: \label{solvrg}
306: \eeq
307: where $d$ is a short distance scale which is of the order of $\lambda$. 
308: However, we will see at the end of Sec. V C that it is necessary to choose 
309: $d$ to be substantially larger than $\lambda$ in order to obtain a good fit 
310: between the numerical results and the expression in Eq. (\ref{solvrg}).
311: [At $l \rightarrow \infty$, Eq. (\ref{solvrg}) has fixed points at $|t|=$ 0 
312: and 1, depending on the sign of $\alpha$; there is no fixed point at 
313: intermediate values of $t$.]
314: 
315: We can check that Eq. (\ref{rg3}) is equivalent to Eqs. (\ref{rg1}) and 
316: (\ref{rg2}) if $\alpha$ is small and $|t|$ is close to 1 and 0 respectively. 
317: If $\alpha$ is small, the Luttinger parameter $K \simeq 1 - \alpha$. 
318: 
319: We note that Eq. (\ref{rg3}) is complementary to Eqs. (\ref{rg1}) and 
320: (\ref{rg2}) obtained from bosonization in the following sense. Eq. 
321: (\ref{rg3}) is valid for all values of $t$ but only small values of $\alpha$.
322: On the other hand, Eqs. (\ref{rg1}) and (\ref{rg2}) are valid for arbitrary 
323: values of the interaction parameter $\alpha$ but only for
324: small values of $V$ or $1/V$, i.e., only when either the reflection 
325: amplitude $r$ or the transmission amplitude $t$ is small. 
326: 
327: \subsection{Spin-$1/2$ electrons}
328: 
329: We now turn to the more realistic case of electrons with spin. For this 
330: situation, we will only discuss the RG equations obtained from the HF 
331: approximation for the case of weak interactions \cite{yue,lal2}. We again 
332: consider a $\delta$-function impurity at one point in the TLL, and a 
333: density-density interaction of the form given in (\ref{hint}); the density 
334: is now given by
335: \beq
336: \rho ~=~ \Psi_\ua^\dagger \Psi_\ua ~+~ \Psi_\da^\dagger \Psi_\da ~,
337: \eeq
338: where $\ua$ and $\da$ denote spin-up and spin-down respectively. Introducing 
339: right and left moving fields $\Psi_{R,\sigma}$ and $\Psi_{L,\sigma}$ as in 
340: Eq. (\ref{psirl}), with $\sigma = \ua$ or $\da$, we obtain the interaction
341: Hamiltonian
342: \bea
343: & & H_{\rm int} = \non \\
344: & & \int dx \sum_{\sigma,\sigma^\prime} [g_1
345: \Psi_{R\sigma}^\dagger \Psi_{L\sigma^\prime}^\dagger \Psi_{R\sigma^\prime}
346: \Psi_{L\sigma}+g_2 \Psi_{R\sigma}^\dagger \Psi_{L\sigma^\prime}^\dagger
347: \Psi_{L\sigma^\prime} \Psi_{R\sigma} \non \\
348: & & ~~~~~~~~ + \frac{1}{2} g_4 ( \Psi_{R\sigma}^\dagger
349: \Psi_{R\sigma^\prime}^\dagger \Psi_{R\sigma^\prime} \Psi_{R\sigma} +
350: \Psi_{L\sigma}^\dagger \Psi_{L\sigma^\prime}^\dagger \Psi_{L\sigma^\prime}
351: \Psi_{L\sigma} )], \non \\
352: & &
353: \eea
354: where $g_1 = {\tilde U} (2k_F)$, and $g_2 = g_4 = {\tilde U} (0)$. (We 
355: have ignored umklapp scattering terms here; they only arise if the model
356: is defined on a lattice and we are at half-filling). It is known that 
357: $g_1$, $g_2$ and $g_4$ satisfy some RG equations \cite{solyom}; 
358: the solutions of the lowest order RG equations are given by \cite{yue}
359: \bea
360: g_1 (l) &=& \frac{{\tilde U} (2k_F)}{1 ~+~ \frac{{\tilde U} (2k_F)}{\pi v_F}
361: \ln l} ~, \non \\
362: g_2 (l) &=& {\tilde U} (0) ~-~ \frac{1}{2} ~{\tilde U} (2k_F) ~+~ \frac{1}{2}~
363: \frac{{\tilde U} (2k_F)}{1 ~+~ \frac{{\tilde U} (2k_F)}{\pi v_F} \ln l} ~, 
364: \non \\
365: g_4 (l) &=& {\tilde U} (0) ~.
366: \label{g124}
367: \eea
368: 
369: Next, we can use the HF approximation and the existence of Friedel
370: oscillations in the density to derive the RG equation of the transmission
371: amplitude $t$. We find that \cite{yue,lal2}
372: \beq
373: \frac{dt}{d \ln l} ~=~ -~ \frac{g_2 (l) ~-~ 2 g_1 (l)}{2 \pi v_F} ~t ~|r|^2 ~.
374: \label{rg4}
375: \eeq
376: Due to the RG flow of the couplings $g_1$ and $g_2$, the solution of Eq. 
377: (\ref{rg4}) is more complicated than the expression for spinless fermions 
378: given in Eq. (\ref{solvrg}) \cite{yue}. We will therefore not attempt to
379: make a quantitative comparison between our numerical results and the RG 
380: equations for the case of spin-1/2 electrons.
381: 
382: It is interesting to note that unlike in the spinless case, Eq. (\ref{rg4})
383: allows the possibility of $|t|$ increasing for a while and then decreasing (or
384: vice versa) \cite{yue}. This is due to the flow of the couplings $g_1$ and 
385: $g_2$; it happens if ${\tilde U} (0) - 2 {\tilde U} (2k_F)$ and ${\tilde U} 
386: (0) - (1/2) {\tilde U} (2k_F)$ have opposite signs. This is precisely the 
387: situation for the Hubbard model in which ${\tilde U} (0)$ and ${\tilde U} 
388: (2k_F)$ are both equal to the Hubbard parameter $U$. 
389: 
390: \section{Lattice models and the Hartree-Fock approximation}
391: 
392: Although the system of interest may be defined in the continuum, it is
393: convenient to approximate it by a lattice model in order to do numerical 
394: calculations. (We should of course ensure that physical quantities
395: like the wire length are much larger than the lattice spacing, so that
396: the lattice approximation does not introduce significant errors).
397: Let us discuss the form of the lattice Hamiltonians 
398: that we will consider. The Hamiltonian has a hopping term
399: \beq
400: H_0 ~=~ - \gamma ~\sum_{n,\sigma} ~[~ c_{n,\sigma}^\dagger c_{n+1,\sigma} ~+~
401: c_{n+1,\sigma}^\dagger c_{n,\sigma} ~] ~,
402: \label{hhop}
403: \eeq
404: where $c_{n,\sigma}$ annihilates an electron with spin $\sigma$ at site $n$,
405: and the hopping amplitude $\gamma$ will be taken to be positive for 
406: convenience. Next, we will place an impurity at one site, say, $n=0$, so that
407: \beq
408: H_V ~=~ V \sum_\sigma ~c_{0,\sigma}^\dagger c_{0,\sigma} ~.
409: \label{hv}
410: \eeq 
411: Finally, we will introduce an interaction between the electrons in
412: the wire (but not in the leads which will be considered in the next section).
413: For the spin-1/2 case, the interaction will be taken to be of the Hubbard form,
414: \beq
415: H_{\rm int} ~=~ U ~\sum_n ~c_{n,\ua}^\dagger c_{n,\ua} ~c_{n,\da}^\dagger 
416: c_{n,\da} ~.
417: \label{int1}
418: \eeq
419: For the case of spinless electrons, 
420: the spin index $\sigma$ will be dropped in Eqs. (\ref{hhop}) and (\ref{hv}),
421: and the interaction will be taken to be between nearest neighbor sites,
422: \beq
423: H_{\rm int} ~=~ \frac{U}{2} ~\sum_n ~c_n^\dagger c_n ~c_{n+1}^\dagger 
424: c_{n+1} ~.
425: \label{int2}
426: \eeq
427: (For spinless electrons, we cannot have an on-site interaction since
428: $(c_n^\dagger c_n)^2 = c_n^\dagger c_n$).
429: 
430: In the absence of the impurity and interactions, the energy is related to 
431: the wavenumber as
432: \beq
433: E(k) ~=~ -~ 2 \gamma ~\cos k ~,
434: \label{disp}
435: \eeq
436: where $- \pi < k \le \pi$. (We have set the lattice spacing equal to 1). The 
437: velocity is given by $v= dE/dk = 2 \gamma \sin k$. If the chemical potential 
438: is given by $\mu$, with $-2 \gamma < \mu < 2 \gamma$, the ground state is the 
439: one in which the states with momenta $-k_F$ to $k_F$ are filled, where $\mu 
440: = - 2 \gamma \cos k_F$, with $0 < k_F < \pi$. It is convenient to define the 
441: Fermi energy as the difference between the chemical potential and the bottom 
442: of the band, namely, $E_F = \mu + 2 \gamma = 2 \gamma (1 - \cos k_F)$.
443: 
444: Let us now consider the effect of the impurity placed at $n=0$ (we are still 
445: ignoring the interactions). The reflection and transmission amplitudes are 
446: given by
447: \bea
448: r(k) &=& - ~\frac{iV}{2 \gamma \sin k ~+~ iV} ~, \non \\
449: {\rm and} \quad t(k) &=& \frac{2 \gamma \sin k}{2 \gamma \sin k ~+~ iV} ~.
450: \label{rt2}
451: \eea
452: In addition, there is a bound state if $V < 0$; the bound state energy is
453: given by $E_b = - 2 \gamma \cosh k_b$, where $V = - 2 \gamma \sinh k_b$,
454: with $k_b > 0$. The normalized bound state wave function is given by
455: \beq
456: \psi_{b,n} ~=~ (\tanh k_b)^{1/2} ~e^{-k_b |n|} 
457: \label{psib}
458: \eeq
459: for all values of $n$.
460: 
461: It is interesting to compute the electron density 
462: $\rho_n \equiv \rho_{n,\ua} = \rho_{n,\da}$, as a function of the 
463: site index $n$. The contribution from the scattering states is given by
464: \bea
465: \rho_n &=& \int_0^{k_F} ~\frac{dk}{2\pi} ~[~ |e^{-ik|n|} ~+~ r(k) 
466: e^{ik|n|}|^2 ~ +~ |t(k)|^2 ~] \non \\
467: &=& {\bar \rho} ~+~ \int_0^{k_F} ~\frac{dk}{2\pi} ~[~ r(k) e^{i2k|n|} ~+~
468: r^*(k) e^{-i2k|n|} ~] ~, \non \\
469: & &
470: \eea
471: where ${\bar \rho} = k_F /\pi$ is the value of the density very far
472: from the impurity. In the limit $|n| \rightarrow \infty$, we find that
473: the density has oscillations given by
474: \bea
475: & & \rho_n ~-~ {\bar \rho} \non \\
476: & & \simeq~ -~ \frac{i}{4 \pi |n|} [~ r (k_F) e^{i2k_F|n|} ~-~ r^* (k_F) 
477: e^{-i2k_F|n|} ~] ~.
478: \label{rhon}
479: \eea
480: 
481: Now we will introduce interactions and consider the HF approximation for the 
482: cases of spin-1/2 and spinless electrons. This will give us another way of
483: understanding the RG equations discussed in Sec. II.
484: 
485: \subsection{Spin-1/2 electrons}
486: 
487: For the interaction given in Eq. (\ref{int1}), the HF approximation takes 
488: the form
489: \beq
490: H_{\rm int,HF} ~=~ U ~\sum_n ~[~ \rho_n ~(c_{n,\ua}^\dagger c_{n,\ua} ~+~ 
491: c_{n,\da}^\dagger c_{n,\da}) ~-~ \rho_n^2 ~]~,
492: \label{int3}
493: \eeq
494: where we have set $<c_{n,\ua}^\dagger c_{n,\ua}>= <c_{n,\da}^\dagger 
495: c_{n,\da}>= \rho_n$. [This is called the restricted HF. One can also make
496: an unrestricted HF approximation in which one allows $<c_{n,\ua}^\dagger 
497: c_{n,\ua}>$ to differ from $<c_{n,\da}^\dagger c_{n,\da}>$.]
498: 
499: At this point, we would like to impose the condition that the interaction 
500: should have no effect on the conductance if there is no impurity, i.e., 
501: if $\rho_n$ is equal to the constant $\bar \rho$ at all sites. The reason
502: for imposing this condition is that we do not want the interactions by
503: themselves to lead to any scattering; that would change the conductance
504: from the ideal value of $2e^2 /h$ in the absence of any impurities which is
505: an undesirable effect. (This will become clearer in the next section when
506: we describe our model for the wires and the leads). We will therefore 
507: modify Eq. (\ref{int3}) to the form
508: \bea
509: H_{\rm int,HF} &=& U ~\sum_n ~(\rho_n ~-~ {\bar \rho}) ~(c_{n,\ua}^\dagger 
510: c_{n,\ua} ~+~ c_{n,\da}^\dagger c_{n,\da}) ~, \non \\
511: & &
512: \label{int4}
513: \eea
514: where we have dropped the constant $\rho_n^2$ which turns out to have no 
515: effect on the conductance. (The terms proportional to $\bar \rho$ in Eq.
516: (\ref{int4}) are equivalent to adding a chemical potential).
517: 
518: To gain an insight into the effect of Eq. (\ref{int4}), let us consider the 
519: Born approximation. In order to use this approximation, we will assume here
520: that both $V$ and $U$ are small. The total potential seen by either spin-up or
521: spin-down electrons at site $n$ is
522: \beq
523: {\cal V}_n ~=~ V \delta_{n,0} ~+~ U (\rho_n ~-~ {\bar \rho}) ~.
524: \eeq
525: For small values of $|V|/v_F$ (where $v_F = 2 \gamma \sin k_F$ is the 
526: Fermi velocity), $r(k_F) = -iV/v_F$ from Eq. (\ref{rt2}), and we have 
527: \beq
528: {\cal V}_n ~=~ V \delta_{n,0} ~-~ \frac{VU}{4 \pi v_F |n|} ~ (e^{i2k_F|n|} ~
529: +~ e^{-i2k_F|n|}) ~,
530: \eeq
531: where the second term is valid only for $|n| >> \pi /k_F$.
532: 
533: The Born approximation for the reflection amplitude in one dimension for
534: a lattice model is given by
535: \beq
536: r_B (k_F) ~=~ - \frac{i}{v_F} ~\sum_n {\cal V}_n ~e^{i2k_F n} ~.
537: \eeq
538: If the region in which the electrons interact with each other only
539: extends from $n=-L/2$ to $L/2$, then for $L>> \pi /k_F$, we get 
540: \beq
541: r_B (k_F) ~=~ - ~\frac{iV}{v_F} ~[~ 1 ~-~ \frac{U}{2 \pi v_F} ~\ln 
542: \frac{L}{2} ~]~.
543: \label{rb1}
544: \eeq
545: If $U>0$, we see that the conductance $G = (2e^2 /h) [1-|r_B (k_F)|^2]$ 
546: increases with $L$ till it reaches a maximum when $\ln (L/2) \simeq 2 \pi 
547: v_F/U$. However, the Born approximation that we are using here cannot be 
548: trusted up to such large length scales because we are not self-consistently 
549: modifying the densities at the different sites in response to the 
550: renormalization of the reflection amplitudes (produced by the interaction 
551: $U$). Because of this lack of self-consistency, we can only trust the Born 
552: approximation results up to length scales where the reflection amplitude 
553: has not changed very much from the value of $-iV/v_F$ that it has in the 
554: non-interacting theory. 
555: 
556: We note that Eq. (\ref{rb1}) is consistent with Eq. (\ref{rg4}) for
557: small values of $V$ and $(U/2\pi v_F) \ln l$, since $(g_2 (l)-2g_1 (l))/2\pi 
558: v_F = - U/2\pi v_F$ for such values of $l$.
559: 
560: \subsection{Spinless electrons}
561: 
562: For the interaction given in Eq. (\ref{int2}), the HF approximation takes 
563: the form
564: \bea
565: H_{\rm int,HF} ~=~ \frac{U}{2} ~\sum_n ~[~ & & \rho_{n+1} ~c_n^\dagger c_n ~
566: +~ \rho_n c_{n+1}^\dagger c_{n+1} \non \\
567: & & -~ <c_n^\dagger c_{n+1}> c_{n+1}^\dagger c_n \non \\
568: & & -~ <c_{n+1}^\dagger c_n> c_n^\dagger c_{n+1} ~]~,
569: \label{int5}
570: \eea
571: where $\rho_n \equiv <c_n^\dagger c_n>$, and we have dropped some constants.
572: In the absence of any impurities, $\rho_n = k_F /\pi$ and $<c_{n+1}^\dagger 
573: c_n> = (\sin k_F)/\pi$ for all values of $n$.
574: 
575: Once again, we impose the condition that the interaction should have no 
576: effect on the conductance if there is no impurity. We therefore 
577: modify Eq. (\ref{int5}) to the form
578: \bea
579: H_{\rm int,HF} ~=~ \frac{U}{2} ~\sum_n ~[ & & (\rho_{n+1} ~+~ \rho_{n-1} ~-~ 
580: 2 {\bar \rho}) ~c_n^\dagger c_n \non \\
581: & & -~ (<c_n^\dagger c_{n+1}> - \frac{\sin k_F}{\pi}) ~c_{n+1}^\dagger c_n 
582: \non \\
583: & & -~ (<c_{n+1}^\dagger c_n> - \frac{\sin k_F}{\pi}) ~c_n^\dagger c_{n+1} ~]~,
584: \non \\
585: & & 
586: \label{int6}
587: \eea
588: where ${\bar \rho} = k_F /\pi$ as before.
589: 
590: Let us again use the Born approximation to understand the effect of Eq. 
591: (\ref{int6}), assuming that both $V$ and $U$ are small. If one adds
592: a perturbation of the form
593: \bea
594: H_{\rm pert} ~=~ \sum_n ~[~ & & {\cal V}_n ~c_n^\dagger c_n \non \\
595: & & +~ {\cal V}_{n+1/2} ~c_{n+1}^\dagger c_n ~+~ {\cal V}_{n+1/2}^* ~
596: c_n^\dagger c_{n+1} ~] \non \\
597: & & 
598: \eea
599: to the Hamiltonian in Eq. (\ref{hhop}), the Born approximation for the
600: reflection amplitude is given by
601: \bea
602: r_B (k_F) = - \frac{i}{v_F} \sum_n ~[ & & {\cal V}_n ~e^{i2k_F n} \non \\
603: & & + ({\cal V}_{n+1/2} + {\cal V}_{n+1/2}^*) ~e^{ik_F (2n+1)} ]~. \non \\
604: & &
605: \label{rb2}
606: \eea
607: In our case,
608: \bea
609: {\cal V}_n &=& V \delta_{n,0} ~+~ \frac{U}{2} ~[ ~\rho_{n+1} ~+~ \rho_{n-1} ~
610: -~ 2 {\bar \rho} ~] ~, \non \\
611: {\cal V}_{n+1/2} &=& -~ \frac{U}{2} ~[~ <c_n^\dagger c_{n+1}> ~-~ 
612: \frac{\sin k_F}{\pi} ~]~.
613: \eea
614: Hence,
615: \bea
616: {\cal V}_n ~=~ V \delta_{n,0} ~-~ \frac{iU \cos(2k_F)}{4 \pi |n|} ~ & &
617: [~ r (k_F) e^{i2k_F|n|} \non \\ 
618: & & ~-~ r^* (k_F) e^{-i2k_F|n|} ~] ~, \non \\
619: & &
620: \eea
621: where the second term is only valid for $|n| \rightarrow \infty$, and
622: \bea
623: {\cal V}_{n+1/2} &=& \frac{iU}{8\pi |n|} ~[~ r (k_F) ~e^{i2k_F |n| + 
624: i {\rm sgn} (n) k_F} \non \\
625: & & ~~~~~~~~~-~ r^* (k_F) ~e^{-i2k_F |n| - i {\rm sgn} (n) k_F} ~] ~, 
626: \non \\
627: & &
628: \eea
629: for $|n| \rightarrow \infty$; here ${\rm sgn} (n) \equiv n /|n|$. 
630: For small values of $|V|/v_F$, $r (k_F) = 
631: - i V /v_F$. Substituting everything in the Born approximation (\ref{rb2}), 
632: and assuming that the region in which the electrons interact with each other 
633: extends from $n=-L/2$ to $L/2$, where $L>> \pi /k_F$, we find that
634: \beq
635: r_B (k_F) ~=~ - ~\frac{iV}{v_F} ~[~ 1 ~+~ \frac{U}{2 \pi v_F} ~(1 - \cos 
636: (2k_F)) ~\ln \frac{L}{2} ~]~.
637: \label{rb3}
638: \eeq
639: Eq. (\ref{rb3}) is consistent with Eq. (\ref{rg3}) for small values of $V$ 
640: and $(U/2\pi v_F) \ln l$, since $\alpha = (U/ 2\pi v_F)[1 - \cos (2k_F)]$.
641: 
642: \section{Non-equilibrium Green's function formalism}
643: 
644: In this section, we will introduce the NEGF formalism which will allow us to 
645: study the conductance of a wire with any length, both short (of the order 
646: of $\pi /k_F$) and long (where a continuum description and RG analysis may be 
647: expected to be reliable). 
648: 
649: \subsection{Self-energy, density and conductance}
650: 
651: An important concept in the NEGF formalism is a ``self-energy" which 
652: describes the amplitude for an electron to leave or enter the wire from the 
653: leads (reservoirs) which are maintained at some 
654: chemical potentials and temperature \cite{datta,meir,datta2,dhar,tsukada}.
655: The self-energy is a non-Hermitian term in the single-particle Hamiltonian of 
656: the wire. To see how this arises, let us begin by modeling
657: one of the reservoirs by a tight-binding Hamiltonian
658: \beq 
659: H ~=~ -\gamma ~\sum_{n,\sigma} ~(~c_{n,\sigma}^\dagger c_{n+1,\sigma} ~ 
660: +~ c_{n+1,\sigma}^\dagger c_{n,\sigma} ~) ~. 
661: \eeq
662: The energy of an electron in the reservoir is related to its wavenumber as 
663: $E = - 2 \gamma \cos k$. The reservoir has a chemical potential $\mu$ and 
664: an inverse temperature $\beta = 1/(k_B T)$. The reservoir is semi-infinite;
665: the last site at one end of the reservoir is coupled to the first site of
666: the wire by a hopping amplitude $\gamma'$ (see Fig. 1).
667: 
668: \begin{figure}[htb]
669: \begin{center}
670: \epsfig{figure=sigma.ps,width=8.5cm}
671: \end{center}
672: \caption{(Color online) Picture of a semi-infinite reservoir going from site
673: number 0 to $- \infty$ and a wire beginning from site number 1.}
674: \end{figure}
675: 
676: The reservoir gives rise to a self-energy at the first site of the wire of 
677: the form \cite{datta,dhar}
678: \beq 
679: \Sigma (E) ~=~ \sigma (E) ~\sum_\sigma ~c_{1,\sigma}^\dagger c_{1,\sigma} ~,
680: \label{self}
681: \eeq
682: where
683: \beq 
684: \sigma (E) ~=~ -~ \frac{\gamma'^2}{\gamma} ~e^{ik} 
685: \eeq
686: for $-2\gamma \le E = - 2 \gamma \cos k \le 2\gamma$ (with $0 \le k \le \pi$),
687: \beq 
688: \sigma (E) ~=~ -~ \frac{\gamma'^2}{\gamma} ~e^{-k} ~-~ i \eta 
689: \eeq
690: for $E = - 2 \gamma \cosh k \le -2\gamma$ (with $k \ge 0$), and
691: \beq 
692: \sigma (E) ~=~ \frac{\gamma'^2}{\gamma} ~e^{-k} ~-~ i \eta
693: \eeq
694: for $E = 2 \gamma \cosh k \ge 2\gamma$ (with $k \ge 0$). Here $\eta$ is
695: an infinitesimal positive number which appears only if $E$ lies 
696: outside the range $[-2\gamma, 2\gamma]$. For $-2 \gamma < E < 2 \gamma$,
697: the self-energy already has an imaginary piece, so it is not necessary to
698: add an $i \eta$ term.
699: 
700: \begin{figure}[htb]
701: \epsfig{figure=negf.ps,width=8.5cm}
702: \caption{(Color online) A quantum wire system with a $N$-site wire in the 
703: middle and semi-infinite reservoirs on its left and right.}
704: \end{figure}
705: 
706: Now let us consider the complete system which consists of a wire with $N$ sites
707: in the middle, and reservoirs on its left and right as shown in Fig. 2. If the
708: wire is also modeled by a tight-binding Hamiltonian (with a hopping amplitude
709: $\gamma_1$), the Green's function of the wire at energy $E$ is given by a 
710: $N \times N$ matrix
711: \bea 
712: & & G (E) ~=~ [~ E ~I ~-~ H ~-~ \Sigma_L (E) ~-~ \Sigma_R (E) ~]^{-1} ~, 
713: \non \\ 
714: & & \non \\
715: & & H ~+~ \Sigma_L (E) ~+~ \Sigma_R (E) ~= \non \\
716: & & \non \\
717: & & -~ \left( \begin{array}{ccccc} ~\\
718: \sigma (E)~ & ~ \gamma_1~ & ~0~ & ~\cdots~ & ~\cdots~ \\
719: ~\gamma_1~ & ~0~ & ~\gamma_1~ & ~0~ & ~\cdots~ \\
720: ~0~ & ~\gamma_1~ & ~0~ & ~\gamma_1~ & ~0~ \\
721: ~\cdots~ & ~\cdots~ & ~\cdots~ & ~\cdots~ & ~\cdots~ \\
722: ~\cdots~ & ~\cdots~ & ~0~ & ~\gamma_1~ & ~\sigma (E)~ \end{array} \right) ~. 
723: \eea
724: (The effects of the reservoirs on the wire is completely taken into account
725: by the self-energy terms). We would like to emphasize that the relation 
726: between $E$ and $k$ is given entirely by the dispersion in the reservoirs 
727: (we assume that the dispersion is the same in both the reservoirs), and 
728: {\it not} on the form of the Hamiltonian $H$ inside the wire.
729: 
730: The density matrices due to electrons coming in from 
731: the left and right reservoirs are given by \cite{datta,meir,datta2,dhar}
732: \beq 
733: \int ~\frac{dE}{2\pi} ~G \Gamma_L G^\dagger ~f_L \quad {\rm and} \quad 
734: \int ~\frac{dE}{2\pi} ~G \Gamma_R G^\dagger ~f_R 
735: \label{density}
736: \eeq
737: respectively, where $\Gamma_a (E) = i (\Sigma_a - \Sigma_a^\dagger),$
738: \beq f_a (E) ~=~ [e^{\beta (E - \mu_a)} + 1]^{-1} \eeq
739: is the Fermi function, and $\mu_a~$ is the chemical potential in reservoir $a$.
740: We note that if $E$ lies outside the range $[-2\gamma,2\gamma]$, the matrix
741: $\Gamma_a (E) \sim 2 \eta$ is infinitesimal. In that case, the density matrix 
742: can still receive a contribution from certain states; these are typically bound
743: states which have a discrete set of energies $E_b$. The reason that such states
744: can make a finite contribution in Eq. (\ref{density}) even though $\Gamma_a$
745: is infinitesimal is that $G$ takes the form $1/(E - E_b +i \eta)$, and
746: \beq
747: {\rm Lim}_{\eta \rightarrow 0^+} ~\frac{2~\eta}{(E-E_b)^2 ~+~ 
748: \eta^2} ~=~ 2~\pi~\delta (E-E_b) ~.
749: \eeq 
750: 
751: Finally, the current is given by the expression
752: \beq 
753: I ~=~ - ~\frac{e}{h} ~\int ~dE ~{\rm trace} (G \Gamma_L G^\dagger \Gamma_R) ~
754: [ f_L (E) - f_R (E) ] ~, 
755: \label{current}
756: \eeq
757: and the conductance is $G ~=~ e ~I / (\mu_L ~-~ \mu_R) ~.$ If $E$ lies outside
758: the range $[-2\gamma,2\gamma]$, $\Gamma_L (E)$ and $\Gamma_R (E)$ are both 
759: infinitesimal, and the current does not get any contribution from states lying
760: in that energy range. (The difference between the density matrix and the
761: current for such states is that the density matrix has only one factor of 
762: $\eta$ in the numerator while the current has two factors of $\eta$).
763: 
764: We will be interested below in the case of linear response, i.e, the limit
765: $\mu_L \rightarrow \mu_R = \mu$. In that limit, the Fermi wavenumber $k_F$ is
766: given by $\mu = - 2 \gamma \cos k_F$. Further, the conductance takes the form
767: \bea
768: G &=& \frac{e^2}{h} ~\int_{-2\gamma}^{2\gamma} ~dE ~{\rm trace} (G \Gamma_L 
769: G^\dagger \Gamma_R) ~\frac{\beta}{[2 ~\cosh \frac{\beta}{2} (E - \mu)]^2}. 
770: \non \\
771: &&
772: \eea
773: (We have to multiply the above expression by 2 for spin-half electrons).
774: 
775: \subsection{Interactions}
776: 
777: Let us now consider how interactions can be studied within the NEGF formalism.
778: Note that this formalism works with a one-particle Hamiltonian, e.g., the 
779: self-energy in Eq. (\ref{self}) is given in terms of the energy of a single 
780: electron which is entering or leaving the wire. Hence, the only way to deal 
781: with interactions is to do a HF decomposition as shown in Eqs. (\ref{int3}) 
782: and (\ref{int5}); these take the form of corrections to the on-site chemical 
783: potential or the hopping amplitude.
784: 
785: However, there is a difficulty in using the expressions in 
786: (\ref{int3}) or (\ref{int5}). We want to have 
787: interactions between the electrons only in the wire, not in the reservoirs.
788: The HF decompositions shown in (\ref{int3}) and (\ref{int5}) will then lead to
789: one-body terms only in the wire; these terms will back-scatter electrons coming
790: from the reservoirs, and hence reduce the conductance from its
791: ideal value of $2e^2 /h$ or $e^2 /h$. We would like to ensure that there is 
792: no such scattering if there are no impurities inside the wire and if
793: the leads connect adiabatically to the wires, i.e., if the entire system with
794: wire and reservoirs is translation invariant. We know that 
795: two-body scatterings between electrons conserve momentum and therefore do not
796: affect the conductance in the absence of impurities and scattering from the
797: lead-wire junctions. We therefore modify the form of the HF decomposition from
798: Eqs. (\ref{int3}) and (\ref{int5}) to Eqs. (\ref{int4}) and (\ref{int6}) 
799: respectively; the extra terms in those equations ensure that the HF terms 
800: vanish identically if there are no impurities. [An alternative way of ensuring
801: adiabaticity between the leads and the wires is to turn on the interaction 
802: strength $U$ smoothly from zero in the leads to a finite value inside 
803: the wire. This, however, is harder to implement numerically. Subtracting the 
804: mean values as in (\ref{int4}) and (\ref{int6}) is easier to implement, and
805: it serves the same purpose because the subtracted quantities approach zero 
806: near the ends of the wires (this will become clear from the numerical 
807: results presented below).]
808: 
809: In the presence of interactions, we will implement a self-consistent NEGF 
810: calculation as follows:
811: \begin{itemize} 
812: \item{Start with the Hamiltonian with no interactions, and calculate the 
813: density matrix. The diagonal elements of the density matrix give the 
814: densities at different sites.} 
815: \item{Use the HF approximation to compute the Hamiltonian with interactions, 
816: and use that to calculate the density matrix again.} 
817: \item{Repeat the previous step till the density changes no further,
818: i.e., between two successive iterations, the maximum change in density on 
819: any lattice site is less than about $10^{-4}$.} 
820: \item{Use the converged density to compute the conductance.} 
821: \end{itemize}
822: 
823: \section{Numerical results}
824: 
825: In this section, we will present the results obtained by the NEGF formalism, 
826: first for non-interacting electrons, and then for interacting spin-1/2 and 
827: spinless electrons, using the procedure described in Sec. IV B. We will study 
828: the dependence of the conductance $G$ on the wire length $L$, the inverse 
829: temperature $\beta$, the impurity strength $V$, and the interaction parameter 
830: $U$. Finally, we will make a quantitative comparison between our numerical
831: results and the RG equations for the case of spinless electrons.
832: 
833: We present some details about our choice of parameters for the calculations.
834: We always take the wire to have an odd number of sites, and the impurity to be
835: on the middle site. The hopping amplitudes in the reservoirs and in the wire 
836: is chosen to be the same ($\gamma = \gamma_1 = \gamma' =1$); the values of
837: the inverse temperature $\beta = 1/(k_B T)$ is quoted in units of $1/\gamma$.
838: (We set both $\hbar$ and the lattice spacing equal to 1). We choose $k_F = 
839: \pi /10$; hence, $v_F = 2 \sin k_F = 0.618$, $\mu = - 2 \cos k_F = -1.902$, 
840: and $E_F = 2 (1 - \cos k_F) = 0.098$. The thermal coherence length is
841: given by $L_T = 0.618 \beta$.
842: In the energy integrations for the current and conductance, we take the energy
843: step size to be $dE=2 \times 10^{-4}$. In the integration outside the energy 
844: range $[-2\gamma , 2 \gamma]$, we take the quantity $\eta$ to be 5 times $dE$.
845: Finally, in all the figures, the conductance $G$ is expressed in units of 
846: $2e^2 /h$ and $e^2 /h$ for the spin-1/2 and spinless cases respectively.
847: 
848: Our choice of parameters was dictated partly by experiments on quantum wires 
849: in semiconductor heterojunctions \cite{tarucha,liang,bkane,yacoby,aus,reilly},
850: and partly by our numerical limitations (which prevent us from going to
851: very large wire lengths). Experimentally, the ratio of the wire length to
852: the de Broglie wavelength of the electrons ($Lk_F /\pi$) ranges from 
853: about 20 to 200, the ratio of the wire length to the inverse temperature 
854: $L/\beta$ ranges from about 1/2 to 10, and the parameter $U/(2 \pi v_F)$ 
855: appearing in the Eq. (\ref{rb1}) is about $0.2$ - $0.3$ \cite{lal1}. In our 
856: calculations, $Lk_F /\pi$ goes from about 1 to 30, $L/\beta$ goes from about 
857: 1/40 to 6, and the interaction parameters take the values $U/(2\pi v_F) = 
858: 0.0773$ in the spin-1/2 case (Eq. (\ref{rb1})) and $(U/ 2 \pi v_F)[1 - 
859: \cos(2k_F)] = 0.0148$ in the spinless case (Eq. (\ref{rb3})) for $U=0.3$.
860: The much smaller value of the interaction parameter in the spinless case 
861: leads to smaller changes in the conductance in that case as we will see.
862: 
863: \subsection{Non-interacting electrons}
864: 
865: Let us first discuss some properties of a system in which there is a 
866: $\delta$-function impurity of strength $V$ in the middle of the wire, and 
867: there are no interactions between the electrons. Fig. 3 shows the Friedel 
868: oscillations in the density $<c_n^\dagger c_n>$ for $V=0.3$ for two different 
869: temperatures. Note that the density at the middle site is much lower than the 
870: mean value of ${\bar \rho} = k_F /\pi = 1/10$ due to the repulsive nature of 
871: the impurity. We also observe that the oscillations die out beyond a length 
872: scale of order $L_T = 0.618 \beta$ in the lower figure. This is a simple 
873: illustration of the idea of a thermal coherence length. 
874: 
875: \begin{figure}[htb]
876: \begin{center}
877: \epsfig{figure=rho1_ni.eps,width=8.5cm}
878: \epsfig{figure=rho2_ni.eps,width=8.5cm}
879: \end{center}
880: \caption{(Color online) Friedel oscillations in the density of 
881: non-interacting electrons caused by an impurity with $V=0.3$ placed in the 
882: middle of a wire with 151 sites. The upper and lower figures have $\beta =$ 
883: 300 and 50 respectively.}
884: \end{figure}
885: 
886: \begin{figure}[htb]
887: \begin{center}
888: \epsfig{figure=rho3b_ni.eps,width=8.5cm}
889: \end{center}
890: \caption{(Color online) Friedel oscillations in the density of 
891: non-interacting electrons caused by an impurity with $V=-0.3$ placed in 
892: the middle of a wire with 151 sites and $\beta = 300$.} 
893: \end{figure}
894: 
895: \begin{figure}[htb]
896: \begin{center}
897: \epsfig{figure=cndcnp1a.eps,width=8.5cm}
898: \epsfig{figure=cndcnp1b.eps,width=8.5cm}
899: \end{center}
900: \caption{(Color online) Plots of $\rho_1 \equiv <c_n^\dagger c_{n+1}>$ 
901: for non-interacting electrons in the presence of an impurity with $V=0.3$ 
902: (upper figure) and $-0.3$ (lower figure) placed in the middle of a wire with 
903: 151 sites, with $\beta = 300$.} 
904: \end{figure}
905: 
906: \begin{figure}[htb]
907: \begin{center}
908: \epsfig{figure=GvsT_ni.eps,width=8.6cm}
909: \end{center}
910: \caption{(Color online) Conductance versus $\beta$ for non-interacting 
911: electrons in the presence of an impurity with $V=0.3$ placed in the middle 
912: of a wire with 201 sites.} 
913: \end{figure}
914: 
915: \begin{figure}[htb] 
916: \begin{center}
917: \epsfig{figure=Gvsl1_sf.eps,width=8.5cm}
918: \epsfig{figure=Gvsl2_sf.eps,width=8.5cm}
919: \end{center}
920: \caption{(Color online) Conductance versus wire length for interacting 
921: spin-1/2 electrons for three different values of $\beta$, for $V=0.3$. $U=$ 
922: 0.3 and -0.3 in the upper and lower figures respectively.}
923: \end{figure}
924: 
925: \begin{figure}[htb]
926: \begin{center}
927: \epsfig{figure=GvsT2_sf.eps,width=8.5cm}
928: \end{center}
929: \caption{(Color online) Conductance versus $\beta$ for interacting spin-1/2 
930: electrons for three different wire lengths, for $V=0.3$ and $U=0.3$.}
931: \end{figure}
932: 
933: \begin{figure}[htb]
934: \begin{center}
935: \epsfig{figure=GvsV1_sf.eps,width=8.5cm}
936: \end{center}
937: \caption{(Color online) Conductance $G$ versus $V$ for interacting spin-1/2 
938: electrons for $U$ = 0.3, 0 and -0.3, with $L=101$ and $\beta = 100$.}
939: \end{figure}
940: 
941: \begin{figure}[htb]
942: \begin{center}
943: \epsfig{figure=Gvsu1_sf.eps,width=8.5cm}
944: \end{center}
945: \caption{(Color online) Conductance $G$ versus $U$ for interacting spin-1/2 
946: electrons for $V$ = 1 and -1, with $L=101$ and $\beta = 100$.}
947: \end{figure}
948: 
949: Fig. 4 shows the Friedel oscillations in the density for $V=-0.3$ for
950: $\beta = 300$. The density in the middle site is now much higher than 
951: the mean value; this is because of the presence of a bound state whose wave 
952: function has a peak at that site. Indeed, in the absence of interactions,
953: the difference in the site densities for $V=0.3$ and $-0.3$ is given entirely 
954: by the bound state. We see from Figs. 3 and 4 that the density difference 
955: at the site of the impurity is about 0.15; this agrees well with the
956: value of $\tanh k_b = 0.148$ given by Eq. (\ref{psib}). (For $V=-0.3$, we
957: have $k_b = 0.149$).
958: 
959: Finally, Fig. 5 shows the quantity $\rho_1 \equiv <c_n^\dagger c_{n+1}>$ for
960: $V=0.3$ and $-0.3$ for $\beta = 300$. There is a strong similarity between 
961: these and the density plots shown in Figs. 3 (upper figure) and 4. This
962: is because the electrons at the Fermi energy (which dominate the long 
963: distance properties of the system) have a wavelength which is $\pi /k_F = 10$
964: times longer than the lattice spacing; hence there is very little difference 
965: between $<c_n^\dagger c_n>$ and $<c_n^\dagger c_{n+1}>$.
966: 
967: For non-interacting electrons, the conductance depends on the temperature, 
968: but not on the length of the wire. For a $\delta$-function impurity of 
969: strength $V$, the conductance for spinless electrons is given by
970: \bea
971: G &=& \frac{e^2}{h} ~\int_{-2}^2 ~dE ~{\cal T} (E)~
972: \frac{\beta}{[2 ~\cosh \frac{\beta}{2} (E - \mu)]^2} ~, \non \\
973: {\cal T} (E) &=& \frac{4 \sin^2 k}{4 \sin^2 k + V^2}~.
974: \label{cond}
975: \eea
976: where $E= - 2 \cos k$. For $k_B T << E_F = \mu + 2$, Eq. (\ref{cond}) has 
977: a Sommerfeld expansion of the form \cite{ashcroft}
978: \beq
979: G ~=~ \frac{e^2}{h} ~[~ T(\mu) ~+~ \frac{\pi^2}{6} ~(k_B T)^2 ~T''(\mu) ~+~ 
980: \cdots ~]~.
981: \label{g0}
982: \eeq
983: For $k_F = \pi /10$ and $V=0.3$, $T(\mu) = 0.809$ and $(\pi^2 /6) T''(\mu) = 
984: 42.1$; this implies a significant temperature dependence of the conductance. 
985: Fig. 6 shows that the conductance changes appreciably with $\beta$ till 
986: $\beta$ reaches about 100. We have to keep this in mind when studying the 
987: temperature dependence of the conductance of a system of interacting electrons.
988: 
989: \begin{figure}[h!]
990: \begin{center}
991: \epsfig{figure=rho1a_sf.eps,width=8.5cm}
992: \epsfig{figure=rho2a_sf.eps,width=8.5cm}
993: \epsfig{figure=rho3a_sf.eps,width=8.5cm}
994: \end{center}
995: \caption{(Color online) Friedel oscillations for spin-1/2 electrons for 
996: $U =$ 0.5, 0 and -0.5 (upper, middle and lower figures respectively) for 
997: $V=0.3$, $L=101$ and $\beta = 100$.}
998: \end{figure}
999: 
1000: \subsection{Spin-1/2 electrons}
1001: 
1002: We now consider the spin-1/2 model with an on-site interaction between the
1003: electrons. In the absence of interactions, the conductance at zero temperature
1004: is independent of the wire length and is given by Eq. (\ref{g0}), with a 
1005: factor of 2 for spin; namely, 
1006: \beq 
1007: G_0 ~=~ \frac{2e^2}{h} ~\frac{4 \sin^2 k_F}{4 \sin^2 k_F ~+~ V^2} ~.
1008: \label{gvsv}
1009: \eeq
1010: For $k_F = \pi /10$ and $|V| = 0.03$, $G_0 = (2e^2 /h) ~0.809$.
1011: 
1012: In Fig. 7, we show the conductance as a function of the wire length $L$ for 
1013: three different value of $\beta$, with $U=0.3$ in the upper figure and 
1014: $U=-0.3$ in the lower figure. The trends in Fig. 7 are in accordance with 
1015: Eq. (\ref{rb1}). Firstly, the conductance increases with the length scale if 
1016: $U > 0$ and decreases if $U < 0$. For $U=0.3$, we have $U/ (2 \pi v_F) = 
1017: 0.0773$; we expect the HF approximation to be reasonable for such a small 
1018: value. [Due to the RG flows of $g_1$ and $g_2$ in Eq. (\ref{rg4}), the 
1019: conductance should start decreasing for positive $U$ and increasing for 
1020: negative $U$ at a very large length scale of the order of $L \sim 
1021: \exp(1/0.0773) \sim 420000$, provided that $L_T$ is also larger than $420000$. 
1022: Such a wire length is beyond our numerical capability.]
1023: 
1024: Secondly, for fixed $\beta$, $G$ increases till $L$ reaches a value of the 
1025: order of $L_T = 0.618 \beta$ beyond which $G$ stops changing. This happens 
1026: because there is complete thermal decoherence once $L$ exceeds $L_T$; hence 
1027: $G$ does not vary any more with $L$ in that regime. Thirdly, the conductance 
1028: for two different values of $\beta$, say, $\beta_1$ and $\beta_2$ (where 
1029: $\beta_1 < \beta_2$), start separating from each other at a value of the wire 
1030: length $L_c$ such that $L_c /\beta_1$ is much less than $0.618$. This happens
1031: because if the inverse temperature is $\beta_1$ or higher, there is complete 
1032: thermal coherence, and $G$ does not depend on $\beta$ any more. Thus, for any 
1033: temperature $\beta$, there is a range of wire lengths going from $L_c$ up to 
1034: $L_T$ where there is partial thermal coherence, and the conductance evolves 
1035: slowly in this range of lengths. We therefore conclude that the RG flow does 
1036: not suddenly stop at a length scale which is the smaller of the wire length 
1037: $L$ and the thermal coherence length $L_T$; rather, the flow continues to 
1038: occur (but slows down) in an intermediate range of length scales.
1039: 
1040: Finally, Fig. 7 shows that even for the smallest values of $L$, the conductance
1041: differs from the non-interacting value of $0.809$ by an appreciable amount. 
1042: This is due to the large deviations of the density from the mean value at the
1043: sites close to the impurity. In that region, the density does not satisfy the
1044: $1/|n|$ form given in Eq. (\ref{rhon}). Since that form is intimately tied 
1045: to the RG results (the sum over $1/|n|$ gives $\ln (L)$), we do not expect 
1046: the numerically obtained value of the conductance to agree well with the RG
1047: analysis for short wire lengths.
1048: 
1049: In Fig. 8, we show the conductance as a function of $\beta$ for three 
1050: different value of $L$, with $U=0.3$. Once again, we see the same trends as 
1051: in Fig. 7. Namely, for any two values of $L$, there is 
1052: a $\beta$ where the values of $G$ start separating from each other. Then 
1053: there is a higher value of $\beta$ where $G$ stops changing. The region 
1054: between the two is where there is only partial thermal coherence, and $G$
1055: varies relatively slowly in that region.
1056: 
1057: In Fig. 9, we show the conductance as a function of $V$ for three different
1058: values of $U$, with $L=101$ and $\beta = 100$. Note that for $U=0$, the 
1059: conductance is an even function of $V$ given by Eq. (\ref{gvsv}). But in the 
1060: presence of interactions, $G$ is no longer precisely an even function of $V$,
1061: particularly for large values of $V$.
1062: This is more clearly visible in Fig. 10 which shows the conductance as a 
1063: function of $U$ for $V=$ 1 and -1, with $L=101$ and $\beta = 100$. For any 
1064: value of $U$, we find that the conductance deviates less from the 
1065: non-interacting value for $V>0$ compared to $V<0$, for the same value of $|V|$.
1066: For small values of $U$, this can be explained as follows. We saw in Figs. 3 
1067: and 4 that the density at the site of the impurity deviates from the mean 
1068: density of ${\bar \rho} = 1/10$ by a smaller amount for $V>0$ compared to 
1069: $V<0$; the numerical values for $\rho_0 - 1/10$ are given by -0.046 and 0.095 
1070: for $V=$ 0.3 and -0.3 respectively. The effective impurity strength is given by
1071: \beq
1072: V_{\rm eff} ~=~ V ~+~ U ~( \rho_0 ~-~ {\bar \rho}) ~,
1073: \eeq
1074: which equals $0.3 - 0.046 U$ and $-0.3 + 0.095 U$ for $V=0.3$ and $-0.3$ 
1075: respectively. For $U$ small and positive, this means that the magnitude of 
1076: the effective impurity strength and therefore the reflection probability is 
1077: larger for $V=0.3$; hence the conductance is smaller for $V=0.3$. The opposite
1078: statement is true if $U$ is small and negative. To conclude, the different 
1079: values of the density deviation at the impurity site for positive and negative
1080: values of $V$ are responsible for the asymmetry in $G$ as a function of $V$.
1081: (This asymmetry is more prominent for the case of spinless electrons as we 
1082: will see below).
1083: 
1084: The fact that the deviation of the conductance from the non-interacting 
1085: value is smaller for $V>0$ than for $V<0$ is clearly 
1086: a result of the bound state which raises the density
1087: at the impurity site for $V<0$; hence this is a short distance effect.
1088: [The contribution of the scattering states to the reflection probability
1089: $|r|^2$ is an even function of $V$ as one can see from Eq. (\ref{rb1}).
1090: The RG equations which are derived from continuum theories take into account 
1091: only long distance effects; these come from the scattering states only.]
1092: 
1093: In Fig. 11, we show the Friedel oscillations for three different values of $U$
1094: with $V=0.3$, $L=101$ and $\beta = 100$. We see that a repulsive interaction 
1095: ($U>0$) suppresses the Friedel oscillations, while an attractive interaction
1096: ($U<0$) with the same magnitude enhances the oscillations by a much larger
1097: amount. This is because attractive interactions tend to lead to the formation
1098: of a charge density wave; if a charge density wave is already present (due to
1099: the impurity), attractive interactions enhance it. Since the scattering from 
1100: the Friedel oscillations renormalize the scattering from the impurity, this 
1101: difference in the magnitude of the oscillations helps to explain why the 
1102: deviation of the conductance $G$ from the non-interacting value is larger 
1103: for $U<0$ compared to $U>0$, as can be seen in Fig. 7.
1104: 
1105: \begin{figure}[htb] 
1106: \begin{center}
1107: \epsfig{figure=Gvsl1_sl.eps,width=8.5cm}
1108: \end{center}
1109: \caption{(Color online) Conductance versus wire length for interacting 
1110: spinless electrons for three different values of $\beta$, for $V=0.3$ and 
1111: $U=0.3$.}
1112: \end{figure}
1113: 
1114: \begin{figure}[h!]
1115: \begin{center}
1116: \epsfig{figure=GvsT2_sl.eps,width=8.5cm}
1117: \end{center}
1118: \caption{(Color online) Conductance versus $\beta$ for interacting spinless 
1119: electrons for three different values of $U$, for a wire with 201 sites and 
1120: $V=0.3$.}
1121: \end{figure}
1122: 
1123: \begin{figure}[htb]
1124: \begin{center}
1125: \epsfig{figure=rgfit2.eps,width=8.6cm}
1126: \end{center}
1127: \caption{(Color online) Conductance $G$ versus $L$ for interacting spinless 
1128: electrons compared with the RG expression in Eq. (\ref{solvrg}), for $V = 0.3$,
1129: $U=0.3$, and $\beta = \infty$. Line A (dash dot line) corresponds to $d=11, ~
1130: |t(d)|^2 = 0.805$ and $\alpha = 0.0148$. Line B (dashed line) corresponds 
1131: to $d=51, ~|t(d)|^2 = 0.795$ and $\alpha = 0.0148$. Line C (solid line) 
1132: corresponds to $d=51, ~|t(d)|^2 = 0.795$ and $\alpha = 0.016$. The asterisks 
1133: show the numerical results.}
1134: \end{figure}
1135: 
1136: \begin{figure}[h!]
1137: \begin{center}
1138: \epsfig{figure=GvsV2_sl.eps,width=8.5cm}
1139: \end{center}
1140: \caption{(Color online) Conductance $G$ versus $V$ for interacting spinless 
1141: electrons for $U$ = 0.3, 0 and -0.3, with $L=101$ and $\beta = 100$.}
1142: \end{figure}
1143: 
1144: \begin{figure}[htb]
1145: \begin{center}
1146: \epsfig{figure=Gvsu2_sl.eps,width=8.5cm}
1147: \end{center}
1148: \caption{(Color online) Conductance $G$ versus $U$ for interacting spinless 
1149: electrons for $V$ = 1 and -1, with $L=101$ and $\beta = 100$.}
1150: \end{figure}
1151: 
1152: For spin-1/2 electrons, we can do a self-consistent HF calculation in two ways,
1153: restricted and unrestricted. In a restricted HF calculation, the site densities
1154: of spin-up and spin-down electrons are taken to be equal at all stages. 
1155: In a unrestricted HF calculation, spin-up and spin-down electrons are allowed 
1156: to have different densities. We have done both kinds of calculations. For the 
1157: range of the interaction $-0.5 \le U \le 0.5$, we find that they give the same
1158: results; the spin-up and spin-down densities converge to the same values even 
1159: if one begins with different initial values for them. Thus we do not find any 
1160: spin density wave within our range of parameters.
1161: 
1162: \subsection{Spinless electrons} 
1163: 
1164: We now consider a model of spinless electrons with nearest neighbor 
1165: interactions between the electrons. Comparing the forms of Eqs. (\ref{rb1}) 
1166: and (\ref{rb3}), we see that the behaviors of the spin-1/2 model for $U>0$ 
1167: and $U<0$ should be similar to that of the spinless model for $U<0$ and $U>0$ 
1168: respectively. In Fig. 12, we show the conductance as a function of the wire 
1169: length $L$ for three different value of $\beta$, with $V=0.3$ and $U=0.3$. 
1170: The trends are in agreement with Eq. (\ref{rb1}), with $\alpha = U (1 - \cos 
1171: (2k_F))/ (2 \pi v_F) = 0.0148$. For fixed $\beta$, $G$ decreases till $L$ 
1172: reaches a value of the order of $L_T = 0.618 \beta$.
1173: 
1174: In Fig. 13, we show the conductance as a function of $\beta$ for three 
1175: different value of $U$, with $V=0.3$ and $L=201$. As we saw earlier in 
1176: Fig. 6, $G$ has a significant dependence on $\beta$ for small values of
1177: $\beta$ even for non-interacting electrons. It is only when we go to
1178: large values of $\beta$ that we see the trend expected from the RG 
1179: equations for interacting electrons; namely, as $\beta$ increases, the 
1180: conductance decreases if $U>0$ and increases if $U<0$.
1181: 
1182: Since $G$ has an appreciable dependence on $\beta$ even for non-interacting 
1183: electrons, it is easier to consider the dependence of $G$ on the wire 
1184: length $L$ at zero temperature in order to see how well the numerical
1185: results compare with the RG expression given in Eq. (\ref{solvrg}). That
1186: equation has two parameters, namely, the interaction parameter $\alpha$
1187: and the short distance scale $d$ (with its corresponding transmission 
1188: probability $|t(d)|^2$). To begin, let us choose $\alpha = 0.0148$ as given 
1189: by the analytical expression in Eq. (\ref{rb3}) in terms of $U$ and $v_F$.
1190: In Fig. 14, we show a comparison between the expression in (\ref{solvrg}) for 
1191: two values of $d$, namely, $d=11$ with $|t(d)|^2 = 0.805$ (the dash dot line 
1192: A), and $d=51$ with $|t(d)|^2 = 0.795$ (the dashed line B), with our numerical
1193: results shown by astersisks. (The values of $|t(d)|^2$ for $d=$ 11 and 51 have
1194: themselves been obtained by our numerical calculations). We see that the RG 
1195: expression with $d=51$ fits the numerical results better than the expression 
1196: with $d=11$.
1197: This difference between the lines A and B shows that the interactions change 
1198: the conductance by an appreciable amount between $d=11$ and $d=51$, and this 
1199: change is not captured accurately by the expression in Eq. (\ref{solvrg}).
1200: [The conductance at short distances is affected substantially by quantities
1201: such as the large value of $\rho - {\bar \rho}$ at the site of the impurity. 
1202: The RG analysis, on which (\ref{solvrg}) is based, does not take such effects
1203: into account.] Since the wavelength of the Friedel oscillations is given by 
1204: $\lambda = 10$, we conclude that the RG results agree reasonably well with 
1205: the numerical results only if we begin the integration of the RG flows from a 
1206: distance which is significantly larger than $\lambda$. However, we observe that
1207: even line B corresponding to $d=51$ starts deviating from the numerical results
1208: at large length scales. We have therefore tried varying the parameter $\alpha$ 
1209: also, keeping $d$ and $|t(d)|^2$ fixed at $51$ and $0.795$ respectively. We 
1210: find that $\alpha = 0.016$ (the solid line C in Fig. 14) gives an excellent 
1211: fit to the numerical results.
1212: 
1213: In Fig. 15, we show the conductance as a function of $V$ for three different
1214: values of $U$, with $L=101$ and $\beta = 100$. We see that $G$ is not an even 
1215: function of $V$ in the presence of interactions.
1216: In Fig. 16, we show the conductance as a function of $U$ for $V=$ 1 and -1, 
1217: with $L=101$ and $\beta = 100$. For any value of $U$, we again find that 
1218: the conductance deviates less from the non-interacting value for $V=1$ 
1219: compared to $V=-1$. The explanation for this is similar to that for the 
1220: similar phenomenon in the spin-1/2 model, except that the roles of $U$ and 
1221: $-U$ are interchanged in the two models. The larger change in the conductance 
1222: for $V=-1$ is again due to the presence of a bound state.
1223: 
1224: \section{Discussion}
1225: 
1226: We have used the NEGF formalism to study the dependence of the 
1227: conductance of a quantum wire with both spin-1/2 and spinless electrons on 
1228: various parameters such as the wire length, temperature, impurity potential, 
1229: and the strength of the interactions between the electrons. The advantage of
1230: the NEGF formalism is that it can be used to compute the density and
1231: conductance at any length scale. At large length scales, our numerical 
1232: results agree with those obtained by an RG analysis of continuum theories. 
1233: We find that the trends of the RG results can be understood using the Born 
1234: approximation for scattering from a weak impurity and from the density 
1235: oscillations produced by the impurity. 
1236: 
1237: Our numerical results differ in detail in two ways from those obtained 
1238: analytically from the RG equations. Firstly, the dependence of $G$ on the wire
1239: length and temperature fits the expression in Eq. (\ref{solvrg}) only if we 
1240: take the starting point of the RG equation to be significantly larger than 
1241: the short distance scale $\lambda$, and we allow $\alpha$ to be a little 
1242: different from its analytically obtained value. Secondly, the conductance is 
1243: not an even function of the impurity potential $V$ (as one expects from the 
1244: RG analysis); this is due to the existence of a bound state for an attractive 
1245: impurity. These differences between the numerical results and 
1246: the results based on the RG analysis seem to be due to effects at short 
1247: distances, where the density deviates significantly from the mean density. 
1248: 
1249: Before ending, we would like to briefly compare our work with that of Refs. 
1250: \cite{enss,ander,barnabe}. Using the functional RG technique, the 
1251: authors of those papers have shown that there is excellent agreement at very 
1252: large length scales between the numerical results and the asymptotic scaling 
1253: forms given by the RG equations derived from continuum theories.
1254: We have not been able to go up to such large length scales using the NEGF
1255: formalism. However, even at the length scales studied by us, our numerical 
1256: results agree quite well with the RG equations if we start at a short 
1257: distance scale which is about 5 times larger than $\lambda$. 
1258: 
1259: As mentioned towards the beginning of Sec. V, the length scales $L$ and $L_T$ 
1260: that we have considered are comparable to those studied experimentally. Our 
1261: observations about the short distance effects may therefore have implications
1262: for fitting experimental data to expressions obtained by an RG analysis.
1263: 
1264: \vskip .5 true cm
1265: \centerline{\bf Acknowledgments}
1266: \vskip .5 true cm
1267: 
1268: We thank Supriyo Datta, Avik Ghosh and V. Ravi Chandra for many stimulating 
1269: discussions. We thank the Department of Science and Technology, India for 
1270: financial support under projects SR/FST/PSI-022/2000 and SP/S2/M-11/2000.
1271: 
1272: \begin{thebibliography}{99}
1273: 
1274: \bib{tarucha} S. Tarucha, T. Honda, and T. Saku, Sol. St. Comm. {\bf 94}, 413
1275: (1995).
1276: 
1277: \bib{liang} C. -T. Liang, M. Pepper, M. Y. Simmons, C. G. Smith, and D. A. 
1278: Ritchie, Phys. Rev. B {\bf 61}, 9952 (2000).
1279: 
1280: \bib{bkane} B. E. Kane, G. R. Facer, A. S. Dzurak, N. E. Lumpkin, R. G. Clark,
1281: L. N. Pfeiffer, and K. W. West, App. Phys. Lett. {\bf 72}, 3506 (1998).
1282: 
1283: \bib{yacoby} A. Yacoby, H. L. Stormer, N. S. Wingreen, L. N. Pfeiffer, K. W. 
1284: Baldwin, and K. W. West, Phys. Rev. Lett. {\bf 77}, 4612 (1996).
1285: 
1286: \bib{aus} O. M. Auslaender, A. Yacoby, R. de Picciotto, K. W. Baldwin, L. N.
1287: Pfeiffer, and K. W. West, Phys. Rev. Lett. {\bf 84}, 1764 (2000).
1288: 
1289: \bib{reilly} D. J. Reilly, G. R. Facer, A. S. Dzurak, B. E. Kane, R. G. Clark,
1290: P. J. Stiles, J. L. O'Brien, N. E. Lumpkin, L. N. Pfeiffer, and K. W. West, 
1291: Phys. Rev. B {\bf 63}, 121311(R) (2001).
1292: 
1293: \bib{kane} C. L. Kane and M. P. A. Fisher, Phys. Rev. B {\bf 46}, 15233 (1992).
1294: 
1295: \bib{furusaki} A. Furusaki and N. Nagaosa, Phys. Rev. B {\bf 47}, 4631 (1993).
1296: 
1297: \bib{lal1} S. Lal, S. Rao and D. Sen, Phys. Rev. Lett. {\bf 87}, 026801 (2001);
1298: Phys. Rev. B {\bf 65}, 195304 (2002).
1299: 
1300: \bib{datta} S. Datta, {\it Electronic transport in mesoscopic systems}
1301: (Cambridge University Press, 1995).
1302: 
1303: \bib{imry} Y. Imry, {\it Introduction to Mesoscopic Physics} (Oxford
1304: University Press, 1997).
1305: 
1306: \bib{chu} C. S. Chu and R. S. Sorbello, Phys. Rev. B {\bf 40}, 5941 (1989);
1307: Y. B. Levinson, M. I. Lubin, and E. V. Sukhorukov, Phys. Rev. B {\bf 45}, 
1308: 11936 (1992); E. Granot, Europhys. Lett. 68, 860 (2004).
1309: 
1310: \bib{yue} D. Yue, L. I. Glazman, and K. A. Matveev, Phys. Rev. B {\bf 49}, 1966
1311: (1994); K. A. Matveev, D. Yue, and L. I. Glazman, Phys. Rev. Lett. {\bf 71},
1312: 3351 (1993).
1313: 
1314: \bib{lal2} S. Lal, S. Rao and D. Sen, Phys. Rev. B {\bf 66}, 165327 (2002); S.
1315: Das, S. Rao and D. Sen, Phys. Rev. B {\bf 70}, 085318 (2004).
1316: 
1317: \bib{polyakov} D. G. Polyakov and I. V. Gornyi, Phys. Rev. B {\bf 68}, 035421
1318: (2003).
1319: 
1320: \bib{enss} T. Enss, V. Meden, S. Andergassen, X. Barnabe-Theriault, W. Metzner,
1321: and K. Sch\"onhammer, Phys. Rev. B {\bf 71}, 155401 (2005).
1322: 
1323: \bib{ander} S. Andergassen, T. Enss, V. Meden, W. Metzner, U. Schollw\"ock, and
1324: K. Sch\"onhammer, Phys. Rev. B {\bf 70}, 075102 (2004).
1325: 
1326: \bib{barnabe} X. Barnabe-Theriault, A. Sedeki, V. Meden, and K. Sch\"onhammer,
1327: Phys. Rev. B {\bf 71}, 205327 (2005), and Phys. Rev. Lett. {\bf 94}, 136405 
1328: (2005).
1329: 
1330: \bib{meden} V. Meden and U. Schollw\"ock, Phys. Rev. B {\bf 67}, 035106 (2003);
1331: V. Meden, W. Metzner, U. Schollw\"ock, O. Schneider, T. Stauber, and K. 
1332: Sch\"onhammer, Eur. Phys. J. B {\bf 16}, 631 (2000).
1333: 
1334: \bib{meir} Y. Meir and N. S. Wingreen, Phys. Rev. Lett. {\bf 68}, 2512 (1992). 
1335: 
1336: \bib{datta2} S. Datta, Superlattices and Microstructures {\bf 28}, 253 (2000).
1337: 
1338: \bib{dhar} A. Dhar and B. S. Shastry, Phys. Rev. B {\bf 67}, 195405 (2003).
1339: 
1340: \bib{tsukada} M. Tsukada, K. Tagami, K. Hirose, and N. Kobayashi, J. Phys. Soc.
1341: Jpn. {\bf 74}, 1079 (2005).
1342: 
1343: \bib{gogolin} A. O. Gogolin, A. A. Nersesyan, and A. M. Tsvelik, {\it
1344: Bosonization and Strongly Correlated Systems} (Cambridge University Press,
1345: Cambridge, 1998); S. Rao and D. Sen, in {\it Field Theories in Condensed
1346: Matter Physics}, edited by S. Rao (Hindustan Book Agency, New Delhi, 2001);
1347: T. Giamarchi, {\it Quantum Physics in One Dimension} (Oxford University Press,
1348: Oxford, 2004).
1349: 
1350: \bib{merzbacher} E. Merzbacher, {\it Quantum Mechanics} (John Wiley \& Sons,
1351: Singapore, 1999).
1352: 
1353: \bib{solyom} J. Solyom, Adv. Phys. {\bf 28}, 201 (1979).
1354: 
1355: \bib{ashcroft} N. W. Ashcroft and N. D. Mermin, {\it Solid State Physics} 
1356: (Holt, Rinehart and Winston, New York, 1976).
1357: 
1358: \end{thebibliography}
1359: 
1360: \end{document}
1361: 
1362: