1: \documentclass[prb,aps,amssymb,nofootinbib,floatfix,twocolumn,eqsecnum,showpacs]{revtex4}
2: \usepackage{graphicx,amsmath}% Include figure files
3: \newcommand{\be}{\begin{equation}}
4: \newcommand{\ee}{\end{equation}}
5: \newcommand{\bea}{\begin{eqnarray}}
6: \newcommand{\eea}{\end{eqnarray}}
7: \newcommand{\GG}{\mathcal{G}}
8:
9: \begin{document}
10: \preprint{First draft}
11:
12: \title{Fermi Liquid Theory for the Persistent Current
13: Past a Side-Coupled Quantum Dot}
14:
15: \author{Ian Affleck}
16: \affiliation{Department of Physics and Astronomy, University of British Columbia, Vancouver, British Columbia, Canada V6T 1Z1}
17:
18: \author{Erik~S.~S\o rensen}%
19: \affiliation{Department of Physics and Astronomy, McMaster University, Hamilton,
20: ON, L8S 4M1 Canada }
21:
22: % \altaffiliation[Also at ]{Physics Department, XYZ University.}%Lines break automatically or can be forced with \\
23: \date{\today}% It is always \today, today,
24: % but any date may be explicitly specified
25:
26:
27: \begin{abstract}
28: A Fermi Liquid theory is developed for the persistent current past a side coupled
29: quantum dot yielding analytical predictions for the behavior of the first two
30: harmonics of the persistent current as a function of applied magnetic flux.
31: The quantum dot is assumed weakly coupled to a ring of non-interacting electrons
32: and thus appropriately described as a Kondo impurity. The theory is valid at weak Kondo couplings in the regime where the
33: system size, $L$, is much larger than the size of the Kondo screening cloud, $\xi_K$.
34: The predictions of the Fermi Liquid theory are compared to exact diagonalization results for the persistent
35: current that lend support to the existence of a regime correctly described by this theory.
36: The finite temperature conductance, at $T\ll T_K$ is also calculated using Fermi liquid theory allowing
37: the definition of a ``Wilson ratio" relating the conductance and the persistent current.
38: \end{abstract}
39:
40: \pacs{72.10.Fk, 73.23.Ra, 72.15.Qm, 73.23.-b}% PACS, the Physics and Astronomy
41: % Classification Scheme.
42: %\keywords{Suggested keywords}%Use showkeys class option if keyword
43: %display desired
44: \maketitle
45: \section{Introduction}
46: The recent experimental observation of the Kondo effect and
47: related physical phenomena in
48: quantum dots,\cite{Goldhaber,GG,Cronenwett,Wiel} quantum corrals~\cite{Manoharan}
49: molecular electronics devices~\cite{Park,Liang}
50: and carbon nanotubes~\cite{Nygaard} as well as related nano-structures has led to significant renewed
51: interest in this fundamental effect. For a review of recent progress see Ref.~\onlinecite{KouwenGlazman}.
52: The Kondo effect in ordinary metals is usually associated with an increase in
53: the resistance at low temperatures but experiments performed on these
54: nano-structures~\cite{Goldhaber,Nygaard,Manoharan,Wiel} have shown that the Kondo effect
55: can also lead to an increase in the conductance that in some settings can reach the unitary limit ($G=2e^2/h$),
56: completely overcoming the Coulomb blockade. The fundamental and technological interest in the complete
57: understanding of this phenomenon is therefore considerable. The experiments are typically performed
58: on semi-conductor quantum dots where the electron
59: occupation number on the dot is controllable by a gate voltage. When the occupation
60: number is odd the quantum dot can act as a spin $s=1/2$ impurity screened by the electrons in
61: the leads. The screening of the impurity spin by the conduction electrons is, in some circumstances,
62: associated with the formation of a ``screening cloud" of size $\xi_K=v_F/T_K$ surrounding
63: the impurity.~\cite{Nozieres,Sorensen96,Barzykin96,Barzykin98,Cornaglia,Hallberg,Borda} Here $v_F$ is the Fermi velocity and $T_K$ the Kondo temperature.
64:
65: We note that the picture of a screening cloud of size
66: $v_F/\xi_K$ is valid in a one dimensional system and also
67: in an infinite two or three dimensional
68: system where only one channel (such as the s-wave) couples
69: most strongly to the Kondo impurity. However the behavior
70: may be rather different in a disordered
71: two or three dimensional box in which all states couple
72: strongly to the Kondo impurity. For the 3-dimensional
73: model considered in [\onlinecite{Thimm}], for instance, the effective
74: screening cloud is much smaller, $\propto T_K^{-1/3}$.
75: See also [\onlinecite{Hand}].
76:
77: The experimental observation of the screening cloud has proven elusive but recent
78: experiments~\cite{Wiel} have suggested that it might be observable in finite-size
79: properties of the conductance when the dot is attached to a single quantum wire
80: that forms an Aharonov-Bohm ring of length $L$. A natural extension of these experiments~\cite{Wiel} is to consider
81: not only the conductance but also finite-size effects observable in the persistent current generated in the
82: presence of a magnetic flux, $\Phi$.
83:
84:
85: The persistent current in quantum dot systems operated in the Kondo limit
86: has been the topic of a series of recent theoretical
87: studies.\cite{
88: ZvyaginJLT,
89: ZvyaginSchlottmann,
90: Ferrari,
91: Kang,
92: Cho,
93: Eckle,
94: SAprl,
95: SAprb,
96: Zvyaginprl,
97: Hu,
98: Aligia,
99: prl.persistent}
100: Usually, two fundamentally different geometries are considered;\cite{SAprb} the embedded quantum dot (EQD)
101: and the side-coupled quantum dot (SCQD) and it has been clearly established that the persistent current
102: has a strong dependence on the parity of the total number of electrons, $N$, in the system.
103: Here we shall exclusively be concerned with the SCQD.
104: Suppressing electron spin indices, the Hamiltonian is in this case:
105: \begin{equation}
106: H_{\rm SCQD}=-t\sum_{j=0}^{L-1}\left(c^{\dagger}_jc_{j+1}+{\rm h.c.}\right)+H_K,\ \ (c_L\equiv c_0)\label{eq:H}
107: \end{equation}
108: where $H_K=J_K{\vec S}\cdot c_0^{\dagger}
109: \frac{\vec\sigma}{2}c_0$, and
110: $c_{j\sigma}$ is the electron annihilation operator at site $j$ for spin $\sigma$
111: and $S^a$ are S=1/2 spin operators. In the following we set $t=1$. We always include the impurity spin as an electron in
112: our definition of $N$ but do not count the impurity site in our definition of $L$. Hence, at half-filling we have $N=L+1$.
113: The dimensionless magnetic flux, $\alpha\equiv e\Phi/c$ is introduced by appropriately changing the phases
114: of the hopping terms at the corresponding sites. (We set $\hbar =1$.)
115: The persistent current is given
116: by $j=-edE_0/d\alpha$ where $E_0$ is the ground-state energy. Despite the simplicity
117: of this geometry, a complete understanding of the behavior of the persistent current
118: both at half-filling and off half-filling has proved exceedingly difficult to achieve.
119: In previous work~\cite{SAprb,SAprl} by one of us and P. Simon it was predicted that,
120: $jL/(ev_F)\to 0$ when $\xi_K/L\to 0$ for both parities of $N$ in contrast to the results of Refs~\onlinecite{Eckle,Cho}.
121: However, subsequent work~\cite{prl.persistent} confirmed this prediction at half-filling but showed
122: both analytical and numerical evidence for additional terms off half-filling for $N$ even.
123:
124: In the present paper we develop a Fermi liquid theory (FLT) for the persistent current
125: past a SCQD, valid for $1\ll \xi_K\ll L$. The theory provides independent
126: confirmation of the above mentioned prediction at half-filling~\cite{SAprb,SAprl} as
127: well as additional analytical evidence for the additional terms off half-filling for $N$ even.
128: The theory yields detailed predictions about the finite-size $L$-dependence of the
129: first two harmonics of the persistent current, using a single parameter, $1/T_K$.
130: Section~\ref{sec:flt} outlines the FLT in considerable detail while section~\ref{sec:half}
131: presents our numerical results at half-filling and section~\ref{sec:offhalf} our numerical
132: results off half-filling. In Sec.~\ref{sec:conductance} we calculate the finite
133: temperature conductance past the quantum dot, at $T\ll T_K$.
134: This enables us to calculate a generalized ``Wilson ratio''
135: relating the persistent current for a finite ring at $T=0$
136: with the conductance at finite $T$ for the same quantum dot
137: side-coupled to infinite leads. This could be useful
138: in an experimental context since a prior infinite leads
139: measurement of the low $T$ conductance would fix
140: the precise numerical coefficient in the finite size
141: dependence of the persistent current at $T=0$.
142:
143: \section{Fermi Liquid Theory for the Persistent Current\label{sec:flt}}
144: Kondo physics at low energy scales and long length scales,
145: $T\ll T_K$, $L\gg \xi_K$,
146: can be described by an effective Hamiltonian not containing the
147: impurity spin, since it is screened by the conduction electrons
148: and the energy scale associated with breaking this spin singlet
149: is of $O(T_K)$. The low energy quasi-particles propogate freely
150: except that they obey a modified boundary condition which
151: can be thought of as arising because
152: they must stay in wave-functions orthogonal to that
153: of the screening electron. This modified boundary condition
154: is equivalent to a $\pi /2$ phase shift for even-channel
155: electrons (or for s-wave electrons
156: in the three dimensional, spherically symmetric version
157: of the model). In addition to the
158: modified boundary condition an interaction term is generated
159: in this effective Hamiltonian which is only non-vanishing
160: near the impurity location, at $|x|<\xi_K$. In a long
161: wavelength effective Hamiltonian this interaction appears right at
162: the origin. This interaction is
163: irrelevant in the renormalization
164: group sense; treating it in perturbation theory, its effects
165: become weaker and weaker at lower energies and longer lengths.
166: Building on earlier work by Wilson,~\cite{wilson} Nozi\`eres~\cite{Nozieres}
167: developed an elegant description of this physics, following
168: the Landau approach to Fermi liquid theory, by writing
169: an expression for the phase shift which depends on the
170: energy of the electron being scattered and also on the
171: local density of electons at the origin. In [\onlinecite{AL}]
172: this approach was transcribed into an effective Hamiltonian,
173: an approach which is simpler to deal with in some situations
174: and more in accord with modern renormalization group (RG)
175: techniques. (In a similar
176: way, the Landau approach to Fermi liquids can be recast in
177: terms of an effective Hamiltonian which can be studied
178: by RG methods. See, for example, [\onlinecite{shankar}].) Furthermore,
179: in [\onlinecite{AL}] it was shown how this
180: effective Hamiltonian could be uniquely determined, up to one
181: overall coupling constant with dimensions of inverse energy, $1/T_K$,
182: by using the one-dimensional (1D) formulation of
183: the Kondo model and 1D spin-charge separation. Essentially {\it all} low
184: energy long distance properties of the Kondo model can be
185: calculated by perturbation theory in this Fermi liquid interaction.
186: They are generally proportional to the first or second power
187: of the coupling constant, $1/T_K$ or $1/T_K^2$. It is not an
188: easy matter to determined the precise value of $T_K$ for
189: a given microscopic model. However, for a weak bare Kondo
190: coupling, $\nu J\ll 1$ (where $\nu$ is the density of states) it
191: is known that $T_K$ is exponentially small, $T_K\propto D\exp [-1/(J\nu )]$,
192: where $D$ is a cut-off or bandwidth scale. Furthermore, when the
193: bare coupling is weak, the physics is universal at all energy
194: scales $\ll D$. This implies, in particular, that
195: the ratios of various physical quantities calculated in
196: perturbation theory in $1/T_K$ are universal. The most famous
197: such universal ratio, between the impurity susceptibility
198: and specific heat, is known as the Wilson ratio. However, many
199: more such universal ratios can be readily calculated. In this
200: section we extend extend this approach to the persistent current.
201:
202: \subsection{The Boundary Conditions}
203: Since Fermi liquid theory applies at low energies and long distances,
204: it is convenient to
205: linearize the spectrum around
206: the Fermi points $\pm k_F$ and to introduce left (L) and right (R) moving
207: chiral fields:
208: \begin{equation}
209: c_x\sim e^{ik_F x}\psi_R(x)+e^{-ik_Fx}\psi_L(x).
210: \label{eq:2.10}
211: \end{equation}
212: The non-interacting part of the continuum Hamiltonian becomes:
213: \begin{equation} H_0=\int_{0}^L\left[ \psi_R^{\dagger}i\partial_x\psi_R
214: -\psi_L^{\dagger}i\partial_x\psi_L\right] .\end{equation}
215: We have set $v_F=1$.
216: The Kondo interaction becomes:
217: \begin{equation} H_K\to J_K\vec S\cdot [\psi^\dagger_R(0)+\psi^\dagger_L(0)]
218: {\vec \sigma \over 2} [\psi_R(0)+\psi_L(0)].\end{equation}
219: The strong coupling fixed point of the Kondo problem corresponds to
220: the boundary condition:
221: \begin{equation} c_0\propto \psi_L(0)+\psi_R(0)=0.\label{bc0}\end{equation}
222: We may think of the sceening electron as sitting at the origin, $j=0$
223: and the other electrons (low energy quasi-particles) must not
224: enter or leave the origin in order to avoid breaking up the Kondo singlet.
225: Considering the model defined on a ring, and ignoring, for the moment,
226: the magnetic flux, there is a second boundary condition at the
227: strong coupling fixed point:
228: \begin{equation} c_L\propto e^{ik_F L}\psi_R(L)+e^{-ik_FL}\psi_R(L)=0.\label{bcL}\end{equation}
229: This strong coupling fixed point corresponds to an open chain
230: of $L-1$ sites [$j=1,2,3,\ldots (L-1)$] with no impurity spin.
231: Initially we consider half-filling.
232: In this case, for such an open chain, $k_F=\pi /2$
233: regardless of the parity of $L$. Since the boundary conditions
234: of Eqs. (\ref{bc0} ,\ref{bcL}) are true at all times, while
235: $\psi_{R/L}$ is a function of $t\mp x$ only, it follows from Eq. (\ref{bc0}) that
236: we can regard $\psi_R(x)$, for $x>0$ as the analytic continuation of $\psi_L(x)$
237: to the negative $x$ axis:
238: \begin{equation} \psi_R(x)=-\psi_L(-x),\ \ (x>0).\label{ac1}\end{equation}
239: Likewise, Eq. (\ref{bcL}) implies that we can also regard
240: $\psi_R(x)$ as the analytic continuation of $\psi_L(x)$:
241: \begin{equation} \psi_R(L-x)=-e^{-2ik_FL}\psi_L(L+x).\ \ (x>0)\label{ac2}\end{equation}
242: Letting $x\to L$ in Eqs. (\ref{ac2}), we see that:
243: \begin{equation} \psi_L(2L)=e^{2ik_FL}\psi_L(0).\end{equation}
244: At half-filling this becomes:
245: \begin{equation} \psi_L(2L)=(-1)^{L}\psi_L(0),\label{pbc}\end{equation}
246: periodic for $L$ even and anti-periodic for $L$ odd. We
247: may formulate the model in terms of left-movers only with
248: these periodic or anti-periodic boundary conditions.
249:
250: \subsection{The Fermi Liquid Interaction with Zero Flux}
251: We now wish to write the Fermi liquid interaction in terms of
252: this left-moving formulation of the model. This must involve
253: the even channel fermions (even under $x\to L-x$) only since only the
254: even channel appears in the Kondo interaction. Furthermore,
255: it only involves this channel near $x=0$. Consider for example
256: the even channel fermions in the lattice model at a distance
257: of 1 site from the origin:
258: \begin{widetext}
259: \begin{equation} c_1+c_L \propto e^{ik_F}\psi_R(1)+e^{-ik_F}\psi_L(0)
260: +e^{ik_F(L-1)}\psi_R(L-1)+e^{-ik_F(L-1)}\psi_L(L-1).\end{equation}
261: \end{widetext}
262: We now use the facts that $k_F=\pi /2$, the continuum fields $\psi_L(x)$
263: and $\psi_R(x)$ are slowly varying on the lattice scale and
264: the boundary conditions of Eq. (\ref{bc0}, \ref{bcL}) to write this
265: in the purely left-moving formulation as:
266: \begin{equation} c_1+c_L\propto e^{-i\pi /2}\psi_L(0)+e^{-i\pi (L-1)/2}\psi_L(L).\label{psie}\end{equation}
267: We expect {\it only} this combination of fields to appear
268: in the Fermi liquid interaction, up to higher dimension operators
269: involving derivatives of the fermion fields. This follows from
270: observing, using the boundary conditions of
271: Eq. (\ref{bc0}, \ref{bcL}) and the
272: fact that the continuum fields vary slowly, that {\it any} non-vanishing even lattice fermion field
273: near the origin is proportional to this one.
274: For $j$ {\it even} and $j\ll \xi_K$ one finds:
275: \begin{equation}
276: c_j+c_{L-j}\approx 0.
277: \end{equation}
278: While for $j$ {\it odd} and $j\ll \xi_K$ we have:
279: \begin{eqnarray}
280: \lefteqn{c_j+c_{L-j}\propto e^{-i\pi j/2}\psi_L(0)+e^{-i\pi (L-j)/2}\psi_L(L)}&&\nonumber\\
281: &&=(-1)^{(j-1)/2}\left[e^{-i\pi /2}\psi_L(0)+e^{-i\pi (L-1)/2}\psi_L(L) \right],\nonumber\\
282: \end{eqnarray}
283: which is proportional to Eq.~(\ref{psie}).
284:
285: Once we have written the Kondo interaction in terms of
286: the even sector fermions only, it is convenient
287: to bosonize, introducing separate spin and charge bosons.
288: Importantly, the Kondo interaction involves
289: only the spin bosons in the even sector.
290: It
291: then follows that the Fermi liquid interactions
292: can be written in terms of these bosons only. The only possible
293: dimension 2 operator, in the spin sector, which respects the SU(2)
294: symmetry is $\vec J_e^2(0)$, the square of the spin
295: density operator. Going back
296: to the fermion representation:
297: \begin{equation}
298: \vec J_e(0)\equiv \psi^\dagger_e(0){\vec \sigma \over 2}
299: \psi_e(0).
300: \end{equation}
301: where, from Eq. (\ref{psie}),
302: \begin{equation} \psi_e(0)\propto {\psi_L(0)-e^{-i\pi L/2}\psi_L(L)\over \sqrt{2}}.\end{equation}
303: Thus we write the leading irrelevant Fermi liquid interaction,
304: for the case, $L$ even as:
305: \begin{eqnarray}
306: H_{int}&=&{-8\pi \over 3T_K}\vec J_e^2(0)\nonumber\\
307: &=&{-2\pi \over 3T_K}
308: \{[\psi_L(0)-e^{-i\pi L/2} \psi_L(L)]^\dagger {\vec \sigma \over 2}\nonumber\\
309: &\times&
310: [\psi_L(0)-e^{-i\pi L/2}\psi_L(L)]\}^2.
311: \label{Hint}
312: \end{eqnarray}
313: We have written the coupling constant in front of this operator
314: as $8\pi /3T_K$. It follows from standard scaling arguments
315: that this coupling constant should be of order the RG cross-over
316: scale. In general, it should be written as $1/T_K$ times
317: a dimensionless constant of O(1). All low energy propertites
318: of the system can be determined from this interaction term,
319: including the impurity susceptibility and specific heat,
320: for example. The sign in Eq. (\ref{Hint}) is known
321: to be the correct one since it gives the correct (positive)
322: sign for these two quantities. The factor of $8\pi/3$ is
323: purely a matter of a convention, or a precise definition
324: of what is meant by $T_K$. Unfortunately, a great number of
325: different conventions for $T_K$ are in current use. We have
326: chosen here to use the same convention as used by
327: Nozi\`eres~\cite{Nozieres} (who also
328: referred to $1/T_K$ as $\alpha$) and by Glazman and Pustilnik~\cite{GP}.
329: In Appendix B we briefly review other definitions of $T_K$ in popular
330: use and the constant factors relating them to each other.
331:
332: Strictly speaking, one other operator of dimension 2 is permitted,
333: $J_e(0)^2$, the square of the charge current, rather
334: than the spin current. As mentioned above, naively the charge sector
335: decouples from the Kondo interactions so that this operator
336: should apparently not be generated. However, irrelevant
337: operators (at the weak coupling fixed point) couple
338: charge and spin sectors together ultimately permitting this
339: operator to occur. However, the coefficient of this
340: operator is expected to be $O(1/D)$, where $D$ is the
341: bandwidth ($t$ in the tight-binding model), rather than $O(1/T_K)$.
342: In the weak coupling limit, where $T_K\ll D$, this other
343: interaction can be ignored.
344:
345: \subsection{The Fermi Liquid Interaction with Finite Flux}
346: So far we haven't mentioned the flux, $\Phi = c\alpha /e$.
347: The phases in the hopping terms corresponding
348: to any desired flux can be inserted anywhere in the ring. They
349: can be moved around freely by phase redefinitions of
350: the electron fields (gauge transformations).
351: For purposes of understanding the strong coupling
352: fixed point it is convenient to imagine that they
353: are inserted far from the impurity compared to $\xi_K$
354: so that the strong coupling boundary conditions, discussed
355: above, are unaffected by the flux. Of course, this
356: is only possible for $L\gg \xi_K$ but it is precisely
357: that limit which we are now considering. The Fermi
358: liquid theory only applies in that limit. It can
359: be seen that adding phases to hopping terms neccessarily
360: couples even and odd sectors together, since it
361: breaks parity. For this reason, it is more convenient
362: to go back to left and right movers and then ultimately
363: to left movers only on a ring of length $2L$, as
364: discussed above. It is also most convenient to
365: put the phase at the origin {\it after} establishing
366: the strong coupling b.c. This amounts to:
367: \begin{equation}
368: \psi_L(L)\to e^{-i\alpha}\psi_L(L).
369: \end{equation}
370: Hence the Ferm liquid interaction becomes:
371: \begin{eqnarray}
372: H_{int}
373: &=&{-2\pi \over 3T_K}
374: \{ [\psi_L(0)-e^{-i(\pi L/2+\alpha )} \psi_L(L)]^\dagger
375: {\vec \sigma \over 2}\nonumber\\
376: &\times&
377: [\psi_L(0)-e^{-i(\pi L/2+\alpha )} \psi_L(L)]\}^2.
378: \label{flt}
379: \end{eqnarray}
380:
381: \subsection{Perturbation Theory}
382: The remaining calculations are quite straightforward.
383: We simply do first order perturbation theory in $1/T_K$
384: for the groundstate energy, imposing periodic
385: or anti-periodic b.c.'s, Eq. (\ref{pbc}),
386: on the purely left-moving fermion fields.
387:
388: Now we turn to calculating the expectation value of
389: $H_{int}$. From Eq.~(\ref{flt}), recalling that
390: $L=N-1$ at half-filling, it is clear that it is advantageous to define
391: a shifted flux $\tilde\alpha$ in the following manner:
392: \begin{eqnarray}
393: \tilde\alpha &=& \alpha+\pi(N-1)/2 \ \ {\rm (N\ odd)}\nonumber\\
394: \tilde\alpha &=& \alpha+\pi N/2 \ \ {\rm (N\ even)}.
395: \label{eq:alpha}
396: \end{eqnarray}
397: [This definition can been seen largely as a way of defining the coefficient in
398: front of $\sin(\tilde\alpha)$ in the persistent current to be positive.]
399: Using this definition,
400: we first rewrite $H_{int}$ as:
401: \begin{widetext}
402: \begin{eqnarray}
403: H_{int}&=&{ \pi \over 6T_K}\bigl\{ e^{2i\tilde\alpha}[\psi^\dagger (L)\vec \sigma \psi (0)]^2
404: -2ie^{i\tilde \alpha} \psi^\dagger (L) \vec \sigma \psi (0)
405: \cdot\left[\psi^\dagger (0) \vec \sigma \psi (0)
406: +\psi^\dagger (L) \vec \sigma \psi (L)\right]
407: +h.c. + \ldots \bigr\}, \ \ {\rm (N\ even)}\nonumber \\
408: H_{int}&=&{ \pi \over 6T_K}\bigl\{- e^{2i\tilde\alpha}[\psi^\dagger (L)\vec \sigma \psi (0)]^2
409: +2e^{i\tilde \alpha} \psi^\dagger (L) \vec \sigma \psi (0)
410: \cdot\left[\psi^\dagger (0) \vec \sigma \psi (0)
411: +\psi^\dagger (L) \vec \sigma \psi (L)\right]
412: +h.c. + \ldots \bigr\}. \ \ {\rm (N\ odd)}\nonumber\\
413: \label{hint2}
414: \end{eqnarray}
415: \end{widetext}
416: Here we have dropped the $L$ subscripts and also dropped terms
417: which are independent of $\alpha$ and hence won't contribute
418: to the current in first order in $1/T_K$.
419: The first term can be rewritten:
420: \begin{equation}
421: H_{int}^{(1)}={\pi (-1)^N\over T_K}e^{2i\tilde\alpha}\psi^{\uparrow \dagger}(L)
422: \psi_{\uparrow}(0)\psi^{\downarrow \dagger}(L)
423: \psi_{\downarrow}(0)+ h.c.
424: \end{equation}
425: We now evaluate
426: the various terms using Wick's theorem. For this we need the
427: equal time propogator for a left mover with anti-periodic and periodic boundary conditions
428: on an interval of length $2L$.
429:
430: \subsubsection{$N$ Even at Half-filling}
431: For the case $N$ even, with anti-periodic
432: boundary conditions, this is:
433: \begin{eqnarray}
434: <\psi^{\dagger \alpha}(x) \psi_\beta (0)>&=&\delta^\alpha_\beta
435: {1\over 2L}\sum_0^\infty e^{-i\pi (m+1/2)x/L}\nonumber\\
436: &=&{-i
437: \delta^\alpha_\beta \over 4L\sin (\pi x/2L)}.
438: \label{prop}
439: \end{eqnarray}
440: In particular:
441: \begin{equation}
442: <\psi^{\dagger \alpha}(L) \psi_\beta (0)>={-i
443: \delta^\alpha_\beta \over 4L}.\label{prope}
444: \end{equation}
445: Of course,
446: \begin{equation}
447: <\psi^{\dagger \alpha}(L) \psi_\beta (0)>=
448: -<\psi_\beta (0) \psi^{\dagger \alpha} (L)>.\label{ob}
449: \end{equation}
450: We also need an expression for $<\psi^{\dagger\alpha} (0)\psi_\beta (0)>$.
451: The sum in Eq. (\ref{prop}) is ultraviolet divergent so we introduce
452: a (dimensionless) cut-off, an upper bound on the summation variable
453: $m$: $m < D$, giving:
454: \begin{equation}
455: <\psi^{\dagger\alpha} (0)\psi_\beta (0)>={\delta^{\alpha}_\beta D\over 2L}.
456: \label{D}
457: \end{equation}
458: We will see that our results for the current do not depend
459: on the value of $D$. It follows from PH symmetry that:
460: \begin{equation}
461: <\psi_\beta (0) \psi^{\dagger\alpha} (0)>={\delta^\alpha_\beta
462: D\over 2L}.\label{PH}
463: \end{equation}
464: Furthermore
465: \begin{equation}
466: <\psi_\beta (0) \psi^{\dagger\alpha}(0)>=
467: <\psi_\beta (L) \psi^{\dagger\alpha}(L)>, \label{0L}
468: \end{equation}
469: as follows from the reflection symmetry of the model which
470: takes $x\to L-x$.
471:
472: The $\cos 2\alpha$ term is:
473: \begin{equation}
474: E_0^{(2)}={-\pi \over T_K}e^{2i\tilde\alpha}\left( {-i\over 4L}\right)^2+h.c.
475: ={\pi \over 8T_KL^2}\cos 2\tilde\alpha ,
476: \end{equation}
477: With a current:
478: \begin{equation}
479: j^{(2)}={-edE_0/d\tilde\alpha}={e\pi \over 4T_KL^2}\sin 2\tilde\alpha\ (N\ {\rm even}).\label{2}
480: \end{equation}
481:
482: Now consider the $\cos \tilde\alpha$ term in the energy. This is actually zero
483: for $N$ even but let's go through some steps which are
484: useful also for $N$ odd. We use Wick's theorem to write this as:
485: \begin{eqnarray}
486: &&<\psi^\dagger (L)\vec \sigma \psi (0)\cdot
487: [\psi^\dagger (0)\vec \sigma \psi (0)+\psi^\dagger (L)\vec \sigma \psi (L)]>\nonumber\\
488: &&=2<\psi^\dagger (L)\vec \sigma \psi (0)>\cdot
489: <\psi^\dagger (0)\vec \sigma \psi (0)>\nonumber \\
490: &&+<\psi^{\alpha \dagger}(L)\psi_{\delta}(0)><\psi_\beta (0)
491: \psi^{\gamma \dagger}(0)>\vec \sigma^\beta_\alpha \cdot
492: \vec \sigma^{\delta}_\gamma \nonumber\\
493: &&+ <\psi^{\alpha \dagger}(L)\psi_{\delta}(L)><\psi_\beta (0)
494: \psi^{\gamma \dagger}(L)>\vec \sigma^\beta_\alpha \cdot
495: \vec \sigma^{\delta}_\gamma. \nonumber\\
496: \label{wick}
497: \end{eqnarray}
498: For $N$ even it is easy to see that the first term in Eq. (\ref{wick})
499: vanishes and the last two cancel using Eqs. (\ref{prope})-(\ref{0L}).
500:
501: \subsubsection{$N$ Odd at Half-filling}
502: Now consider $N$ odd. The strong coupling groundstate has
503: spin -1/2; we choose the state with total $S^z=1/2$. The
504: propogators are now different for spin up or down. We find:
505: \begin{equation}
506: <\psi^{\uparrow \dagger}(x)\psi_{\uparrow} (0)>=
507: {1\over 2L}\sum_{m=0}^\infty e^{-i\pi mx/L}={-i\over 4L}
508: {e^{i\pi x/2L}\over \sin (\pi x/2L)}.
509: \end{equation}
510: On the other hand, the $\downarrow$ propagator is the
511: same except that the $m=0$ term is omitted from the sum,
512: subtracting $1/2L$ to the final result. Thus:
513: \begin{equation}
514: <\psi^{\alpha \dagger}(L)\psi_\beta(0)>={(\sigma^z)^\alpha_\beta
515: \over 4L}.\label{propodd}
516: \end{equation}
517: Again we impose a cut-off so:
518: \begin{equation}
519: <\psi^{\uparrow \dagger}(0)\psi_\uparrow (0)>={D\over 2L}.
520: \end{equation}
521: Now we have:
522: \begin{equation}
523: <\psi_\uparrow (0) \psi^{\uparrow \dagger}(0)>={D-1\over 2L}.\label{PHoup}
524: \end{equation}
525: The $(-1)$ reflects the breaking of particle-hole (P-H) symmetry.
526: We have one unpaired spin up electron, sitting right
527: at the Fermi surface. Furthermore,
528: \begin{equation}
529: <\psi^{\downarrow \dagger} (0) \psi_{\downarrow }(0)>={D-1\over 2L}
530: \end{equation}
531: and
532: \begin{equation}
533: <\psi_\downarrow (0) \psi^{\downarrow \dagger}(0)>={D\over 2L}.\label{PHod}
534: \end{equation}
535: We may combine these as:
536: \begin{eqnarray}
537: <\psi^{\alpha \dagger} (0)\psi_\beta (0)>&=&{D-1/2\over 2L}
538: \delta^{\alpha}_{\beta} + {1\over 4L}(\sigma^z)^{\alpha}_{\beta}
539: \nonumber \\
540: <\psi_\alpha (0)\psi^{\beta \dagger} (0)>&=&{D-1/2\over 2L}
541: \delta_{\alpha}^{\beta} - {1\over 4L}(\sigma^z)^{\alpha}_{\beta}.
542: \end{eqnarray}
543: We now substitute this into Eq. (\ref{wick}). It is
544: easy to see that the terms containing $D$ all cancel, for
545: the same reasons that the entire expression vanishes for $N$ even.
546: The remaining terms are:
547: \begin{eqnarray}
548: && <\psi^\dagger (L)\vec \sigma \psi (0)\cdot
549: [\psi^\dagger (0)\vec \sigma \psi (0)+\psi^\dagger (L)\vec \sigma \psi (L)]>\nonumber\\
550: &&=\left({1\over 4L}\right)^2[2tr (\vec \sigma \sigma^z) \cdot
551: tr (\vec \sigma \sigma^z)-2tr (\vec \sigma \sigma^z\cdot \vec \sigma \sigma^z)]\nonumber\\
552: &&= \left({1\over 4L}\right)^2[8+4]\nonumber\\
553: &&={3\over 4L^2}.
554: \end{eqnarray}
555: Thus the $\cos \tilde\alpha$ term in the energy, for $N$ odd, is:
556: \begin{equation}
557: E^{(1)}_0={\pi \over 6T_K}2e^{i\tilde\alpha}{3\over 4L^2}+c.c.={\pi \over 2T_KL^2}
558: \cos \tilde \alpha .
559: \end{equation}
560: The term in the current is:
561: \begin{equation}
562: j^{(1)}={e\pi \over 2T_KL^2}\sin \tilde \alpha\ (N\ {\rm odd}).\label{eq:sin}
563: \end{equation}
564:
565: The $\sin (2\tilde\alpha)$ in the current for odd $N$ is given again by
566: Eq. (\ref{2}), with the opposite sign:
567: \begin{equation}
568: j^{(2)}=-{e\pi \over 4T_KL^2}\sin 2\tilde\alpha\ (N\ {\rm odd}).
569: \end{equation}
570: Note that the minus
571: sign which follows from spin up and down Green's functions having
572: opposite sign for $N$ odd, Eq. (\ref{propodd}), replaces
573: the minus sign due to the factors of $i$ in the Green's function
574: for $N$ even, Eq. (\ref{prope}), leaving only the minus
575: sign from Eq. (\ref{hint2}).
576:
577: We have so far set $v_F=1$. We summarize our results on the persistent current, for $\xi_K\ll L$
578: below, reinserting a factor of $v_F^2$ by dimensional analysis and replacing
579: $v_F/T_K$ by $\xi_K$:
580: \bea j_e &\to& {ev_F\over L}{\xi_K\over L}{\pi \over 4}\sin 2\tilde \alpha\nonumber \\
581: j_o &\to & {ev_F\over L}{\xi_K\over L}\left[ {\pi \over 2}\sin \tilde \alpha -{\pi \over 4}\sin 2\tilde \alpha
582: \right].\label{jfinal}\eea
583: It was argued earlier\cite{SAprl}, that both $j_e$ and $j_o$ can be written as $ev_F/L$ times
584: universal scaling functions of $\xi_K/L$ and $\tilde \alpha$. Eq. (\ref{jfinal}) indeed has that form.
585:
586: \subsubsection{Off Half-filling}
587: Now consider the case away from 1/2-filling, where P-H
588: symmetry is broken.
589: Here we refer to particle-hole symmetry breaking by
590: an amount of $O(1)$ not an amount of $O(1/L)$
591: as occurs, even at 1/2-filling for $N$ even.
592: Note that we continue to choose the reduced band to be symmetric around $k_F$:
593: $k_F-\Lambda \leq k \leq k_F+\Lambda$.
594: However, the
595: process of integrating out wave-vectors further away from $k_F$ is
596: not P-H symmetric, when $k_F$ does not have the P-H symmetric value,
597: so we expect to generate PH symmetry breaking terms
598: in $H_{int}$. The most important such term,
599: which can lead to a current of $O(1/L)$, is:
600: \begin{equation}
601: H_2 = -e^{i\pi (N-1)/2}\lambda e^{i \alpha} \psi^\dagger (L)\psi (0)
602: + h.c.
603: \label{H2}\end{equation}
604: where $\lambda$ is a real coupling constant. Here
605: the phase in this expression is determined by parity, i.e.
606: by the fact that
607: it arises from a term $-2\lambda \psi^\dagger_e\psi_e$ using
608: Eq. (\ref{psie}).
609: This term only occurs when P-H symmetry is broken. We expect $\lambda \propto (J_K/t)^2$
610: where $t$ is the original bandwidth.
611:
612: More generally, we
613: might start with a microscopic model which includes
614: second neighbor tunnelling past the quantum dot:
615: \begin{equation}
616: \delta H = -t_2(c_{L-1}^\dagger c_1+h.c.).\end{equation}
617: Such a term breaks P-H symmetry even at half-filling
618: and leads to a term in the low energy effective Hamiltonian
619: of the form of $H_2$ in Eq. (\ref{H2}) with $\lambda \propto t_2$.
620:
621: If the system is very close to half-filling, $n-1=\delta n\ll 1/\xi_K$,
622: and we begin with a pure Kondo model with no potential scattering
623: or direct tunnelling, then we can calculate the value of
624: $\lambda$, in $H_2$ directly from the Fermi liquid interaction
625: $H_{int}$ of Eq. (\ref{hint2}). To do this, it is convenient
626: to integrate out wave-vectors in the RG transformation symmetrically
627: around $k=\pi /2$ even thought this is no longer $k_F$, which has
628: the value:
629: \begin{equation} k_F=(\pi /2)(1+\delta n).\end{equation}
630: In this
631: way, $\lambda$ is not generated during the RG transformation,
632: and we obtain only the interaction $H_{int}$. However, the
633: low energy effective theory inherits an asymmetric cut-off:
634: \begin{equation} k_F-\Lambda -(\pi /2)\delta n < k < k_F-(\pi /2)\delta n+\Lambda .\end{equation}
635: We may now generate $H_2$ from $H_{int}$ by simply normal
636: ordering $H_{int}$. This follows since Eqs. (\ref{D}, \ref{PH})
637: are now modified to:
638: \bea <\psi^{\dagger \alpha}(0)\psi_\beta (0)>&=&
639: \delta^\alpha_\beta \left[ {D\over 2L}+{\delta n\over 4}\right]
640: \nonumber \\
641: <\psi_\beta (0)\psi^{\dagger \alpha}(0)>&=&
642: \delta^\alpha_\beta \left[ {D\over 2L}-{\delta n\over 4}\right]
643: \label{noPH}\eea
644: We now write $H_{int}$ as a normal ordered part plus
645: a correction of the form of $H_2$. This follows using:
646: \begin{widetext}
647: \bea &&\psi^\dagger (L)\vec \sigma \psi (0)\cdot \left[ \psi^\dagger (0)\vec \sigma \psi (0)
648: +\psi^\dagger (L)\vec \sigma \psi (L)\right]
649: =:\psi^\dagger (L)\vec \sigma \psi (0)\cdot \left[ \psi^\dagger (0)\vec \sigma \psi (0)
650: +\psi^\dagger (L)\vec \sigma \psi (L)\right] :
651: \nonumber \\
652: &+&\left[ \psi^{\alpha \dagger}(L)\psi_{\delta}(0)<0|\psi_\beta (0)\psi^{\gamma \dagger}(0)|0>
653: -\psi^{\gamma \dagger}(L)\psi_\beta (0)<0|\psi^{\alpha \dagger}(L)\psi_\delta (L)\right]
654: (\vec \sigma^\beta_\alpha \cdot \vec \sigma^\delta_\gamma ).\label{normord}\eea
655: \end{widetext}
656: Using Eq. (\ref{noPH}) we see that the two terms in the second line of Eq. (\ref{normord})
657: don't cancel, away from half-filling, yielding instead:
658: \begin{widetext}
659: \begin{equation} \psi^\dagger (L)\vec \sigma \psi (0)\cdot \left[ \psi^\dagger (0)\vec \sigma \psi (0)
660: +\psi^\dagger (L)\vec \sigma \psi (L)\right]
661: =:\psi^\dagger (L)\vec \sigma \psi (0)\cdot \left[ \psi^\dagger (0)\vec \sigma \psi (0)
662: +\psi^\dagger (L)\vec \sigma \psi (L)\right] :-{3\delta n\over 2}\psi^\dagger (L)\psi (0).
663: \end{equation}
664: \end{widetext}
665:
666: Inserting this expression into Eq. (\ref{hint2})
667: gives a quadratic term of the form of $H_2$ with:
668: \begin{equation} \lambda = {\pi \delta n\over 2T_K},\label{lambdan}\end{equation}
669: for $N$ even.
670:
671: We may now evaluate the additional terms in the current
672: arising from $H_2$, using perturbation
673: theory in $H_{int}$. We only discuss first order perturbation
674: theory.
675: We evaluate $<H_2>$ using Eq. (\ref{prope}). This gives an
676: additional current:
677: \begin{equation}
678: \delta j = (e\lambda /L)\sin \tilde \alpha ,\label{jPH}
679: \end{equation}
680: for $N$ even. For $N$ odd, there
681: is no extra contribution to first order in $\lambda$ since
682: $<\psi^{\dagger \alpha}(L)\psi_\beta (0)>\propto (\sigma^z)^\alpha_\beta$.
683: Hence, to first order, we expect that in the limit of small $\delta n$ the current for $N$ odd,
684: $j_o$, remain unchanged with respect to it's value at half-filling.
685:
686: Note however, that we have only evaluated the contributions
687: of $H_{int}$ to the current in first order perturbation theory in
688: $H_2$. Thus we haven't ruled out the possibility of
689: a term of $O(1/L)$ in the current for $N$ odd, with broken PH symmetry.
690:
691: Adding the universal result for the PH symmetric case to this
692: correction from PH symmetry breaking (and reinserting the factor of $v_F$ previously set to one) gives a persistent current,
693: for $N$ even:
694: \begin{equation} j_e = {ev_F\over L}\left[{\pi \xi_K\over 4L} \sin (2\tilde \alpha )+\lambda \sin (\tilde \alpha )
695: \right].\end{equation}
696: The non-universal term arising from PH symmetry breaking, proportional to $\lambda$ will always dominate at sufficiently
697: large $L$. However, provided that the dimensionless coupling, $\lambda \ll 1$,
698: as we expect for small bare Kondo coupling and not too strong direct tunnelling
699: across the quantum dot, the first, universal, term will dominate over
700: the range of lengths:
701: \begin{equation} \xi_K\ll L \ll \xi_K/\lambda .\end{equation}
702:
703: The presence of a term in $j_e$ proportional to $\sin(\tilde\alpha)$ in the absence of PH symmetry
704: was shown in Ref.~\onlinecite{prl.persistent} using strong coupling perturbation theory. The above derivation
705: provides further independent analytical evidence for such a term.
706:
707: \section{Numerical Results\label{sec:numres}}
708: The Fermi liquid theory results for the persistent current developed in the previous section are parameterized
709: in terms of $T_K$ and should be valid in the regime $T\ll T_K$, $1\ll \xi_K\ll L$.
710: For a useful numerical test of these results it is therefore necessary to study
711: ratios of the Fourier components of the persistent current that become {\it independent} of $T_K$ and
712: can be tested at $T=0$ using exact diagonalization (ED) methods. Furthermore, in order for $1\ll \xi_K$ to hold,
713: $J_K$ should not be too large. A more severe constraint is that
714: the system sizes should satisfy $L\gg\xi_K$. With an exponentially diverging $\xi_K$ at small $J_K$
715: we see that we quickly leave the regime of validity of Fermi liquid theory as $J_K\to 0$ for the
716: rather modest systems sizes we can treat using ED. Clearly, the interesting regime is then intermediate values
717: for $J_K$.
718: Our approach is then to calculate the different
719: Fourier components, $a_n$, of the persistent current, $j$, as a function of the flux, $\alpha$.
720: Our convention for the
721: Fourier components are $j(\tilde\alpha)=\sum_n a_n \sin(n\tilde\alpha)$.
722: Note that, a very precise determination of
723: the $\alpha$ dependence is necessary in order to determine the Fourier components reliably and we typically
724: use 200 values for $\alpha$.
725: The numerical results can then be compared to
726: the predictions of the Fermi liquid theory as well as results previously obtained using strong
727: coupling perturbation.\cite{prl.persistent}
728:
729: \subsection{At Half-filling\label{sec:half}}
730: We start by discussing our numerical ED results obtained at half-filling
731: with systems sizes of $L=12-15$. Due to the lack of symmetry in these systems it is difficult to reach
732: larger sizes and the largest system $(L=14,N=15)$ required a diagonalization of a matrix of size
733: $20,796,633$ for each value of $\alpha$ after symmetry reductions.
734: \begin{figure}[t]
735: \begin{center}
736: \includegraphics[clip,width=8cm]{Fig1.eps}
737: \caption{The ratio, $a_2^o/a_1^o$, of the two first Fourier coefficients of the current for
738: the SCQD at 1/2-filling, for $N$ odd. $\circ$ indicates ED results with $L=12$ and $N=13$ while
739: $\bullet$ indicates ED results with $L=14$ and $N=15$. The dashed line (barely visible) is the strong coupling
740: perturbative result, Eq.~(\ref{eq:stronghalf}), for this ratio with $L=12$, the solid line is $L\to\infty$
741: the strong coupling perturbative result, Eq.~(\ref{eq:thermostronghalf}), the dashed dotted line the fermi liquid
742: result, Eq.~(\ref{eq:flthalf}), and the dotted line
743: the ideal ring result, Eq.~(\ref{eq:zerohalf}).
744: \label{fig:HalfRatOddOdd}}
745: \end{center}
746: \end{figure}
747:
748: In Ref~\onlinecite{prl.persistent} strong coupling perturbative results were given for the
749: persistent current at half-filling past the SCQD. It was shown that,
750: for odd or even $N$,
751: \begin{eqnarray}
752: j_oL/e&\approx &\frac{32}{9J_K^2}[
753: \tan (\frac{\pi}{2L})+
754: \tan (\frac{3\pi}{2L})
755: ]\sin \tilde \alpha
756: \nonumber \\
757: &&+\frac{128}{3J_K^3L}[2\sin \tilde \alpha -\sin (2\tilde \alpha )]
758: \nonumber \\
759: j_eL/e&\approx &\frac{32}{3J_K^3L}[1+1/\cos (\pi /L)]^2\sin 2\tilde\alpha .
760: \label{eq:jlarge}
761: \end{eqnarray}
762: As elsewhere, $L$ does {\it not} include
763: the impurity site while $N$ {\it does} include the impurity electron.
764: The definition of $\tilde\alpha$ is also the same: For
765: $N$ odd $\tilde\alpha = \alpha+\pi(N-1)/2$ and for $N$ even $\tilde\alpha = \alpha+\pi N/2$.
766: If we concentrate on the first two Fourier coefficients of the current we see
767: from the above strong coupling results of Eq.~(\ref{eq:jlarge}) that:
768: \begin{eqnarray}
769: \frac{a_2^o}{a_1^o}
770: &=&\frac{-12}{J_K L[\tan({\pi}/{2L})+\tan({3\pi}/{2L})]+24}\nonumber\\
771: \frac{a_2^o}{a_2^e}
772: &=&\frac{-4}{(1+{1}/{\cos({\pi}/{L^e})})^2}\frac{L^{e}}{L^{o}}.
773: \label{eq:stronghalf}
774: \end{eqnarray}
775: Here the superscripts o and e refer to $N$ odd or even, respectively.
776: In the limit $L\to\infty$ we see that the strong coupling expansion gives:
777: \begin{equation}
778: \frac{a_2^{o}}{a_1^{o}}
779: \to\frac{-6}{\pi J_K+12}, \ \ \frac{a_2^{o}}{a_2^{e}}\to -1.
780: \label{eq:thermostronghalf}
781: \end{equation}
782: The Fermi liquid theory developed in the preceeding section,
783: valid for $1\ll \xi_K\ll L$,
784: predicts, using Eqs.~(\ref{2}) and (\ref{eq:sin}):
785: \begin{equation}
786: \frac{a_2^{o}}{a_1^{o}}
787: =-\frac{1}{2},\ \
788: \frac{a_2^{o}}{a_2^{e}}
789: =-1.\label{eq:flthalf}
790: \end{equation}
791: \begin{figure}[t]
792: \begin{center}
793: \includegraphics[clip,width=8cm]{Fig2.eps}
794: \caption{The ratio, $a_2^o/a_2^e$, of the second Fourier coefficients of the current for
795: the SCQD at 1/2-filling, for $N$ odd and $N$ even. $\circ$ indicates ED results with $L=11$, $N=12$ and $L=12$, $N=13$ while
796: $\bullet$ indicates ED results with $L=13$, $N=14$ and $L=14$, $N=15$. The dashed line is the strong coupling
797: perturbative result, Eq.(\ref{eq:stronghalf}), for this ratio with $L^{\rm even}=11, L^{\rm odd}=12$, the dashed dotted line the fermi liquid
798: result, Eq.~(\ref{eq:flthalf}), and the dotted line
799: the ideal ring result, Eq.~(\ref{eq:zerohalf}).
800: \label{fig:HalfRatOddEven}}
801: \end{center}
802: \end{figure}
803:
804: When $J_K$ is exactly zero we obtain the behavior of an ideal ring, in this case the current has a
805: charateristic saw-tooth like form~\cite{ProkofevGogolin,SAprl} where the same Fourier coefficients
806: can be trivially calculated. One finds at $J_K=0:$
807: \begin{equation}
808: \frac{a_2^{o}}{a_1^{o}}
809: =\frac{1}{2},\ \
810: \frac{a_2^{o}}{a_2^{e}}
811: =1.
812: \label{eq:zerohalf}
813: \end{equation}
814: Note that this result is dramatically different from the result of Eq.~(\ref{eq:flthalf}),
815: valid for $1\ll \xi_K\ll L$. The ratio of the Fourier coefficients appear discontinuous in the limit $J_K\to 0$.
816:
817: We begin by discussing our results for ${a_2^o}/{a_1^o}$ which are shown in
818: Fig.~\ref{fig:HalfRatOddOdd} for system sizes of $L=12$ and $L=14$, for a range
819: of different $J_K$. The theoretical strong coupling result of
820: Eq.~(\ref{eq:stronghalf}) for $L=12$ is shown as the dashed line and differs
821: only slightly from the $L\to\infty$ result of Eq.~(\ref{eq:thermostronghalf})
822: shown as the solid line. The numerical results agrees with the strong
823: coupling results once $J_K>10$. In the opposite limit, $J_K\to 0$, we see
824: that the numerical results quickly approach the ideal ring result of
825: Eq.~(\ref{eq:zerohalf}), shown as the dotted line. This is very reasonable,
826: since the very limited system sizes of $L=12,14$ certainly no longer
827: satisfies the equality $L\gg\xi_K$ for Fermi liquid theory to be valid once
828: $\xi_K$ diverges as $J_K\to 0$. The interesting region is therefore the
829: intermediate coupling region, $0.1<J_K<10$. A very rapid crossover from the
830: ideal ring result of 1/2, at small $J_K$, to a negative value is observed, consistent with an
831: exponentially diverging $\xi_K$. Furthermore, ${a_2^o}/{a_1^o}$ develops a
832: pronounced plateau (``dip") around $J_K\sim 4$,
833: reaching negative values, even for these
834: rather modest system sizes. As we increase the system size from $L=12$ to $L=14$ this
835: ``dip" becomes slightly more pronounced.
836: This implies that the cross-over between the ideal ring result for this ratio of 1/2, at small $J_K$, to
837: the $J_K\to\infty$ result of 0, {\it can not} be monotonic even in the thermodynamic limit.
838: We take this to be indicative of the validity of
839: the Fermi Liquid theory. We expect that numerical results for
840: ${a_2^o}/{a_1^o}$ in the thermodynamic limit would roughly follow the strong
841: coupling result given by Eq.~(\ref{eq:thermostronghalf}) (the solid line in
842: Fig.~\ref{fig:HalfRatOddOdd}) out to $J_K\sim 0.1$ and then attain the value -1/2
843: rather quickly, jumping discontinuously for $J_K=0$, in accordance with the
844: Fermi Liquid theory.
845: In the subsequent section, where we discuss our results
846: away from half-filling, we therefore exclusively focus on the value of
847: $J_K=4$ where the ``dip" occurs, since this would appear to be the most promising value for $J_K$ for
848: observing FLT behavior with the available system sizes.
849:
850: Results for $a_2^o/a_2^e$ are shown in Fig.~\ref{fig:HalfRatOddEven}. In this
851: case the behavior as a function of $J_K$ is monotonic and the evidence for a
852: regime described by FLT perhaps less obvious. The dotted, dashed and
853: dashed-dotted lines are the ideal ring results of Eq.~({\ref{eq:zerohalf}), the
854: strong coupling result of Eq.~(\ref{eq:stronghalf}) and the FLT result of
855: Eq.~(\ref{eq:flthalf}). Again an extremely rapid crossover is seen at
856: intermediate values of $J_K$. Furthermore, as the system size is increased
857: from $L=12$ to $L=14$ this crossover moves to smaller values of $J_K$
858: approaching the predictions of Fermi Liquid theory.
859:
860:
861: \subsection{Off Half-filling\label{sec:offhalf}}
862: \begin{figure}[t]
863: \begin{center}
864: \includegraphics[clip,width=8cm]{Fig3.eps}
865: \caption{The first Fourier coefficient scaled by $L$, $a_1^e L$, of the current for
866: the SCQD away from 1/2-filling, for $N$ even as a function of $\delta n=(N-1)/L-1$. In all cases $J_K=4.0$. $\circ$ indicates
867: ED results with $L=13$ and $N=14,12,10,8,6,4$, $\triangle$ indicates ED results with $L=14$ and $N=14,12,10,8,6,4$, finally
868: $\bullet$ indicates ED results with $L=15$ and $N=16,14,12,10,8,6$. The lines are guides to the
869: eye.
870: \label{fig:a1Nevenoffhalf}}
871: \end{center}
872: \end{figure}
873: Next, we turn to a discussion of our results for off half-filling.
874: With our definitions of $N$ and $L$ we define, as elsewhere, $\delta n = (N-1)/L -1$.
875: A typical Hilbert space size is $64,414,350$ for $(L=15,N=14)$ after symmetry reductions.
876: In Ref~\onlinecite{prl.persistent} strong coupling perturbative results for
877: the current past the SCQD away from half-filling were also given. At large
878: $J_K$ it was found that for $N$ odd and $N$ even:
879: \begin{eqnarray}
880: \lefteqn{j_oL/e\approx\frac{32}{9J_K^2}\sin (\tilde \alpha )}&&\nonumber\\
881: &&\times\left[\sin \frac{\pi (N-1)}{2L}\tan \frac{\pi}{2L}
882: -\sin \frac{3\pi (N-1)}{2L}\tan \frac{3\pi}{2L}\right]\nonumber\\
883: \lefteqn{j_eL/e \approx \frac{32}{9J_K^2} \sin (\tilde \alpha )}&&\nonumber\\
884: &&\times\left[ \frac{\cos (\pi (N-1)/(2L))}{\cos (\pi /(2L))}-
885: \frac{\cos (3\pi (N-1)/(2L))}{\cos (3\pi/(2L))}\right]\nonumber\\
886: \label{eq:scqdoff}
887: \end{eqnarray}
888: This results predicts that as $L\to\infty$, $\delta n \to 0$, $j_oL/e$ is changed
889: from the value at half-filling, Eq.~(\ref{eq:jlarge}), only by a term proportional
890: to $-(\delta n)^2\sin(\tilde\alpha)/(J_K^2L)$. If we only work to linear order in $\delta n$
891: this correction can be neglected.
892: However in this limit, $j_e L/e\sim \sin(\tilde\alpha)64\pi\delta n/9 J_K^2$, that is, a $\sin(\tilde\alpha)$ term proportional
893: to $\delta n$. At half-filling this term is clearly 0 for $j_e$, and the $\sin(2\tilde\alpha)$ term
894: in Eq.~(\ref{eq:jlarge}) dominates.
895: We note that this limit of the strong coupling results agrees with the previously
896: developed Fermi liquid theory that, in the limit $\delta n\to 0$, predicted no change in $j_o$
897: but the generation of a $\sin(\tilde\alpha)$ term proportional to $\delta n/L$ for $j_e$,
898: Eq.~(\ref{lambdan}), (\ref{jPH}). In the following, we therefore focus exclusively on the first Fourier
899: coefficient, $a_1$, determining its dependence on $\delta n$.
900: \begin{figure}[t]
901: \begin{center}
902: \includegraphics[clip,width=8cm]{Fig4.eps}
903: \caption{The first Fourier coefficient scaled by $L^2$, $a_1^o L^2$, of the current for
904: the SCQD away from 1/2-filling, for $N$ odd as a function of $\delta n=(N-1)/L-1$. In all cases $J_K=4.0$. $\circ$ indicates
905: ED results with $L=12$ and $N=13,11,9,7,5,3$,
906: $\triangle$ indicates ED results with $L=13$ and $N=13,11,9,7,5,3$, finally
907: $\bullet$ indicates ED results with $L=14$ and $N=15,13,11,9,7,5$. The lines are guides to the
908: eye.
909: \label{fig:a1Noddoffhalf}}
910: \end{center}
911: \end{figure}
912:
913: Our results for $a_1^e L$ for $N$ even are shown in Fig.~\ref{fig:a1Nevenoffhalf} for system sizes $L=13,14,15$
914: for a range of $\delta n$. All results are for an intermediate coupling of $J_K=4$, the most promising
915: coupling to show clear indications of FLT behavior for our limited system sizes. Two conclusions are
916: immediately evident; since the results fall on a single curve this term in the current is indeed proportional
917: to $1/L$ as predicted by theory. Secondly, for small $\delta n$, $a_1L$ increases approximately linearly with $\delta n$,
918: again consistent with the theoretical predictions. At larger $\delta n$ the results strongly deviate from linear
919: behavior as one would expect.
920:
921: Finally, we show results for $a_1^o L^2$ for $N$ odd in Fig.~\ref{fig:a1Noddoffhalf} for system sizes of $L=12,13,14$,
922: for a range of $\delta n$. The results are again for an intermediate coupling of $J_K=4$. As was the case for $a_1^e L$,
923: the results for different system sizes again follow a single curve, validating the scaling $a_1^o\sim 1/L^2$ predicted
924: by theory. A term proportional to $1/L$ can definitely be excluded for this Fourier coefficient.
925: We also see that, for $\delta n\to 0$, our results are consistent with deviations from the result at half-filling
926: being of higher than linear order in $\delta n$, although a definite conclusion is hard to obtain due to the limited
927: system sizes available.
928:
929: \section{Finite Temperature Conductance\label{sec:conductance}}
930: We now
931: calculate the conductance for a quantum dot, side-coupled
932: to infinitely long leads,
933: at a low finite temperature, $T<<T_K$.
934: We do this using our Fermi liquid Hamiltonian, of Eq. (\ref{flt}). We choose $L$ and $L/2$
935: even, for convenience but we are now taking the $L\to \infty$ limit
936: so this choice is immaterial. Working with left-movers only, and taking into account
937: that the point $x=L$ corresponds to $x=0^-$, $x=0$ to $0^+$, we write the Fermi liquid
938: interaction as:
939: \bea H_{int}={-8\pi \over 3T_K}:[\psi^\dagger
940: {\vec \sigma \over 2} \psi]^2:
941: -{4i\over 3 T_K}:\psi^\dagger
942: \overleftrightarrow{d\over dx}\psi :
943: +\hbox{const}.\label{FLHC}\eea
944: Here
945: \begin{equation} \psi \equiv {1\over \sqrt{2}}\left[\psi_L(0^+)-\psi_L(0^-)e^{-i\alpha}
946: \right],\label{psidef}\end{equation}
947: the $:\ :$ denotes normal ordering and we define:
948: \begin{equation} f(x)\overleftrightarrow{d\over dx}g(x)\equiv f{dg\over dx}-
949: {df\over dx}g.\end{equation}
950: The second, derivative term in Eq. (\ref{FLHC}), arises from
951: a point splitting procedure. The second term is referred to as the elastic
952: part of the interaction, corresponding to a single electron impuirity scattering
953: process, while the first term is referred to as the inelastic part,
954: corresonding to an electron-electron interaction at the origin.
955: (For a derivation of the Fermi liquid interaction
956: in this form, see [\onlinecite{AL}], but note that the fermion fields are
957: defined with an unconventional normalization there so that they are larger by
958: a factor of $\sqrt{2\pi}$.)
959: Now the phase $\alpha$, is regarded as a time-dependent vector potential, related to the
960: potential energy drop across the quantum dot by:
961: \begin{equation} d\alpha /dt= -\int_{0^-}^{0^+} dx E(x)=\Delta V.\label{V}\end{equation}
962: (In this section we set the electron charge, $e=1$, reinstating it
963: at the end.)
964: The conductance through a quantum dot, treating the leads as ballistic, is insensitive
965: to where the electric field is applied so we have chosen, for convenience, to
966: apply it right at the junction, over a region of width of order $\xi_K$.
967:
968:
969: We write the current operator as the rate of change of one-half the
970: difference of the total number of electrons on
971: the right hand side ($x>0$) of the system minus the total number on the left hand side:
972: \begin{equation} I= (1/2)d\hat (N_+-\hat N_-)/dt = (1/2)i[H,\hat N_+-\hat N_-],\end{equation}
973: where
974: \begin{equation} \hat N_+\equiv \int_0^\infty [\psi^\dagger_L(x)\psi_L(x)+\psi^\dagger_R(x)\psi_R(x)],\end{equation}
975: and similarly for $\hat N_-$.
976: As discussed above, we can use the perfectly reflecting strong-coupling b.c. of Eqs. (\ref{bc0}),
977: (\ref{bcL}))
978: to rewrite the left and right movers at $x>0$ in terms of left-movers only on the entire real line.
979: Thus we define:
980: \bea \psi_{L+}(x) &\equiv& \psi_L(x),\ \ (x>0)\nonumber \\
981: &=& \psi_R(-x) \ \ (x<0).\eea
982: Here the subscript $+$ indicates that this field derives from the $x>0$ region. Similarly,
983: we define another left-moving field on the entire real line, $\psi_{L-}(x)$ which
984: derives from the region $x<0$:
985: \bea \psi_{L-}(x) &\equiv& \psi_L(x),\ \ (x<0)\nonumber \\
986: &=& \psi_R(-x) \ \ (x>0).\eea
987: In this notation:
988: \begin{equation} \hat N_\pm = \int_{-\infty}^\infty \psi^\dagger_{L\pm}(x)\psi_{L\pm}(x)\end{equation}
989: and the operator appearing in the Fermi liquid interaction,
990: defined in Eq. (\ref{psidef}) becomes:
991: \begin{equation} \psi \equiv{1\over \sqrt{2}}\left[\psi_{L+}(0)-\psi_{L-}(0)e^{-i\alpha}
992: \right].\end{equation}
993:
994: We wish to calculate $<I>$ in linear response to a time dependent
995: vector potential $\alpha (t)$. At this point, there are two
996: ways of proceeding. We may use the expression $I=(1/2)d(\Delta \hat N)/dt$
997: where
998: \begin{equation} \Delta \hat N\equiv \hat N_+-\hat N_-,\end{equation}
999: and ultimately relate the conductance to the Green's function
1000: of $\Delta \hat N$, or we may calculate explicitly
1001: the commutator $[H,\Delta \hat N] $ and relate the conductance to
1002: the Green's function of that operator. In both
1003: approaches the calculation must be carried out to second order
1004: in the Fermi liquid interaction and the amount of work is
1005: roughly the same either way. However, it is actually much
1006: more convenient to work with $d \Delta \hat N/dt$ because the
1007: needed Green's function can be related to the ${\cal T}$-matrix,
1008: which was calculated previously.\cite{Nozieres,AL} This
1009: approach was used in [\onlinecite{GP}] for instance, to
1010: calculate the conductance through an {\it embedded} quantum dot.
1011: The calculation is very similar in the side-coupled case and
1012: we may obtain the answer by only a slight modification of
1013: their result. This approach is used in the rest of
1014: this section. It is also of some interest to calculate
1015: the conductance using the commutator method and this is
1016: done in Appendix A, yielding precisely the same result.
1017:
1018: In the $d\Delta \hat N/dt$ method, we use the fact that, to $O(\alpha )$,
1019: $H=H(0)+\alpha (t)I$, to obtain the conductance:
1020: \begin{equation} C=\lim_{\omega \to 0} {1\over \omega}\int_0^\infty dt e^{i\omega t}<[I(t),I(0)]>.\end{equation}
1021: Integrating by parts:
1022: \begin{equation} C=(1/4)\lim_{\omega \to 0}\omega \int_0^\infty d\omega e^{i\omega t}
1023: <[\Delta \hat N(t),\Delta \hat N(0)]>,\end{equation}
1024: where the retarded Green's function is evaluated at $\alpha =0$. It is now
1025: convenient to go to the even-odd basis:
1026: \begin{equation} \psi_{e/o}={1\over \sqrt{2}}(\psi_{L+}\mp \psi_{L-}).\end{equation}
1027: Since the interactions only involve $\psi_e$, $C$ factorizes into
1028: a free Green's function for $\psi_o$ multiplied by the non-trivial Green's function
1029: for $\psi_e$. This latter Green's function can be expressed in terms of
1030: the ${\cal T}$-matrix, giving a formula for the conductance:\cite{PG}
1031: \begin{equation} C={e^2\over h}\sum_s \int d\epsilon (-df/d\epsilon )
1032: [-\pi \nu \hbox{Im} {\cal T}_s (\epsilon )],\label{C-T}\end{equation}
1033: where $s=\pm 1$ labels the fermion spin,
1034: $f(\epsilon )$ is the Fermi distribution function at temperature $T$ and
1035: $\nu$ is the density of states. As shown in [\onlinecite{GP}] following
1036: [\onlinecite{Nozieres}, \onlinecite{AL}], for the embedded quantum dot:
1037: \begin{equation} -\pi \nu T_s (\epsilon )={1\over 2i}\left[\exp [2i\delta_s (\epsilon )]-1\right]
1038: +\exp [2i\delta_s (\epsilon )][-\pi \nu \tilde{\cal T}_{in}(\epsilon )],\end{equation}
1039: where
1040: \begin{eqnarray} \delta_s (\epsilon )&=& s\pi /2+\tilde \delta (\epsilon)\nonumber \\
1041: \tilde \delta (\epsilon ) &=& \omega /T_K \nonumber \\
1042: -\pi \nu \tilde T_{in}(\epsilon )&=&i\frac{(\epsilon^2+\pi^2T^2)}{2T_K^2}.\end{eqnarray}
1043: $\tilde \delta (\epsilon )$ and $\tilde T_{in}(\epsilon )$ are the
1044: phase shift and inelastic part of the ${\cal T}$-matrix calculated in
1045: perturbation theory in the Fermi liquid interactions. The extra
1046: $s\pi /2$ term in $\delta_s$ arises from the $\pm \pi /2$ phase shift
1047: characterizing the strong coupling fixed point. This $\pm \pi /2$ phase
1048: shift reflects the perfect transmission ($C=2e^2/h$) at $T=0$
1049: for the embedded quantum dot. On the other hand, for
1050: the side-coupled quantum dot, $C=0$ at the zero temperature fixed point,
1051: implying the absence of this extra phase shift. Thus, for the side-coupled
1052: quantum dot we have:
1053: \begin{equation} -\pi \nu T_s (\epsilon )={1\over 2i}\left[\exp [2i\tilde \delta_s (\epsilon )]-1\right]
1054: +\exp [2i\tilde \delta_s (\epsilon )][-\pi \nu {\cal T}_{in}(\epsilon )].\end{equation}
1055: To order $\omega^2$, $T^2$, the difference between ${\cal T}$-matrices for
1056: the embedded and side-coupled quantum dots is just a change in sign in
1057: the first and third terms. This implies:
1058: \bea -\pi \nu \hbox{Im}{\cal T} (\epsilon )&=&1-\frac{(3\epsilon^2+\pi^2T^2)}{2T_K^2}\ \ (\hbox{embedded})\nonumber \\
1059: -\pi \nu \hbox{Im}{\cal T} (\epsilon )&=&\frac{(3\epsilon^2+\pi^2T^2)}{2T_K^2}\ \ (\hbox{side-coupled})\eea
1060: Inserting these expressions into Eq. (\ref{C-T}) for the conductance, and restoring
1061: $\hbar$ and $e$ which were previously set to one, gives:
1062: \bea C&=&{2e^2\over h}\left[ 1-{\pi^2T^2\over T_K^2}\right] ,\ \
1063: \ \ (\hbox{embedded})\nonumber \\
1064: &=&{2e^2\over h}{\pi^2T^2\over T_K^2}\ \ (\hbox{side-coupled})\label{condfin}\eea
1065:
1066: We may now write universal Wilson-type ratios between the low temperature conductance, $C$,
1067: through a side-coupled quantum dot and the zero temperature persistent current, $j_e$, $j_o$, through
1068: the same quantum dot inserted into a ring of size $L\gg \xi_K$:
1069: \bea
1070: {L^2j_e/ev_F^2\over \sqrt{hC/(2e^2T^2)}}&
1071: \to &{1\over 4}\sin 2\tilde \alpha \nonumber \\
1072: {L^2j_o/ev_F^2\over \sqrt{hC/(2e^2T^2)}}&
1073: \to&{1\over 2}\sin\tilde \alpha -{1\over 4}\sin 2\tilde \alpha .
1074: \eea
1075:
1076: When PH symmetry is broken the conductance also gets a contribution of
1077: second order in $\lambda$ the coupling constant in the term, $H_2$
1078: of the effective Hamiltonian, given in Eq. (\ref{H2}). The simplest
1079: way of evaluating this contribution is to observe that, since
1080: the Hamiltonian is non-interacting, ignoring the other
1081: term $H_{int}$ in Eq. (\ref{flt}), we can evaluate the conductance
1082: at zero temperature
1083: using the Landauer formula, $C=(2e^2/h)T_r$ where $T_r$ is the transmission
1084: probability through the quantum dot. For small $\lambda$, $T_r=\lambda^2$.
1085:
1086: Including both terms, the conductance takes the form at low temperatures:
1087: \begin{equation} C = \left[ {\pi^2T^2\over T_K^2}+\lambda^2\right] {2e^2\over h}.
1088: \end{equation}
1089: At sufficiently low $T$ the second, non-universal term,
1090: arising from PH symmetry breaking always dominates. However
1091: provided that the PH symmetry breaking (and, in particular,
1092: the direct tunnelling across the quantum dot) is small,
1093: the first, universal, term dominates for:
1094: \begin{equation} \lambda T_K<<T<<T_K.\end{equation}
1095: Note that this situation is very analogous to that
1096: for the persistent current, with $T$ replaced by $v_F/L$.
1097:
1098:
1099:
1100: \section{Conclusion}
1101: We have developed a Fermi liquid theory of the persistent current in a side coupled quantum dot.
1102: This theory should correctly describe the limit $a\ll \xi_K\ll L$ where $a$ is a microscopic
1103: length scale. Numerical results are
1104: largely in agreement with the existence of a regime correctly described by this Fermi liquid picture.
1105: However, probably due to the limited system sizes available and the requirement that $L\gg\xi_K$,
1106: the numerical evidence cannot be described as strong. Our Fermi liquid theory confirms
1107: the existence of a term in the persistent current proportional to $\sin(\tilde\alpha)\delta n/J_K^2L$ for $N$ even,
1108: absent at half-filling. We have also calculated the conductance through a side-coupled quantum
1109: dot at low $T\ll T_K$ and calculated a universal ``Wilson'' ratio relating the conductance to the
1110: persistent current.
1111:
1112:
1113: \appendix
1114: \section{Conductance by Commutator Method}
1115: In this appendix we repeat the calculation of the conductance, writing the
1116: current as $I=i[H,\Delta \hat N]/2$. This provides a check on the previous
1117: Fermi liquid calculations involving the ${\cal T}$-matrix in [\onlinecite{Nozieres,AL,GP,PG}]
1118: and also provides an instructive example of how universal information can be
1119: extracted from a cut-off dependent result.
1120:
1121: Since $[\hat N_++\hat N_-,H]=0$, we may equivalently use $I=i[H,\hat N_+]$.
1122: Note that $\hat N_+$ commutes with the non-interacting part of the Hamiltonian which
1123: is simply that of free fermions with perfectly reflecting b.c.'s at the origin, since
1124: this Hamiltonian does not transmit any electrons between left and right sides of the system.
1125: The commutator of $\hat N_+$ with $H_{int}$ is readily calculated. Since all fields in the
1126: commutator are left-movers sitting at $x=0$, we drop the $L$ subscript and the $(0)$ argument
1127: for simplicity. The result is:
1128: \begin{widetext}
1129: \bea I=i[H_{int},N_+] &=& -{2\pi i\over 3T_K}\{[e^{-i\alpha}\psi_+^\dagger{\vec \sigma \over 2}\psi_- -h.c.],
1130: [\psi_+-e^{-i\alpha}\psi_-]^\dagger{\vec \sigma \over 2}[\psi_+-e^{-i\alpha}\psi_-]\}\nonumber \\
1131: && +{1\over 2T_K}\left[\psi_+^\dagger{d\over dx}\psi_-e^{-i\alpha}-{d\over dx}\psi^\dagger_+\psi_-e^{-i\alpha}
1132: +h.c.\right]
1133: .\eea
1134: \end{widetext}
1135: We note that this is, equivalently, $I=dH_{int}/d\alpha$.
1136: We expand the current operator up to first order in the vector potential, $\alpha$:
1137: \begin{equation} I = I_0+\alpha I_1 + O(\alpha^2).\end{equation}
1138: Here:
1139: %\begin{widetext}
1140: \bea I_0 &=& {4\pi i\over 3T_K}[(\psi_+^\dagger {\vec \sigma \over 2}\psi_-)^2-\psi_+^\dagger {\vec \sigma \over 2}\psi_-\cdot
1141: (\vec J_++\vec J_-)-h.c.]\nonumber \\&&
1142: +{1\over 2T_K}\left[\psi_+^\dagger\overleftrightarrow{d\over dx}\psi_- +h.c. \right]
1143: \equiv I_{in}+I_{el},
1144: \eea
1145: %\end{widetext}
1146: where
1147: \begin{equation} \vec J_{\pm} \equiv \psi_{\pm}^\dagger {\vec \sigma \over 2}\psi_{\pm}\end{equation}
1148: and we have defined the inelastic and elastic terms in the current operator
1149: which are, respectively, quartic and quadratic in fermion operators.
1150: The precise form of $I_1$ will not be needed.
1151: The conductivity is obtained from $<I>$, calculated to first order in $\alpha$.
1152: Thus we obtain:
1153: \begin{equation} I(\omega ) = [G_R(\omega)+<I_1>]\alpha (\omega ),\end{equation}
1154: where $G_R$ is the retarded Green's function:
1155: \begin{equation} G_R(\omega ) = -i\int_0^\infty e^{i\omega t}<[I_0(t),I_0(0)]>_T.\end{equation}
1156: It can be seen that $I=0$ when $\alpha (t)$ is a constant, independent of $t$.
1157: A non-zero $I$ in this case would correspond to a persistent current, but this
1158: must vanish for infinite $L$ since a constant phase $\alpha$ in $H_{int}$ can
1159: be eliminated by a gauge transformation in that limit. Thus we conclude that
1160: $<I_1>=-G_R(0)$, so that:
1161: \begin{equation} I(\omega )=[G_R(\omega )-G_R(0)]\alpha (\omega ).\end{equation}
1162: From Eq. (\ref{V}), we see that:
1163: \begin{equation} I(\omega )=C(\omega )\Delta V(\omega ),\end{equation}
1164: where the conductance, $C(\omega )$, is given by:
1165: \begin{equation} C(\omega ) = {i\over \omega }[G_R(\omega )-G_R(0)],\end{equation}
1166: where $\omega_n=2\pi nT$ and $\beta \equiv 1/T$. (We set Boltzmann's constant
1167: equal to one.)
1168: The dc conductance is:
1169: \begin{equation} C=\lim_{\omega \to 0}{i\over \omega }[G_R(\omega )-G_R(0)].\label{CDC}\end{equation}
1170: We calculate the retarded Green's function of $I_0$ by analytic continuation
1171: from the imaginary time, Matsubara Green's function:
1172: \begin{equation} \GG(i\omega_n )=-\int_0^\beta d\tau e^{i\omega_n\tau}<I_0(\tau )I_0(0)>.\end{equation}
1173: The free fermion Matsubara
1174: Green's function, at late times, using the same normalization as in Eq. (\ref{prop}), that
1175: was used in the calculation of the persistent current is:
1176: \begin{equation} <\psi^{\dagger \alpha}_i(\tau )\psi_{\gamma j}(0)>\to {\delta^\alpha_\gamma
1177: \delta_{ij}\over 2\beta \sin (\pi \tau /\beta )},\end{equation}
1178: where $i$, $j=\pm$ labels right and left sides of the junction, $x=0^{\pm}$
1179: and $\beta$ is the inverse temperature. Using Wick's theorem, the identity:
1180: \begin{equation} \sum_{a,b}[tr \sigma^a\sigma^b~tr\sigma^a\sigma^b-tr\sigma^a\sigma^b\sigma^a\sigma^b]=18,\end{equation}
1181: and collecting the various terms, we obtain the long-time behavior of the Matsubara Green's function:
1182: \begin{equation} <I_{in}(\tau )I_{in}(0)>\to {16\pi^2\over T_K^2}{1\over (2\beta )^4 \sin^4 (\pi \tau /\beta )}.\end{equation}
1183:
1184: To evaluate $<I_{el}(\tau )I_{el}(0)>$, we use the derivatives of the free fermion Green's function:
1185: %\begin{widetext}
1186: \begin{eqnarray}
1187: {d\over dx}\left\{ {1\over \sin [\pi (\tau + ix)/\beta ]}\right\} &=&
1188: -{i\pi \cos [\pi (\tau +ix)/\beta ]\over \beta \sin^2 [\pi (\tau +ix)/\beta ]},
1189: \nonumber \\
1190: {d^2\over dx^2} {1\over \sin [\pi (\tau + ix)/\beta ]} &=&
1191: -{\pi^2\over \beta^2}{1+\cos^2[\pi (\tau + ix)/\beta ]\over \sin^3 [\pi (\tau + ix)/\beta ]}.\nonumber\\
1192: \end{eqnarray}
1193: %\end{widetext}
1194: The terms proportional to $\cos^2(\pi \tau /\beta )$ cancel leaving:
1195: %\begin{widetext}
1196: \begin{eqnarray}
1197: <I_{el}(\tau )I_{el}(0)>&\to& {8\pi^2\over T_K^2}{1\over (2\beta )^4\sin^4(\pi \tau /\beta)}\nonumber\\
1198: &=&(1/2) <I_{in}(\tau )I_{in}(0)>.
1199: \end{eqnarray}
1200: %\end{widetext}
1201: We now consider the Fourier transform of $G(\tau )$ at low frequencies and temperatures. The
1202: analytic continuation of $\GG(i\omega_n)$ ($\omega_n\equiv 2\pi n/\beta$) to real frequencies gives the retarded Green's function:
1203: \begin{equation} \GG(i\omega_n\to \omega +i\delta)=G_R(\omega ),\end{equation}
1204: where $\delta$ is a postive infinitesimal.
1205: We will find that
1206: \begin{equation} \GG(i\omega_n)\to A+B|\omega_n| + O(\omega_n^2),\end{equation}
1207: for $\omega_n<<D$. The analytic continuation gives:
1208: %\begin{widetext}
1209: \begin{eqnarray} |\omega_n|&=&\omega_n\int_{-\infty}^\infty {d\tau \over 2\pi i}{e^{i\omega_n\tau}
1210: -e^{-i\omega_n\tau}\over \tau -i\epsilon}\nonumber\\
1211: &\to& (-i\omega)
1212: \int_{-\infty}^\infty {d\tau \over 2\pi i}{e^{(i\delta -\omega )\tau}
1213: -e^{-i(\delta -i\omega )\tau}\over \tau -i\epsilon}\nonumber\\
1214: &=& -i\omega
1215: .\label{omegacon}\end{eqnarray}
1216: %\end{widetext}
1217: $G(\omega_n)$, which is an even function of $\omega_n$, in our assumed particle-hole
1218: symmetric model, contains non-universal ultraviolet cut-off dependent terms. Introducing
1219: a cut-off, $\tau_0$ with dimensions of time, and of order $1/D$, we find, at low $\omega_n$:
1220: %\begin{widetext}
1221: \begin{eqnarray}
1222: \lefteqn{\GG(i\omega_n)\to}&&\nonumber\\
1223: &&{1\over T_K^2}\left[{a\over \tau_0^3}+b{1\over \tau_0\beta^2}
1224: +c{|\omega_n|\over \beta^2}+d{\omega_n^2\over \tau_0}+e|\omega_n|^3+\ldots \right],
1225: \nonumber\\
1226: \label{loww}
1227: \end{eqnarray}
1228: %\end{widetext}
1229: where $a,b,c,d,e$ are dimensionless numbers.
1230: While the terms of O$(\omega_n^0)$ and $(\omega_n^2)$ are non-universal and cut-off
1231: dependent, we expect that the terms $\propto |\omega_n|$ and $|\omega_n|^3$ are not. This is related to the
1232: fact that only these terms are singular functions of $\omega_n$. This singularity
1233: arises from the universal long-time behavior of $G(\tau )$. Note that the
1234: dc conductance is completely determined by the universal $c$-term $\propto |\omega_n|$
1235: and so is a universal quantity in the sense that it is independent of the
1236: details of the ultraviolet cut-off.
1237: To verify these assertions, we consider a particularly simple ultraviolet cut-off.
1238: The $\tau$ integral defining the Fourier transform of
1239: the current Green's function is restricted to $\tau_0<\tau <\beta - \tau_0$
1240: Thus we begin with:
1241: \begin{equation} \GG(i\omega_n)\equiv {-24\pi^2\over T_K^2(2\beta )^4}\int_{\tau_0}^{\beta -\tau_0}
1242: {d\tau e^{i\omega_n\tau}\over \sin^4(\pi \tau /\beta )}.
1243: \label{int-cutoff}\end{equation}
1244: This integral can be straightforwardly
1245: developed in an expansion in $\tau_0/\beta$, which has the form of Eq. (\ref{loww}).
1246: This can be conveniently done by deforming the $\tau$-integral into the complex plane.
1247: The original integral is equal to the sum of an integral along the straight line from $\tau = i\tau_0$
1248: to $\tau = i\tau_0+\beta$ plus the integral on two quarter-circle contours from
1249: $\tau = \tau_0$ to $\tau = i\tau_0$ and from $\beta + i\tau_0$ to $\tau = \beta -\tau_0$, which we label $C_1$ and $C_2$ respectively. [See fig.
1250: (\ref{fig:contour}).]
1251: \begin{figure}
1252: \begin{center}
1253: \includegraphics[clip,width=8cm]{Fig5.eps}
1254: \caption{Contour used to evaluate the integral in Eq. (\ref{int-cutoff}).}
1255: \label{fig:contour}
1256: \end{center}
1257: \end{figure}
1258: The integral along the straight line can be written:
1259: \begin{equation} K=\int_0^\beta {d\tau e^{i\omega_n(\tau +i\tau_0)}\over \sin^4[\pi (\tau +i\tau_0)/\beta ]}.\label{int_st}\end{equation}
1260: We may now Taylor expand the denominator:
1261: \begin{equation} {1\over \sin^4[\pi (\tau +i\tau_0)/\beta ]}=16e^{4\pi i(\tau +i\tau_0)/\beta}
1262: \sum_{m=0}^\infty a_me^{2\pi im(\tau +i\tau_0)/\beta}.\end{equation}
1263: Note that the sum converges due to the factors $e^{-2\pi \tau_0/\beta}$. Here the $a_m$
1264: are the binomial coefficients:
1265: \begin{equation} a_m= (m+1)(m+2)(m+3)/6.\end{equation}
1266: We now integrate over $\beta$ term by term. The result is:
1267: \begin{equation} \int_0^\beta d\tau e^{i[\omega_n+(2+m)(2\pi /\beta )]\tau}=\beta \delta_{n,-(2+m)}.\end{equation}
1268: (Recall that $\omega_n\equiv 2\pi n/\beta$.) Thus we obtain:
1269: \begin{equation} K=-{4\beta^2|\omega_n|\over 3\pi}\left[
1270: 1-\left({\beta \omega_n \over 2\pi}\right)^2\right]
1271: \theta (-\omega_n).\end{equation}
1272: Alternatively, the integral of Eq. (\ref{int_st}) can be evaluated
1273: by introducing the complex variable $z\equiv e^{i2\pi (\tau+i\tau_0)/\beta}$.
1274: The integration contour for $z$ is a circle of radius $e^{-2\pi \tau_0/\beta}<1$.
1275: This contour encloses a pole of order $|n|-1$ when $n<-1$.
1276:
1277: We also
1278: must perform the integration over the two quarter-circle contours:
1279: \begin{equation} K_{12}\equiv \sum_{i=1}^2
1280: \int_{C_i}dz{e^{i\omega_nz}\over \sin^4(\pi z/\beta)}.\end{equation}
1281: On $C_1$ we write $z\equiv \tau_0e^{i\theta}$ where $\theta$ goes from
1282: $0 \to \pi /2$ along the contour. On $C_2$ we write $z=\beta +\tau_0
1283: e^{i\theta}$ where now $\theta$ goes from $\pi /2\to \pi$. Since
1284: the integrand is invariant under a translation: $z\to z+\beta$,
1285: we may combine these two terms to write:
1286: \begin{equation} K_{12}=i\tau_0\int_0^\pi d\theta {e^{i\theta}\exp
1287: [i\omega_n\tau_0e^{i\theta}]\over \sin^4(\pi \tau_0e^{i\theta}/\beta )}.\end{equation}
1288: Next we expand $\exp [i\omega_n\tau_0e^{i\theta}]$ in powers of
1289: $\omega_n\tau_0$ and expand $\sin^{-4}(\pi \tau_0e^{i\theta}/\beta )$
1290: in powers of $\tau_0/\beta$, giving:\begin{widetext}
1291: \begin{equation} K_{12}=i\tau_0\left({\beta \over \pi \tau_0}\right)^4\int_0^\pi
1292: e^{i\theta}e^{-4i\theta}\left[1+i\omega_n\tau_0e^{i\theta}-
1293: {\omega_n^2\tau^2\over 2}e^{2i\theta}-{i\omega_n^3\tau_0^3\over 6}e^{3i\theta}
1294: +\ldots \right]\left[1-{\pi^2 \tau_0^2\over 6\beta^2}e^{2i\theta}
1295: +{\pi^4\tau_0^4\over 120 \beta^4}e^{4i\theta}+\ldots \right]^{-4}.\end{equation}
1296: \end{widetext}
1297: We Taylor expand, collect terms, and integrate term by term using:
1298: \bea \int_0^\pi d\theta e^{im\theta}&=&\pi \ \ (m=0)\nonumber \\
1299: &=& 0\ \ (m \ \hbox{even},\ m\geq 2)\nonumber \\
1300: &=& 2i/m,\ \ (m \ \hbox{odd}).\eea
1301: Note that non-zero terms from the expansion must be proportional to
1302: $e^{3i\theta}$ or else $e^{2im\theta}$ to give
1303: a non-zero contribution. The first type of terms are odd functions of
1304: $\omega_n$, proportional to $\omega_n$ and $\omega_n^3$. All other
1305: terms are even functions of $\omega_n$. Keeping only the terms
1306: which are non-vanishing as $\tau_0\to 0$ gives:
1307: \begin{equation} K_{12}\approx {-2\omega_n\beta^2\over 3\pi}+{\omega_n^3\beta^4\over 6\pi^3}
1308: +{2\beta^4\over 3\pi^4\tau_0^3}-{\omega_n^2\beta^4\over \pi^4\tau_0}
1309: +{4\beta^2\over 3\pi^2\tau_0}+\ldots \end{equation}
1310: The omitted terms are all even functions of $\omega_n$ and vanish
1311: as $\tau_0\to 0$.
1312:
1313: Note that when we add $K+K_{12}$ we regain an even function of $\omega_n$,
1314: as we must:
1315: %\begin{widetext}
1316: \begin{eqnarray} K+K_{12}&=&{-2\beta^2\over 3\pi}|\omega_n|\left[1-
1317: \left({\beta \omega_n\over 2\pi}\right)^2\right]\nonumber\\
1318: &+&{2\beta^4\over 3\pi^4\tau_0^3}-{\omega_n^2\beta^4\over \pi^4\tau_0}
1319: +{4\beta^2\over 3\pi^2\tau_0}+\ldots .\end{eqnarray}
1320: %\end{widetext}
1321: The first two terms are singular functions of $\omega_n$
1322: at $\omega_n=0$. All remaining terms are non-singular, even
1323: powers of $\omega_n$. The singular terms are cut-off independent,
1324: unlike the non-singular ones.
1325:
1326: Thus we may write the imaginary frequency Green's function for the current
1327: operator as:
1328: \begin{equation} \GG(i\omega_n)\to \GG(0)+{\pi T^2|\omega_n|\over T_K^2}+O(\omega_n^2).\end{equation}
1329: The analytic continuation to real frequency is straightforward, using
1330: Eq. (\ref{omegacon}):
1331: \begin{equation} G_R(\omega ) \to G_R(0)-{i\pi T^2\omega \over T_K^2}+O(\omega^2).\end{equation}
1332: Thus, we obtain the dc conductance from Eq. (\ref{CDC}):
1333: \begin{equation} C = {\pi T^2\over T_K^2}\to {2e^2\over h}{\pi^2T^2\over T_K^2},\end{equation}
1334: the same result obtained from the derivative method, Eq. (\ref{condfin}).
1335: (A factor of $e^2/\hbar$, previously
1336: set equal to one was inserted in the last step.)
1337:
1338: In Sec. V, we evaluated the conductance from the P-H symmetry violating
1339: tunneling term using the Landauer formalism.
1340: It is instructive to evaluate the conductance instead using the Kubo formula, and the commutator approach,
1341: as done earlier in this Appendix. In this case, the new term in the ($\alpha =0$) current operator is:
1342: \begin{equation} \delta I_0 = i\lambda e^{i\pi (N-1)/2}\psi^\dagger_+\psi_- + h.c.\end{equation}
1343: The Green's function for $I_0$, for large imaginary times, now picks up a correction:
1344: \begin{equation} \delta <I_0(\tau )I_0(0)>=4\lambda^2{1\over (2\beta )^2\sin^2(\pi \tau /\beta )}.\end{equation}
1345: We may evaluate the Fourier transform using the same methods as above. However,
1346: we focus immediately on the $T=0$ limit, for simplicity:
1347: \begin{equation} \delta \GG (\omega_n)=-4\lambda^2\int_{-\infty}^\infty
1348: {d\tau e^{i\omega_n\tau}\over (2\pi \tau )^2}.\label{deltaG}\end{equation}
1349:
1350: This integral requires an ultraviolet cut off, as before. Cutting off the
1351: integral at $|\tau |>\tau_0$, gives:
1352: \begin{equation} \delta \GG = {-2\lambda^2\over \pi^2}\int_{\tau_0}^\infty {d\tau \over \tau^2}
1353: \cos (\omega_n\tau ) .\end{equation}
1354: It is convenient to integrate by parts, giving:
1355: \begin{equation} \delta \GG = {-2\lambda^2\over \pi^2\tau_0}+{2\lambda^2\omega_n\over \pi^2}
1356: \int_{\tau_0}^\infty {d\tau \over \tau}
1357: \sin (\omega_n\tau ).\end{equation}
1358: [We used $\cos (\omega_n\tau_0)\approx 1$ in the first term.]
1359: Since this integral now converges, we may take $\tau_0\to 0$, assuming that
1360: $|\omega_n|\tau_0\ll 1$. We may then rescale the integral giving:
1361: \begin{equation} \delta \GG = {-2\lambda^2\over \pi^2\tau_0}+
1362: {2\lambda^2|\omega_n|\over \pi^2}\int_0^\infty {ds\over s}\sin s + O(\tau_0).\end{equation}
1363: Evaluating the integral gives:
1364: \begin{equation} \delta \GG = {-2\lambda^2\over \pi^2\tau_0}+
1365: {2\lambda^2|\omega_n|\over 2\pi}+O(\tau_0).\end{equation}
1366: Note that we again obtain a term which is even in $\omega_n$ but singular,
1367: involving an absolute value. Analytically continuing to real
1368: frequency, using Eq. (\ref{omegacon}) we obtain the zero temperature
1369: DC conductance from
1370: Eq. (\ref{CDC}):
1371: \begin{equation} C={\lambda^2\over \pi}\to {2e^2\over h}\lambda^2,\end{equation}
1372: the same result than is obtained from the Landauer formula. (Again
1373: a factor of $e^2/\hbar$ was inserted at the last step.)
1374:
1375: Alternatively, we could cut off the integral in Eq. (\ref{deltaG}), by
1376: a finite bandwidth, $2D$. Then the fermion propogator is modified to:
1377: \begin{equation} <\psi^{\dagger \alpha}_i(\tau )\psi_{\beta j}(0)>\to \delta^\alpha_\beta
1378: \delta_{ij}{1-e^{-D|\tau |}\over 2\pi \tau},\end{equation}
1379: giving:
1380: \begin{equation} \delta G(\omega_n)=4\lambda^2\int_{-\infty}^\infty
1381: {d\tau e^{i\omega_n\tau}\left( 1-e^{-D|\tau |}\right)^2 \over (2\pi \tau )^2},\label{deltaG2}\end{equation}
1382: where the integration region is now the entire real line. This can be expressed in
1383: terms of the exponential integral function, $E_i(x)$, giving a term of $O(D)$,
1384: the same universal term, $-\lambda^2|\omega_n|/\pi$, plus terms that
1385: vanish when $1/D\to 0$. This confirms that the term $\propto |\omega_n|$, which
1386: determines the conductance, is
1387: indeed universal.
1388:
1389: \section{Definitions of $T_K$}
1390: The definition of $T_K$ used in this paper, and defined
1391: in Eq. (\ref{Hint}), is the one first introduced by Nozi\`eres~\cite{Nozieres}
1392: in the original paper on local Fermi Liquid Theory for the Kondo model. It
1393: corresponds to an arbitrary choice but does have the advantage that
1394: the coefficient of $T^2/T_K^2$ in the conductance
1395: has the relatively simple value of $\pi^2$. [See Eq. (\ref{condfin}).]
1396: Another popular definition is the one adopted by Wilson, which
1397: we refer to as $T_K^W$. This is fixed by the requirement that
1398: the impurity susceptibility, at $T\gg T_K$ have a particular
1399: form:
1400: \begin{eqnarray}
1401: \lefteqn{\chi_{imp}\to}&&\nonumber\\
1402: &&{(g\mu_B)^2\over 4T} \bigg[1-{1\over \ln (T/T_K^W)}
1403: -{(1/2)\ln \left[\ln (T/T_K^W)\right]\over \ln^2(T/T_K^W)}\nonumber\\
1404: &&+O\left[\ln^2\left[\ln (T/T_K^W)\right]/\ln^3(T/T_K^W)\right] \bigg].
1405: \end{eqnarray}
1406: At low temperatures,
1407: \be \chi_{imp}\to {(g\mu_B)^2w^2\over 4T_K^W},\ee
1408: where the Wilson number, $w$ has the value:
1409: \be w =e^{C+1/4}/\pi^{3/2}\approx .4128 \ee
1410: and $C$ is Euler's constant ($\approx .577216$).
1411: These definitions of $T_K$ are related by:
1412: \be T_K^W=(\pi w/4)T_K.\ee
1413: Rather accurate results have been determined for the conductance
1414: at all temperatures using numerical renormalization group~\cite{Costi}
1415: and integrability.\cite{Konik} These authors generally use
1416: a Kondo temperature, $T_K^C$, related to the other ones by:
1417: \be T_K^C=T_K^W/w=(\pi /4)T_K,\ee
1418: motivated, perhaps, by the fact that the zero temperature
1419: impurity susceptibility now takes the simple form
1420: $\chi_{imp}\to (g\mu_B)^2/(4T_K^C)$.
1421: Some experimental papers~\cite{GG,Wiel}
1422: on the Kondo effect in embedded quantum dots define
1423: a Kondo temperature by the condition that the conductance take
1424: half its zero temperature value at $T=T_K^e$:
1425: \be C(T=T_K^e)/C(T=0)=1/2.\ee
1426: The theoretical results in [\onlinecite{Konik}] indicate
1427: that $T_K^e\approx T_K^C$.
1428: In [\onlinecite{AL}] we wrote the Fermi liquid interaction:
1429: \be H_{int}=-\lambda \left (\psi_e^\dagger {\vec \sigma \over 2}\psi_e
1430: \right)^2,\ee
1431: (where $\lambda$ is not to be confused with the coefficient
1432: of the particle-hole breaking term in the current paper)
1433: and we normalized our fermions in an unconventional way
1434: so that $\psi_{AL}=\sqrt{2\pi}\psi$. Thus:
1435: \be \lambda = {2\over 3\pi T_K}.\ee
1436: Numerous other definitions of $T_K$ are also in use
1437: including, for example, the frequency scale
1438: at which Im${\cal T}(\omega ,T=0)$ is reduced by 1/2 from
1439: its maximum at zero frequency.
1440:
1441:
1442:
1443:
1444: \acknowledgments
1445: This research is supported by NSERC, CFI, SHARCNET and CIAR.
1446: IA acknowledges interesting conversations with P. Simon.
1447: \bibliography{flt}
1448:
1449: \end{document}
1450: