1: As we have seen in the last section, there are infrared stable fixed
2: points that cannot be identified with either weak- or strong- hopping
3: limit in the bosonized representation Eq.~(\ref{eq:SB-set-up})
4: with the Klein factors.
5:
6: The presence of the phase factors due to the fermionic statistics,
7: encoded in the Klein factors, makes the
8: analysis more difficult than that for the standard problems with only
9: bosonic terms in the boundary action. The extra phases due to the
10: fermionic statistics appear, for example, if one attempts a
11: perturbative expansion in the hopping amplitude $\Gamma$. The
12: coefficients appearing in each order in perturbation theory are
13: different than those obtained without the phase factors, and hence
14: they alter the nature of the strong hopping limit.
15:
16: In this section, we develop an alternative representation,
17: in which the fermionic statistics of electrons are encoded
18: without using the Klein factors.
19: Namely, we will show that the perturbative series in the
20: hopping amplitude $\Gamma$ for the Y-junction problem is identical to
21: the series obtained for the dissipative Hofstadter model (quantum
22: motion of a single particle under a magnetic field and a periodic
23: potential, subject to dissipation) studied by Callan and Freed in
24: Ref.~\cite{CF}. In the dissipative Hofstadter action, phase factors
25: arise from the motion in a magnetic field, but the action contains
26: solely bosonic terms, and hence it is amenable to standard methods
27: for determining the strong coupling behavior, such as instanton
28: expansions.
29:
30: \subsection{Generalized Coulomb gas from the Y-junction}
31:
32: We consider the perturbation theory in the hopping parameter
33: $\Gamma$, starting from the bosonized
34: representation Eq.~(\ref{eq:SB-set-up}) with the Klein factor.
35: The partition function $Z_Y$ of the system can be expanded as
36: \begin{widetext}
37: \begin{eqnarray}
38: Z_Y=\sum_{n}
39: %\sum_{n=0}^\infty
40: %\frac{1}{n!}\;
41: \;\Gamma^n
42: \int_{\tau_{n-1}}^\infty \!\!\!\!\! d\tau_{n}\;
43: \int_{\tau_{n-2}}^{\tau_{n}} \!\!\!\!\! d\tau_{n-1}\;
44: \cdots
45: \int_{-\infty}^{\tau_{2}} \!\!\!\! d\tau_{1}\;
46: \sum_{\{L_j\}}
47: %\left[
48: %\zeta_{\vec L_n}(\tau_n)
49: %\cdots
50: %\zeta_{\vec L_2}(\tau_2)\;
51: %\zeta_{\vec L_1}(\tau_1)
52: %\right]
53: \left[
54: \zeta_{\vec L_n}
55: \cdots
56: \zeta_{\vec L_2}\;
57: \zeta_{\vec L_1}
58: \right]
59: \;
60: \langle
61: e^{i \vec{L}_{n} \cdot \vec{\Phi}(\tau_n)}
62: \ldots
63: e^{i \vec{L}_{2} \cdot \vec{\Phi}(\tau_2)}
64: e^{i \vec{L}_{1} \cdot \vec{\Phi}(\tau_1)}
65: \rangle_0
66: \;,
67: \label{eq:Z_Y}
68: \end{eqnarray}
69: \end{widetext}
70: where ${\vec L}_j=q_j{\vec K}_{a_j}$ is one of the six vectors $\pm
71: \vec{K}_{1,2,3}$ ($q_j=\pm 1$ and $a_j=1,2,3$), and $\langle \
72: \rangle_0$ denotes the expectation value with the Neumann boundary
73: condition on $\vec{\Phi}$. The $\zeta_{\vec
74: L_j}=(-i)^{q_j}\,e^{iq_j\phi/3}\,\eta_{a_j}$ in the $[\ldots]$ term in
75: Eq.~(\ref{eq:Z_Y}) keep track of the phase factors.
76:
77: Now, each term in the expansion Eq.~(\ref{eq:Z_Y}) corresponds to a
78: closed directed path
79: formed by the vectors
80: ${\vec L}_1,{\vec L}_2,\ldots, {\vec L}_n$,
81: on the triangular lattice
82: spanned by the primitive vectors ${\vec K}_1$ and ${\vec K}_2$.
83: An
84: example of a closed path is shown in Fig.~\ref{fig:pathex}. A charge vector
85: ${\vec L}_j$ is associated to each vertex operator $e^{i{\vec L}_j\cdot
86: \vec\Phi}$, and the condition that the sequence of vectors must form a
87: closed path is a consequence of the charge neutrality condition of the
88: Coulomb gas expansion.
89:
90: \begin{figure}
91: \begin{center}
92: \includegraphics[width=0.45\linewidth]{pathex.eps}
93: \caption{A typical closed path occurring in perturbative expansion
94: of partition function}
95: \label{fig:pathex}
96: \end{center}
97: \end{figure}
98:
99: Let us now argue that the total phase factor for a given path ${\cal
100: L}=({\vec L}_1,{\vec L}_2,\ldots, {\vec L}_n)$,
101: \begin{equation}
102: e^{i\Upsilon_{\cal L}}=
103: \left[
104: \zeta_{\vec L_n}
105: \cdots
106: \zeta_{\vec L_2}\;
107: \zeta_{\vec L_1}
108: \right]
109: \; ,
110: \end{equation}
111: is the sum of the phase contributions from all the elementary
112: triangles enclosed by the loop. The phase contributions from ``up''
113: and ``down'' triangles, however, are different. The phase factor for a
114: counterclockwise loop on the up triangle [see part (a) of
115: Fig.~\ref{fig:updown}] is determined to be
116: \begin{equation}
117: e^{i \upsilon_{\bigtriangleup}}=
118: (-i \eta_3) e^{i \phi/3} (-i \eta_2) e^{i \phi/3} (-i \eta_1) e^{i \phi/3}
119: = e^{i \phi}\label{up}
120: \end{equation}
121: while that for the counterclockwise loop on the down triangle
122: [see part (b) of Fig.~\ref{fig:updown}] is
123: \begin{equation}
124: e^{i \upsilon_{\bigtriangledown}}=
125: (i \eta_3) e^{- i \phi/3} (i \eta_2) e^{- i \phi/3}
126: (i \eta_1) e^{- i \phi/3} = e^{i (\pi - \phi)}.
127: \label{down}\end{equation}
128: Namely, loops on the triangular lattice pick up a phase as if there is
129: a staggered magnetic flux of $\phi$ and $\pi - \phi$ in each
130: elementary triangle. The total phase accumulated due to the loop
131: ${\cal L}$ is thus
132: \begin{equation}
133: \Upsilon_{\cal L}=
134: N_\bigtriangleup({\cal L})\;\upsilon_\bigtriangleup+
135: N_\bigtriangledown({\cal L})\;\upsilon_\bigtriangledown
136: \; ,
137: \label{phaseY}\end{equation}
138: where $N_\bigtriangleup({\cal L})$ and $N_\bigtriangledown({\cal L})$
139: are the {\it net} numbers of up and down triangles enclosed by the path
140: (a triangle contributes $\pm 1$ to $N_\bigtriangleup$
141: and $N_\bigtriangledown$ depending on whether
142: it is traversed clockwise or counter-clockwise).
143:
144: \begin{figure}
145: \begin{center}
146: \includegraphics[width=0.4\linewidth]{up-down-triangle.eps}
147: \caption{(a) a counter-clockwise loop around an ``up triangle''.
148: (b) A counter-clockwise loop around a ``down triangle''.}
149: \label{fig:updown}
150: \end{center}
151: \end{figure}
152:
153: The argument why the phase for a given path ${\cal L}$ is the sum of
154: the phases due to the enclosed elementary triangles is constructed
155: inductively. First consider the special case where the closed path
156: consists of a single non-intersecting
157: closed loop. This implies that each vertex on the loop is visited
158: a single time except for the first vertex which is visited twice,
159: at the beginning and end of the loop. The phase from this
160: loop is equal to the sum of phases of elementary triangular
161: loops into which it can be decomposed.
162: Say it is so for all paths enclosing up to $m$ triangles,
163: and then consider a loop enclosing $m+1$ triangles; we must argue that
164: the phase accumulated around the loop enclosing $m+1$ triangles is
165: equal to the sum of the phases enclosed by two smaller area paths
166: with $m_1$ and $m_2$ triangles ($m_1+m_2=m+1, m_{1,2}\le m$).
167:
168: Consider then an arbitrary non-intersecting loop ${\cal L}=({\vec L}_1,{\vec
169: L}_2,\ldots, {\vec L}_n)$, which we break into two loops by adding an
170: internal segment that splits the loop into two
171: loops. Let the segment be the
172: sequence of vectors ${\vec L'}_1,{\vec L'}_2,\dots,{\vec L'}_p$.
173: It is easy to show
174: that
175: \begin{equation}
176: \left[
177: \zeta_{-\vec L'_1}\;
178: \zeta_{-\vec L'_2}\;
179: \cdots
180: \zeta_{-\vec L'_p}
181: \right]
182: \left[
183: \zeta_{\vec L'_p}
184: \cdots
185: \zeta_{\vec L'_2}\;
186: \zeta_{\vec L'_1}
187: \right]
188: =\openone
189: \; ,
190: \end{equation}
191: so that
192: \begin{widetext}
193: \begin{equation}
194: e^{i\Upsilon_{\cal L}}=
195: \left[
196: \zeta_{\vec L_n}
197: \cdots
198: \zeta_{\vec L_{j+2}}\;
199: \zeta_{\vec L_{j+1}}\;
200: \;
201: \zeta_{-\vec L'_1}\;
202: \zeta_{-\vec L'_2}\;
203: \cdots
204: \zeta_{-\vec L'_p}
205: \right]
206: \left[
207: \zeta_{\vec L'_p}
208: \cdots
209: \zeta_{\vec L'_2}\;
210: \zeta_{\vec L'_1}
211: \zeta_{\vec L_j}
212: \cdots
213: \zeta_{\vec L_2}\;
214: \zeta_{\vec L_1}
215: \right]
216: =e^{i\Upsilon_{{\cal L}_1}}\;e^{i\Upsilon_{{\cal L}_2}}
217: \; ,
218: \end{equation}
219: \end{widetext}
220: for two loops ${\cal L}_{1,2}$ that encircle areas smaller than the
221: original loop ${\cal L}$ , {\it i.e.}, $l_{1,2}\le m$ and with
222: $l_{1}+l_{2}=m+1$, as required by the induction argument.
223: An example of a decomposition of a non-crossing loop into
224: elementary triangles is given in Fig.~\ref{fig:loopdec}.
225: \begin{figure}
226: \begin{center}
227: \includegraphics[width=0.85\linewidth]{loopdec.eps}
228: \caption{Decomposition of a non-crossing loop into
229: elementary triangles}
230: \label{fig:loopdec}
231: \end{center}
232: \end{figure}
233:
234:
235: Finally we must argue that an arbitrary closed path can be
236: decomposed into a set of non-intersecting loops such that the
237: total phase of the path is the sum of phases for each
238: non-intersecting loop. This can be seen as follows.
239: Consider following an arbitrary oriented path. Consider the first
240: time that any vertex on the path is visited for a second time.
241: Label this vertex $V_1$.
242: The sequence of edges between the first and second visit to $V_1$
243: defines a non-intersecting closed loop, $L_1$. (This follows since
244: the first intersection on the path occurs at $V_1$ at the
245: instant of the second visit.) The phase due to this closed
246: non-intersecting loop may be calculated unambiguously and
247: makes an additive contribution to the total phase of
248: the path. This follows since the product of the Klein factors
249: around a closed loop is always proportional to the identity matrix.
250: Now excise the closed loop, $L_1$, from the path. The remaining path
251: remains closed. Consider the first time that any vertex
252: on this excised path is visited twice, at some other
253: vertex, $V_2$. This defines another closed loop $L_2$ which
254: we again excise. We continue in this way until we finally
255: revisit the first point on the path. This final revisitation
256: defines a final closed loop $L_n$.
257: An example of the decomposition of a path into
258: non-crossing loops is given in Fig.~\ref{fig:pathdec}.
259: Note that when a path (or an
260: excised path)
261: makes a U-turn and retraces a link this counts as a closed
262: loop of length 2 and zero area. The total phase
263: associated with the path is the sum of phases of each
264: non-intersecting closed loop. These phases can each be
265: expressed as a sum of phases of elementary triangles into
266: which they can be decomposed. Adding up the net number of
267: up triangles and down triangles in all the non-intersecting loops
268: gives us Eq. (\ref{phaseY}) for the total phase.
269: \begin{figure}
270: \begin{center}
271: \includegraphics[width=0.65\linewidth]{pathdec.eps}
272: \caption{Decomposition of a path into
273: non-crossing loops}
274: \label{fig:pathdec}
275: \end{center}
276: \end{figure}
277:
278:
279: Let us now turn to the correlation function entering in Eq.~(\ref{eq:Z_Y}):
280: \begin{eqnarray}
281: \lefteqn{
282: \langle
283: e^{i \vec{L}_{n} \cdot \vec{\Phi}(\tau_n)}
284: \ldots
285: e^{i \vec{L}_{2} \cdot \vec{\Phi}(\tau_2)}
286: e^{i \vec{L}_{1} \cdot \vec{\Phi}(\tau_1)}
287: \rangle_0 = } \nonumber \\
288: & \delta_K(\sum_j \vec{L}_j)&
289: \exp{\left[ \frac{1}{g}\sum_{j>k} \vec{L}_j \cdot \vec{L}_k
290: \ln{|\tau_j - \tau_k|}^2 \right]},
291: \label{eq:loop}
292: \end{eqnarray}
293: where $\delta_K(\vec{L}) = 1$ if $\vec{L}=0$, and
294: $\delta_K(\vec{L}) = 0$ otherwise.
295: % MO -- edited 5 Sep 2005
296: The factor $\delta_K(\sum_j \vec{L}_j)$ enforcing the charge
297: vectors ${\vec L}_j$ must form a closed path, as we mentioned above,
298: explicitly follows from of charge conservation in the Coulomb gas
299: expansion.
300:
301: Using this expression, in conjunction with the phase factor
302: we discussed above, we can rewrite the partition function as:
303: \begin{widetext}
304: \begin{eqnarray}
305: Z_Y=\sum_{n}
306: %\sum_{n=0}^\infty
307: %\frac{1}{n!}\;
308: \;\Gamma^n
309: \int_{\tau_{n-1}}^\infty \!\!\!\!\! d\tau_{n}\;
310: \int_{\tau_{n-2}}^{\tau_{n}} \!\!\!\!\! d\tau_{n-1}\;
311: \cdots
312: \int_{-\infty}^{\tau_{2}} \!\!\!\! d\tau_{1}\;
313: \sum_{{\cal L}=(\vec L_1,\dots,\vec L_n)}
314: %\left[
315: %\zeta_{\vec L_n}(\tau_n)
316: %\cdots
317: %\zeta_{\vec L_2}(\tau_2)\;
318: %\zeta_{\vec L_1}(\tau_1)
319: %\right]
320: \delta_K(\sum_j \vec{L}_j)\;\;e^{i\Upsilon_{\cal L}}
321: \;\;
322: %\exp{[
323: e^{\frac{1}{g}\sum_{j>k} \vec{L}_j \cdot \vec{L}_k
324: \ln{|\tau_j - \tau_k|}^2 }
325: %]}
326: \;.
327: \label{eq:Z_Y2}
328: \end{eqnarray}
329: \end{widetext}
330: Thus, the partition function $Z_Y$ is written in a similar form
331: to the so-called Coulomb gas (particles interacting with the
332: logarithmic potential)
333: in one dimension.
334: The difference from the standard Coulomb gas is in the
335: phase factor $e^{i\Upsilon_{\cal L}}$, which is indeed the consequence
336: of the Fermi statistics of the electron and of the Aharonov-Bohm phase.
337: We may call this a ``generalized Coulomb gas.''
338: In fact, below we show that this is indeed identical to the generalized
339: Coulomb gas which appeared in a quite different problem --
340: the dissipative Hofstadter model on a triangular lattice.
341:
342: \subsection{Generalized Coulomb gas from the dissipative Hofstadter model}
343:
344: Dissipative quantum mechanics (also known as
345: quantum Brownian motion)
346: is the problem
347: of the quantum motion of a single particle subject to
348: dissipation (see Ref.~\cite{Leggett-etal} for a review).
349: In the presence of any periodic potential, single-particle quantum mechanics
350: implies that the particle never localizes due to the tunneling effect.
351: However, if the dissipation is sufficiently strong, the
352: tunneling can be suppressed and the particle may be localized.
353: When the dissipation is Ohmic, there is a surprising mapping
354: of the problem to the boundary problem of the free boson field
355: theory. Here, the bulk of the free boson is related to
356: the heat bath in the dissipative quantum mechanics, and only the
357: boundary field is related to the original particle.
358: By integrating over the bulk degrees of freedom, the boundary
359: problem can be reduced to the one-dimensional Coulomb gas.
360:
361: Dissipative quantum mechanics in the presence of the periodic potential and
362: the magnetic field
363: is known as the dissipative Hofstadter problem.
364: It is again related to the boundary problem of the free boson,
365: or the following one-dimensional generalized Coulomb gas.
366: The ``free'' action for the dissipative Hofstadter model, namely
367: the effective action in the absence of the potential (but with
368: the magnetic field and the dissipation), is
369: \begin{equation}
370: S_0[\vec{X}] = \frac{1}{2} \int \frac{d\omega}{(2\pi)^2}
371: \left[
372: \alpha |\omega| \delta_{\mu \nu} + \beta \omega \epsilon_{\mu \nu} \right]
373: X^*_{\mu}(\omega) X_{\nu}(\omega),
374: \label{eq:S_0}\end{equation}
375: where $\mu,\nu=1,2$ and $\epsilon_{12}=-\epsilon_{21}=1$,
376: $\epsilon_{11}=\epsilon_{22}=0$. $\alpha$ and $\beta$ are
377: related to the dissipation and the magnetic field, respectively.
378: This determines the propagator:
379: \begin{eqnarray}
380: \lefteqn{
381: D_{\mu \nu}(\tau) = \langle X_{\mu}(\tau) X_{\nu}(0) \rangle = }
382: \nonumber \\
383: && \!\!\!\!\!\!
384: - \frac{\alpha}{\alpha^2 + \beta^2}\ln{\tau^2} \;\delta_{\mu \nu}
385: + i \pi \frac{\beta}{\alpha^2+\beta^2}\; \sgn{\,\tau} \;\epsilon_{\mu \nu} .
386: \label{eq:D}\end{eqnarray}
387: We now introduce a potential term with the periodicity of the
388: triangular lattice (notice the difference from the ``rectangular'' one
389: in Ref.~\cite{CF}), as
390: \begin{equation}
391: S_V[\vec{X}]
392: = - V e^{i \delta/3}\int d\tau \sum_{a=1}^3
393: e^{i \vec{K}_a \cdot \vec{X}} + \mbox{c.c.},
394: \label{eq:CFpot}
395: \end{equation}
396: where $V$ and $\delta$ are chosen to be real.
397: Expanding the partition function in powers of $V$, one obtains
398: \begin{widetext}
399: \begin{eqnarray}
400: Z_\shbox{DHM}=\sum_{n}
401: %\sum_{n=0}^\infty
402: %\frac{1}{n!}\;
403: \;V^n
404: \int_{\tau_{n-1}}^\infty \!\!\!\!\! d\tau_{n}\;
405: \int_{\tau_{n-2}}^{\tau_{n}} \!\!\!\!\! d\tau_{n-1}\;
406: \cdots
407: \int_{-\infty}^{\tau_{2}} \!\!\!\! d\tau_{1}\;
408: \sum_{\{L_j\}}
409: \left[
410: \kappa_{\vec L_n}
411: \cdots
412: \kappa_{\vec L_2}\;
413: \kappa_{\vec L_1}
414: \right]
415: \;
416: \langle
417: e^{i \vec{L}_{n} \cdot \vec{X}(\tau_n)}
418: \ldots
419: e^{i \vec{L}_{2} \cdot \vec{X}(\tau_2)}
420: e^{i \vec{L}_{1} \cdot \vec{X}(\tau_1)}
421: \rangle_0
422: \;,
423: \label{eq:Z_DHM}
424: \end{eqnarray}
425: \end{widetext}
426: where again
427: ${\vec L}_j=q_j{\vec K}_{a_j}$
428: is one of the six vectors
429: $\pm \vec{K}_{1,2,3}$
430: ($q_j=\pm 1$ and $a_j=1,2,3$), and the
431: $\kappa_{\vec L_j}=e^{iq_j\delta/3}$ in the $[\ldots]$ term keep track
432: of phase factors due to $\delta$. The correlation function is given by
433: \begin{eqnarray}
434: \lefteqn{\langle
435: e^{i \vec{L}_n \cdot \vec{X}(\tau_n)}
436: \ldots
437: e^{i \vec{L}_2 \cdot \vec{X}(\tau_2)}
438: e^{i \vec{L}_1 \cdot \vec{X}(\tau_1)}
439: \rangle_0 = } \nonumber \\
440: &\delta_K(\sum_j \vec{L}_j) & \exp{\Big[ \frac{\alpha}{\alpha^2+\beta^2}
441: \sum_{j>k} \vec{L}_j \cdot \vec{L}_k
442: \ln{|\tau_j - \tau_k|}^2 }
443: \nonumber \\
444: && \!\!\!\!\!\!\!\! \!\!\!\!\! \!\!\!\!\!
445: -i \pi \frac{\beta}{\alpha^2+\beta^2} \sum_{j>k}\vec{L}_j \times \vec{L}_k
446: \;\sgn(\tau_j - \tau_k) \Big].
447: \label{eq:loopCF}
448: \end{eqnarray}
449: (Because of the time ordering, notice that $\sgn(\tau_j - \tau_k)=1$
450: for all terms in the sum.)
451:
452: Notice that the $\kappa_{{\vec L}_j}$ in the $[\ldots]$ of
453: Eq.~(\ref{eq:Z_DHM}) commute with one another, in contrast to the
454: Klein factors $\zeta_{{\vec L}_j}$ in the $[\ldots]$ of
455: Eq.~(\ref{eq:Z_Y}). However, the $\langle\ \rangle_0$ term in
456: Eq.~(\ref{eq:loopCF}) contains phase factors not present in
457: Eq.~(\ref{eq:loop}). While the phases entering the perturbative
458: expansions for $Z_Y$ and $Z_\shbox{DHM}$
459: are arising from different sources,
460: their net effect turns out to be identical, as we show below.
461:
462: The phases in the expansion of the dissipative Hofstadter model can
463: also be associated with the closed paths ${\cal L}=({\vec L}_1,{\vec
464: L}_2,\ldots, {\vec L}_n)$. The contribution from the
465: $\vec{L}_j\times\vec{L}_k$ terms amounts to a flux through the
466: oriented area ${\cal A}_{\cal L}$ enclosed by ${\cal L}$.
467: To see this we again decompose any closed path into a set of
468: closed non-intersecting loops. The oriented area of such
469: a non-intersecting loop is
470: \begin{equation}
471: {\cal A}_{\cal L}=-\frac{1}{2}\sum_{j>k}\vec{L}_j\times\vec{L}_k
472: \; .
473: \end{equation}
474: Thus, the accumulated phase $\Lambda_{\cal L}$ due to the magnetic
475: flux is
476: \begin{eqnarray}
477: \Lambda_{\cal L}&&=2\pi \frac{\beta}{\alpha^2+\beta^2} \;{\cal A}_{\cal L}
478: \\
479: &&= 2\pi \frac{\beta}{\alpha^2+\beta^2}\; A \left[
480: N_\bigtriangleup({\cal L})+
481: N_\bigtriangledown({\cal L})
482: \right]
483: \;,
484: \end{eqnarray}
485: where ${A}=\frac{\sqrt{3}}{4}$
486: is the area of an elementary triangle,
487: and $N_\bigtriangleup({\cal L})$, $N_\bigtriangledown({\cal L})$
488: are the {\it net} numbers of enclosed elementary up and down triangles,
489: as before.
490:
491: Now we turn to the phase due to the $[\dots]$ in
492: Eq.~(\ref{eq:Z_DHM}). The total phase factor for a given closed path ${\cal
493: L}=({\vec L}_1,{\vec L}_2,\ldots, {\vec L}_n)$,
494: \begin{equation}
495: e^{i\;\Xi_{\cal L}}=
496: \left[
497: \kappa_{\vec L_n}
498: \cdots
499: \kappa_{\vec L_2}\;
500: \kappa_{\vec L_1}
501: \right]
502: \; ,
503: \end{equation}
504: is the sum of the phase contributions from all the elementary
505: triangles enclosed by the path. The argument for this is carried out
506: by induction, identically to what we have used for the products of the
507: $\zeta_{\vec{L}_j}$ above.
508:
509: The phase contributions from ``up'' and ``down'' triangles are again
510: different. The phase factor for a counterclockwise loop on the
511: up triangle is determined determined to be
512: \begin{equation}
513: %e^{i \xi_{\bigtriangleup}}=
514: e^{i \delta/3}\;e^{i \delta/3}\;e^{i \delta/3}=
515: e^{i \delta}\label{upDHM}
516: \end{equation}
517: while that for the counterclockwise loop on the down triangle is
518: \begin{equation}
519: %e^{i \xi_{\bigtriangledown}}=
520: e^{-i \delta/3}\;e^{-i \delta/3}\;e^{-i \delta/3}=
521: e^{-i \delta}\label{downDHM}
522: \;.
523: \end{equation}
524: Namely, loops on the triangular lattice pick up a phase as if there is
525: a staggered magnetic flux of $\delta$ and $-\delta$ in each elementary
526: triangle. The total phase accumulated due to the loop ${\cal L}$ is
527: thus
528: \begin{equation}
529: \Xi_{\cal L}=
530: N_\bigtriangleup({\cal L})\;\delta-
531: %\xi_\bigtriangleup+
532: N_\bigtriangledown({\cal L})\;\delta
533: %\xi_\bigtriangledown
534: \; ,
535: \end{equation}
536: where $N_\bigtriangleup({\cal L})$ and $N_\bigtriangledown({\cal L})$
537: are the net numbers of up and down triangles enclosed by the loop.
538:
539: The total phase factor due to the loop ${\cal L}$ is the sum
540: $\tilde\Upsilon_{\cal L}=\Lambda_{\cal L}+\Xi_{\cal L}$:
541: \begin{equation}
542: \tilde\Upsilon_{\cal L}=
543: N_\bigtriangleup({\cal L})\;\tilde\upsilon_\bigtriangleup+
544: N_\bigtriangledown({\cal L})\;\tilde\upsilon_\bigtriangledown
545: \; ,
546: \end{equation}
547: where
548: \begin{eqnarray}
549: \tilde\upsilon_\bigtriangleup&&=2\pi \frac{\beta}{\alpha^2+\beta^2} A+\delta
550: \label{vup}\\
551: \tilde\upsilon_\bigtriangledown&&=2\pi \frac{\beta}{\alpha^2+\beta^2} A-\delta
552: \; .\label{vdown}
553: \end{eqnarray}
554:
555: We can now rewrite the partition function for the dissipative
556: Hofstadter model on the triangular lattice as:
557: \begin{widetext}
558: \begin{eqnarray}
559: Z_\shbox{DHM}=\sum_{n}
560: %\sum_{n=0}^\infty
561: %\frac{1}{n!}\;
562: \;V^n
563: \int_{\tau_{n-1}}^\infty \!\!\!\!\! d\tau_{n}\;
564: \int_{\tau_{n-2}}^{\tau_{n}} \!\!\!\!\! d\tau_{n-1}\;
565: \cdots
566: \int_{-\infty}^{\tau_{2}} \!\!\!\! d\tau_{1}\;
567: \sum_{{\cal L}=(\vec L_1,\dots,\vec L_n)}
568: %\left[
569: %\zeta_{\vec L_n}(\tau_n)
570: %\cdots
571: %\zeta_{\vec L_2}(\tau_2)\;
572: %\zeta_{\vec L_1}(\tau_1)
573: %\right]
574: \delta_K(\sum_j \vec{L}_j)\;\;e^{i\tilde\Upsilon_{\cal L}}
575: \;\;
576: %\exp{[
577: e^{\frac{\alpha}{\alpha^2+\beta^2}\sum_{j>k} \vec{L}_j \cdot \vec{L}_k
578: \ln{|\tau_j - \tau_k|}^2 }
579: %]}
580: \;.
581: \label{eq:Z_DHM2}
582: \end{eqnarray}
583: \end{widetext}
584: Following Callan and Freed~\cite{CF},
585: we may regard the DHM partition function
586: as the partition function of a generalized classical 1-dimensional Coulomb
587: gas. $\tau_j$ is the position of the j$^{th}$ particle
588: and $\vec L_j$ is its generalized ``charge'', a vector quantity
589: taking on six possible values.
590:
591: \subsection{Mapping of the Y-junction to the dissipative Hofstadter model}
592: \label{subsec:mapping}
593: The expression for $Z_\shbox{DHM}$ in Eq.~(\ref{eq:Z_DHM2}) is identical to
594: that for $Z_{Y}$ in Eq.~(\ref{eq:Z_Y2}),
595: provided the following conditions are satisfied.
596: % Edited by MO -- 5 Sep 2005.
597: First we need the relation
598: \begin{equation}
599: \frac{\alpha}{\alpha^2+\beta^2} = \frac{1}{g},
600: \label{eq:alpha}
601: \end{equation}
602: for the scaling dimension of the leading perturbation to be matched
603: between the two theories.
604: Furthermore, in order to match the phase factor,
605: we require
606: $\tilde\Upsilon_{\cal L}\equiv\Upsilon_{\cal L} \mod{2\pi}$, or
607: equivalently
608: \begin{eqnarray}
609: &&\tilde\upsilon_\bigtriangleup \equiv
610: \upsilon_\bigtriangleup \; \mod{2\pi},
611: \label{eq:upcond}
612: \\
613: &&\tilde\upsilon_\bigtriangledown \equiv
614: \upsilon_\bigtriangledown \; \mod{2\pi}
615: \label{eq:downcond}
616: \;,
617: \end{eqnarray}
618: i.e.
619: \bea
620: {\sqrt{3}\pi \over 2}{\beta \over \alpha^2+\beta^2}+\delta
621: &=&\phi \ \ \mod{2\pi} , \nonumber \\
622: {\sqrt{3}\pi \over 2}{\beta \over \alpha^2+\beta^2}-\delta
623: &=&\pi -\phi \ \ \mod{2\pi} .
624: \eea
625: The above two equations imply:
626: \bea
627: \sqrt{3} \frac{\beta}{\alpha^2 + \beta^2} &=& (2n - 1) \nonumber \\
628: \delta &=& \phi + \pi (1/2-n).
629: \label{eq:n}\eea
630: for an arbitrary integer, $n$. Because the phases, $\phi$ and $\delta$, are
631: only defined modulo $2\pi$, there is actually an infinite number of
632: different choices of $\alpha$ and $\beta$ labeled by the integer $n$.
633: Eqs. (\ref{eq:alpha}) and (\ref{eq:n}) give:
634: \bea \alpha &=& {3g\over 3+(2n-1)^2g^2}\nonumber \\
635: \beta &=& {\sqrt{3}(2n-1)g^2\over 3+(2n-1)^2g^2}.\label{eq:ndef}\eea
636:
637: To understand the ambiguity in the integer $n$,
638: let us define the Coulomb gas charge density
639: \be \vec \rho (\tau )\equiv \sum_{j=1}^N\vec L_j \; \delta (\tau -\tau_j).\ee
640: The generalized ``energy'', $E$, of an
641: arbitrary classical gas configuration is determined by the $N$
642: positions, $\tau_j$ and charges, $L_j$. This generalized
643: ``energy'' is now a complex quantity and $\exp (-E/T)$
644: (where $T$ is the classical temperature) is given by Eq. (\ref{eq:Z_DHM}).
645: i.e.
646: \begin{widetext}
647: \be -\frac{E}{T}=\sum_{j=1}^n[\ln V +i\delta\, q_j/3] +
648: \frac{\alpha}{\alpha^2+\beta^2}
649: \sum_{j>k} \vec{L}_j \cdot \vec{L}_k
650: \ln |\tau_j - \tau_k|^2
651: -i \pi \frac{\beta}{\alpha^2+\beta^2} \sum_{j>k}\vec{L}_j \times \vec{L}_k
652: \;\sgn(\tau_j - \tau_k),
653: \ee \end{widetext}
654: where $q_j$ is defined by $\vec L_j\equiv q_j\vec K_{a_j}$ as before.
655: Since $\vec L_j\times \vec L_k$ is always $0$ or $\pm \sqrt{3}/2$,
656: $\exp [-E/T]$ is invariant under a shift:
657: \be {\pi \sqrt{3}\over 2}{\beta \over \alpha^2+\beta^2}\to
658: {\pi \sqrt{3}\over 2}{\beta \over \alpha^2+\beta^2}+2\pi .\ee
659: Eqs. (\ref{vup}), (\ref{vdown}) imply that there is also a symmetry under:
660: \bea {\pi \sqrt{3}\over 2}{\beta \over \alpha^2+\beta^2}&\to&
661: {\pi \sqrt{3}\over 2}{\beta \over \alpha^2+\beta^2}+\pi \nonumber \\
662: \delta &\to & \delta - \pi ,\eea
663: which corresponds to shifting the integer $n$ in Eq. (\ref{eq:ndef}) by $1$.
664: Since the energy of any configuration is independent of $n$,
665: it follows that the Coulomb gas charge density correlation function,
666: $\langle \rho_\mu (\tau )\rho_\nu (\tau ')\rangle$, is also
667: $n$-independent.
668:
669: On the other hand, the correlation function $\langle
670: X^\mu (\tau )X^\nu (\tau ')\rangle$, {\it does} depend on $n$.
671: These two correlation functions can be simply related to
672: each other by adding an external source, $\vec a(\tau )$, to the action,
673: coupling to the Coulomb gas charge density. That is,
674: we modify the energy of the Coulomb gas by:
675: \be -E/T \to -E/T -i\int d\tau\; \vec a(\tau )\cdot \vec \rho (\tau ).\ee
676: The Coulomb gas correlation function is given by the functional
677: derivative of the partition function with respect to the source term:
678: \begin{equation}
679: \frac{ \partial^2 \ln{Z}}{\partial a^{\mu}(\omega_n )
680: \partial a^{\nu}(\omega_n' )}
681: = - {1\over (2\pi )^2}\langle \rho_{\mu}(-\omega_n ) \rho_{\nu} (-\omega_n' )\rangle .
682: \label{eq:Zaa}
683: \end{equation}
684: Note that the Coulomb gas energy arises from the functional
685: integral over $\vec X$:
686: \be e^{-E/T}={1\over Z_0}
687: V^n e^{i(\delta /3)\sum_jq_j}
688: \int [d\vec X (\tau )]e^{-S_q(\vec X)}, \ee
689: where:
690: \be S_q \equiv S_0(\vec X)-i\int d\tau \vec X(\tau )\cdot \vec \rho (\tau ).
691: \ee
692: Here $Z_0$ is the functional integral with $S_q$ replaced by $S_0$,
693: which is given by Eq. (\ref{eq:S_0}).
694: Thus, adding a source term to the Coulomb gas corresponds to
695: a modification of the potential term in the DHM action:
696: \begin{equation}
697: S_V[\vec{X}]
698: \to - V e^{i \delta/3}\int d\tau \sum_{a=1}^3
699: e^{i \vec{K}_a \cdot [\vec{X}(\tau )+\vec a(\tau )]} + \mbox{c.c.},
700: \label{eq:CFpot2}\ee
701: Alternatively, we may shift $\vec X (\tau )\to \vec X (\tau )-\vec a(\tau)$,
702: removing the source term from the potential term in the DHM action
703: but inserting it into the non-interaction term, $S_0$:
704: \begin{widetext}
705: \be S_0\to {1\over 2}\int {d\omega_n \over (2\pi )^2}
706: [X^\mu (\omega_n )-a^\mu ]^*(\omega_n )
707: \;\left(D^{-1}\right)_{\mu \nu}(\omega_n )\;
708: [X^\nu (\omega_n )-a^\nu ] (\omega_n ).\ee
709: \end{widetext}
710: Here we have introduced the inverse of the DHM propagator, defined
711: in Eq. (\ref{eq:D}), which is simply the matrix appearing
712: in $S_0$, in Eq. (\ref{eq:S_0}):
713: \be \left( D^{-1}\right)_{\mu \nu}\equiv \alpha \delta_{\mu \nu}|\omega |
714: +\beta \epsilon_{\mu \nu}\omega .\ee
715: Taking a second derivative with respect to $a^{\mu}$ we see that:
716: \begin{widetext}
717: \be
718: \langle \rho_{\mu}(-\omega ) \rho_{\nu} (-\omega ' )\rangle
719: =\delta (\omega +\omega ')\left( D\right)^{-1}_{\mu \nu}(-\omega )
720: -{1\over (2\pi )^2}\left( D^{-1}\right)_{\mu \sigma}(-\omega )
721: \langle X^{\sigma}(-\omega )X^{\lambda}(-\omega ')\rangle
722: \left( D\right)^{-1}_{\lambda \nu}(\omega '),\label{eq:rrXX}
723: \ee
724: %\end{widetext}
725: or, equivalently:
726: %\begin{widetext}
727: \begin{equation}
728: \langle X^{\mu}(-\omega_n )X^{\nu}(-\omega_n ')\rangle
729: =(2\pi )^2\delta (\omega_n+\omega_n')
730: D^{\mu \sigma}(-\omega_n )
731: -(2\pi )^2 D^{\mu \sigma}(-\omega_n )
732: \langle \rho_{\sigma }(-\omega_n ) \rho_{\lambda} (-\omega_n ' )\rangle
733: D^{\lambda \nu}(\omega_n ')\label{eq:XX}
734: \end{equation}
735: \end{widetext}
736: Due to the complicated dependence of $\alpha$ and $\beta$ on $n$,
737: given by Eq. (\ref{eq:ndef}), and hence the complicated dependence of
738: $D^{\mu \nu}$ on $n$, we see that
739: $\langle X^{\mu}X^{\nu}\rangle$ is $n$-dependent.
740:
741: However,
742: what we are interested in here, for obtaining properties of the
743: Y-junction, is only the Coulomb gas correlation functions,
744: which are $n$-independent. Hence we may choose any value of
745: $n$ that we find convenient. It will sometimes be useful to
746: deduce the long time behavior of $\langle X^{\mu}X^{\nu}\rangle$
747: by RG and physical arguments and then use Eq.~(\ref{eq:rrXX})
748: to deduce the Coulomb gas correlation function. In so doing
749: we may choose any value of $n$. In general,
750: the behavior of $\langle X^{\mu}X^{\nu}\rangle$ will
751: depend strongly on $n$ but this $n$-dependence will cancel
752: when we compute $\langle \rho_{\mu}\rho_{\nu}\rangle$ using
753: Eq. (\ref{eq:rrXX}).
754:
755: Let us now use this mapping between the Y-junction problem and the
756: dissipative Hofstadter model and consider the case of $g>1$. Then, as
757: we have already discussed, the electron hopping is a relevant
758: perturbation in the disconnected limit. In the DHM, for any choice of
759: $n$, $V$ is a relevant perturbation for $g>1$. We would like to find
760: the infrared stable fixed point reached in the low energy limit. A
761: simple guess is that it occurs at $V\to \infty$. The stability of the
762: $V\to \infty$ fixed point can be determined using the instanton
763: method~\cite{CF}. Namely, in the strong potential limit, the
764: $\vec{X}$ field is pinned at one of the minima of the potential
765: (\ref{eq:CFpot}). The leading perturbation to this limit is given by
766: a tunneling process between the neighboring minima, represented by an
767: ``instanton.'' This is a classical solution, $\vec X_{\shbox{cl}}(\tau )$
768: which interpolates between two adjacent minima of the potential
769: as $\tau$ goes from $-\infty$ to $\infty$. The nature of
770: these solutions is discussed in Callan and Fried and, in more detail,
771: in Ref.~\cite{Freed-Harvey}. To determine the scaling dimension
772: of the tunneling operator we must consider the classical action
773: of a dilute multi-instanton configuration which tunnels successively
774: from minimum to minimum with each tunneling event well separated
775: in time from the rest. This instanton gas is dilute because,
776: in the limit of large $V$, the action of an instanton goes to
777: infinity, so the density of instantons goes to zero.
778: In the dilute limit, we can approximate
779: such a multi-instanton by a sum of the form:
780: \be \vec X(\tau ) \approx \vec X_0 +
781: \sum_{j=1}^n\vec X_{j}(\tau -\tau_j),\label{eq:inst}\ee
782: where
783: \bea \vec X_{j}(\tau ) &\to & 0,\ \ (\tau \to -\infty)\nonumber \\
784: &\to & \vec M_{j},\ \ (\tau \to \infty ).\eea
785: Here the $j^{\shbox{th}}$ tunneling event takes place at time $\tau_j$
786: and involves the particle tunneling between two
787: adjacent minima of $V(\vec X)$, displaced by $\vec M_{j}$. In
788: the dilute limit each function $\vec X_{j}(\tau )$ is a
789: single instanton solution. In the long time low energy limit
790: of interest to us, we may regard the single instanton solutions
791: as having the form:
792: \be \vec X_{a_j}(\tau ) \approx \vec M_{j} \,f(\tau ),\ee
793: for a real function, $f(\tau )$ which interpolates between
794: $0$ and $1$.
795: The Fourier transform of the multi-instanton
796: solution can then be approximated as:
797: \be \vec X(\omega )\approx \sum_j \vec M_{j}\,e^{i\omega \tau_j}
798: f(\omega ).\ee
799: The low frequency behavior of $f(\omega )$, which
800: follows from the long time asymptotic behavior of the interpolating
801: function $f(\tau)$, is given by:
802: \be f(\omega ) \to {1\over i\omega}
803: \label{eq:fas}.\ee
804: The important interaction term
805: in the instanton action comes simply from $S_0$ and is
806: obtained by substituting Eq. (\ref{eq:inst}) into $S_0$ in
807: Eq. (\ref{eq:S_0}). This gives the
808: instanton interaction term in the action:
809: \begin{widetext}
810: \be S_{\shbox{int}}\approx
811: \sum_{i>j}M_{i,\mu}M_{j,\nu}
812: \int {d\omega \over (2\pi )^2}\;e^{i\omega (\tau_j-\tau_i)}\;
813: \left[ \alpha |\omega |\delta_{\mu \nu}+\beta \epsilon_{\mu \nu}\omega\right]|f(\omega )|^2.
814: \ee \end{widetext}
815: In the dilute limit, where the time separations
816: are large, we may use the asymptotic form of $f(\omega )$ in
817: Eq. (\ref{eq:fas}).
818: In this we recover the Coulomb gas energy:
819: \begin{widetext}
820: \be S_{\shbox{int}}\approx -{\alpha \over (2\pi )^2}
821: \sum_{i>j}\vec M_i\cdot \vec M_j \ln
822: |\tau_i-\tau_j|^2 +{i\pi \beta \over (2\pi )^2}
823: \sum_{i>j}\vec M_i\times \vec M_j\;
824: \hbox{sgn} (\tau_i-\tau_j).\ee \end{widetext}
825: We see that this expansion is equivalent to the weak $V$ expansion
826: and therefore that the tunneling events are relevant only if
827: \be|\vec M_i|^2\alpha^2/(2\pi )^2 <1. \ee
828:
829: Thus we see that the separation of the minima of the potential,
830: $|\vec M_i|$ is a crucial quantity in determining the stability
831: of the localized fixed point. This separation is determined by
832: a simple classical analysis of the potential, $V(\vec X)$
833: in Eq. (\ref{eq:CFpot}). The stationary points of $V(\vec X)$
834: occur at the points:\be
835: {\vec X} = {4\pi \over 3}\sum_{i=1}^2 n_i\;\vec K_i, \ee
836: independent of $\delta$ and forming
837: a triangular lattice. However, the energies (values of $V$)
838: at these stationary
839: points depend on $\delta$. For general values of $\delta$, these
840: energies are different at the 3 sub-lattices into which the
841: triangular lattice can be partitioned:
842: \bea V(\vec 0) &=& -6V \cos (\delta /3)\nonumber \\
843: V\left( {4\pi}\;\vec K_1/3\right)
844: &=& -6V\cos [(2\pi -\delta )/3]\nonumber \\
845: V\left( -{4\pi}\;\vec K_2/3\right)&=& -6V\cos [(2\pi +\delta )/3].\eea
846: The sub-lattice containing $\vec 0$, is a triangular lattice
847: with basis vectors, $2\pi \vec R_i$ ($i=1,2$) with
848: \be \vec R_i\equiv (2/\sqrt{3})\;\vec K_i\times \hat z.
849: \label{eq:Rdef}\ee
850: Of course, the other sub-lattices are simply displaced from this
851: one by the fixed vectors, $(4\pi /3)\vec K_1$ and $(4\pi /3)\vec K_2$.
852: Therefore, the vectors $\vec M_i$ occurring in the instanton
853: solutions are $\pm 2\pi \vec R_1$, $\pm \vec 2\pi R_2$
854: and $\pm 2\pi (\vec R_1+\vec R_2)$ with length $4\pi /\sqrt{3}$.
855: As we vary $\delta$ the relative energies on the three sub-lattices
856: change.
857: % Edited by MO 5 Sep 2005
858: For $\delta =0$, the potential minima occur on the $\vec X=0$
859: sublattice. As we increase $\delta$ from zero,
860: the minima remains on the $\vec X=0$ sublattice until we
861: reach $\delta =\pi$. For $\pi < \delta < 2 \pi$, the minima
862: are rather on the $\vec X=(4\pi /3)\vec K_1$ sub-lattice.
863: Right at $\delta =\pi$, there are degenerate minima on both $\vec X=0$ and
864: $\vec X=(4\pi /3)\vec K_1$ sub-lattices. These two sub-lattices
865: define a honeycomb lattice with lattice spacing $4\pi /3$,
866: smaller by a factor of $1/\sqrt{3}$ compared to
867: that of the triangular lattices.
868:
869: Thus, from the dilute instanton analysis, we conclude that, for
870: general $\delta$,
871: the localized phase is stable when
872: \be \Delta\equiv {\alpha |2\pi R_i|^2\over (2\pi )^2}={4\alpha \over 3}
873: = {4g\over 3+(2n-1)^2g^2}>1.\ee
874: While this does not prove rigorously that
875: the DHM is in a localized phase for small $V$ when
876: this inequality is satisfied, it is
877: reasonable to believe that the system is in a localized
878: phase for large enough $V$, in this region of $\alpha$.
879:
880: For $n\neq 0$ or $1$, the localized phase is never stable,
881: for any value of $g$. This follows since the maximum
882: value of $\Delta$, with respect to $g$, for fixed $n$
883: is $2/[\sqrt{3}|2n-1|]$, which is $<1$ for $n\neq 0$ or $1$.
884: On the other hand, for $n=0$ or $1$, $\Delta <1$, for
885: $1<g<3$. Thus is it reasonable to assume that
886: the behavior of the Y-junction, at least for large $\Gamma$,
887: corresponds to the localized phase of the DHM in
888: the $n=0$, or $1$ representation. This result appears
889: at first peculiar because we might just as well have
890: chosen a different value of $n$, in which case
891: we would conclude that the DHM is {\it not} in the
892: localized phase. For $g>1$, the DHM is not in the
893: freely diffusing phase either, since $V$ is relevant.
894: It therefore must be in some intermediate phase which
895: is neither freely diffusing nor localized. Remarkably,
896: Eq. (\ref{eq:XX}) allows us to determine $\langle XX \rangle$
897: in these intermediate phases by determining $\langle \rho
898: \rho \rangle$ using $n=0$ or $1$~\cite{CF}.
899:
900: However, an ambiguity still remains because both $n=0$ and $1$
901: give stable localized fixed points. However, the localized
902: fixed points for these two values of $n$ predict different
903: $\langle \rho \rho \rangle$. Consistency requires that,
904: while both of these fixed points is stable at large $V$,
905: the RG flow goes all the way to the localized fixed point,
906: starting from a small $V$, for only one of these two
907: values of $n$. For the other value of $n$ the RG
908: flow should go to an intermediate $V$ fixed point, which
909: must give the same $\langle \rho \rho \rangle$.
910:
911: The particular cases $\phi = \pm \pi /2$, which maximally break
912: time-reversal, are instructive in understanding the resolution of this
913: ambiguity. From Eq. (\ref{eq:n}), for $\phi=\pi/2$, the choice $n=1$ ($\beta$
914: positive) gives $\delta=0$, and the potential minima forms a triangular
915: lattice as usual. However the other choice $n=0$ ($\beta$ negative) gives
916: $\delta=\pi$, for which the potential minima form a honeycomb lattice with
917: spacing $4\pi /3$, making the strong potential limit unstable for all $g$.
918: Similarly, for $\phi = - \pi/2$, $n=0$ ($\beta$ negative) is the unique
919: choice that gives a stable fixed point. This suggests that these localized
920: fixed points $\chi_{\pm}$ reflect the breaking of time reversal symmetry due
921: to the magnetic flux $\phi$. Indeed, the conductance at these fixed points
922: exhibits a chiral behavior breaking the time reversal invariance, as we show
923: below.
924:
925: Let us next consider the case when $\phi$ is very close to, but not
926: exactly equal to, $\pi /2$. If one chooses $n=1$, $\delta$ will be
927: close to zero, and the potential minima forms a triangular lattice
928: with a large separation between the true minima and secondary minima.
929: However, if one chooses $n=0$ instead of $n=1$, the minima of
930: the potential $V(\vec X)$ will form a triangular lattice with spacing
931: $4\pi /\sqrt{3}$, but there are local minima with only slightly larger
932: energy such that the distance between a true minimum and the nearest
933: local minima is $4\pi /3$. We could then envision another type of
934: approximate instanton solution where the particle tunnels from a true
935: minimum to a local minimum where it lingers for a long but finite time
936: but must eventually tunnel to a true minimum. The typical time that
937: the particle spends at a minimum is determined by the instanton
938: fugacity, and grows exponentially as $V\to \infty$. Thus this other
939: type of instanton might never be important for sufficiently large $V$.
940: However, for intermediate values of $V$, these instantons could play
941: an important role, effectively preventing the RG flow to go to
942: infinite $V$ for the $n=0$ choice. This argument suggests that the
943: choice of $n=1$ should represent the stable fixed point around a
944: vicinity of $\phi=\pi/2$. Of course, similar statements hold regarding
945: the case $\phi$ very close to $-\pi /2$, for which $n=0$ should
946: represent the stable fixed point.
947:
948: We are thus led to conjecture that the localized fixed point for $n=1$ ($n=0$)
949: corresponds to the stable fixed point of the Y-junction, for $1<g<3$,
950: even for small bare $V$, for a range of $\phi$ around $\pi /2$ ($-\pi /2$).
951: While none of our arguments have proven rigorously that the
952: RG flow really goes all the way from $V=0$ to $V=\infty$ (i.e. to the
953: localized phase) for any values of $g$ and $\phi$, this
954: seems like the most ``economical'' assumption since otherwise there
955: would necessarily be intermediate $V$ fixed points for all values of $n$.
956: On the other hand, for $\phi =0$ or $\pi$, the $n=0$ and $n=1$ fixed points
957: appear equally stable. Since, as we show in the next sub-section,
958: the conductance exhibits broken time-reversal symmetry for these localized
959: fixed points, we expect that neither of them is stable at $\phi =0$. There
960: must be some other stable fixed point, corresponding to intermediate $V$,
961: in this case. We refer to this fixed point as M (for mysterious).
962:
963: It is convenient to now think about the RG flow in the space of the
964: physical parameters of the Y-junction, $\Gamma$ and $\phi$. We have
965: proposed (unique) stable fixed points for $\phi$ near $\pi /2$ ($\chi_+$)
966: or $-\pi /2$ ($\chi_-$) and another fixed point, M, stable for $\phi =0$.
967: While the $\chi_{\pm}$ fixed points are stable against a small change in
968: the flux, we do not know whether the M fixed point has such a stability.
969: The simplest assumption is that it is does not. In this case the RG flows
970: might go to the
971: $\chi_{\pm}$ fixed points for any non-zero $\phi$, starting from
972: arbitrarily small $V$. Alternatively, it is possible that
973: the M fixed point is stable against adding a small flux. In that
974: case, there would have to be additional unstable fixed points
975: defining the boundaries between the basins of attraction of the
976: M, $\chi_+$ and $\chi_-$ fixed points. So again, an ``economy''
977: principle, suggests the simple picture with only 3 stable fixed points,
978: for $1<g<3$, as sketched in Fig.~\ref{fig:Chiflow}.
979:
980:
981: \subsection{Calculation of the conductance tensor}
982:
983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
984:
985: Let us calculate the conductance tensor $G_{jk}$
986: as defined in Eq.~(\ref{eq:def-of-conduc-tensor}) at
987: the chiral fixed points $\chi_{\pm}$.
988: There are two convenient ways of introducing the voltages drops. They
989: can either occur right at the junction, or else can be spread over finite
990: regions of the wires which are long compared
991: to microscopic length scales like $k_F^{-1}$
992: but could still be short compared to the length of the wires
993: and other long length scales such as an inelastic scattering length.
994: Both lead to the same results. The former
995: approach has the advantage that it allows the conductance tensor to be
996: expressed in terms of the Coulomb gas correlation function. The
997: latter approach is useful when the BCFT formalism is used and is the
998: only method at our disposal in the case of the Dirichlet fixed
999: point. We discuss each in turn, one in this section and the other in
1000: the next. For convenience, we set the electron charge, $e$, to $-1$. (In
1001: our convention the actual electron charge $e<0$.)
1002: We also set $\hbar =1$.
1003:
1004: We first consider the case where the voltage drop occurs right at the
1005: junction. We then modify the tunneling Hamiltonian by time-dependent
1006: phases, $A_j(t)$:
1007: \be H_T\to -\Gamma e^{i\phi /3}\sum_j\psi^\dagger_j\psi_{j-1}e^{i(A_j-A_{j-1})}
1008: +h.c.\label{HTA}\ee
1009: Note that, in the case where $A_j(t)=-V_jt$, that this
1010: corresponds to the replacement:
1011: \be \psi^\dagger_j(t)\to e^{-iV_jt}\psi^\dagger_j(t).\ee
1012: This corresponds to a potential $V_j$ on wire $j$, so we see that:
1013: \be dA_j/dt = -V_j.\ee
1014: It is convenient to define $I_j$ as
1015: the current on wire $i$, headed towards the junction:
1016: \be I_j= -dN_j/dt=i[N_j,H].
1017: \ee
1018: Here $N_j$ is the total number of particles on wire $j$. Since
1019: only the tunneling terms destroy conservation of particle
1020: number on each wire separately, we readily obtain:
1021: \begin{widetext}
1022: \be I_j=-i\Gamma\left[ e^{i\phi /3}\,e^{i(A_j-A_{j-1})}\,
1023: \psi^\dagger_j\psi_{j-1}
1024: +e^{-i\phi /3}\,e^{i(A_j-A_{j+1})}\,
1025: \psi^\dagger_j\psi_{j+1}\right] +h.c.\ee
1026: Expanding to first order in the $A_j$'s gives:
1027: \be \langle I_j \rangle = \langle I_j^0 \rangle
1028: -[(A_j-A_{j-1})+(A_j-A_{j+1})]\;
1029: E_T/3.\label{Ij}\ee
1030: Here $E_T$ is the tunneling energy:
1031: \be E_T=-\Gamma e^{i\phi /3}\sum_j
1032: \langle \psi_j^\dagger \psi_{j-1} \rangle +c.c.
1033: =-3\Gamma e^{i\phi /3} \langle \psi_j^\dagger \psi_{j-1} \rangle +c.c.\ee
1034: where the $Z_3$ symmetry was used in the last step. $I_j^0$ is
1035: the current operator to zero$^{\shbox{th}}$ order in the $A_k$:
1036: \be I_j^0\equiv -i\Gamma \left[e^{i\phi /3}\psi^\dagger_j\psi_{j-1}
1037: +e^{-i\phi /3}\psi^\dagger_j\psi_{j+1}\right] + h.c.\ee
1038: $ \langle I_j^0 \rangle $ is calculated to first order in $A_j(t)$ using:
1039: \be
1040: H\approx H_0+H'=H_0+\sum_j I_j^0\;A_j.
1041: \ee
1042: \be \langle I_j^0(t) \rangle =-i\int_{-\infty}^t dt'\; \langle [I_j^0(t),H'(t')] \rangle
1043: =-i\sum_k\int_{-\infty}^tdt'\; \langle [I_j^0(t),I_k^0(t')] \rangle \;A_k(t')
1044: =\sum_k \int_{-\infty}^\infty dt'\;K_{jk}^{\shbox{ret}}(t-t')\,A_k(t').\ee
1045: \end{widetext}
1046: Here $K_{ij}^{\shbox{ret}}$ is the retarded Green's function
1047: of the $I_j^0$ operators. Fourier transforming,
1048: \be A_j(\omega )\equiv \int dt \;e^{i\omega t}\;A_j(t),\ee
1049: and including the ``diamagnetic term'' in Eq. (\ref{Ij}), we obtain
1050: the Kubo formula in the form:
1051: \be \langle I_j \rangle (\omega )=\sum_k G_{jk}(\omega )\,V_k(\omega ).\,\ee
1052: where the conductance matrix is given by:
1053: \be G_{jk}(\omega )={1\over i\omega }\left[K_{jk}^{\shbox{ret}}
1054: (\omega )-C_{jk}\right],
1055: \label{eq:asp-cond}
1056: \ee
1057: and the $\omega$-independent symmetric matrix, $C_{jk}$ is given by:
1058: \bea C_{jj}&=&2E_T/3,\nonumber \\
1059: C_{jk}&=& -E_T/3\ \ (j\neq k).\eea
1060: Note that
1061: \be
1062: \sum_jC_{jk}=\sum_kC_{jk}=0
1063: \;,
1064: \ee
1065: which combined with Eqs.~(\ref{eq:conduc-aspen}) and (\ref{eq:asp-cond})
1066: implies
1067: \be \sum_jK_{jk}^{\shbox{ret}}
1068: =\sum_kK_{jk}^{\shbox{ret}}=0.\ee
1069: The retarded Green's function may be obtained by analytic continuation
1070: from the Matsubara Green's function:
1071: \be K^M_{jk}(\omega_n)\equiv -\int d\tau \;e^{i\omega_n\tau}\;
1072: \langle {\cal T}\, I^0_j(\tau )\,I^0_k(0) \rangle .\ee
1073: Here $\omega_n=2\pi nT$ for integer $n$,
1074: at temperature $T$, $\tau$ is imaginary
1075: time and ${\cal T}$ represents time ordering. The retarded Green's
1076: function is obtained, as usual, by:
1077: \be K^{\shbox{ret}}_{jk}(\omega)=\lim_{\delta \to 0^+}K^M_{jk}(\delta
1078: -i\omega).\ee
1079: The conductivity can be conveniently derived from the imaginary
1080: time path integral, with a vector potential, $A_j(\tau )$, defined
1081: as in Eq. (\ref{HTA}) which
1082: depends on imaginary time. The imaginary time partition function,
1083: $Z_I[A_j(\tau )]$, is obtained from the path integral of
1084: $\exp [-S_0-\int d\tau \;H_T(\tau )]$, where $S_0$ is the imaginary
1085: time action
1086: for the three independent leads. We see that:
1087: \be {\delta \ln Z_I\over \delta A_j(\tau )}=- \langle I_j(\tau ) \rangle \ee
1088: and, at $A_j\to 0$,
1089: \be {\delta^2 \ln Z_I\over \delta A_j(\tau )\,\delta A_k(\tau ')}
1090: =-K^{M}_{jk}(\tau -\tau ')+\delta (\tau -\tau ')C_{jk}.\ee
1091: Fourier transforming:
1092: \be {\delta^2 \ln Z_I\over \delta A_j(\omega_n) \,\delta A_k(\omega_n ')}
1093: ={\delta (\omega_n +\omega_n ')\over 2\pi}[-K^M_{jk}(\omega_n)+C_{jk}].\ee
1094: The conductivity can be obtained from this by analytic continuation.
1095: Explicitly, writing:
1096: \be {\delta^2 \ln Z_I\over \delta A_j(\omega_n ) \,\delta A_k(\omega_n ')}
1097: ={1\over (2\pi )^2}\delta (\omega_n +\omega_n ')Z_{jk}(\omega_n),\ee
1098: the dc conductance is given by:
1099: \be
1100: G_{jk}=\lim_{\omega_n\to 0}{1\over 2\pi
1101: \omega_n}Z_{jk}(\omega_n).\label{Zcond}\ee
1102: Here the limit $\omega_n\to 0$ must be taken by first continuing
1103: $\omega_n$ into the complex plane and then taking $\omega_n\to 0$
1104: along the imaginary axis with a small positive real part.
1105: This version of the Kubo formula is convenient because this second
1106: derivative of $\ln Z_I$ is proportional to
1107: the Coulomb gas density correlation function in the DHM. To make
1108: this connection it is convenient
1109: to rewrite the three vector potentials, $A_j(\tau )$
1110: in terms of a pair of vector potentials, $a^\mu (\tau )$ such that:
1111: \be A_j-A_k=\sum_l\epsilon_{jkl}\;\vec K_l\cdot \vec a.\label{a}\ee
1112: Here $\epsilon_{jkl}$ is the antisymmetric unit tensor
1113: ($\epsilon_{123}=1$ etc.) and the Greek indices run over $1$ and $2$ only.
1114: Then the tunneling term is modified to:
1115: \be H_T\to i\Gamma e^{i\phi /3}\sum_{j=1}^3\eta_ie^{i\vec K_j\cdot
1116: (\vec \Phi +\vec a)}+ h.c.\ee
1117: The corresponding modification of the DHM action is the shift:
1118: \be \vec X\to \vec X +\vec a.\ee
1119: This is precisely the source term coupled to the Coulomb gas
1120: density discussed above, which appeared in Eq. (\ref{eq:CFpot2}).
1121:
1122: We now see that the Y-junction conductivity is directly related to the Coulomb
1123: gas correlation function. To make the connection we use:
1124: \be \sum_jK^{\mu}_jK^{\mu}_j=(3/2)\delta^{\mu \nu}\ee
1125: to invert (\ref{a}):
1126: \be \vec a=\sum_{j,k,l}\vec K_j\epsilon_{jkl}(A_k-A_l)/3.\label{aA}\ee
1127: This implies:
1128: \be {\partial a^\mu \over \partial A_k}={2\over 3}\sum_{j,l}K^\mu_j
1129: \epsilon_{jkl}={2\over 3}\sum_jK^\mu_j\epsilon_{jk}.\ee
1130: Here we use an anti-symmetric $3\times 3$ tensor, $a_{jk}$, defined by:
1131: \be \epsilon_{ij}\equiv \sum_k\epsilon_{ijk}\ee
1132: with $\epsilon_{12}=\epsilon_{23}=\epsilon_{31}=1$.
1133: We may thus write:
1134: \begin{widetext}
1135: \begin{equation}
1136: \frac{ \partial^2 \ln{Z}}{\partial A_j(\omega_n ) \,\partial A_k(\omega_n')}
1137: = \frac{4}{9} \sum_{m,n,\mu,\nu}
1138: \epsilon_{jm} \epsilon_{kn} K_m^{\mu} K_n^{\nu}
1139: \; \frac{ \partial^2 \ln{Z}}{\partial a^{\mu}(\omega_n) \,\partial a^{\nu}
1140: (\omega_n')}.
1141: \label{eq:ZAA}
1142: \end{equation}
1143: The derivative on the right hand side of Eq. (\ref{eq:ZAA}) is proportional
1144: to the Coulomb gas correlation function. $\langle \rho_\mu (\omega_n)
1145: \rho_\nu (\omega_n')\rangle$ as we observed in Eq. (\ref{eq:Zaa}).
1146:
1147: There are two simple limits of the DHM. One is the case
1148: where the potential, $V$, is irrelevant. Then the correlators
1149: of the $X$-fields are determined by $S_0$:
1150: \be \langle X^{\sigma}(-\omega_n )X^{\lambda}(-\omega_n ') \rangle
1151: =(2\pi )^2 \delta (\omega_n +\omega_n')
1152: \;(D^{-1})^{\sigma \lambda}(-\omega_n ),\ee
1153: \end{widetext}
1154: and so the Coulomb gas correlation function vanishes. This corresponds
1155: to the limit of disconnected wires and zero conductance.
1156: The other simple limit the corresponds to the potential, $V$,
1157: renormalizing to infinity and pinning the $X$ co-ordinates at one
1158: of its minima. The Coulomb gas correlation function is then
1159: given by the first term only in Eq. (\ref{eq:rrXX}).
1160: Thus,
1161: \be Z_{jk}(\omega_n)=
1162: -\frac{4}{9} \sum_{m,n,\mu,\nu}
1163: \epsilon_{jm} \epsilon_{kn} K_m^{\mu} K_n^{\nu}\;
1164: D_{\mu \nu}(-\omega_n)
1165: .\ee
1166: Using
1167: this gives:
1168: \be Z_{jk}(\omega_n)={2\over 3}(3\delta_{jk}-1)\,\alpha \,|\omega_n|
1169: +{2\over \sqrt{3}}\beta \,\omega_n \,\epsilon_{jk}.\ee
1170: Thus, from Eq. (\ref{Zcond}), the dc conductivity is given by:
1171: \be G_{jk}={e^2\over h}\left[{2\over 3}(3\delta_{jk}-1)\,\alpha +
1172: {2\over \sqrt{3}}\,\beta\,
1173: \epsilon_{jk}\right] .\ee
1174: We have restored the factors of $e^2$ and $\hbar$ which we had previously set
1175: equal to one. The values of $\alpha,\beta$ for the chiral fixed points are
1176: obtained from Eq.~(\ref{eq:ndef}) with $n=1$ and $n=0$, and lead to the
1177: conductance tensor
1178: \begin{equation}
1179: G_{jk}^{\pm} =\frac{e^2}{h}\;
1180: {2g\over (3+g^2)} [(3 \delta_{jk} -1)
1181: \pm g \,\epsilon_{jk}],
1182: \label{eq:GDGA}
1183: \end{equation}
1184: with $\pm$ corresponding to the fixed points $\chi_{\pm}$. It is interesting
1185: to check that the same result can be obtained by applying electric fields
1186: uniformly to the entire wire $j$, which we do in the next section, where we
1187: address the Y-junction problem using BCFT.
1188:
1189:
1190:
1191:
1192:
1193:
1194:
1195:
1196: