1: \documentstyle[aps,eqsecnum,prb,multicol,epsf]{revtex}
2: %
3: %
4: \newcommand{\beq}{\begin{equation}}
5: \newcommand{\eeq}{\end{equation}}
6: \newcommand{\beqa}{\begin{eqnarray}}
7: \newcommand{\eeqa}{\end{eqnarray}}
8: \newcommand{\w}{\omega}
9: \newcommand{\W}{\Omega}
10: \newcommand{\nn}{{\bf n}}
11: \renewcommand{\r}{{\bf r}}
12: \newcommand{\q}{{\bf q}}
13: \newcommand{\p}{{\bf p}}
14: \newcommand{\Q}{{\bf Q}}
15: \newcommand{\sgn}{\mbox{sgn}}
16: \renewcommand{\>}{\rangle}
17: \newcommand{\<}{\langle}
18: \newcommand{\St}{\mbox{St}}
19: \renewcommand{\Re}{\mbox{Re} }
20: \renewcommand{\Im}{\mbox{Im} }
21: %
22: \begin{document}
23: %
24: %
25:
26: \title{Interaction corrections at intermediate temperatures:
27: Longitudinal conductivity and kinetic equation}
28: \author{G\'abor Zala, B.N. Narozhny, and I.L. Aleiner}
29: \address{Department of Physics and Astronomy, SUNY at Stony Brook,
30: Stony Brook, NY, 11794}
31: \date{\today}
32: %
33: \maketitle
34: \begin{abstract}
35: It is well known that electron-electron interaction in two dimensional
36: disordered systems leads to logarithmically divergent Altshuler-Aronov
37: corrections to conductivity at low temperatures ($T\tau\ll 1$; $\tau$
38: is the elastic mean-free time). This paper is devoted to the fate of
39: such corrections at intermediate temperatures $T\tau \gtrsim 1$. We
40: show that in this (ballistic) regime the temperature dependence of
41: conductivity is still governed by the same physical processes as the
42: Altshuler-Aronov corrections - electron scattering by Friedel
43: oscillations. However, in this regime the correction is linear in
44: temperature; the value and even the {\em sign} of the slope depends on
45: the strength of electron-electron interaction. (This sign change may
46: be relevant for the ``metal-insulator'' transition observed recently.)
47: We show that the slope is directly related to the renormalization of
48: the spin susceptibility and grows as the system approaches the
49: ferromagnetic Stoner instability. Also, we obtain the temperature
50: dependence of the conductivity in the cross-over region between the
51: diffusive and ballistic regimes. Finally, we derive the quantum
52: kinetic equation, which describes electron transport for arbitrary
53: value of $T\tau$.
54: \end{abstract}
55: \draft
56: \pacs{PACS numbers: 72.10.-d, 71.30.+h, 71.10.Ay }
57:
58:
59: \begin{multicols}{2}
60:
61: \section{Introduction}
62:
63: Temperature dependent corrections to conductivity due to
64: electron-electron interactions has been a subject of theoretical
65: \cite{aar,fin,Lee,stern,dstern,gdl,das}
66: and experimental \cite{aae,Bergman,cwh} studies for
67: more than two decades. Recently the interest in the matter was renewed
68: with appearance of new data \cite{new,nfl} showing a sign change in the
69: temperature dependence of conductivity in two dimensions
70: (2D). Theoretical discussions \cite{nfl} that followed emphasized the
71: question of whether that data indicated a non-Fermi liquid behavior.
72: However, the experiments were performed in a regime where the
73: temperature $T$ was of the same order of magnitude as the inverse
74: scattering time $\tau^{-1}$ (obtained from the Drude conductivity),
75: while pre-existing calculations were focused on the two limiting
76: cases: the diffusive regime \cite{aar,fin,Lee} $T\tau\ll 1$, and the
77: ballistic regime \cite{gdl,das} $T\tau\gg 1$.
78:
79: In the diffusive limit one finds\cite{aar,fin,Lee} for the logarithmically
80: divergent correction to the diagonal conductivity $\delta \sigma$:
81: \beq
82: \delta\sigma=-
83: \frac{e^2}{2\pi^2\hbar}\ln\left(\frac{\hbar}{T\tau}\right)
84: \Big[1 + 3\left(1-\frac{\ln(1+F_0^\sigma)}{F_0^\sigma}
85: \right) \Big],
86: \label{introdiff}
87: \eeq
88: where $F_0^\sigma$ is the interaction constant in the triplet
89: channel which depends on the interaction strength. It is clear,
90: that the sign of this logarithmically divergent correction may
91: be positive (metallic) or negative (insulating), depending on
92: the value of $F_0^\sigma$\cite{footnote1}.
93:
94: The result\cite{gdl,das}
95: for the ballistic region frequently cited in literature
96: reads
97: \beq
98: \delta\sigma= - \frac{e^2}{\pi\hbar}\left(\frac{T\tau}{\hbar}\right)
99: f(r_s),
100: \label{introbal}
101: \eeq where $f(r_s)$ is a positive function of the gas parameter of the
102: system, $r_s$. In a sharp contrast to Eq.~(\ref{introdiff}), equation
103: (\ref{introbal}) predicts always metallic sign of the interaction
104: correction.
105:
106: The absence of a rigorous calculation at intermediate temperatures,
107: $T\tau/\hbar \simeq 1$, may have contributed to the notion that those
108: two limits are governed by different physical processes. In this paper
109: we prove that notion erroneous: the results (\ref{introdiff}) and
110: (\ref{introbal}) are due to the same physical process, namely elastic
111: scattering of electrons by the self-consistent potential created by
112: all the other electrons. Therefore, these two different expressions
113: are in fact the two limits of a single interaction correction. We
114: calculate the correction within assumptions of the Fermi liquid theory
115: (other limitations of our approach we discuss below) and present the
116: cross-over function between the diffusive and ballistic limits.
117:
118: Moreover, we show that the existing theory for the ballistic limit
119: (\ref{introbal}) is incomplete. First, the results of
120: Ref.~\onlinecite{stern,dstern,gdl,das} account only for Hartree-like
121: interaction terms missing the exchange or Fock terms. Second, this
122: theory essentially employs a perturbative expansion in terms of the
123: interaction strength, which breaks down for stronger coupling. Both
124: issues lead to the change in the theoretical prediction even on a
125: qualitative level.
126:
127: The consequence of the first point is that the correction to
128: conductivity (\ref{introbal}) is always negative unlike the correction
129: in the diffusive limit that changes sign depending on the value of
130: $F_0^\sigma$. This sign change is due to competition between the
131: universal (and positive) Fock correction and the coupling-specific
132: (and negative) Hartree contribution. If the Fock potential (or, to be
133: more precise, singlet channel) is properly taken into account, then
134: the sign of the correction in the ballistic limit is also not
135: universal (being positive for weak interaction in contrast to
136: Refs.~\onlinecite{stern,dstern,gdl,das}).
137:
138: As follows from the second point, for the stronger interaction the
139: Hartree correction should be modified to include higher order
140: processes. For this case we show (see Section~\ref{diagrams}) that in
141: fact it should be replaced by the triplet channel correction, which is
142: characterized by the Fermi liquid constant $F_0^\sigma$. This constant
143: measures the strength of the spin-exchange interaction. If $F_0^\sigma
144: < 0$, the spin-exchange interaction tends to align electron spins and
145: (if it is strong enough) leads to the ferromagnetic Stoner
146: instability\cite{footnote1}. Even though this constant is unknown, it
147: can be found experimentally by means of independent measurement of the
148: spin susceptibility of the system. As a function of temperature the
149: interaction correction to conductivity is almost always monotonous,
150: except for a narrow region of parameters (where it is so small that it
151: can hardly be observed).
152:
153: The remainder of the paper is organized as follows: the following
154: section is devoted to qualitative discussion of the physics
155: involved. In the same section we summarize our results. Then we
156: present two alternative approaches to the microscopic calculation. In
157: Section~\ref{diagrams} we use the traditional perturbation
158: theory~\cite{aar} to derive the results presented in
159: Section~\ref{qualitative}, while in Section~\ref{eilenberger} the same
160: results are obtained using the quantum kinetic equation that we
161: derive. The advantage of the kinetic equation approach is that it can
162: be readily used to discuss the temperature behavior of quantities
163: other than conductivity. These results are advertised in Conclusions
164: and will be published elsewhere \cite{pre}.
165:
166:
167:
168:
169: \section{Qualitative discussion and results}
170: \label{qualitative}
171:
172:
173:
174: In this section we describe the scattering processes contributing to
175: the temperature dependence of conductivity. We show that unlike the
176: standard Fermi liquid $T^2$ corrections, the leading correction to
177: conductivity is accumulated at large distances, of the order
178: $v_F/min(T,\sqrt{T/\tau})$. In the ballistic limit such correction is
179: linear in temperature and we derive this result here using a text-book
180: quantum mechanical approach. The diffusive limit is discussed in
181: detail in Ref.~\onlinecite{aar}. The resulting correction
182: $\delta\sigma\sim\ln T$ seems to be rather different from the linear
183: one, but we show that both corrections arise due to the same physics -
184: coherent scattering by Friedel oscillations. Throughout the paper we
185: keep the units such that $\hbar = 1$, except for the final answers.
186:
187:
188:
189: \subsection{Scattering by Friedel oscillations}
190: \label{friedel}
191:
192:
193:
194: We start with the simplest case of a weak short-range interaction
195: $V_0(\vec r_1 - \vec r_2)$ and show how one can obtain the correction
196: to conductivity in the ballistic limit, i.e. due to a single
197: scatterer. This discussion is similar to that of
198: Ref.~\onlinecite{rag}, where the correction to the one-particle
199: density of states (DoS) was discussed, and also of
200: Ref.~\onlinecite{aag}, which describes the correction to the
201: conductivity in the diffusive limit.
202:
203: Consider a single impurity localized at some point, taken as the
204: origin. The impurity potential $U(\vec r)$ induces a modulation of
205: electron density close to the impurity. The oscillating part of the
206: modulation is known as the Friedel oscillation, which in 2D can be
207: written as
208:
209: \begin{equation}
210: \delta\rho(\vec r) = - \frac{\nu\lambda}{2\pi r^2} \sin(2k_F r).
211: \label{frie}
212: \end{equation}
213:
214: \noindent
215: Here $r$ denotes the distance to the impurity and its potential is
216: treated in the Born approximation $\lambda=\int U(\vec r)d\vec r$. In
217: 2D the free electron DoS is given by $\nu=m/\pi\hbar^2$ and $m$ is the
218: electron mass, $k_F$ is the Fermi momentum.
219:
220: Taking into account electron-electron interaction $V_0(\vec r_1 - \vec
221: r_2)$ one finds additional scattering potential due to the Friedel
222: oscillation Eq.~(\ref{frie}). This potential can be presented as a sum
223: of the direct (Hartree) and exchange (Fock) terms \cite{kit}
224:
225: \begin{mathletters}
226: \begin{equation}
227: \delta V(\vec r_1, \vec r_2) = V_H(\vec r_1)
228: \delta(\vec r_1 - \vec r_2)-V_F(\vec r_1, \vec r_2);
229: \end{equation}
230: \begin{equation}
231: V_H(\vec r_1) = \int d\vec r_3
232: V_0(\vec r_1 - \vec r_3)\delta\rho(\vec r_3);
233: \label{hp}
234: \end{equation}
235: \begin{equation}
236: V_F(\vec r_1, \vec r_2) = \frac{1}{2}V_0(\vec r_1 - \vec r_2)
237: \delta n(\vec r_1, \vec r_2),
238: \label{fp}
239: \end{equation}
240: \label{hf}
241: \end{mathletters}
242:
243: \noindent
244: where by $\rho(\vec r)$ we denote diagonal elements of the one
245: electron density matrix $n$,
246:
247: \begin{equation}
248: n(\vec r_1, \vec r_2) = \sum\limits_k \Psi^*_k(\vec r_1)
249: \Psi_k(\vec r_2).
250: \label{dm}
251: \end{equation}
252:
253: \noindent
254: The factor $1/2$ indicates that only electrons with the same spin
255: participate in exchange interaction. As a function of distance from
256: the impurity the Hartree-Fock energy $\delta V$ oscillates similarly
257: to Eq.~(\ref{frie}).
258:
259: {
260: \narrowtext
261: \begin{figure}[ht]
262: \vspace{0.5 cm}
263: \epsfxsize=6 cm
264: \centerline{\epsfbox{fock1.eps}}
265: \vspace{0.5cm}
266: \caption{Scattering by the Friedel oscillation. Interference between
267: the two paths A and B contributes mostly to backscattering. The Friedel
268: oscillation is created due to backscattering on the impurity, path C.}
269: \label{2}
270: \end{figure}
271: }
272:
273: The leading correction to conductivity is a result of interference
274: between two semi-classical paths depicted on Fig.~\ref{2}. If an
275: electron follows path ``A'', it scatters off the Friedel oscillation
276: created by the impurity and path ``B'' corresponds to scattering by
277: the impurity itself. Interference is most important for scattering
278: angles close to $\pi$ (or for backscattering), since the extra phase
279: factor accumulated by the electron on path ``A'' ($e^{i2kR}$ with $R$
280: being the length of the extra path interval relative to ``B'' and $2k$
281: being the difference between initial and final momenta for that extra
282: path interval) is canceled by the phase of the Friedel oscillation
283: $e^{-i2k_FR}$ so that the amplitudes corresponding to the two paths
284: are coherent. As a result, the probability of backscattering is
285: greater than the classical expectation (taken into account in the
286: Drude conductivity). Therefore, taking into account interference
287: effects leads to a correction to conductivity. We note that the
288: interference persists to large distances, limited only by temperature
289: $R\approx 1/|k-k_F|\le v_F/T$. Thus there is a possibility for the
290: correction to have a non-trivial temperature dependence. The sign of
291: the correction depends on the sign of the coupling constant that
292: describes electron-electron interaction.
293:
294: To put the above argument onto more rigorous footing and to find the
295: temperature dependence of the correction consider now a scattering
296: problem in the potential Eq.~(\ref{hf}). Following the textbook
297: approach \cite{llq}, we write a particle's wave function as a sum of
298: the incoming plane wave and the out-coming spherical wave (in 2D it is
299: given by a Bessel function, which we replace by its asymptotic form)
300:
301: \begin{eqnarray}
302: \Psi = e^{i\vec k\cdot\vec r}
303: + if(\theta)\sqrt{\frac{2\pi}{kr}}e^{ikr}.
304: \label{wv}
305: \end{eqnarray}
306:
307: \noindent
308: Here $f(\theta)$ is the scattering amplitude, which we will discuss in
309: the Born approximation. For the impurity potential itself the
310: amplitude $f(\theta)$ weakly depends on the angle. At zero temperature
311: it determines the Drude conductivity $\sigma_D$, while the leading
312: temperature correction is proportional to $T^2$, as is usual for Fermi
313: systems. We now show that this is not the case for the potential
314: Eq.~(\ref{hf}). In fact, taking into account Eq.~(\ref{hf}) leads to
315: enhanced backscattering and thus to the conductivity correction that
316: is linear in temperature.
317:
318: First, we discuss the Hartree potential Eq.~(\ref{hp}). Far from the
319: scatterer the wave function of a particle can be found in the first
320: order of the perturbation theory as $\Psi = e^{i\vec k\cdot\vec r} +
321: \delta\Psi (\vec r)$, where the correction is given by \cite{llq}
322:
323: \begin{equation}
324: \delta\Psi (\vec r) = i\int d\vec r_1 V_H(\vec r_1)
325: e^{i\vec k\cdot\vec r_1}
326: \sqrt{\frac{2\pi}{k|\vec r -\vec r_1|}}e^{ik|\vec r -\vec r_1|}.
327: \label{dpsi}
328: \end{equation}
329:
330: \noindent
331: Here $|\vec r -\vec r_1|\approx r - \vec r \cdot\vec r_1/r$, since we
332: are looking at large distances. Substituting the form of the potential
333: Eq.~(\ref{hp}) and introducing the Fourier transfer of the
334: electron-electron interaction $V_0$ we can rewrite Eq.~(\ref{dpsi}) as
335:
336: \begin{equation}
337: \delta\Psi (\vec r) = -i\frac{\nu\lambda}{\sqrt{2\pi}}V_0(q)
338: \frac{e^{ikr}}{\sqrt{kr}}\int \frac{d\vec r_1}{r_1^2} \sin (2k_Fr_1)
339: e^{i\vec q\cdot\vec r_1},
340: \label{dp1}
341: \end{equation}
342:
343: \noindent
344: where
345:
346: \[
347: \vec q = \vec k - k\vec r/r, \quad |q| =
348: 2k\sin(\theta /2),
349: \]
350:
351: \noindent
352: with $\theta$ being the angle of scattering. Comparing to
353: Eq.~(\ref{wv}) we find the scattering amplitude as a function of
354: $\theta$ (it also depends on the electron's energy $\epsilon =
355: k^2/2m$)
356:
357: \begin{equation}
358: f(\theta) = - \frac{\nu\lambda}{2\pi}V_0(q)
359: \int \frac{d\vec r}{r^2} \sin(2k_Fr)
360: e^{i\vec q\cdot\vec r}.
361: \label{ft1}
362: \end{equation}
363:
364: \noindent
365: The integral can be evaluated exactly\cite{gri} and the result is
366: given by
367:
368: \begin{equation}
369: f(\theta) = - \frac{\nu\lambda}{2\pi}V_0(q)
370: \left\{
371: \matrix{
372: \frac{\pi}{2}, & |q|<2k_F ;\cr
373: \arcsin\left(\frac{2k_F}{q}\right), & |q|>2k_F .\cr
374: }
375: \right.
376: \label{ft}
377: \end{equation}
378:
379: \noindent
380: Let us examine this expression more closely. Since ${|q|\le 2k}$, the
381: scattering amplitude Eq.~(\ref{ft}) for small $k$ weakly depends on
382: the angle through the Fourier component of the interaction $V_0(q)$,
383: see background value of $f(\theta)$ on Fig.~\ref{1}. However, we are
384: dealing with electronic excitations close to the Fermi surface, so in
385: fact $k$ is close to $k_F$, $|k-k_F|/k_F \ll 1$. If $k>k_F$, then the
386: scattering amplitude Eq.~(\ref{ft}) has a non-trivial angular
387: dependence around $\theta = \pi$ shown on Fig.~\ref{1}.
388:
389: According to Eq.~(\ref{ft}) such dependence is only possible in the
390: region $|q| > 2k_F$. This translates into the condition $|\theta -
391: \pi| < [2(k-k_F)/k_F]^{1/2}$, which determines the singular dependence
392: of the width of the feature in the scattering amplitude on the energy
393: of the scattered electron. Finally, using the fact that $\arcsin(1-x)
394: = \pi/2 - \sqrt{2x}$, we find that the dependence of the height of the
395: feature in the scattering amplitude is also singular: $\delta
396: f(\theta) \simeq [(k-k_F)/k_F]^{1/2}$.
397:
398: {
399: \narrowtext
400: \begin{figure}[ht]
401: \vspace{0.5 cm}
402: \epsfxsize=8 cm
403: \centerline{\epsfbox{fock2.eps}}
404: \vspace{0.5cm}
405: \caption{Scattering amplitude. The singularity for backscattering
406: is due to interference of paths ``A'' and ``B'' on Fig. 1.}
407: \label{1}
408: \end{figure}
409: }
410:
411: The transport scattering rate $\tau^{-1}$ is determined by the
412: scattering cross-section and can be found with the help of the
413: amplitude Eq.~(\ref{ft}), as well as the constant amplitude $f_0$ of
414: the scattering by the impurity itself
415:
416: \begin{equation}
417: \tau^{-1}(\epsilon) = \int \frac{d\theta}{2\pi}
418: (1-\cos\theta) |f_0 + f(\theta)|^2.
419: \label{t1}
420: \end{equation}
421:
422: \noindent
423: The leading energy dependence of $\tau^{-1}$ comes from the
424: interference term, which is proportional to $f(\theta)$. Then
425: integration around $\theta=\pi$ is dominated by the feature of
426: $f(\theta)$ resulting in a term of order $(\epsilon -
427: \epsilon_F)/\epsilon_F$. It is this term that gives rise to the linear
428: temperature dependence we are after. Since we are interested in this
429: leading correction only, in all other terms we can set $k\approx k_F$
430: and write the scattering rate as
431:
432: \begin{equation}
433: \tau^{-1}(\epsilon) = \tau^{-1}_0 + \frac{\nu\lambda}{2}V_0(2k_F)
434: \frac{\epsilon - \epsilon_F}{\epsilon_F}
435: \eta(\epsilon - \epsilon_F)f_0.
436: \label{te}
437: \end{equation}
438:
439: \noindent
440: Here $\eta(x)$ is the Heaviside step function and $\tau^{-1}_0$ is the
441: zero-temperature rate that determines the Drude conductivity (indeed,
442: the $\theta = \pi$ feature in $f(\theta)$ only exists for $k>k_F$ and
443: at $T=0$ there are no electrons with $k>k_F$).
444:
445: To obtain the scattering time we have to integrate the
446: energy-dependent rate Eq.~(\ref{te}) with the derivative of the Fermi
447: distribution function $n_F(\epsilon)$
448:
449: \begin{eqnarray*}
450: \tau = \int d\epsilon \tau(\epsilon)
451: \frac{\partial}{\partial\epsilon} n_F(\epsilon).
452: \end{eqnarray*}
453:
454: \noindent
455: Then the second term in Eq.~(\ref{te}) leads to a linear correction to
456: the Drude conductivity, small as $T/\epsilon_F$. However this is not
457: the only contribution to the temperature dependence. At finite
458: temperatures we also have to modify the Friedel oscillation
459: Eq.~(\ref{frie}) as follows:
460:
461: \begin{eqnarray*}
462: \delta\rho(\vec r) = -
463: \frac{\nu\lambda T^2}{2\pi v_F^2 \sinh^2\left(\frac{rT}{v_F}\right)}
464: \sin(2k_F r).
465: \end{eqnarray*}
466:
467: \noindent
468: Consequently, the scattering amplitude Eq.~(\ref{ft1}) becomes
469: temperature dependent
470:
471: \begin{eqnarray}
472: &&f(\theta) = - \frac{\nu\lambda}{2\pi}V_0(q)
473: \int \frac{d\vec r_2}{r_2^2} e^{-2r_2\frac{T}{v_F}}\sin(2k_Fr_2)
474: e^{i\vec q\cdot\vec r_2}
475: \nonumber\\
476: &&
477: \nonumber\\
478: &&\quad\quad\quad\quad\quad
479: = - \frac{\nu\lambda}{2\pi}V_0(q)
480: \arcsin\left(\frac{4k_F}{p}\right);
481: \label{ftt}
482: \\
483: &&
484: \nonumber\\
485: &&
486: p=\sqrt{\left(\frac{2T}{v_F}\right)^2 + (q+2k_F)^2}+
487: \sqrt{\left(\frac{2T}{v_F}\right)^2 + (q-2k_F)^2}.
488: \nonumber
489: \end{eqnarray}
490:
491: \noindent
492: Neglecting the small temperature dependent term in the denominator in
493: Eq.~(\ref{ftt}) brings us back to Eq.~(\ref{ft}). Keeping this term
494: leads to the same feature in $f(\theta)$ as the one on Fig.~\ref{1},
495: only now its width and magnitude are proportional to $\sqrt{T}$. The
496: resulting correction to the conductivity is therefore similar to the
497: one discussed above. Up to a numerical coefficient,
498:
499: \begin{equation}
500: \frac{\delta\sigma}{\sigma_D} = - 2 \nu V_0(2k_F)\frac{T}{\epsilon_F}.
501: \label{ds1}
502: \end{equation}
503:
504: The conductivity Eq.~(\ref{ds1}) is the same correction as the one
505: calculated in Ref.~\onlinecite{gdl}, see also Eq.~(\ref{introbal}), up
506: to a numerical factor. It is also clear that Eq.~(\ref{ds1}) is not
507: the full story. We have forgotten about the Fock part of the potential
508: Eq.~(\ref{hf})! Substituting Eq.~(\ref{wv}) into Eq.~(\ref{dm}), we
509: find the perturbation of the density matrix [which appears in the Fock
510: potential Eq.~(\ref{fp})] $\delta n(\vec r_1, \vec r_2) \approx
511: \delta\rho[(\vec r_1+\vec r_2)/2]$. Then the argument can be repeated.
512: The only difference is that the leading temperature correction comes
513: from the Fourier component at $q=0$, rather than $q=2k_F$. What is
514: most important, the Fock potential enters with the opposite
515: sign. Therefore the expression for the conductivity Eq.~(\ref{ds1})
516: has to be corrected
517:
518: \begin{equation}
519: \sigma = \sigma_D \left[1 - \nu \Big(2V_0(2k_F)-V_0(0)\Big)
520: \frac{T}{\epsilon_F}\right].
521: \label{ds}
522: \end{equation}
523:
524: \noindent
525: The sign of the correction is thus not universal and depends on the
526: details of electron-electron scattering. If the weak interaction is
527: reasonably long-ranged, then $V_0(0)\gg V_0(2k_F)$, so that
528: the correction in Eq.~(\ref{ds}) has the sign opposite to that in
529: Ref.~\onlinecite{gdl}.
530:
531: {
532: \narrowtext
533: \begin{figure}[ht]
534: \vspace{0.5 cm}
535: \epsfxsize=7 cm
536: \centerline{\epsfbox{fock3.eps}}
537: \vspace{0.5cm}
538: \caption{Scattering process with two impurities and the Friedel
539: oscillation. Scattering to all angles is affected by interference.
540: The relevant Friedel oscillation is created by the self-intersecting
541: path C.}
542: \label{3}
543: \end{figure}
544: }
545:
546: So far we have considered the effect of a single impurity. The
547: extension of the above arguments to the case of many impurities is
548: straightforward. In particular, one can consider a scattering process,
549: which involves two impurities and the Friedel oscillation shown on
550: Fig.~\ref{3}. It is clear that this process contributes to the
551: scattering amplitude at any angle, and not just for backscattering as
552: the single impurity process on Fig.~\ref{2} (which is typical for the
553: diffusive motion of electrons). Such processes were discussed in
554: detail (although using a slightly different language) in
555: Ref.~\onlinecite{aag}. Scattering by Friedel oscillations created by
556: multiple impurities results in a conductivity correction
557: (\ref{introdiff}) that is logarithmic in temperature and is typical
558: for 2D diffusive systems \cite{aar}.
559:
560: Comparing the scattering processes on Figs.~\ref{2} and \ref{3}, one
561: can clearly see that conductivity corrections, which arise from these
562: processes are governed by the same physics: coherent scattering by the
563: Friedel oscillation, which means that the ballistic and diffusive
564: regions should be analyzed on the same footing. In the next
565: subsection we present the results of such anlysis, postponing
566: the actual calculations until Secs.~\ref{diagrams} and
567: \ref{eilenberger}.
568:
569:
570:
571: \subsection{Results}
572: \label{results}
573:
574:
575:
576: Let us first consider the case of a weak, short range interaction
577: potential. Then the interaction can be treated in the lowest order of
578: perturbation theory, so that the resulting correction is proportional
579: to the interaction constant:
580:
581: \begin{eqnarray}
582: \delta\sigma_w = \frac{e^2}{\pi\hbar}
583: \left[
584: %\sigma_D\frac{T}{E_F}
585: \gamma_1 \frac{T\tau}{\hbar}
586: \left[ 1 -\frac{3}{8} w(T\tau)\right]
587: -\frac{\gamma_2}{4\pi}\ln\frac{E_F}{T}\right].
588: \label{sw}
589: \end{eqnarray}
590:
591: \noindent
592: Similarly to Eq.~(\ref{ds}), it can be written as a sum of Hartree and
593: Fock contributions (similar expression for $\gamma_1$ in one-dimensional
594: systems was obtained in Ref.~\onlinecite{gla}):
595:
596: \begin{eqnarray}
597: &&\gamma_1 = \nu \Big[V_0(0)-2V_0(2k_F)\Big],
598: \nonumber\\
599: &&
600: \nonumber\\
601: &&
602: \gamma_2 = \nu \Big[V_0(0)-2\langle V_0(k)\rangle_{FS}\Big],
603: \label{u}
604: \end{eqnarray}
605:
606: \noindent
607: where $\langle\dots\rangle_{FS}$ stands for the average over the Fermi
608: surface. Here we kept the notation for the electron-electron
609: interaction adopted in the previous Section. Then the Hartree
610: correction is proportional to the Fourier component of $V_0(q)$ at
611: $q=2k_F$, while for the Fock correction $q=0$. The two corrections
612: have different sign as we discussed above. The extra factor of $2$ in
613: the Hartree correction is due to electron spin degeneracy.
614:
615: Note, that Eq.(\ref{sw}) is defined only up to a
616: temperature-independent constant which is determined by the
617: ultraviolet contribution. We have chosen the argument of the
618: logarithm to be $E_F/T$ instead of the usual $1/T\tau$ to emphasize
619: that contrary to the naive expectations the logarithmic term persists
620: up to temperatures much larger than $1/\tau$, see also
621: Ref.~\onlinecite{rag}.
622:
623: The different expressions for the Hartree terms in $\gamma_1$ and
624: $\gamma_2$ are related to the fact that the single impurity
625: scattering, see Fig.~\ref{2}, and multiple impurity case, see
626: Fig.~\ref{3}, allow for different possible scattering angles.
627:
628: The dimensionless function $w(T\tau)$ describes the crossover between
629: ballistic and diffusive regimes. In the ballistic limit $T\tau\gg 1$
630: it vanishes as
631:
632: \begin{eqnarray*}
633: w(x\gg 1) \approx \frac{8}{3\pi x}\left[\ln(2x)-\frac{1}{4}
634: (\ln x -1)(6\ln 2 -1)\right].
635: %\label{wb}
636: \end{eqnarray*}
637:
638: \noindent
639: In the opposite limit $T\tau\ll 1$ it approaches a constant value
640: (${\cal C}\approx 0.577\dots$ is the Euler's constant and
641: $\zeta^\prime(x)$ is a derivative of the Riemann zeta function),
642:
643: \begin{eqnarray*}
644: w(x\ll 1) \approx && 1 +
645: \frac{2\pi x}{9}
646: \left(\ln x - \ln 2 - {\cal C} + \frac{3}{4}
647: + 6\zeta^\prime(2)\right),
648: %\label{wd}
649: \end{eqnarray*}
650:
651: \noindent
652: so that the linear correction does not completely vanish in the
653: diffusive limit, but competes with the logarithmic term and in
654: semiconductor structures with low Fermi energy it might be important
655: except for the lowest temperatures. The full function $w(x)$ is
656: plotted on Fig~\ref{plot:w}.
657:
658: {
659: \narrowtext
660: \begin{figure}[ht]
661: \vspace{0.5 cm}
662: \epsfxsize=7 cm
663: \centerline{\epsfbox{f4.eps}}
664: \vspace{0.5cm}
665: \caption{Dimensionless function $w(x)$, which is defined so that
666: $w(0)=1$.}
667: \label{plot:w}
668: \end{figure}
669: }
670:
671:
672: If the Coulomb interaction is considered, then the lowest order in
673: interaction is not sufficient since for the long range interaction
674: $\nu V_0(q \approx 0) \gg 1$. Although the interaction itself is
675: still independent of the electron spin, summation of the perturbation
676: theory depends on the spin state of the two electrons involved. In the
677: first order correction Eq.~(\ref{sw}) all spin channels gave identical
678: contributions. The total number of channels is $4$ and they can be
679: classified by the total spin of the two electrons: one state with the
680: total spin zero (``singlet'') and three states with the total spin $1$
681: (``triplet''; the three states differ by the value of the
682: $z$-component of the total spin). For long range interaction the
683: perturbation theory for the Hartree correction singlet and triplet
684: channels is different. It is known \cite{aar,fin} (see also
685: Section~\ref{diagrams}), that the singlet channel contribution should
686: be combined with the Fock correction as a renormalization of the
687: coupling constant. However, the final result is universal due to
688: dynamical screening: the singlet channel modification of the coupling
689: does not affect the result. What remains of the Hartree term is the
690: triplet channel contribution, which now depends on the corresponding
691: Fermi liquid constant $F_0^\sigma$. Thus, the total correction to the
692: conductivity can be written as a sum of the ``charge'' (which combines
693: Fock and singlet part of Hartree) and triplet contributions
694:
695: \begin{mathletters}
696: \label{all}
697: \begin{equation}
698: \sigma = \sigma_D + \delta\sigma_T + \delta\sigma_C,
699: \end{equation}
700:
701: \noindent
702: where the charge channel correction is given by
703:
704: \begin{equation}
705: \delta\sigma_C =
706: \frac{e^2}{\pi\hbar} \frac{T\tau}{\hbar}
707: \left[ 1 -\frac{3}{8}f(T\tau)\right]
708: -\frac{e^2}{2\pi^2\hbar}\ln\frac{E_F}{T}
709: ,
710: \label{fc}
711: \end{equation}
712:
713: \noindent
714: and the triplet channel correction is
715:
716: \begin{eqnarray}
717: \delta\sigma_T = &&
718: \frac{3F_0^\sigma}{(1+F_0^\sigma)}
719: \frac{e^2}{\pi\hbar}\frac{T\tau}{\hbar}
720: \left[ 1 -\frac{3}{8}t(T\tau; F_0^\sigma)\right]
721: \nonumber\\
722: &&
723: \nonumber\\
724: &&
725: -3\left(1-\frac{1}{F_0^\sigma}
726: \ln(1+F_0^\sigma)\right)
727: \frac{e^2}{2\pi^2\hbar}\ln\frac{E_F}{T}.
728: \label{tc}
729: \end{eqnarray}
730: \label{s}
731: \end{mathletters}
732:
733: \noindent
734: here the factor of three in the triplet channel correction
735: Eq.~(\ref{tc}) is due to the fact that all three components of the
736: triplet state contribute equally. We reiterate that the corrections
737: Eqs.(\ref{s}) are defined only up to a temperature independent
738: (however not necessarily Fermi-liquid constant independent) term,
739: see also discussion after Eq.~(\ref{sw}).
740:
741: {
742: \narrowtext
743: \begin{figure}[ht]
744: %\vspace{0.5 cm}
745: \epsfxsize=9 cm
746: \centerline{\epsfbox{f5.eps}}
747: \vspace{0.2cm}
748: \caption{Dimensionless function $f(x)$, defined so that $f(0)=1$}
749: \label{plot:f}
750: \end{figure}
751: }
752:
753: We should warn the reader here, that we describe the interaction in
754: the triplet channel by one coupling constant $F_0^\sigma$. For the
755: weak coupling limit, it corresponds to the approximation $V_0(2k_F)
756: \simeq \langle V_0(k) \rangle_{FS}$. This approximation overestimates
757: the triplet channel contribution to the ballistic case for $r_s =
758: \sqrt{2}e^2/(\kappa\hbar v_F) \ll 1$. However, in this limit
759: contribution itself is much smaller than the singlet one. For better
760: estimates in this regime one should use
761:
762: \[
763: F_0^\sigma \to - \frac{1}{2}\frac{r_s}{r_s + \sqrt{2}}
764: \]
765:
766: \noindent
767: in the first line of Eq.~(\ref{tc}) and
768:
769: \begin{eqnarray*}
770: F_0^\sigma \to - \frac{1}{2\pi}
771: \frac{r_s}{ \sqrt{2-r_s^2} }
772: \ln \left(
773: \frac{\sqrt{2}+\sqrt{2-r_s^2} }{\sqrt{2} - \sqrt{2-r_s^2} }
774: \right),
775: \ \
776: && r_s^2 < 2;
777: \\
778: F_0^\sigma \to - \frac{1}{\pi}\frac{r_s}{\sqrt{r_s^2-2}}
779: \arctan\sqrt{\frac{1}{2}r_s^2-1},
780: \ \
781: && r_s^2 > 2
782: \end{eqnarray*}
783:
784: \noindent
785: in the second line. For $r_s \gtrsim 1$ our replacament is well
786: justified even within weak coupling scheme.
787:
788: Similar to Eq.~(\ref{sw}) the dimensionless functions $f(x)$ and
789: $t(x;F_0^\sigma)$ describe the cross-over between ballistic and
790: diffusive limits. They are plotted on Figs.~\ref{plot:f} and
791: \ref{plot:t} and full expressions are given by Eqs.~(\ref{f2}). The
792: universal function $f(x)$ has the following limits
793:
794: \begin{mathletters}
795: \begin{equation}
796: f(x\gg 1) \approx -\frac{1}{3\pi x}\Big(2(\ln x - 1)\ln 2 -
797: \frac{7}{2}\ln (2x)\Big);
798: \end{equation}
799: \begin{eqnarray}
800: &&f(x\ll 1) \approx 1 - \gamma_1 x
801: + \frac{\pi}{6} x \ln x;
802: \\
803: &&
804: \nonumber\\
805: &&
806: \gamma_1 = -\frac{\zeta^\prime(2)}{\pi}+\frac{\pi}{6}
807: \left({\cal C} +\frac{1}{3}\ln 2 \right)
808: \approx 0.7216;
809: \nonumber
810: \end{eqnarray}
811: \label{fl}
812: \end{mathletters}
813:
814:
815: {
816: \narrowtext
817: \begin{figure}[ht]
818: %\vspace{0.2 cm}
819: \epsfxsize=9 cm
820: \centerline{\epsfbox{f6.eps}}
821: \vspace{0.2cm}
822: \caption{Dimensionless function $t(x,F_0^\sigma)$ defined so that
823: $t(0,F_0^\sigma)=1$.}
824: \label{plot:t}
825: \end{figure}
826: }
827:
828: The function $t(x;F_0^\sigma)$ depends on the coupling constant and
829: therefore its asymptotic form also depends on $F_0^\sigma$. For very
830: small $x\ll 1+F_0^\sigma$ the asymptotic form is
831:
832: \begin{eqnarray}
833: &&t(x\ll 1+F_0^\sigma) \approx 1 -\gamma_2 x
834: +\frac{\pi}{18} x \ln x \left(3+\frac{1}{1+F_0^\sigma}\right);
835: \nonumber\\
836: &&
837: \nonumber\\
838: && \quad
839: \gamma_2 = -\frac{\zeta^\prime(2)}{3\pi}
840: \left(3+\frac{1}{1+F_0^\sigma}\right)
841: -\frac{\pi\gamma_3}{9(1+F_0^\sigma)}
842: \label{tsx}\\
843: &&
844: \nonumber\\
845: && \quad\quad\quad
846: +\frac{\pi}{18}
847: \left[{\cal C}\left(3+\frac{1}{1+F_0^\sigma}\right )
848: +\ln 2 \left(1+\frac{3}{1+F_0^\sigma}\right)\right];
849: \nonumber\\
850: &&
851: \nonumber\\
852: && \quad
853: \gamma_3=1-\frac{5F_0^\sigma -3}{1+F_0^\sigma}
854: - \left(\frac{5}{2} - 2F_0^\sigma
855: \right)\frac{\ln (1+F_0^\sigma)}{F_0^\sigma}.
856: \nonumber
857: \end{eqnarray}
858: Notice that at $T\tau \to 0$, Eqs. (\ref{all}) reproduce the known
859: result (\ref{introdiff}).
860: Let us point out that for numerical reasons contributions
861: of scaling functions $w,\ f,\ t$ change the result only by few
862: percents and they can be neglected for all the practical purposes.
863:
864:
865: Notice that while the charge channel correction Eq.~(\ref{fc}) is
866: universal, the triplet channel correction Eq.~(\ref{tc}) is
867: proportional to $F_0^\sigma$, which might be negative. That leads to
868: the conclusion, that the overall sign of the total correction
869: Eqs.~(\ref{s}) depends on value of
870: $F_0^\sigma$: it can be either positive or negative, see
871: Fig.~\ref{sigma}.
872:
873:
874: {
875: \narrowtext
876: \begin{figure}[ht]
877: %\vspace{0.1 cm}
878: \epsfxsize=5 cm
879: %\centerline{\epsfbox{f7.eps}}
880: %The figure was changed!!! Do not change it back!!!
881: {\epsfbox{Fig7.eps}}
882: \vspace{0.2cm}
883: \caption{Total interaction correction to conductivity. The divergence
884: at $T\tau/\hbar\protect\rightarrow 0$ is due to the usual logarithmic
885: correction
886: \protect\cite{aar}.
887: Curve $F_0^\sigma=0$ corresponds to the universal behavior of
888: completely spin polarized electron gas.
889: The correction is defined up to a temperature
890: independent part, see
891: Eq.~(\protect\ref{ultra}) and discussion after Eq.~(\protect\ref{sw}).}
892: \label{sigma}
893: \end{figure}
894: }
895:
896:
897: Combining together all of the above results we plot the total
898: correction to the conductivity on Fig.~\ref{sigma} for different
899: values of $F_0^\sigma$. The divergence at low temperature is due to
900: the usual logarithmic correction \cite{aar}. Although the exact value
901: of $F_0^\sigma$ can not be calculated theoretically (in particular,
902: its relation to the conventional measure of the interaction strength,
903: $r_s$, is unknown for $r_s>1$), in principle it can be found from a
904: measurement of the Pauli spin susceptibility
905:
906: \begin{equation}
907: \chi = \frac{\nu}{1+F_0^\sigma},
908: \label{pauli}
909: \end{equation}
910:
911: \noindent
912: where the density of states $\nu$ should be obtained from a
913: measurement of the specific heat (at $\tau^{-1}\ll T \ll E_F$). The
914: constant $F_0^\sigma$ is the only parameter in our theory which
915: describes all the data, including the Hall coefficient and the
916: magneto-resistance in the parallel field. The theory for interaction
917: corrections in the magnetic field will be addressed in the forthcoming
918: paper \cite{pre}.
919:
920: The correction in Fig.~\ref{sigma} is almost always monotonous, except
921: for a narrow region $-0.45 < F_0^\sigma < -0.25$. A typical curve in
922: this region is shown in Fig.~\ref{nonmon}. Note, however, that the
923: overall magnitude of the correction in the range of $T\tau$ in
924: Fig.~\ref{sigma} is so small that it can hardly be observed.
925:
926: {
927: \narrowtext
928: \begin{figure}[ht]
929: %\vspace{0.1 cm}
930: \epsfxsize=5 cm
931: %\centerline{\epsfbox{f8.eps}}
932: %The figure was changed!!! Do not change it back!!!
933: {\epsfbox{Fig8.eps}}
934: \vspace{0.2cm}
935: \caption{The non-monotonous correction to conductivity. Note the difference
936: in the overall scale relative to the previous figure.}
937: \label{nonmon}
938: \end{figure}
939: }
940:
941:
942: When the interaction becomes so strong that the system approaches the
943: Stoner instability, $F_0^\sigma$ ceases to be a constant and becomes
944: momentum-dependent. Thus the result Eq.~(\ref{s}) is no longer
945: valid. Although the simple condition $\delta\sigma_T < \sigma_D$
946: suggests that this happens at $T\approx (1+F_0^\sigma)E_F$, the more
947: detailed analysis (see Section~\ref{triplet}) shows that it happens
948: much earlier. In fact, the approximation of the constant $F_0^\sigma$
949: is valid in the parameter region defined by the inequality
950:
951: \begin{equation}
952: \frac{T}{E_F} < (1+F_0^\sigma)^2,
953: \label{lim}
954: \end{equation}
955:
956: \noindent
957: see Section~\ref{triplet} for the origin of this inequality. We
958: were not able to make a reliable calculation of
959: $\delta \sigma(T)$ at higher temperatures.
960:
961: \section{Perturbation theory}
962: \label{diagrams}
963:
964: In this section we show how the announced results Eq.~(\ref{s}) can be
965: obtained with the help of the traditional perturbation theory. We try
966: to explain the most important points of the calculation in detail.
967: The comprehensive review of the diagrammatic technique for disordered
968: systems can be found in Ref.~\onlinecite{aar}. We start by a brief
969: discussion of the case of a weak, short-range interaction
970: potential. Although this case is artificial and is unrelated to any
971: experiment, it is governed by the same physics as the general problem,
972: and it is simple enough to allow a transparent presentation. To
973: generalize to stronger coupling, we need to recall the basic ideas of
974: the Landau Fermi liquid theory and to identify the soft modes in the
975: system. Then we present the calculation leading to Eq.~(\ref{s}).
976: Finally, to establish the relation of our results to existing
977: literature, we briefly discuss scattering on a single impurity (this
978: discussion is completely analogous to the one in
979: Section~\ref{qualitative} but uses the language of diagrams).
980:
981:
982: \subsection{Hartree-Fock considerations.}
983: \label{h-f}
984:
985: %\subsubsection{Short-range potential}
986:
987: The static conductivity of a system of electrons is given by the Kubo
988: formula
989:
990: \begin{eqnarray}
991: &&\sigma_{\alpha\beta} =
992: \label{kubo}\\
993: &&
994: \nonumber\\
995: &&
996: -
997: \lim_{\omega \to 0}
998: {\rm Re}
999: \left[
1000: \frac{1}{\Omega_n}
1001: \int\limits_{0}^{1/T} d\tau \langle {\bf T_\tau} \hat j_\alpha(\tau)
1002: \hat j_\beta(0) \rangle e^{i\Omega_n\tau}
1003: \right]_{i\Omega_n \to \omega},
1004: \nonumber
1005: \end{eqnarray}
1006:
1007: \noindent
1008: where $\hat j_\alpha(\tau)$ is the operator of the electric current at
1009: imaginary time $\tau$ and the analytic continuation of the function
1010: defined at Matzubara frequencies $\Omega_n = 2\pi Tn$ to function
1011: analytic at ${\rm Im} \omega >0$ is performed.
1012:
1013: {
1014: \narrowtext
1015: \begin{figure}[ht]
1016: \vspace{0.5 cm}
1017: \epsfxsize=8 cm
1018: \centerline{\epsfbox{f9.eps}}
1019: \vspace{0.5cm}
1020: \caption{Interaction correction to conductivity in the lowest order of
1021: perturbation theory. Here solid lines correspond to Matsubara Green's
1022: functions $-G(i\epsilon_n; \vec r_1, \vec r_2)$ and the wavy line
1023: represents the interaction potential, $-V(\vec r_1- \vec r_2)$.}
1024: \label{sr}
1025: \end{figure}
1026: }
1027:
1028: Assuming that electrons interact by means of a weak, short-range
1029: interaction (range shorter than $v_F min(\tau, 1/T )$, $V(r)$ it is
1030: sufficient to consider the lowest order of the perturbation theory.
1031: The perturbation theory can be conveniently expressed in terms of
1032: Feinman diagrams. The lowest order diagrams for the interaction
1033: correction to the conductivity are shown on Fig.~\ref{sr}. The Hartree
1034: term corresponds to the diagrams ``a'', while the Fock contribution
1035: corresponds to diagrams ``b''. Evaluation of the correction consists
1036: of two main steps: (i) analytic continuation to real time, and (ii)
1037: disorder averaging. While these two steps can be performed in any
1038: order without affecting the result, it is more convenient (for
1039: technical reasons) to start with step (i).
1040:
1041: Although analytic continuation in Eq.~(\ref{kubo}) is now a textbook
1042: task, we include a brief discussion of the standard procedure in the
1043: Appendix to make the paper self-contained. After the continuation any
1044: physical quantity is expressed in terms of exact (i.e. not averaged
1045: over disorder) retarded and advanced Green's functions of the
1046: electronic system, which are defined as
1047:
1048: \begin{eqnarray}
1049: G^{R(A)}_{12}(\epsilon)=\sum\limits_j
1050: {{\Psi^*_j(\vec{r}_1)\Psi_j(\vec{r}_2)}
1051: \over{\epsilon-\epsilon_j\pm\imath 0}},
1052: \label{gra}
1053: \end{eqnarray}
1054:
1055: \noindent
1056: where $j$ labels the exact eigenstates of the system and $\epsilon_j$
1057: are the exact eigenvalues, counted from the Fermi energy:
1058:
1059: \[
1060: \left(\frac{-\nabla^2}{2m} + U(\vec r) \right) \Psi_j(\vec{r}) =
1061: (\epsilon_j + \epsilon_F) \Psi_j(\vec{r}).
1062: \]
1063:
1064: \noindent
1065: Here $U(\vec r)$ is the disorder potential.
1066:
1067: The resulting expression for the correction to the symmetric part of
1068: the conductivity (the Hall conductivity will be discussed in a
1069: separate publication \cite{pre}) can be written as \cite{aag}
1070:
1071:
1072: \begin{eqnarray}
1073: \delta&&\sigma_{\alpha\beta} = \int\limits_{-\infty}^{\infty}
1074: \frac{d\Omega}{8\pi^2}
1075: \left[\frac{\partial}{\partial\Omega}
1076: \left(\Omega\coth\frac{\Omega}{2T}\right)\right]
1077: \int d^2 r_3 d^2 r_4
1078: \label{cc}
1079: \\
1080: &&
1081: \nonumber\\
1082: &&
1083: \times
1084: {\rm Im} \Bigg\{ V(\vec r_3- \vec r_4)
1085: \Big ( B^{\alpha\beta}_F (\Omega; \vec r_3, \vec r_4)
1086: -2 B^{\alpha\beta}_H (\Omega; \vec r_3, \vec r_4)
1087: \nonumber\\
1088: &&
1089: \nonumber\\
1090: && \hspace{5 cm}
1091: + \{\alpha\leftrightarrow\beta\}
1092: \Big)\Bigg\},
1093: \nonumber
1094: \end{eqnarray}
1095:
1096: \noindent
1097: where the extra factor of $2$ in the Hartree term is due to the
1098: summation over electron spin. Here we denoted products of four
1099: Green's functions as $B_{F(H)}$. For the Fock term we have
1100:
1101: \begin{mathletters}
1102: \begin{eqnarray}
1103: B&&^{\alpha\beta}_F(\Omega; \vec r_3, \vec r_4) =
1104: \int \frac{d^2 r_1 d^2 r_5}{{\cal V}}
1105: \nonumber\\
1106: &&
1107: \nonumber\\
1108: &&
1109: \times\Big\{
1110: \hat{J}_1^\alpha G^R_{15}(\epsilon)
1111: \hat{J}_5^\beta G^A_{53}(\epsilon)
1112: G^R_{34}(\epsilon-\Omega) G^A_{41}(\epsilon)
1113: \label{f11}
1114: \\
1115: &&
1116: \nonumber\\
1117: &&\quad
1118: + \hat{J}_1^\alpha G^A_{15}(\epsilon)
1119: \hat{J}_5^\beta G^R_{53}(\epsilon)
1120: G^R_{34}(\epsilon-\Omega) G^R_{41}(\epsilon)
1121: \label{f12}
1122: \\
1123: &&
1124: \nonumber\\
1125: &&\quad
1126: + 2 \hat{J}_1^\alpha G^R_{13}(\epsilon)
1127: G^R_{35}(\epsilon-\Omega) \hat{J}_5^\beta
1128: G^R_{54}(\epsilon-\Omega) G^A_{41}(\epsilon)
1129: \label{f32}
1130: \\
1131: &&
1132: \nonumber\\
1133: &&\quad
1134: - \hat{J}_1^\alpha G^A_{15}(\epsilon) \hat{J}_5^\beta
1135: G^A_{53}(\epsilon)
1136: G^R_{34}(\epsilon-\Omega) G^A_{41}(\epsilon)
1137: \label{f13}
1138: \\
1139: &&
1140: \nonumber\\
1141: &&\quad
1142: - \hat{J}_1^\alpha G^A_{13}(\epsilon) G^R_{35}(\epsilon-\Omega)
1143: \hat{J}_5^\beta G^R_{54}(\epsilon-\Omega)
1144: G^A_{41}(\epsilon)
1145: \Big\},
1146: \label{f31}
1147: \end{eqnarray}
1148: \label{fck}
1149: \end{mathletters}
1150:
1151: \noindent
1152: where ${\cal V}$ is the area of the system. Equations~(\ref{f32}) and
1153: (\ref{f31}) come from the diagram ``b3'' on Fig.~\ref{sr} and the rest
1154: of Eq.~(\ref{fck}) correspond to diagrams ``b1'' and ``b2''. For the
1155: Hartree term the expression is similar,
1156:
1157: \begin{mathletters}
1158: \begin{eqnarray}
1159: B&&^{\alpha\beta}_H (\Omega; \vec r_3, \vec r_4) =
1160: \int \frac{d^2 r_1 d^2 r_5}{{\cal V}}
1161: \nonumber\\
1162: &&
1163: \nonumber\\
1164: &&
1165: \times\Big\{
1166: \hat{J}_1^\alpha G^R_{15}(\epsilon) \hat{J}_5^\beta
1167: G^A_{53}(\epsilon)
1168: G^R_{44}(\epsilon-\Omega) G^A_{31}(\epsilon)
1169: \label{h11}
1170: \\
1171: &&
1172: \nonumber\\
1173: &&\quad
1174: + \hat{J}_1^\alpha G^A_{15}(\epsilon) \hat{J}_5^\beta
1175: G^R_{53}(\epsilon)
1176: G^R_{44}(\epsilon-\Omega) G^R_{31}(\epsilon)
1177: \label{h12}
1178: \\
1179: &&
1180: \nonumber\\
1181: &&\quad
1182: + 2 \hat{J}_1^\alpha G^R_{13}(\epsilon) G^R_{45}(\epsilon-\Omega)
1183: \hat{J}_5^\beta G^R_{54}(\epsilon-\Omega) G^A_{31}(\epsilon)
1184: \label{h32}\\
1185: &&
1186: \nonumber\\
1187: &&\quad
1188: - \hat{J}_1^\alpha G^A_{15}(\epsilon) \hat{J}_5^\beta
1189: G^A_{53}(\epsilon)
1190: G^R_{44}(\epsilon-\Omega) G^A_{31}(\epsilon)
1191: \label{h13}
1192: \\
1193: &&
1194: \nonumber\\
1195: &&\quad
1196: - \hat{J}_1^\alpha G^A_{13}(\epsilon) G^R_{45}(\epsilon-\Omega)
1197: \hat{J}_5^\beta G^R_{54}(\epsilon-\Omega) G^A_{31}(\epsilon)
1198: \Big\}.
1199: \label{h31}
1200: \end{eqnarray}
1201: \label{hrt}
1202: \end{mathletters}
1203:
1204: \noindent
1205: Again, Eqs.~(\ref{h32}) and (\ref{h31}) correspond to the diagram
1206: ``a3'' in Fig.~\ref{sr}. The current operator is defined as
1207:
1208: \begin{eqnarray}
1209: f_1(\vec r) \hat{\vec J} f_2(\vec r) && = \frac{ie}{2m}
1210: \left[ \left( \vec\nabla f_1 \right) f_2 -
1211: \left( f_1\vec\nabla f_2 \right) \right]
1212: \nonumber\\
1213: &&
1214: \nonumber\\
1215: &&
1216: - \frac{e\vec A(\vec r)}{m} f_1(\vec r) f_2(\vec r).
1217: \label{jop-def}
1218: \end{eqnarray}
1219:
1220: \noindent
1221: In the above expressions terms corresponding to diagrams ``b3'' and
1222: ``a3'' on Fig.~\ref{sr} allow for at least one of the spatial
1223: integrations to be performed with the help of the identity
1224:
1225: \begin{equation}
1226: \int d\vec r_5 G^R_{35}(\epsilon)\hat{J}_5^\beta G^R_{54}(\epsilon)
1227: =-ie(\vec r_3 - \vec r_4)^\beta G^R_{34}(\epsilon).
1228: \label{id}
1229: \end{equation}
1230:
1231: \noindent
1232: Now it is clear that Hartree terms Eqs.~(\ref{h32}) and (\ref{h31})
1233: vanish identically, since there the identity (\ref{id}) should be
1234: applied with coordinates $\vec r_3$ and $\vec r_4$ being equal to each
1235: other. In the Fock terms Eqs.~(\ref{f31}) and (\ref{f32}) one needs to
1236: further multiply the result of Eq.~(\ref{id}) by the interaction
1237: potential $V(\vec r_3 - \vec r_4)$. In the case of the short range
1238: interaction potential this also gives vanishing contribution. Thus we
1239: conclude, that the diagram ``a3'' on Fig.~\ref{sr} does not contribute
1240: for any form of the interaction, while the diagram ``b3'' vanishes for
1241: the short-range interaction.
1242:
1243: The same identity can also be applied to terms Eqs.~(\ref{f13}) and
1244: (\ref{h13}), which also vanish by the same reason. Thus the task of
1245: averaging over disorder is now simplified because we only need to
1246: average two Fock terms Eqs.~(\ref{f11}) and (\ref{f12}) and two
1247: Hartree terms Eqs.~(\ref{h11}) and (\ref{h12}). These expressions
1248: contain only Green's functions of non-interacting electrons and can be
1249: averaged using the standard diagrammatic technique of the theory of
1250: disordered systems (see Ref.~\onlinecite{aar} for review). The
1251: diagrams for averaged quantities can be constructed using the four
1252: ``building blocks'' (we use the momentum representation since
1253: translational invariance is restored after averaging):
1254:
1255: \noindent
1256: (1) the average electronic Green's function (denoted as a solid line;
1257: there should be no confusion with the previous use of the solid line
1258: for exact Green's functions before averaging), which in momentum space
1259: can be written as
1260:
1261: \begin{eqnarray}
1262: \langle G^{R(A)}\rangle (k,\epsilon) =
1263: \frac{1}{\epsilon - \xi_k \pm \frac{i}{2\tau}};
1264: \label{green}
1265: \end{eqnarray}
1266:
1267: \noindent
1268: (2) the disorder potential, which is assumed to be Gaussian with the
1269: correlator
1270:
1271: \begin{eqnarray*}
1272: \langle U(\vec r_1)U(\vec r_2)\rangle =
1273: \frac{1}{2\pi\nu\tau} \delta(\vec r_1 - \vec r_2).
1274: \end{eqnarray*}
1275:
1276: \noindent
1277: In the diagrams this correlator is represented by the dotted line;
1278:
1279: {
1280: \narrowtext
1281: \begin{figure}[ht]
1282: \vspace{0.2 cm}
1283: \epsfxsize=8.5 cm
1284: \centerline{\epsfbox{f10.eps}}
1285: \vspace{0.2cm}
1286: \caption{Dressed interaction vertex.}
1287: \label{8}
1288: \end{figure}
1289: }
1290:
1291: \noindent
1292: (3) the dressed interaction vertex $\Gamma$ ($q$ and $\Omega$ are
1293: momentum and frequency of the interaction propagator), which
1294: represents a geometric series in disorder potential shown on
1295: Fig.~\ref{8};
1296:
1297: \begin{mathletters}
1298: \begin{equation}
1299: \Gamma ( \vec q; \Omega)= 1+ \frac{1/\tau}{S- \frac{1}{\tau}}
1300: \label{gamma}
1301: \end{equation}
1302:
1303: \noindent
1304: where we denote
1305:
1306: \begin{eqnarray}
1307: S=\sqrt{\left(i\Omega + \frac{1}{\tau}\right)^2 + v_F^2 q^2},
1308: \label{sqrt}
1309: \end{eqnarray}
1310: \label{gsd}
1311: \end{mathletters}
1312:
1313: \noindent
1314: (4) the averaged product of a retarded and an advanced Green's
1315: functions (sometimes referred to as the diffuson), where we have
1316: summed up a geometric series shown on Fig.~\ref{9}
1317:
1318: {
1319: \narrowtext
1320: \begin{figure}[ht]
1321: \vspace{0.2 cm}
1322: \epsfxsize=8 cm
1323: \centerline{\epsfbox{f11.eps}}
1324: \vspace{0.2cm}
1325: \caption{Diffuson - geometric series of impurity lines}
1326: \label{9}
1327: \end{figure}
1328: }
1329:
1330: Using these building blocks we can average the products of Green's
1331: functions as shown on Fig.~\ref{hfav}. It is convenient to write the
1332: averaged $B_{F(H)}$ in the momentum representation. The product $B_F$,
1333: which appears in the Fock term, can be viewed as a function of
1334: coordinates of the two interaction vertices and can be transformed to
1335: the momentum space as
1336:
1337: \begin{equation}
1338: \langle B_F (\Omega; \vec r_1,\vec r_2) \rangle
1339: = \int \frac{d^2 q}{(2\pi)^2}
1340: e^{i\vec q (\vec r_3 - \vec r_4)}
1341: \langle B_F (q, \Omega) \rangle.
1342: \label{noname1}
1343: \end{equation}
1344:
1345: \noindent
1346: Using the explicit expressions Eqs.~(\ref{green}) and (\ref{gsd}) we
1347: can write the analytic form of the averaged $B_F$
1348:
1349: \begin{eqnarray}
1350: \frac{1}{\sigma_D}B_F (q, \Omega) =
1351: \frac{(\Gamma^2-1)\tau}{S}
1352: \label{bf}
1353: +\frac{\Gamma(\Gamma+1)}{v_F^2 q^2}
1354: \left(\frac{i\Omega + \frac{1}{\tau}}{S} - 1 \right)^2.
1355: \nonumber
1356: %\label{bf}
1357: \end{eqnarray}
1358:
1359: In the absence of magnetic field,
1360: $B^{\alpha\beta}_{F(H)}=\delta^{\alpha\beta} B_{F(H)}$, which is why
1361: we did not include the Greek indices in Eq.~(\ref{bf}).
1362:
1363: {
1364: \narrowtext
1365: \begin{figure}[ht]
1366: \vspace{0.2 cm}
1367: \epsfxsize=8.5 cm
1368: \centerline{\epsfbox{f12.eps}}
1369: \vspace{0.2cm}
1370: \caption{Averaged product of four Green's functions.
1371: The wavy lines indicate $\delta(\vec p_{in}- \vec p_{out}+ \vec q)$
1372: for the Fock contribution $B_F(q)$
1373: and $2\pi\delta(\vec p_{in}- \vec p_{out}+ \vec q)
1374: \delta[\widehat{\vec{n}\vec p_{in}}]$
1375: for the Hartree contribution $B_H(q, n_1,n_2)$.}
1376: \label{hfav}
1377: \end{figure}
1378: }
1379:
1380: The Hartree contribution is considered analogously.
1381: One can write
1382:
1383: \beqa
1384: \langle B_H(\Omega; \vec r_1, \vec r_2) \rangle =
1385: \int \frac{d \theta_1}{(2\pi)}&&
1386: \frac{d \theta_2}{(2\pi)}
1387: \frac{d^2 q}{(2\pi)^2}
1388: e^{ik_F(\vec n_1-\vec n_2)(\vec r_1-\vec r_2)}
1389: \nonumber\\
1390: &&
1391: \nonumber\\
1392: &&
1393: \times
1394: B_H(\Omega; \vec n_1, \vec n_2, \vec q),
1395: \label{Ht1}
1396: \eeqa
1397:
1398: \noindent
1399: where $\vec n_i = (\cos \theta_i, \sin \theta_i)$ indictates
1400: the direction of the momentum.
1401: Then, disorder averaging of $B_H(\Omega; \vec n_1, \vec n_2, \vec q)$
1402: is performed with the help of the same diagrams (see Fig.~\ref{hfav})
1403: but the expression for the vertices changed as indicated in the figure
1404: caption.
1405:
1406: Accordingly, the expression for the dressed vertex
1407: (\ref{gamma}), see also Fig.~\ref{8}, is changed to
1408:
1409: \beqa
1410: \Gamma_H&&(\vec n,\vec n_k; \vec q; \Omega) =
1411: 2\pi \delta(\widehat{\vec n \vec n_k})
1412: + \frac{1}{S_H}
1413: \frac{S}{ S{\tau}-1},
1414: \label{gammah}
1415: \\
1416: &&
1417: \nonumber\\
1418: &&
1419: S_H(\vec n, \vec q; \Omega)={i\Omega -iv_F \vec q \vec n + 1/\tau},
1420: \nonumber
1421: \eeqa
1422:
1423: \noindent
1424: where $\vec n_k$ corresponds to the direction of the momentum $k$ on
1425: Fig.~\ref{8}. The final expression for $B_H$ is similar to
1426: Eq.~(\ref{bf})
1427:
1428: \begin{eqnarray}
1429: && \frac{1}{\sigma_D} B_H ( \Omega;
1430: {\vec n}_1, {\vec n}_2,
1431: {\vec q}) = -
1432: \frac{-
1433: 2\pi\delta\left(\widehat{\vec n_1\vec n_2}\right)\tau}{S}
1434: \nonumber\\
1435: &&
1436: \nonumber\\
1437: &&
1438: \quad\quad +
1439: \frac{\tau}{S}
1440: \int \frac{d \theta_3}{2\pi}
1441: \Gamma_H(\vec n_1, \vec n_3)\Gamma_H(\vec n_2, n_3)
1442: \label{bh}
1443: \\
1444: &&
1445: \nonumber\\
1446: &&\quad\quad +
1447: \int \frac{d \theta_3}{2\pi}
1448: \int \frac{d \theta_4}{2\pi}
1449: \left(\vec n_3\vec n_4\right)
1450: \frac{\Gamma_H(\vec n_1,n_3)\Gamma_H(\vec n_2,n_4)}
1451: {S_H(\vec n_3)S_H(\vec n_4)}
1452: \nonumber\\
1453: &&
1454: \nonumber\\
1455: &&\quad\quad
1456: +\left(\vec n_1\vec n_2\right)
1457: \frac{\Gamma}{S_H(\vec n_1)S_H(\vec n_2)},
1458: \nonumber
1459: %\label{bf}
1460: \end{eqnarray}
1461:
1462: \noindent
1463: and we suppressed the arguments $q,\Omega$ in the right-hand-side
1464: of the equation.
1465:
1466: We note in passing, that by construction of Eqs.~(\ref{f11}),
1467: (\ref{f12}), (\ref{h11}) and (\ref{h12}) that
1468:
1469: \[
1470: B_F(\Omega; \vec r, \vec r)=B_H(\Omega; \vec r, \vec r),
1471: \]
1472:
1473: \noindent
1474: and, therefore, according to Eqs.~(\ref{noname1}) and (\ref{Ht1})
1475: the relation
1476:
1477: \[
1478: B_F(\Omega; \vec q) =
1479: \int \frac{d \theta_1}{2\pi}\frac{d \theta_2}{2\pi}
1480: B_H(\Omega; \vec n_1, \vec n_2, \vec q)
1481: \]
1482:
1483: \noindent
1484: must hold [this can be easily verified using explicit expresions
1485: (\ref{bf}) and (\ref{bh})].
1486:
1487: We are now prepared to calculate the temperature dependence
1488: of the conductivity from Eq.~(\ref{cc}). We substitute Eqs.~(\ref{bf})
1489: and (\ref{bh}) into Eq.~(\ref{cc}). As we will see, the
1490: main contribution to the temperature dependence is provided by
1491: wave-vectors $q_{T} \simeq max(T,(T/\tau)^{1/2})/v_F$. On the other hand the
1492: potential $V(\vec r)$ has a range much shorter than $1/q_T$.
1493: This enables us to use the following approximations
1494:
1495: \begin{eqnarray*}
1496: \int \frac{d^2 r_3 d^2 r_4}{\cal V}
1497: V(\vec r_3-\vec r_4)e^{i\vec q(\vec r_3-\vec r_4)}
1498: \approx V(0);
1499: \end{eqnarray*}
1500: \begin{eqnarray*}
1501: &&
1502: \int \frac{d^2 r_3 d^2 r_4}{\cal V}
1503: V(\vec r_3-\vec r_4)e^{i k_F (\vec n_1 - \vec n_2) (\vec r_3-\vec r_4)}
1504: \\
1505: &&
1506: \\
1507: &&
1508: \hspace*{4cm}\approx
1509: V\left(2k_F\sin \frac{\widehat{\vec n_1\vec n_2}}{2}
1510: \right),
1511: \end{eqnarray*}
1512: where $V(k)$ in the right-hand-side of the above equations denotes
1513: the Fourier transform of the interaction potential.
1514:
1515: Altogether, we now write the conductivity correction as
1516:
1517: \begin{eqnarray}
1518: &&\delta\sigma = \int\limits_{-\infty}^{\infty}
1519: \frac{d\Omega}{8\pi^2}
1520: \frac{\partial}{\partial\Omega}
1521: \left(\Omega\coth\frac{\Omega}{2T}\right)
1522: \int\frac{d^2 q}{(2\pi)^2}
1523: \label{cc5}\\
1524: &&
1525: \nonumber\\
1526: &&
1527: \times {\rm Im} \Bigg\{
1528: V_0(0) B_F (q,\Omega)
1529: \nonumber\\
1530: &&
1531: \nonumber\\
1532: &&
1533: - 2 \int\frac{d\theta_1}{2\pi} \int\frac{d\theta_2}{2\pi}
1534: V_0\left(2k_F\sin\frac{\widehat{\vec n_1 \vec n_2}}{2}\right)
1535: B_H (\Omega; \vec n_1, \vec n_2, \vec q)\Bigg\}.
1536: \nonumber
1537: \end{eqnarray}
1538:
1539: \noindent
1540: Evaluating this integral (where we only keep the temperature
1541: dependent part, see Section~\ref{single} for details)
1542: one arrives to the same result
1543: Eq.~(\ref{sw}), but with the coefficient in the form Eq.~(\ref{u}), in
1544: agreement with the discussion of Section~\ref{qualitative}.
1545:
1546: Let us now turn to the case of the Coulomb potential, where the scheme
1547: of the calculation (as described so far) breaks down. In the Fock term
1548: we have $V(0)$, which diverges for the Coulomb interaction ($V(q)\sim
1549: 1/q$). To obtain meaningful results one needs to take into account the
1550: effect of dynamical screening. The Hartree term seems to work better
1551: since using just the static screening makes the result
1552: finite. However, this is wrong also, since in this case diagrams with
1553: extra interaction lines do not contain any smallness (see e.g.
1554: Fig.~\ref{cor}; there the correction is $\sim V(2k_F) V(0)$). Thus one
1555: can not justify the perturbation theory in the interaction
1556: potential. The way out of this problem is the standard theory of
1557: Landau Fermi liquid, which we briefly discuss in the following subsection.
1558:
1559: {
1560: \narrowtext
1561: \begin{figure}[ht]
1562: \vspace{0.5 cm}
1563: \epsfxsize=7 cm
1564: \centerline{\epsfbox{f13.eps}}
1565: \vspace{0.5cm}
1566: \caption{(a) Single impurity contribution to the Hartree term, see
1567: Section~\ref{single} for a detailed discussion
1568: (b) Second order correction to the Hartree term (a).}
1569: \label{cor}
1570: \end{figure}
1571: }
1572:
1573:
1574:
1575: \subsection{Soft modes}
1576: \label{soft}
1577:
1578:
1579:
1580: As we already discussed, the main contribution to the temperature
1581: dependence of physical quantities comes from the processes
1582: characterized by spatial scales much larger than the Fermi wave-length
1583: $\lambda_F$. Therefore, there is a scale separation in the problem;
1584: all the Fermi liquid parameters\cite{Fermiliquid} $F_i$ are established
1585: at small distances of the order of $\lambda_F$, and are not affected
1586: by disorder if the relation
1587: $\epsilon_F\tau \gtrsim 1$ holds. On the other hand, all the
1588: temperature and disorder dependence is determined by infrared behavior
1589: of the system where $F_i$ can be considered as fixed.
1590:
1591: Therefore, our first step is to identify the terms in the interaction
1592: Hamiltonian, which may produce the biggest contributions at
1593: temperatures much smaller than the Fermi energy. This procedure
1594: contains nothing new in comparison with the standard identification of
1595: singlet, triplet and Cooper channels, see Ref.~\onlinecite{AGD}, and
1596: we present here the main steps to make the paper self-contained.
1597:
1598: The original interaction Hamiltonian has the form
1599:
1600: \begin{eqnarray*}
1601: %&&
1602: \hat{H}_{int} =
1603: %\label{100}\\
1604: %&&\quad
1605: \sum_{\vec q, \vec p_i}
1606: \frac{V(q)}{2}\psi^\dagger_{\sigma_1}(\vec p_1)
1607: \psi^\dagger_{\sigma_2}(\vec p_2)
1608: \psi_{\sigma_2}(\vec p_2+\vec q)\psi_{\sigma_1}(\vec p_1-\vec q),
1609: %\nonumber
1610: \end{eqnarray*}
1611:
1612: \noindent
1613: and we imply summation over repeated spin indices. Soft modes of the
1614: system correspond to the situation when two of the fermionic operators
1615: have momenta close to each other. The difference of the momenta $q*$
1616: defines the scale $1/q* \gg \lambda_F$, which is the smallest length
1617: scale allowed in the theory. Therefore, we explicitly separate the
1618: Hamiltonian
1619: into a part that contains all the soft modes (first three terms) and a
1620: correction $\delta\hat{H}$, which does not contain such pairs of
1621: fermionic operators:
1622:
1623: \beq
1624: \label{101}
1625: \hat{H}_{int} =
1626: \hat{H}_\rho + \hat{H}_\sigma + \hat{H}_{pp} + \delta\hat{H}.
1627: \label{hamiltonian}
1628: \eeq
1629:
1630: \noindent
1631:
1632:
1633: The explicit expressions for the entries of the Hamiltonian
1634: (\ref{hamiltonian}) are the
1635: following. The interaction in the singlet channel (charge dynamics) is
1636: described by
1637:
1638: \beqa
1639: &&\hat{H}_\rho =
1640: \frac{1}{2}\sum_{|\vec q| < q*, \vec p_i}
1641: \left[V(q)
1642: +\frac{ F^\rho\left(\widehat{\vec n_1 \vec n_2}\right)}{ \nu }
1643: \right] \label{Hrho}\label{102}
1644: \\
1645: &&
1646: \nonumber\\
1647: &&
1648: \times
1649: \left[\psi^\dagger_{\sigma_1}\left(\vec p_1\right)
1650: \psi_{\sigma_1}\left(\vec p_1-{\vec q}\right)
1651: \right]
1652: \left[\psi^\dagger_{\sigma_2}\left(\vec p_2\right)
1653: \psi_{\sigma_2}\left(\vec p_2+{\vec q}\right)
1654: \right],
1655: \nonumber
1656: \eeqa
1657:
1658: \noindent
1659: where $\vec n_i = \vec p_i/|p_i|$, the dimensionless parameter
1660: $F^\rho\left(\widehat{\vec n_1 \vec n_2}\right)$ is related to the
1661: original interaction potential $V(q)$ by
1662:
1663: \beq
1664: \label{103}
1665: F^\rho\left(\theta\right) = - \frac{\nu}{2}
1666: V\left(2k_F\sin\frac{\theta}{2} \right),
1667: \eeq
1668:
1669: \noindent
1670: and $\nu$ is the thermodynamic density of states of non-interacting
1671: electrons (introduced here to make $F^\rho$ dimensionless).
1672:
1673: Interaction in the triplet channel (spin density dynamics) is governed
1674: by
1675:
1676: \beqa
1677: \label{104}
1678: &&\hat{H}_\sigma =
1679: \frac{1}{2}\sum_{\matrix{ { }_{\vec p_i} \cr { }_{|\vec q| < q*} }}
1680: %_{|\vec q| < q*, \vec p_i}
1681: \sum_{ j=x,y,z}
1682: \frac{F^\sigma\left(\widehat{\vec n_1 \vec n_2}\right)}{\nu}
1683: \times
1684: \label{hsigma}
1685: \\
1686: &&
1687: \nonumber\\
1688: &&
1689: \left[\psi^\dagger_{\sigma_1}\left(\vec p_1\right)
1690: \hat\sigma^j_{\sigma_1\sigma_2}
1691: \psi_{\sigma_2}\left(\vec p_1-{\vec q}\right)
1692: \right]
1693: \left[\psi^\dagger_{\sigma_3}\left(\vec p_2\right)
1694: \hat\sigma^j_{\sigma_3\sigma_4}
1695: \psi_{\sigma_4}\left(\vec p_2+{\vec q}\right)
1696: \right]
1697: \nonumber
1698: \eeqa
1699:
1700: \noindent
1701: where parameters
1702: $F^\sigma\left(\widehat{\vec n_1 \vec n_2}\right)$ are
1703:
1704: \beq
1705: \label{105}
1706: F^\sigma\left(\theta\right) = - \frac{\nu}{2}
1707: V\left(2k_F\sin\frac{\theta}{2} \right).
1708: \eeq
1709:
1710: Finally, the Hamiltonian
1711:
1712: \beqa
1713: \label{106}
1714: &&\hat{H}_{pp} =
1715: \sum_{|\vec q| < q*, \vec p_i} \Bigg\{
1716: \frac{F^e\left(\widehat{\vec n_1 \vec n_2}\right)}{\nu}
1717: \left[\psi^\dagger_{\sigma_1}\left(\vec p_1\right)
1718: \hat\sigma^y_{\sigma_1\sigma_2}
1719: \psi^\dagger_{\sigma_2}\left({\vec q}-\vec p_1\right)
1720: \right]
1721: \nonumber\\
1722: &&
1723: \nonumber\\
1724: &&
1725: \quad\quad\quad\quad
1726: \times\left[\psi_{\sigma_3}\left(\vec p_2\right)
1727: \hat\sigma^y_{\sigma_3\sigma_4}
1728: \psi_{\sigma_4}\left({\vec q}-\vec p_2\right)
1729: \right]
1730: \\
1731: &&
1732: \nonumber\\
1733: &&
1734: +
1735: \sum_{ j=x,y,z}
1736: \frac{F^o\left(\widehat{\vec n_1 \vec n_2}\right)}{\nu}
1737: \left[\psi^\dagger_{\sigma_1}\left(\vec p_1\right)
1738: \tilde\sigma^j_{\sigma_1\sigma_2}
1739: \psi^\dagger_{\sigma_2}\left({\vec q}-\vec p_1\right)
1740: \right]
1741: \nonumber\\
1742: &&
1743: \nonumber\\
1744: &&
1745: \quad\quad\quad\quad
1746: \times
1747: \left[\psi_{\sigma_3}\left(\vec p_2\right)
1748: \left(\tilde\sigma^j\right)_{\sigma_4\sigma_3}^\dagger
1749: \psi_{\sigma_4}\left({\vec q}-\vec p_2\right)
1750: \right]\Bigg\}
1751: \nonumber
1752: \eeqa
1753:
1754: \noindent
1755: describes singlet, $F^e$, and triplet, $F^o$, pairing
1756: fluctuations. The parameters in this Hamiltonian are
1757:
1758: \beq
1759: \label{107}
1760: F^{e,o}\left(\theta\right) = \frac{\nu}{4}
1761: \left[
1762: V\left(2k_F\sin\frac{\theta}{2} \right) \pm
1763: V\left(2k_F\cos\frac{\theta}{2} \right)
1764: \right],
1765: \label{fefo}
1766: \eeq
1767:
1768: \noindent
1769: where plus and minus signs correspond to even ($e$) and odd ($o$)
1770: pairing respectively. Here $\hat\sigma_{\sigma_1\sigma_2}^j$ are the
1771: elements of the Pauli matrices in spin space
1772:
1773: \[
1774: \hat\sigma^x =\pmatrix{0 & 1\cr 1&0}, \quad
1775: \hat\sigma^y =\pmatrix{0 & -i\cr i&0}, \quad
1776: \hat\sigma^z =\pmatrix{1 & 0\cr 0&-1},
1777: \]
1778:
1779: \noindent
1780: and $\tilde\sigma^j= \hat{\sigma}^y\hat\sigma^j$.
1781:
1782: Deriving Eqs.~(\ref{Hrho}) -- (\ref{fefo}), we used the condition, $q*
1783: \ll k_F$. This condition allowed us to make the following
1784: approximation
1785:
1786: \begin{eqnarray*}
1787: (\vec p_1 - \vec p_2)^2 \approx 4 k_F^2
1788: \sin^2\left(\frac{\widehat{\vec n_1 \vec n_1}}{2}\right).
1789: \end{eqnarray*}
1790:
1791: \noindent
1792: We also used the identity
1793:
1794: \begin{eqnarray*}
1795: %\label{108}
1796: 2\delta_{\sigma_1\sigma_2}\delta_{\sigma_3\sigma_4}
1797: &=&\delta_{\sigma_1\sigma_3}\delta_{\sigma_2\sigma_4}
1798: + \hat\sigma_{\sigma_1\sigma_3}^j\hat\sigma_{\sigma_4\sigma_2}^j
1799: \\
1800: &&
1801: \\
1802: &=&\hat\sigma_{\sigma_1\sigma_3}^y\hat\sigma_{\sigma_2\sigma_4}^y
1803: + \tilde{\sigma}_{\sigma_1\sigma_3}^j
1804: \left(\tilde{\sigma}^j\right)^\dagger_{\sigma_4\sigma_2}.
1805: \end{eqnarray*}
1806:
1807: So far, the representation (\ref{hamiltonian}) of original interaction
1808: is exact. The only advantage of this representation is that it
1809: explicitly separates the term $\delta H$ which does not contain
1810: coupling to the low energy excitations of the fermionic
1811: system. Therefore, the contribution of $\delta H$ to physical
1812: quantities is regular and not infrared divergent [like
1813: $(T/v_Fq*)^2$]. Therefore, for the electron system with weak short
1814: range interaction, $\delta H$ can be disregarded at all.
1815:
1816: Moreover, even if the interaction is not weak or long range, $\delta
1817: H$ can be treated in all the orders of perturbation theory without
1818: generating a soft mode. If this term does not break the translational
1819: symmetry at short distances, its only effect is to renormalize the
1820: interaction parameters $F$'s in Eqs.~(\ref{102}), (\ref{104}) and
1821: (\ref{106}) and the Fermi velocity in the non-interacting part of the
1822: Hamiltonian. For instance, one obtains for the two-dimensional
1823: electron gas with the Coulomb interaction $V(q)= 2\pi e^2/(\kappa
1824: |q|)$
1825:
1826: \beqa
1827: F^\rho\left(\theta\right) =F^\sigma\left(\theta\right)
1828: = -\frac{1}{2}
1829: \frac{r_s}{\sqrt{2}|\sin \frac{\theta}{2}| + r_s},
1830: \label{108}
1831: \eeqa
1832:
1833: \noindent
1834: where
1835:
1836: \beq
1837: \label{109}
1838: r_s \equiv\frac{\sqrt{2}e^2}{\kappa\hbar v_F}
1839: \eeq
1840:
1841: \noindent
1842: is the conventional parameter characterizing interaction strength and
1843: $\kappa$ is the low frequency dielectric constant of the host
1844: material. Expression (\ref{108}) is applicable only for $r_s \ll 1$,
1845: however, keeping it in denominator is legitimate for small angle
1846: scattering.
1847:
1848: For stronger interaction $r_s \gtrsim 1$, but still far from the
1849: Wigner crystal instability\cite{Wignercrystal}, $r_s \lesssim 37$,
1850: exact calculation of the parameters $F$ from the first principles (as
1851: well as their explicit expressions in terms of $r_s$) is not possible.
1852: Nevertheless, to study the behavior of the system at distances much
1853: larger than $\lambda_F$, one can still disregard the term $\delta H$
1854: in Eq.~(\ref{101}). Then parameters $F$ are no longer bound by
1855: Eqs.~(\ref{103}), (\ref{105}), and (\ref{107}) [or by Eq.~(\ref{108})
1856: for the Coulomb interaction] but rather should be treated as starting
1857: parameters for the low energy theory. The form of Eqs.~(\ref{102}),
1858: (\ref{104}) and (\ref{106}) is guarded by symmetries of the system:
1859: Eq.~(\ref{102}) is guarded by translational symmetry and charge
1860: conservation; Eq.~(\ref{104}) is guarded by translational symmetry and
1861: symmetry with respect to spin rotations; and Eq.~(\ref{106}) is
1862: guarded by all above symmetries and the
1863: electron-hole symmetry, which holds approximately at low energies.
1864:
1865: All the consideration above essentially repeats the basics of the
1866: Landau Fermi-liquid theory\cite{Fermiliquid}. We reiterate, that this
1867: theory does not imply that the interaction is weak; the only
1868: assumption here is that no symmetry is broken at small distances.
1869:
1870:
1871:
1872: \subsection{Disorder averaging}
1873: \label{prelim}
1874: \label{averaging}
1875:
1876:
1877:
1878: To study the interaction correction to conductivity due to charge and
1879: triplet channel interactions introduced in the previous subsection, we
1880: follow the same route as in the case of the short-range
1881: interaction. In particular, the charge channel correction is a direct
1882: generalization of the Fock term. We start however with the discussion
1883: of disorder averaging.
1884:
1885: The correction to conductivity Eq.~(\ref{cc}) represents the first
1886: order perturbation theory in the original potential $V(q)$, valid when
1887: the potential is weak. For stronger coupling we make use of the
1888: effective Hamiltonian Eq.~(\ref{101}). Although the diagrams for
1889: conductivity look similar to the Fock term ``b'' on Fig.~\ref{sr},
1890: their content is now quite different. First, the wavy line now
1891: represents the propagator for one of the soft modes in
1892: Eq.~(\ref{101}). Therefore the expression for the conductivity
1893: Eq.~(\ref{cc}) should be rewritten as
1894:
1895: \begin{eqnarray}
1896: &&\delta\sigma_{\alpha\beta} = - \int\limits_{-\infty}^{\infty}
1897: \frac{d\Omega}{8\pi^2}
1898: \left[\frac{\partial}{\partial\Omega}
1899: \left(\Omega\coth\frac{\Omega}{2T}\right)\right]
1900: \label{cci}
1901: \\
1902: &&
1903: \nonumber\\
1904: &&
1905: \times
1906: {\rm Im}
1907: \int \frac{d^2r_3d^2r_4}{\cal V}
1908: \left\{ \Bigg [{\cal D}^A(\Omega, \vec r_3, \vec r_4) +
1909: {\bf Tr}\widehat{\cal D}_T^A(\Omega, \vec r_3, \vec r_4)
1910: \Bigg ]\right.
1911: \nonumber\\
1912: &&
1913: \nonumber\\
1914: &&
1915: \quad\quad\quad\times
1916: \Big ( B^{\alpha\beta}_F (\Omega, \vec r_3, \vec r_4 )
1917: + \{\alpha\leftrightarrow\beta\}\Big)\Bigg\},
1918: \nonumber
1919: \end{eqnarray}
1920:
1921: \noindent
1922: where ${\cal D}^A$ and $\widehat{\cal D}_T^A$ are advanced propagators
1923: for charge and triplet channels [$\widehat{\cal D}_T$ is a $3\times3$
1924: matrix as follows from Eq.~(\ref{hsigma}), see also
1925: Sec.~\ref{triplet}] and $B_F$ is the product of electronic Green's
1926: functions given by Eq.~(\ref{fck}), the same as in the Fock term.
1927: Deriving Eq.~(\ref{cci}) we assumed that the spin rotational symmetry
1928: is preserved, i.e. no Zeeman splitting or the spin-orbit interaction
1929: is present. We also neglected the dependence of the interaction
1930: propagators on the direction of the electron momenta.
1931: Lifting of those two assumptions is straightforward but it will not
1932: be done in the present paper.
1933: To the leading order in $1/k_Fl$ we can average the propagators
1934: independently of $B_F$ (see e.g. Ref.~\onlinecite{aar}). Here we
1935: proceed with averaging $B_F$ and the discussion of the propagators
1936: follows.
1937:
1938: We have already averaged the product $B_F$ of four Green's functions
1939: for the case of the short-range potential. There the three terms
1940: Eqs.~(\ref{f31}), (\ref{f32}), and (\ref{f13}) vanished due to the
1941: particular form of the potential. Now we have to take these terms into
1942: account and consider the full set of diagrams shown on Figs.~\ref{10}
1943: and \ref{11}. These diagrams can be evaluated in exactly the same way
1944: as those in Section~\ref{h-f} (where we considered a subset of these
1945: diagrams).
1946:
1947:
1948: As a result, the averaged $B_F$ has a form similar to Eq.~(\ref{bf})
1949: and can again be expressed in terms of the dressed vertex $\Gamma$
1950: [see Eq.~(\ref{gsd})]. We are still interested in the longitudinal
1951: conductivity and thus disregard the Hall contribution. Thus, after
1952: averaging the correction Eq.~(\ref{cci}) takes the form
1953:
1954: \begin{eqnarray}
1955: &&\delta\sigma = -e^2 v_F^2 \pi\nu\int\limits_{-\infty}^{\infty}
1956: \frac{d\Omega}{4\pi^2} \frac{\partial}{\partial\Omega}
1957: \left(\Omega\coth\frac{\Omega}{2T}\right) \int \frac{d^2 q}{(2\pi)^2}
1958: \nonumber\\
1959: &&
1960: \nonumber\\
1961: &&\quad\quad\quad
1962: \times{\rm Im} \Bigg\{
1963: \Big[{\cal D}^A(\Omega, q)+{\bf Tr}\widehat{\cal D}_T^A(\Omega, q)\Big]
1964: \tilde B_F(\Omega,q)
1965: \Bigg\};
1966: \nonumber\\
1967: &&
1968: \nonumber\\
1969: &&\tilde B_F(\Omega,q)=-
1970: \frac{2\tau (i\Omega + \frac{1}{\tau})\Gamma}
1971: {S^3}
1972: +\frac{(\Gamma^2-1)\tau^2}{S}
1973: \nonumber\\
1974: &&
1975: \nonumber\\
1976: &&
1977: +\frac{v_F^2 q^2-2\left(i\Omega + \frac{1}{\tau}\right)^2 }
1978: {S^5}\frac{\Gamma^2}{2}
1979: +\frac{\tau\Gamma(\Gamma+1)}{v_F^2 q^2}
1980: \left(\frac{i\Omega + \frac{1}{\tau}}
1981: {S}
1982: - 1 \right)^2
1983: \nonumber\\
1984: &&
1985: \nonumber\\
1986: &&
1987: -\frac{2\Gamma^2}
1988: { S^3}
1989: \left(\frac{i\Omega + \frac{1}{\tau}}
1990: {S}
1991: - 1 \right)
1992: +\frac{\Gamma^3v_F^2 q^2}{
1993: \tau S^6}.
1994: \label{fock-int}
1995: \end{eqnarray}
1996:
1997: \noindent
1998: where quantities $\Gamma$ and $S$
1999: are defined in Eq.~(\ref{gsd}).
2000:
2001:
2002:
2003:
2004: {
2005: \narrowtext
2006: \begin{figure}[ht]
2007: \vspace{0.2 cm}
2008: \epsfxsize=7 cm
2009: \centerline{\epsfbox{f14.eps}}
2010: \vspace{0.2cm}
2011: \caption{Conductivity diagrams, group I. Diagrams (a), (b), and (d)
2012: were evaluated for the short range interaction in Sec.~\protect\ref{h-f}.
2013: In the diffusive regime\protect\cite{aar}
2014: only diagrams (a), (d), (e), were considered at $\omega, qv_F \ll 1/\tau$.
2015: }
2016: \label{10}
2017: \end{figure}
2018: }
2019:
2020: {
2021: \narrowtext
2022: \begin{figure}[ht]
2023: \vspace{0.2 cm}
2024: \epsfxsize=9 cm
2025: \centerline{\epsfbox{f15.eps}}
2026: \vspace{0.2cm}
2027: \caption{Conductivity diagrams, group II.
2028: Diagrams (a) and (b)
2029: were evaluated for the short range interaction in Sec.~\protect\ref{h-f}.
2030: In the diffusive regime\protect\cite{aar}
2031: only diagrams (a), (d), (e), were considered at $\omega, qv_F \ll 1/\tau$.
2032: }
2033: \label{11}
2034: \end{figure}
2035: }
2036:
2037: It is important to emphasize that
2038:
2039: \beq
2040: \tilde B_F(\Omega,q=0) =0,
2041: \label{ginvariance}
2042: \eeq
2043:
2044: \noindent
2045: [to see this one should use explicit
2046: expressions (\ref{gsd}) in Eq.~(\ref{fock-int})].
2047: This property is not accidental -- it is guarded by the gauge invariance
2048: of the system: no interaction with zero momentum transfer can affect
2049: the value of the closed loop.
2050:
2051: To proceed further with the actual
2052: calculation of the correction (\ref{fock-int})
2053: we need to specify the interaction propagator. It will be done
2054: in the following two subsections.
2055:
2056:
2057:
2058:
2059: \subsection{Charge channel}
2060: \label{charge}
2061:
2062:
2063:
2064:
2065: In this section we discuss the charge channel correction,
2066: described by the Hamiltonian (\ref{102}). Because the effective
2067: interaction is characterized by the momentum transfer much smaller
2068: than the Fermi wave vector, the Random Phase Approximation (RPA), see
2069: Fig.~\ref{15}, is applicable.
2070:
2071: To simplify further considerations, we approximate the
2072: Fermi liquid parameter $F^\rho$ by its zero angular harmonic
2073:
2074: \beq
2075: F^\rho(\theta) \approx F^\rho_0,
2076: \label{rhoapprox}
2077: \eeq
2078:
2079: \noindent
2080: this approximation does not affect the final result because of
2081: the long range nature of the Coulomb potential $V(q \to 0) \to
2082: \infty$.
2083:
2084: Consequently, we write the charge channel propagator in the form
2085:
2086: \begin{mathletters}
2087: \label{int-full}
2088: \begin{equation}
2089: {\cal D}^A(\Omega, q) =
2090: - \frac{\nu V(q)+F_0^\rho}{\nu+ (\nu V(q)+F_0^\rho) \Pi^A},
2091: \label{int-dyson}
2092: \end{equation}
2093:
2094: \noindent
2095: where the polarization operator is given by
2096:
2097: \begin{equation}
2098: \Pi^A(\Omega, q) =
2099: \nu \left[ 1 - \frac{i\Omega}{
2100: S-\frac{1}{\tau}}\right],
2101: \label{pol}
2102: \end{equation}
2103:
2104: \noindent
2105: using the notation (\ref{sqrt}).
2106:
2107: The polarization operator (\ref{pol}) differs from the more standard one
2108: (used for instance in Ref.~\onlinecite{aag}) since the diffusion
2109: approximation has not been made yet. Indeed, expanding the
2110: polarization operator in small $\Omega$ and $q$ we can recover the
2111: usual diffusive form. In terms of the scattering time it corresponds
2112: to the limit $T\tau \ll 1$. We do not do that here since we want
2113: to calculate the conductivity for all values of $T\tau$.
2114: \end{mathletters}
2115:
2116: The form of the propagator~(\ref{int-full})
2117: and expression for the conductivity correction (\ref{fock-int})
2118: suggests that there could be two contributions.
2119: First, the propagator Eq.~(\ref{int-full}) has a
2120: pole which corresponds to the 2D plasmon.
2121: However, the plasmon
2122: dispersion relation is
2123:
2124: \[
2125: (v_Fq_{pl})^2 \nu V(q_{pl})
2126: =2\Omega \left(\Omega + \frac{i}{\tau}\right),
2127: \]
2128:
2129: \noindent
2130: i.e. $(v_Fq_{pl})^2 \ll |\Omega \left(\Omega +
2131: \frac{i}{\tau}\right)|$ at all distances larger than the screening
2132: radius. According to the gauge invariance condition
2133: (\ref{ginvariance}) this contribution is strongly suppressed
2134: (by a factor of the order of $max[T,(T/\tau)^{1/2}]d_{sc}/v_F$,
2135: with $d_{sc}$ being the screening radius, $\nu V(1/d_{sc})=1$)
2136: and we will not take it into account.
2137:
2138: Second, at frequencies smaller than the
2139: plasmon frequency we can neglect the unity in the denominator in
2140: Eq.~(\ref{int-full}), which corresponds to the unitary limit, i.e.
2141:
2142: \begin{eqnarray}
2143: {\cal D}^A = -
2144: \frac{1}{\Pi^A} = - \frac{1}{\nu}
2145: \frac{S - \frac{1}{\tau}}{S- \frac{1}{\tau}-i\Omega}.
2146: \label{f-prop}
2147: \end{eqnarray}
2148:
2149: \noindent
2150: Thus the original coupling $V(q)$ as well as the renormalization of the
2151: coupling by the Fermi liquid parameter Eq.~(\ref{102}) does not affect
2152: the resulting propagator. In other words the propagator becomes
2153: universal.
2154:
2155: It is important to emphasize that Eq.~(\ref{f-prop}) gives the upper
2156: bound for the strength of the repulsive interaction. This is
2157: guaranteed by stability of the electron system with respect to
2158: the Wigner crystallization, i.e. by the condition $\nu V(q)+F_0^\rho >
2159: 0$ at $q<q*$. Therefore, we always have
2160:
2161: \begin{eqnarray*}
2162: \frac{\nu V+F_0^\rho}{\nu+ (\nu V+F_0^\rho) \Pi} < \frac{1}{\Pi},
2163: \end{eqnarray*}
2164:
2165: \noindent
2166: so that Eq.~(\ref{f-prop}) is indeed the upper bound for the
2167: propagator Eq.~(\ref{int-dyson}). Note, that the above condition is
2168: satisfied regardless of the sign of $F_0^\rho$.
2169: In particular, it is possible to have
2170: $F_0^\rho<-1$ so that the so-called compressibility of the system
2171: $\nu/(1+F_0^\rho)$ is negative. This fact, however,
2172: has nothing to do with stability of the Fermi liquid and does not
2173: affect transport phenomena \cite{vsi}.
2174:
2175: Using the propagator Eq.~(\ref{f-prop}) in the expression for the
2176: correction Eq.~(\ref{fock-int}) we obtain after momentum integration
2177:
2178: \begin{eqnarray}
2179: \delta\sigma_C =&& -e^2\tau\int\limits_{0}^{\infty}
2180: \frac{d\Omega}{2\pi} \frac{\partial}{\partial\Omega}
2181: \left(\Omega\coth\frac{\Omega}{2T}\right)
2182: \nonumber\\
2183: &&
2184: \nonumber\\
2185: &&
2186: \left\{ \frac{2}{\pi}\arctan\Omega\tau + \frac{1}{\pi\Omega\tau}
2187: + \frac{\Omega\tau}{2\pi} H(\Omega\tau) \ln 2
2188: \right.
2189: \nonumber\\
2190: &&
2191: \nonumber\\
2192: &&
2193: +\frac{1}{\pi}\Big[1+H(\Omega\tau)\Big]\arctan\frac{1}{\Omega\tau}
2194: \nonumber\\
2195: &&
2196: \nonumber\\
2197: &&\left.
2198: +\frac{\Omega\tau}{4\pi}\left[ \frac{1}{2} +H(\Omega\tau)\right]
2199: \ln\left(1+\frac{1}{\Omega^2\tau^2}\right)\right\},
2200: \label{fock-int-3}
2201: \end{eqnarray}
2202:
2203: \noindent
2204: where the dimensionless function $H(x)$ is defined as
2205:
2206: \begin{eqnarray*}
2207: H(x) = \frac{1}{4+x^2}.
2208: \end{eqnarray*}
2209:
2210: In the frequency integral Eq.~(\ref{fock-int-3}) we single out the
2211: first two terms as being dominant in the ballistic and diffusive
2212: limits respectively with the rest being the crossover function. The
2213: diffusive limit is given by
2214:
2215: \begin{eqnarray}
2216: \delta\sigma_C(T\tau\ll 1) =&& -e^2\tau\int\limits_{0}^{\infty}
2217: \frac{d\Omega}{2\pi} \frac{\partial}{\partial\Omega}
2218: \left(\Omega\coth\frac{\Omega}{2T}\right) \frac{1}{\pi\Omega\tau}
2219: =
2220: \nonumber\\
2221: &&
2222: \nonumber\\
2223: &&
2224: = -\frac{e^2}{2\pi^2}\ln\left(\frac{E_F}{T}\right).
2225: \label{cd}
2226: \end{eqnarray}
2227:
2228: \noindent
2229: In the opposite limit we can replace $\arctan\Omega\tau$ by $\pi/2$.
2230: Then the integral is divergent in the ultra-violet, but that large
2231: constant can be incorporated in the definition of $\tau$. This is done
2232: as follows:
2233:
2234: \begin{eqnarray}
2235: \int\limits_{0}^{\infty} d\Omega \frac{\partial}{\partial\Omega}
2236: \left(\Omega\coth\frac{\Omega}{2T}\right)
2237: \rightarrow -2T + E_F\coth\frac{E_F}{2T},
2238: \label{ultra}
2239: \end{eqnarray}
2240:
2241: \noindent
2242: where $E_F$ is put for the upper limit of the integral. This is
2243: consistent with the approximations in momentum integration, where one
2244: typically relies on fast convergence in order to set the integration
2245: limit (otherwise determined by the Fermi energy) to infinity and to
2246: set all momenta in the numerator to the Fermi momentum in magnitude.
2247: Since we are interested in temperatures $T\ll E_F$, the second term is
2248: essentially a temperature independent (although infinite)
2249: constant. The temperature dependent correction to the conductivity is
2250: determined by the first term. As a result
2251:
2252: \begin{eqnarray}
2253: \delta\sigma_C(T\tau\gg 1) =&& -e^2\tau\int\limits_{0}^{\infty}
2254: \frac{d\Omega}{2\pi} \frac{\partial}{\partial\Omega}
2255: \left(\Omega\coth\frac{\Omega}{2T}\right)
2256: =
2257: \nonumber\\
2258: &&
2259: \nonumber\\
2260: &&
2261: = e^2\frac{T\tau}{\pi}.
2262: \label{cb}
2263: \end{eqnarray}
2264:
2265: Integrating the full expression Eq.~(\ref{fock-int-3}) we find the
2266: correction valid at all values of $T\tau$,
2267:
2268: \begin{equation}
2269: \delta\sigma_C = -\frac{e^2}{2\pi^2}\ln\left(\frac{E_F}{T}\right)
2270: + e^2\frac{T\tau}{\pi}
2271: \left[ 1 -\frac{3}{8} f(T\tau)\right],
2272: \label{fc1}
2273: \end{equation}
2274:
2275: \noindent
2276: where the dimensionless function $f(x)$ is defined as a dimensionless
2277: integral:
2278:
2279: \begin{eqnarray}
2280: f(x) = \frac{8}{3}
2281: \int\limits_{0}^{\infty}&& dz \left[ \frac{\partial}{\partial z}
2282: \left(z\coth z\right) - 1\right]
2283: \left\{\frac{xz}{\pi} H(2xz) \ln 2 \right.
2284: \nonumber\\
2285: &&
2286: \nonumber\\
2287: &&
2288: +\frac{1}{\pi}\Big[1+H(2xz)\Big]\arctan\frac{1}{2xz}
2289: \nonumber\\
2290: &&
2291: \nonumber\\
2292: &&
2293: +\frac{xz}{2\pi}\left[\frac{1}{2}+H(2xz)\right]
2294: \ln\left(1+\frac{1}{(2xz)^2}\right)
2295: \nonumber\\
2296: &&
2297: \nonumber\\
2298: &&\left.
2299: +
2300: \frac{2}{\pi}\arctan 2xz - 1\right\}.
2301: \label{f2}
2302: \end{eqnarray}
2303:
2304: \noindent
2305: The factor $3/8$ is introduced for convenience, so that $f(0)=1$. The
2306: integral can be evaluated analytical in the two limiting cases and the
2307: result is given by Eq.~(\ref{fl}). In the intermediate regime the
2308: integral can be evaluated numerically and the result is plotted on
2309: Fig.~\ref{plot:f}.
2310:
2311:
2312:
2313:
2314: \subsection{Triplet channel}
2315: \label{triplet}
2316:
2317:
2318:
2319: %\subsubsection{Triplet channel interaction}
2320:
2321:
2322: In this section we discuss the correction in the triplet channel.
2323: Similar to the case of the charge channel, we need to derive the
2324: interaction propagator in the triplet channel and then use
2325: Eq.~(\ref{fock-int}). As follows from the Hamiltonian
2326: Eq.~(\ref{hsigma}), the triplet channel propagator is now a $3 \times
2327: 3$ matrix. Apart from this minor complication, the propagator can be
2328: found using the same RPA approximation as the one used in
2329: Section~\ref{charge}, see Fig.~\ref{15}.
2330:
2331: {
2332: \narrowtext
2333: \begin{figure}[ht]
2334: \vspace{0.1 cm}
2335: \epsfxsize=8.5 cm
2336: \centerline{\epsfbox{f16.eps}}
2337: \vspace{0.5cm}
2338: \caption{Interaction propagator in the (a) singlet and (b)
2339: triplet channel}
2340: \label{15}
2341: \end{figure}
2342: }
2343:
2344: Similarly to the charge channel, we take the Fermi liquid coupling
2345: $\hat F^\sigma$ to be independent of electron momenta
2346:
2347: \begin{eqnarray}
2348: F^\sigma\left(\theta\right) \approx F^\sigma_0.
2349: \label{appr}
2350: \end{eqnarray}
2351:
2352: \noindent
2353: Unlike the case of the charge channel, this approximation slightly
2354: affects final results (see discussion after Eqs.~(\ref{s}) for the
2355: drawbacks of this approximation as well as for its remedies). Then
2356: the matrix equation for the triplet propagator has the form
2357:
2358: \begin{equation}
2359: \left[{\cal D}_T\right]_{ij} =
2360: -\delta_{ij}\frac{F^\sigma_0}{\nu}
2361: - \frac{F^\sigma_0}{\nu}\sum_{k=x,y,z}\hat
2362: \Pi_{ik}
2363: \left[{\cal D}_T\right]_{kj}
2364: \label{prop-eq}
2365: \end{equation}
2366:
2367: \noindent
2368: where $i,j=x,y,z$.
2369:
2370: In the absence of the magnetic field and spin-orbit scattering each
2371: electronic Green's function is a diagonal matrix in the spin space,
2372: and therefore
2373:
2374: \beq
2375: \Pi_{ik}^A(q,\Omega)=\delta_{ik}\Pi^A(q,\Omega),
2376: \label{psigma}
2377: \eeq
2378:
2379: \noindent
2380: where $\Pi^A(q,\Omega)$ is the polarization operator given by
2381: Eq.~(\ref{pol}).
2382:
2383: Altogether, using Eq.~(\ref{psigma}) in the equation (\ref{prop-eq}),
2384: we find the triplet channel propagator as
2385:
2386: \begin{eqnarray}
2387: \left[{\cal D}_T^A(\Omega, q)\right]_{ij} = - \delta_{ij}
2388: \frac{F^\sigma_0}{\nu + F^\sigma_0 \Pi^A(\Omega, q)}.
2389: \label{trip-prop}
2390: \end{eqnarray}
2391:
2392: Before we continue, let us discuss the validity of the approximation
2393: Eq.~(\ref{appr}). Consider the situation close to the Stoner
2394: instability $F^\sigma_0 \rightarrow -1$. In this case the pole of the
2395: propagator Eq.~(\ref{trip-prop}) describes a magnetic excitation in
2396: the system. In the ballistic case ($q>1/l$) it is a slow, over-damped
2397: spin wave
2398:
2399: \begin{eqnarray*}
2400: -i\omega \approx (1+F^\sigma_0)v_F|q|.
2401: \end{eqnarray*}
2402:
2403: \noindent
2404: The main contribution to the temperature dependent conductivity
2405: correction comes from this pole at $\omega\sim T$. The corresponding
2406: typical momenta are $k^*\sim T / [(1+F^\sigma_0)v_F]$. Although we
2407: are using the momentum independent $F^\sigma_0$, it is known
2408: \cite{bkm} that fluctuations in the triplet channel produce a
2409: non-analytic correction to the spin susceptibility, so up to a
2410: numerical coefficient $F^\sigma\approx F^\sigma_0 (1 - |q|/k_F)$. Such
2411: momentum dependence can only be neglected if $k^*\le
2412: k_F(1+F^\sigma_0)$. This translates into a limitation for the
2413: temperature range where the results listed in Section~\ref{results}
2414: are valid\cite{lar}:
2415:
2416: \begin{eqnarray}
2417: T\ll T^* \approx (1+F^\sigma_0)^2 E_F.
2418: \end{eqnarray}
2419:
2420: \noindent
2421: At higher temperatures $T>T^*$ our theory is not applicable.
2422:
2423: Having discussed the validity of the approach, we proceed with the
2424: straightforward calculation: one has to substitute the propagator
2425: Eq.~(\ref{trip-prop}) into the expression for the correction
2426: Eq.~(\ref{fock-int}) and evaluate the integral. The result of the
2427: momentum integration is given by
2428:
2429: \begin{mathletters}
2430: \begin{eqnarray}
2431: &&\delta\sigma_T = -3e^2\tau
2432: \int\limits_{-\infty}^{\infty}
2433: \frac{d\Omega}{4\pi} \frac{\partial}{\partial\Omega}
2434: \left(\Omega\coth\frac{\Omega}{2T}\right)
2435: \nonumber\\
2436: &&
2437: \nonumber\\
2438: &&
2439: \left\{ \left(1-\frac{1}{F_0^\sigma}\ln(1+F_0^\sigma)\right)
2440: \frac{1}{\pi\Omega\tau}
2441: + \frac{F_0^\sigma}{(1+F_0^\sigma)}\left[
2442: \frac{2}{\pi}\arctan\Omega\tau
2443: \right.\right.
2444: \nonumber\\
2445: &&
2446: \nonumber\\
2447: &&
2448: + \frac{\Omega\tau}{2\pi} \left(\Big[H(\Omega\tau)+h_1(\Omega\tau)\Big] \ln 2
2449: + h_4(\Omega\tau)\right)
2450: \label{fock-int-t}
2451: \\
2452: &&
2453: \nonumber\\
2454: &&
2455: +\frac{1}{\pi}\Big[1+H(\Omega\tau)
2456: + (\Omega\tau)^2h_3(\Omega\tau)\Big]\arctan\frac{1}{\Omega\tau}
2457: \nonumber\\
2458: &&
2459: \nonumber\\
2460: &&\left.\left.
2461: +\frac{\Omega\tau}{4\pi}\left[ \frac{1}{2} +H(\Omega\tau)
2462: + h_2(\Omega\tau)\right]
2463: \ln\left(1+\frac{1}{\Omega^2\tau^2}\right)\right]\right\},
2464: \nonumber
2465: %\label{fock-int-t}
2466: \end{eqnarray}
2467:
2468: \noindent
2469: where we introduce notations:
2470:
2471: \begin{eqnarray}
2472: h_1(x)=\tilde H&&(x; 1+2F_0^\sigma)
2473: \nonumber\\
2474: &&
2475: \nonumber\\
2476: &&\times
2477: \Big[5+6F_0^\sigma-4(2+3F_0^\sigma)H(x)\Big],
2478: \end{eqnarray}
2479: \begin{eqnarray}
2480: h_2(x)=&&h_1(x)+\tilde H(x; F_0^\sigma)
2481: \left[-\frac{1}{2}(1+F_0^\sigma)\right.
2482: \nonumber\\
2483: &&
2484: \nonumber\\
2485: &&\left.
2486: +F_0^\sigma x^2
2487: \left(\frac{1}{2}-(1+F_0^\sigma)\tilde H(x; F_0^\sigma)\right)\right],
2488: \end{eqnarray}
2489: \begin{eqnarray}
2490: h_3(x)=&&\tilde H(x; 1+2F_0^\sigma)
2491: \Big[-1-2F_0^\sigma+(2+3F_0^\sigma)H(x)\Big]
2492: \nonumber\\
2493: &&
2494: \nonumber\\
2495: &&
2496: +\frac{F_0^\sigma}{2}\tilde H(x; F_0^\sigma)
2497: \Big[1+F_0^\sigma x^2\tilde H(x; F_0^\sigma)\Big],
2498: \end{eqnarray}
2499: \begin{eqnarray}
2500: h_4(x)=\tilde H(x; F_0^\sigma)&&\left[\frac{5F_0^\sigma-3}{2}+
2501: \frac{1-(F_0^\sigma)^2}{F_0^\sigma}\ln(1+F_0^\sigma)\right]
2502: \nonumber\\
2503: &&
2504: \nonumber\\
2505: &&
2506: +h_5(x)\frac{1+F_0^\sigma}{F_0^\sigma}\ln(1+F_0^\sigma),
2507: \end{eqnarray}
2508: \begin{eqnarray}
2509: &&h_5(x)=(2F_0^\sigma-1)\tilde H(x; 1+2F_0^\sigma) +
2510: \tilde H^2(x; F_0^\sigma)
2511: \\
2512: &&
2513: \nonumber\\
2514: &&\quad\quad
2515: \times
2516: \left[\left(\frac{1}{2}-2F_0^\sigma\right)(1+F_0^\sigma)^2 -
2517: (F_0^\sigma)^2x^2 \left(\frac{1}{2}+2F_0^\sigma\right)\right].
2518: \nonumber
2519: \end{eqnarray}
2520: \end{mathletters}
2521:
2522: \noindent
2523: Here we introduce a dimensionless function $\tilde H(x; y)$
2524:
2525: \begin{eqnarray*}
2526: \tilde H(x; y) = \frac{1}{(1+y)^2+(xy)^2},
2527: \end{eqnarray*}
2528:
2529: \noindent
2530: which is related to the function $H(x)$ introduced in Section~\ref{charge}
2531: simply by $H(x)=\tilde H(x; 1)$.
2532:
2533: The expression in brackets turns into its counterpart in the charge
2534: channel in the unitary limit ($F^\sigma_0\rightarrow\infty$). Its
2535: first term describes the diffusive limit described in
2536: Ref.~\onlinecite{aar} (the formal difference in the coefficient stems
2537: from the difference in the definition of the coupling constant). The
2538: frequency integral is evaluated in the same way as in Eq.~(\ref{cd}).
2539: Similar to our discussion of the charge channel correction [see e.g.
2540: Eq.~(\ref{cb})], we identify the second term in Eq.~(\ref{fock-int-t})
2541: with the ballistic limit (which we discuss in more detail in the next
2542: Section). The intermediate temperature regime is described by the
2543: expression [which appeared previously in Section~\ref{results},
2544: Eq.~(\ref{tc})]:
2545:
2546: \begin{eqnarray}
2547: \delta\sigma_T = &&-3\left(1-\frac{1}{F_0^\sigma}
2548: \ln(1+F_0^\sigma)\right)
2549: \frac{e^2}{2\pi^2}\ln\left(\frac{E_F}{T}\right)
2550: \nonumber\\
2551: &&
2552: \nonumber\\
2553: &&
2554: +\frac{3F_0^\sigma}{(1+F_0^\sigma)}e^2\frac{T\tau}{\pi}
2555: \left[ 1 -\frac{3}{8}t(T\tau; F_0^\sigma)\right].
2556: \label{tc2}
2557: \end{eqnarray}
2558:
2559: \noindent
2560: where the dimensionless function $t(x; F_0^\sigma)$ is defined as
2561:
2562: \begin{eqnarray}
2563: t(x; &&F_0^\sigma) = \frac{8}{3}
2564: \int\limits_{0}^{\infty} dz \left[ \frac{\partial}{\partial z}
2565: \left(z\coth z\right) - 1\right]
2566: \nonumber\\
2567: &&
2568: \nonumber\\
2569: &&\times
2570: \Bigg\{\frac{xz}{\pi}
2571: \left(\Big[H(2xz)+h_1(2xz)\Big] \ln 2 + h_4(2xz)\right)
2572: \nonumber\\
2573: &&
2574: \nonumber\\
2575: &&\quad
2576: +\frac{1}{\pi}\Big[1+H(2xz)+4x^2z^2h_3(2xz)\Big]\arctan\frac{1}{2xz}
2577: \nonumber\\
2578: &&
2579: \nonumber\\
2580: &&\quad
2581: +\frac{xz}{2\pi}\left[\frac{1}{2}+H(2xz)+h_2(2xz)\right]
2582: \ln\left(1+\frac{1}{(2xz)^2}\right)
2583: \nonumber\\
2584: &&
2585: \nonumber\\
2586: &&\quad
2587: +
2588: \left[\frac{2}{\pi}\arctan(2xz) - 1\right]\Bigg\}.
2589: \label{t2}
2590: \end{eqnarray}
2591:
2592: \noindent
2593: Except for the limiting cases [see Eq.~(\ref{tsx})] the integral in
2594: Eq.~(\ref{t2}) has to be evaluated numerically. We plot the result
2595: for several values of $F_0^\sigma$ in Fig.~\ref{plot:t}.
2596:
2597:
2598:
2599:
2600: \subsection{Single impurity limit}
2601: \label{single}
2602:
2603:
2604:
2605: In the previous Sections we obtained the expression for the correction
2606: to conductivity averaged over disorder. To complete the calculation we
2607: needed to separately average the interaction propagator and use the
2608: result to evaluate the integral in Eq.~(\ref{fock-int}). In doing this
2609: we assumed that the dimensionless conductance of the system is large
2610: or in terms of the scattering time $\tau E_F\gg 1$. We have not,
2611: however, assumed anything about the relative value of the scattering
2612: rate and temperature. In other words, the correction
2613: Eq.~(\ref{fock-int}) is valid in both the diffusive $T\tau\ll 1$ and
2614: ballistic $T\tau\gg 1$ limits. It also describes the cross-over
2615: behavior at intermediate temperatures.
2616:
2617: The temperature behavior of the interaction correction in the limiting
2618: cases can of course be obtained from the general result
2619: Eq.~(\ref{s}). As we pointed out in Section~\ref{results}, in the
2620: diffusive limit our results coincide with the standard theory,
2621: Ref.~\onlinecite{aar}. On the other hand the correction in the
2622: ballistic limit is subject to conflicting claims in literature
2623: \cite{gdl,rei}. Unfortunately, neither result is completely
2624: correct. Therefore we discuss the ballistic limit in some detail,
2625: starting with diagrams before averaging (i.e. diagrams on
2626: Figs.~\ref{sr}). This way we are able to point out exactly which
2627: diagram produces the dominant result and which diagrams were missed in
2628: existing theories.
2629:
2630: We begin by discussing the Hartree term. This contribution was
2631: considered in Ref.~\onlinecite{gdl} in the framework of the
2632: temperature dependent dielectric function. The physical idea was that
2633: electrons tend to screen the charged impurities and thus modify the
2634: scattering rate. In what follows we show which diagrams describe this
2635: process and how to calculate the resulting correction, which appears
2636: to be the same (up to a numerical factor miscalculated in
2637: Ref.~\onlinecite{gdl}; see below for detailed explanation). The
2638: important difference between the two approaches is that the impurity
2639: screening picture described only the direct (Hartree) interaction,
2640: while missing on the exchange part. The latter was later considered in
2641: Ref.~\onlinecite{rei}. We think that this consideration is erroneous,
2642: and we discuss the Fock term in Section~\ref{sif}.
2643:
2644:
2645:
2646: \subsubsection{Single impurity limit for Hartree term}
2647: \label{sih}
2648:
2649:
2650:
2651:
2652: The goal of this discussion is to show which diagrams correspond to
2653: the ballistic limit of the Hartree term (as discussed in
2654: Section~\ref{qualitative}) and how it relates to other interaction
2655: corrections we discuss in this paper.
2656:
2657: The Hartree term corresponds to averaging the two diagrams on
2658: Fig.~\ref{sr}, where the wavy line represents a weak interaction
2659: potential. In this case the diagram ``a3'' of Fig.~\ref{sr} is equal
2660: to zero even before the averaging (as a total derivative) and we only
2661: need to average the diagrams ``a1'' and ``a2''. The rigorous procedure
2662: would involve dressing the interaction vertices according to
2663: Fig.~\ref{8} and adding diffusons Fig.~\ref{9} as it was done in
2664: Section~\ref{h-f} (see Fig.~\ref{hfav}), evaluating the resulting
2665: expression, and finally taking the limit $T\tau\rightarrow\infty$.
2666: However, the same result can be obtained by making the expansion by
2667: noticing that impurity line brings smallness $1/T\tau$. Therefore,
2668: high temperature limit may be studied by considering diagrams on
2669: Fig.~\ref{13} directly. Such approach is completely equivalent to that
2670: of Ref.~\onlinecite{gdl}. The result [which can also be obtained from
2671: the general expression Eq.~(\ref{tc})] is similar to the one obtained
2672: in Ref.~\onlinecite{gdl} (the difference is the extra factor of $\ln
2673: 2$ found in Ref.~\onlinecite{gdl} due to an error in this reference,
2674: which consists in putting the energy of the scattered electron on the
2675: Fermi shell rather than integrating over it):
2676:
2677: \begin{eqnarray}
2678: \delta\sigma_H = - 4\sigma_D\left(\frac{T}{E_F}\right)
2679: \left[-\nu{\cal D}(2k_F)\right]
2680: \label{hart-si}
2681: \end{eqnarray}
2682:
2683: \noindent
2684: (for weak coupling ${\cal D}(2k_F) \equiv - V(2k_F)$ The factor of $4$
2685: in Eq.~(\ref{hart-si}) can be interpreted as a result of a summation
2686: over four spin configurations. Although correct for weak coupling,
2687: this factor should be modified when stronger interaction is
2688: considered, see discussion above.
2689:
2690:
2691: {
2692: \narrowtext
2693: \begin{figure}[ht]
2694: \vspace{0.5 cm}
2695: \epsfxsize=7 cm
2696: \centerline{\epsfbox{f17.eps}}
2697: \vspace{0.5cm}
2698: \caption{Single impurity diagrams for Hartree channel.}
2699: \label{13}
2700: \end{figure}
2701: }
2702:
2703:
2704: \subsubsection{Fock contribution}
2705: \label{sif}
2706:
2707:
2708:
2709: In the similar manner one can discuss the single impurity contribution
2710: to the Fock term. Again, for weak interaction we could simply expand
2711: the result of disorder averaging for the Fock term Eq.~(\ref{bf})
2712: to the leading order in $1/T\tau$. For Coulomb interaction we would
2713: expand Eq.~(\ref{fock-int}), since in Eq.~(\ref{bf}) the special form
2714: of the delta-function potential was utilized to eliminate the diagram
2715: ``b3'' on Fig.~\ref{sr}. Diagrammatically, such expansions equivalent
2716: to direct evaluation of diagrams without impurity lines (but with
2717: averaged electron Green's functions) shown in Fig.~\ref{3} and
2718: diagrams with only one impurity line shown on Figs.~\ref{4}, \ref{5},
2719: and \ref{6}.
2720:
2721: The evaluation of the single impurity diagrams for the Fock term is
2722: straight-forward and is completely analogous to the Hartree term
2723: discussed in the previous subsection. The result can be written as
2724:
2725: \begin{equation}
2726: \sigma_F = \frac{e^2\tau T}{\pi}.
2727: \label{fock-si}
2728: \end{equation}
2729:
2730: \noindent
2731: This result contradicts (even in sign) that of
2732: Ref.~\onlinecite{rei}. Here we briefly discuss the reason for this
2733: contradiction. We notice that one has to be careful to keep track of
2734: gauge invariance while evaluating diagrams for the Fock term. Gauge
2735: invariance manifests itself in the fact that any interaction at zero
2736: momentum gives no contribution to physical quantities, which are
2737: expressed diagrammatically as closed loops, see
2738: Eq.~(\ref{ginvariance}). This is indeed the case for
2739: Eq.~(\ref{fock-int}), where we summed up all the diagrams. On the
2740: other hand, any individual diagram is not gauge invariant. In
2741: particular, each subset of diagrams in Figs.~\ref{33}-\ref{6} is not
2742: gauge invariant. Therefore to obtain the result Eq.~(\ref{fock-si})
2743: from these diagrams one has to disregard terms which contain higher
2744: than second powers of the scattering rate $1/\tau$.
2745:
2746: {
2747: \narrowtext
2748: \begin{figure}[ht]
2749: \vspace{0.2 cm}
2750: \epsfxsize=7.2 cm
2751: \centerline{\epsfbox{f18.eps}}
2752: \vspace{0.2cm}
2753: \caption{Fock channel diagrams without impurity lines.}
2754: \label{33}
2755: \end{figure}
2756: }
2757:
2758: As we already mentioned, the contribution from the plasmon pole is
2759: small due to the condition (\ref{ginvariance}). However, in
2760: Ref.~\onlinecite{rei} it was claimed otherwise. Namely, diagrams in
2761: Fig.~\ref{33} were claimed to be important for the plasmon correction
2762: and to give a large result, while diagrams in Figs.~\ref{4}-\ref{6}
2763: were alleged to be not important for the plasmon correction. This
2764: claim explicitly violates gauge invariance and leads to incorrect
2765: conclusions. In particular, the plasmon contribution to the
2766: conductivity was overestimated by a factor of order of $(v_F/d_s T)
2767: \simeq (E_F/ T)$.
2768:
2769:
2770: {
2771: \narrowtext
2772: \begin{figure}[ht]
2773: \vspace{0.1 cm}
2774: \epsfxsize=6.5 cm
2775: \centerline{\epsfbox{f19.eps}}
2776: \vspace{0.2cm}
2777: \caption{Single impurity diagrams for Fock channel with the impurity
2778: line dressing one interaction vertex.}
2779: \label{4}
2780: \end{figure}
2781: }
2782: {
2783: \narrowtext
2784: \begin{figure}[ht]
2785: \vspace{0.1 cm}
2786: \epsfxsize=6.5 cm
2787: \centerline{\epsfbox{f20.eps}}
2788: \vspace{0.2cm}
2789: \caption{Single impurity diagrams for Fock channel with the impurity
2790: line connecting a retarded and an advanced Green's functions across the
2791: diagram.}
2792: \label{5}
2793: \end{figure}
2794: }
2795: {
2796: \narrowtext
2797: \begin{figure}[ht]
2798: \vspace{0.2 cm}
2799: \epsfxsize=2.7 cm
2800: \centerline{\epsfbox{f21.eps}}
2801: \vspace{0.1cm}
2802: \caption{The single impurity diagram for Fock channel with the
2803: impurity line connecting two advanced Green's functions.}
2804: \label{6}
2805: \end{figure}
2806: }
2807:
2808:
2809:
2810:
2811: \section{Kinetic equation approach}
2812: %\section{Semi-classical kinetic theory}
2813: \label{eilenberger}
2814:
2815:
2816:
2817:
2818: Our purpose in this Section is to put the treatment of the interaction
2819: effects in disordered systems into the framework of the kinetic
2820: equation. Even though at this point this will not produce any further
2821: physical results, this proves to be more convenient for practical
2822: calculations of more sophisticated quantities, such as the Hall
2823: coefficient, the thermal conductivity, energy relaxation, etc. The
2824: kinetic equation approach is also applicable for the description of
2825: non-linear effects. The main technical advantage of the kinetic
2826: equation is that it operates with gauge invariant quantities from the
2827: very beginning, unlike the perturbation theory, where each diagram
2828: taken separately is not gauge invariant (and may produce non-physical
2829: divergences).
2830:
2831: We will present the final form of the kinetic equation in subsection
2832: \ref{kinetic}, and show how to operate with this equation for the
2833: conductivity calculation in subsection \ref{cond}. The derivation of
2834: this equation based on the Keldysh technique for non-equilibrium
2835: systems \cite{Keldysh} is presented in subsection \ref{derivation}.
2836:
2837:
2838:
2839:
2840: \subsection{Final form of the kinetic equation}
2841: \label{kinetic}
2842:
2843:
2844:
2845:
2846: As usual in the kinetic equation approach, averages of observable
2847: quantities are expressed as certain integrals of the distribution
2848: function $f(t; \epsilon,\vec r,\vec n)$. For instance the averaged
2849: density is
2850:
2851: \begin{mathletters}
2852: \label{eq:4.1}
2853: \beq
2854: \rho (t,\vec r) = \nu\int\limits_{-\infty}^{\infty} d\epsilon
2855: \langle f(t; \epsilon , \vec r, \vec n)\rangle_n
2856: \eeq
2857:
2858: \noindent
2859: and the average current is
2860:
2861: \beq
2862: {\vec J}(t,\vec r) =e\nu v_F \int\limits_{-\infty}^{\infty} d\epsilon
2863: \langle \vec n f(t; \epsilon , \vec r ,\vec n)\rangle_n
2864: \label{current}
2865: \eeq
2866:
2867: \noindent
2868: and so on. Here $\nu$ is the density of states (entering into linear
2869: specific heat of the clean system) at the Fermi surface and $v_F$ is
2870: the Fermi velocity, $\vec n = (\cos\theta, \sin\theta)$ is the unit
2871: vector in the direction of the electron momentum and angular averaging
2872: is introduced as
2873:
2874: \[
2875: \langle \dots \rangle_n =
2876: \int
2877: \frac{d\theta}{2\pi} \dots
2878: \]
2879:
2880: \end{mathletters}
2881:
2882: The Boltzmann-like equation for the distribution function has the form
2883:
2884: \beq
2885: \label{eq:4.2}
2886: \left[ \partial_t + v_F \vec n \vec\nabla +
2887: e v_F (\vec n {\vec E})\frac{\partial}{\partial \epsilon}
2888: +\vec\omega_c
2889: %\mbox{ \boldmath $\w$}_c
2890: \left(\vec n \times \frac{\partial}{\partial \vec n} \right)
2891: \right] f =
2892: \St\left\{f\right\},
2893: \label{Boltzmann}
2894: \eeq
2895:
2896: \noindent
2897: where ${\vec E}$ denotes the external electric field and
2898: %$\mbox{\boldmath $\w$}_c$
2899: $\vec\omega_c$ is a vector with the magnitude equal to the cyclotron
2900: frequency corresponding to an external magnetic field perpendicular to
2901: the plane and the direction along the field.
2902:
2903:
2904: Equations (\ref{eq:4.1}) and (\ref{Boltzmann}) neglect energy
2905: dependence of the velocity of electrons, which makes it inapplicable
2906: for quantities associated with electron-hole asymmetry, such as the
2907: thermopower. On the other hand, any component of the thermal and
2908: electrical conductivities is still within our description.
2909:
2910: All of the interaction effects are taken into account in the collision
2911: integral
2912:
2913: \beq
2914: \St\left\{f\right\} = \St_{el}\left\{f\right\}
2915: + \St_{in}\left\{f\right\}.
2916: \label{eq:4.3}
2917: \eeq
2918:
2919: The elastic part of the collision integral describes scattering of
2920: electrons by static impurities (we assume point-like scattering;
2921: generalization to the finite range is straightforward) as well as by
2922: the self-consistent field generated by all the other electrons:
2923:
2924: \beqa
2925: \St_{el}\left\{f\right\}=&& -\frac{f(t; \epsilon,\vec r,\vec n)
2926: -\langle f(t; \epsilon, \vec r,\vec n)\rangle_n}{\tau}
2927: \nonumber\\
2928: &&
2929: \nonumber\\
2930: &&
2931: + \frac{8}{\tau}I_0(t; \epsilon,\vec n, \vec r)
2932: \langle f(t; \epsilon, \vec r,\vec n)\rangle_n
2933: \label{eq:4.4}
2934: \\
2935: &&
2936: \nonumber\\
2937: &&
2938: + \frac{8}{\tau}
2939: n_\alpha I_1^{\alpha\beta} (t; \epsilon, \vec r) \langle n_\beta
2940: f(t;\epsilon, \vec r,\vec n)\rangle_n.
2941: \nonumber
2942: \eeqa
2943:
2944: \noindent
2945: The effect of the self-consistent field is described by the last two
2946: terms, where we introduce notations:
2947:
2948: \begin{mathletters}
2949: \label{eq:4.5}
2950: \beqa
2951: &&I_0(t; \epsilon,\vec n, \vec r)= -
2952: \int \frac{d\omega}{2\pi}
2953: \Big\{n_\alpha K_0^{\alpha\beta}(\omega)
2954: \langle n_\beta f(t; \epsilon-\omega, \vec r,\vec n)\rangle_n
2955: \nonumber \\
2956: &&
2957: \nonumber \\
2958: &&
2959: +\frac{n_\alpha L_0^{\alpha\beta}(\w)}{2}
2960: \left[\nabla_\beta + e E_\beta\frac{\partial}{\partial \epsilon} \right]
2961: \langle f(t; \epsilon-\omega, \vec r,\vec n)\rangle_n
2962: \Big\}
2963: \label{eq:4.5a}
2964: \eeqa
2965: \beqa
2966: I_1^{\alpha\beta}(t; \epsilon, \vec r)= -
2967: \int \frac{d\w}{2\pi} K_1^{\alpha\beta}(\omega)
2968: \langle f(t; \epsilon-\omega, \vec r,\vec n)\rangle_n.
2969: %\nonumber \\
2970: %&&{}
2971: \label{eq:4.5b}
2972: \eeqa
2973: \end{mathletters}
2974:
2975: \noindent
2976: The collision integral Eq.~(\ref{eq:4.4}) preserves the number of
2977: particles on a given energy shell: integrating $\St_{el}\{f(t;
2978: \epsilon,\vec r,\vec n)\}$ over directions of $\vec n$ gives zero for
2979: any value of $\epsilon$ [see also Eq.(\ref{elastic2})].
2980:
2981: The term $I_1$ expresses enhanced momentum relaxation due to static
2982: disorder. The physics of this term was discussed in detail in Section
2983: ~\ref{qualitative}. The term $I_0$ describes electron scattering by
2984: non-equilibrium non-local Fock like potential created by all other
2985: electrons. This process is responsible for generation of the finite
2986: drift velocity of electrons. One can easily see that $I_0$
2987: vanishes in the equilibrium situation $f(\epsilon, \vec n, \vec r)=
2988: f[\epsilon + e\varphi(\vec r)],\ \nabla_\alpha \varphi = - E_\alpha$.
2989:
2990: The kernels $K_0$, $K_1$, and $L_0$ entering into Eqs.~(\ref{eq:4.5})
2991: can be expressed in terms of interaction propagators and the
2992: propagators describing
2993: semi-classical dynamics of non-interacting electrons. Explicitly:
2994:
2995: \begin{mathletters}
2996: \label{eqs4.6}
2997: \beqa
2998: &&K_1^{\alpha\beta}(\omega) =
2999: \Im \int \frac{d^2q}{(2\pi)^2} {\cal{D}}^R(\omega,\vec q)
3000: \label{K1}
3001: \\
3002: &&
3003: \nonumber\\
3004: &&\quad \times\left\{ \< n_\alpha D \> \< D n_\beta \> -
3005: \frac{\delta_{\alpha \beta}}{2} \left( \< D \> \< D \> +
3006: i \frac{\partial}{\partial \omega} \< D \> \right) \right\}
3007: \nonumber
3008: \eeqa
3009: \beqa
3010: &&
3011: K_0^{\alpha\beta}(\omega) =
3012: \Im \int \frac{d^2q}{(2\pi)^2} {\cal{D}}^R(\omega,\vec q)
3013: \label{K0}\\
3014: &&
3015: \nonumber\\
3016: &&
3017: \quad \times
3018: \left\{ \< n_\alpha D n_\beta \>\< D \>
3019: -\frac{i}{v_F}\frac{\partial}{\partial q_\alpha} \< D n_\beta \>
3020: - \< D n_\alpha \>\< D n_\beta \>
3021: \right\}
3022: \nonumber
3023: \eeqa
3024: \beqa
3025: &&
3026: L_0^{\alpha\beta}(\omega) =
3027: - \Re \int \frac{d^2q}{(2\pi)^2} {\cal{D}}^R(\omega,\vec q)
3028: \label{L0}
3029: \label{eq:4.6c}\\
3030: &&
3031: \nonumber\\
3032: &&
3033: \quad\times
3034: \left\{\< D \> \frac{\partial}{\partial q_\beta} \< n_\alpha D \>
3035: -\< D n_\alpha \>\frac{\partial}{\partial q_\beta}\< D \>
3036: - \< D n_\alpha \frac{\partial}{\partial q_\beta} D \>\right\}
3037: \nonumber
3038: \eeqa
3039: \label{kernels}
3040: \end{mathletters}
3041:
3042: Here, ${\cal D}^R(\omega)={\cal D}^A(-\omega)$
3043: denotes the retarded interaction propagator [see
3044: i.e. Eq.~(\ref{f-prop})] and we introduce the
3045: short hand notation for the angular averaging
3046:
3047: \begin{eqnarray*}
3048: \<a D b\> &\equiv& \int\frac{d \theta d\theta'}{(2\pi)^2}
3049: a(\vec n)
3050: D(\vec n, \vec n'; \omega, q)b(\vec n'),
3051: \\
3052: &&
3053: \nonumber\\
3054: \<a D b D c\> &\equiv& \int\frac{d \theta d\theta'd\theta''}{(2\pi)^3}
3055: a(\vec n)
3056: D(\vec n, \vec n')b(\vec n')D(\vec n', \vec n'')c(\vec n'')
3057: \end{eqnarray*}
3058:
3059: \noindent
3060: for arbitrary functions $a,b$. The function $D(\vec n, \vec n';
3061: \omega, \vec q)$ describes the classical motion of a particle on the
3062: energy shell $\epsilon_F$ in a magnetic field:
3063:
3064: \beqa
3065: &&\Big[-i\omega + i v_F \vec n \vec q +
3066: %\mbox{ \boldmath $\w$}_c
3067: \vec \omega_c
3068: \left(\vec n \times
3069: \frac{\partial}{\partial \vec n} \right)\Big]
3070: D(\vec n, \vec n'; \omega, \vec q)
3071: \label{Diffuson}
3072: \\
3073: &&
3074: \nonumber\\
3075: &&
3076: +\frac{1}{\tau} \Big( D(\vec n, \vec n'; \omega, \vec q)
3077: - \langle D(\vec n, \vec n'; \omega, \vec q) \rangle_n \Big) =
3078: 2\pi \delta(\widehat{\vec n \vec n'}).
3079: \nonumber
3080: \eeqa
3081:
3082: As we have already mentioned, the elastic part of the collision
3083: integral is nulled by a distribution function of the form
3084: $f[\epsilon +e\varphi(\vec r)]$ for an arbitrary $f$.
3085: It is the inelastic term
3086: that is responsible for establishing the local thermal equilibrium
3087: and it has the standard form
3088:
3089: \begin{mathletters}
3090: \label{eq:4.8}
3091: \beqa
3092: \St_{in}&&\left\{f\right\} =
3093: \int d\omega \int d \epsilon_1 A(\omega)
3094: f( \epsilon_1)
3095: \left[1 - f(\epsilon_1 - \omega)\right]
3096: \label{inelastic}
3097: \\
3098: &&
3099: \nonumber\\
3100: &&
3101: \times
3102: \Big\{ - f(\epsilon)
3103: \left[1 - f( \epsilon + \omega)\right]
3104: + \left[1 - f( \epsilon)\right]f( \epsilon - \omega)
3105: \Big\}
3106: \nonumber
3107: \eeqa
3108: \begin{eqnarray*}
3109: f(\epsilon) = \langle f(t; \epsilon, \vec r,\vec n)\rangle_n.
3110: \end{eqnarray*}
3111:
3112: \noindent
3113: The kernel $A(\omega)$ describes matrix elements for inelastic
3114: processes in both ballistic and diffusive limits. The explicit
3115: expression for this kernel is
3116:
3117: \beq A(\omega) = \frac{2\nu}{\pi} \int\frac{d^2q}{(2\pi)^2}
3118: \left[ \Re \langle D \rangle \right]^2
3119: \left|{\cal{D}}^R(\omega, \vec q) \right|^2,
3120: \label{eq:A}
3121: \eeq
3122: \end{mathletters}
3123:
3124: \noindent
3125: where $ \langle D \rangle$ is given by the solution to
3126: Eq.~(\ref{Diffuson}) averaged over angles.
3127:
3128: The above equations are written for the interaction in the singlet
3129: channel only. In a situation where both triplet and singlet channels
3130: are present, but the distribution function does not have a spin
3131: structure (no Zeeman splitting or non-equilibrium spin occupation
3132: present), one has to replace
3133:
3134: \beq
3135: {\cal D}^R \to {\cal D}^R + {\bf Tr} \widehat{\cal D}^R_T
3136: \label{r1}
3137: \eeq
3138:
3139: \noindent
3140: in Eqs.~(\ref{eqs4.6}) and
3141:
3142: \beq
3143: |{\cal D}^R|^2 \to |{\cal D}^R|^2 + {\bf Tr}
3144: \left\{ \widehat{\cal D}^R_T
3145: \left[\widehat{\cal D}^R_T\right]^\dagger\right\}
3146: \label{r2}
3147: \eeq
3148:
3149: \noindent
3150: in Eq.~(\ref{eq:A}).
3151:
3152: Equations (\ref{eq:4.2}) - (\ref{eq:4.8}) constitute the complete
3153: system of transport equations with the leading interaction corrections
3154: taken into account. They may be used to study both linear and
3155: non-linear response. We reiterate that they do not include effects of
3156: electron-hole asymmetry and in this form can not produce finite
3157: thermopower. The Hall effect, the thermal conductivity, and energy
3158: relaxation, however, are included and will be studied in a
3159: subsequent publication \cite{pre}. In the following Subsection we
3160: apply the kinetic equation approach to study the interaction
3161: correction to the conductivity at intermediate and low temperatures
3162: and reproduce the results obtained in Section~\ref{diagrams} by means
3163: of diagrammatic technique. The reason for doing so is to show how the
3164: kinetic equation works and to demonstrate explicitly that both
3165: approaches are equivalent.
3166:
3167: Closing our description of the structure of the kinetic equation, we
3168: discuss the range of its applicability. Any kinetic equation implies
3169: that the distribution function changes slowly
3170: on the spatial scale of the
3171: Fermi wave length $\lambda_F$ and on the time scale $1/\epsilon_F$. In
3172: our case, the conditions are more restrictive. First, in the
3173: interaction correction to the elastic collision integral we take
3174: into account only the effect of the interaction on the zeroth and
3175: first angular harmonics of the distribution function. This implies
3176: that the equation gives the correct description for the interaction
3177: effects on the conductivity and diffusion, whereas it is not correct
3178: for description of the quantities involving higher angular
3179: harmonics. Second, we made a gradient expansion in Eq.~(\ref{eq:4.5a})
3180: and only took into account terms linear in the electric field. This
3181: implies that the distribution function changes slowly on the spatial
3182: scale $L_T = {\rm min}(\hbar v_F/T,\ v_F (\hbar\tau/T)^{1/2}),$ and
3183: on the time scale of the order of $\hbar/T$. The electric field
3184: expansion is justified by the condition $eE L_T \ll T$. One can check
3185: that both these conditions are satisfied, if the energy relaxation
3186: time is much longer than the time for the elastic collisions.
3187: We also did not include quantum effects of the magnetic field.
3188: This is justified at $\omega_c \ll max(T/\hbar, \tau^{-1})$.
3189:
3190: Finally, the interaction part of Eq.~(\ref{eq:4.4}) is calculated in
3191: the first loop approximation. It means, that it has to be considered
3192: as the first order correction to $1/\tau$. If one is interested in the
3193: next order interaction correction to the elastic part, one should
3194: take into account the second loop correction, which is not considered
3195: in the present paper. On the contrary, the inelastic part (\ref{eq:4.8})
3196: can be considered in all orders to find the zero angular momentum
3197: part of the distribution function; the only assumption here is the
3198: validity of the Fermi liquid description at energies smaller than
3199: $\epsilon_F$.
3200:
3201:
3202:
3203:
3204: \subsection{Conductivity calculation}
3205: \label{cond}
3206:
3207:
3208:
3209:
3210: In order to calculate the conductivity at zero magnetic field $\omega_c
3211: =0$ we look for the solution of
3212: Eqs.~(\ref{eq:4.2}) - (\ref{eq:4.4}) in a form
3213:
3214: \beq
3215: f(\vec n,\epsilon) = f_F(\epsilon) + \vec n
3216: %\mbox{\boldmath $\Gamma$}
3217: \vec \Gamma
3218: (\epsilon),
3219: \label{eq:4.9}
3220: \eeq
3221:
3222: \noindent
3223: where $f_F(\epsilon)=1/(e^{\epsilon/T}+1)$ is the Fermi distribution
3224: function (all the energies are counted from the Fermi level), and
3225: $\Gamma$ is the quantity to be found
3226: and it is proportional to the electric field.
3227:
3228: We substitute Eq.~(\ref{eq:4.9}) into Eqs.~(\ref{eq:4.2}),
3229: (\ref{eq:4.4}), (\ref{eq:4.5}) and (\ref{eq:4.8}). The inelastic part
3230: of the collision integral [see Eq.~(\ref{eq:4.8})] obviously vanishes,
3231: as effects of the heating are proportional to at least the second
3232: power of the electric field. As a result, we obtain an equation for
3233: $\Gamma$:
3234:
3235: \beqa
3236: e v_F E_\alpha \frac{\partial f_F(\epsilon) }{\partial \epsilon}
3237: &&= - \frac{ \Gamma_\alpha (\epsilon)}{\tau}
3238: %- \frac{4}{\tau}\int\frac{d\w}{2\pi}
3239: \label{eq:4.10}\\
3240: &&
3241: \nonumber\\
3242: &&
3243: %&\times&
3244: - \frac{4}{\tau}\int\frac{d\omega}{2\pi}
3245: \Big[ K_1^{\alpha\beta}(\omega)
3246: f_F(\epsilon-\omega)\Gamma_\beta (\epsilon)
3247: \nonumber\\
3248: &&
3249: \nonumber\\
3250: &&
3251: \quad\quad\quad\quad+
3252: K_0^{\alpha\beta}(\omega) f_F(\epsilon)\Gamma_\beta (\epsilon-\omega)
3253: \Big]
3254: \nonumber\\
3255: &&
3256: \nonumber\\
3257: &&
3258: -\frac{4f_F(\epsilon)}{\tau}\int\frac{d\omega}{2\pi}
3259: L_0^{\alpha\beta}(\omega)
3260: e E_\beta\frac{\partial}{\partial \epsilon}
3261: f_F(\epsilon-\omega)
3262: \nonumber
3263: \eeqa
3264:
3265: \noindent
3266: We solve Eq.~(\ref{eq:4.10}) by iterations. As usual for kinetic
3267: equations, the solution is expressed in terms of the unperturbed
3268: distribution function $f_F(\epsilon)$ and the kernels, which in this
3269: case are given by Eq.~(\ref{kernels}):
3270:
3271: \beqa
3272: \Gamma_\alpha (\epsilon)
3273: &&
3274: = - e v_F\tau E_\alpha \frac{\partial f_F(\epsilon) }{\partial \epsilon}
3275: \label{eq:4.11}
3276: \\
3277: &&
3278: \nonumber\\
3279: &&
3280: + {4 e v_F\tau}\int\frac{d\w}{2\pi}
3281: \left[ K_1^{\alpha\beta}(\w)
3282: f_F(\epsilon-\w)\frac{\partial f_F(\epsilon) }{\partial \epsilon}
3283: \right.
3284: \nonumber\\
3285: &&
3286: \nonumber\\
3287: &&
3288: \left.
3289: \quad\quad\quad\quad\quad\quad+
3290: K_0^{\alpha\beta}(\w)
3291: f_F(\epsilon)\frac{\partial f_F(\epsilon-\w) }{\partial \epsilon}
3292: \right] E_\beta
3293: \nonumber\\
3294: &&
3295: \nonumber\\
3296: &&
3297: - {4f_F(\epsilon)}\int\frac{d\w}{2\pi}
3298: L_0^{\alpha\beta}(\w)
3299: e E_\beta\frac{\partial}{\partial \epsilon}
3300: f_F(\epsilon-\w)
3301: \nonumber
3302: \eeqa
3303:
3304: \noindent
3305: Substituting Eqs.~(\ref{eq:4.11}) into Eq.~(\ref{eq:4.9}) and
3306: the result into Eq.~(\ref{current}), we
3307: integrate over $\epsilon$ and find
3308: the conductivity
3309:
3310: \begin{mathletters}
3311: \beqa
3312: \sigma = \sigma_D + \delta \sigma,
3313: \label{eq:4.12}
3314: \eeqa
3315: \begin{eqnarray}
3316: \frac{\delta\sigma}{\sigma_D}=
3317: \int\limits_{-\infty}^\infty\frac{d\omega}{\pi}
3318: \frac{\partial}{\partial\omega} &&
3319: \left(\omega\coth \frac{\omega}{2T}\right)
3320: \nonumber\\
3321: &&
3322: \nonumber\\
3323: &&\times
3324: \left[K_0(\omega) - K_1(\omega) -\frac{L_0(\omega)}{v_F\tau}\right],
3325: \label{ccke}
3326: \end{eqnarray}
3327: \end{mathletters}
3328:
3329: \noindent
3330: where the Drude conductivity is $\sigma_D = e^2\nu v_F^2\tau/2$.
3331: Here we used the fact that in the absence of the magnetic field
3332: all the kernels are diagonal, $K^{\alpha\beta}_{i} =
3333: \delta_{\alpha\beta}K_i$,
3334: $L^{\alpha\beta}_{0} = \delta_{\alpha\beta}L_0$.
3335: We also used the identities
3336:
3337: \begin{eqnarray*}
3338: 2\int\limits_{-\infty}^\infty d \epsilon f_F(\epsilon)
3339: \frac{\partial f_F(\epsilon-\omega) }{\partial \epsilon}
3340: =-1+\frac{\partial}{\partial\omega}
3341: \left(\omega\coth \frac{\omega}{2T}\right),
3342: \end{eqnarray*}
3343: \begin{eqnarray*}
3344: \int\limits_{-\infty}^\infty
3345: d\omega K_{i}(\omega)= \int\limits_{-\infty}^\infty
3346: d\omega L_{0}(\omega) = 0.
3347: \end{eqnarray*}
3348:
3349: In order to derive explicit expressions for the kernels $K_i$ and $L_0$
3350: we have to solve Eq.~(\ref{Diffuson}) for the function $D$ in the absence
3351: of the magnetic field.
3352: The result can be written as
3353:
3354: \beqa \label{eq:4.13}
3355: D(\vec n,\vec n';\omega,\vec q) && =
3356: 2\pi\delta(\widehat{\vec n \vec n'})
3357: D_0(\vec n,\omega,\vec q)
3358: \\
3359: &&
3360: \nonumber\\
3361: &&
3362: +D_0(\vec n,\omega,\vec q)D_0(\vec n',\omega,\vec q)
3363: \frac{C}{C\tau-1},
3364: \nonumber
3365: \eeqa
3366:
3367: \noindent
3368: where $D_0$ denotes the solution of Eq.~(\ref{Diffuson}) without the
3369: angular averaged term (and in the absence of the magnetic field)
3370:
3371: \beqa
3372: D_0(\vec n,\omega,\vec q)=
3373: \frac{1}{ -i\omega + i v_F \vec n \vec q+1/\tau}.
3374: \nonumber
3375: \eeqa
3376:
3377: \noindent
3378: Here we used the short-hand notation
3379:
3380: \[
3381: C=\sqrt{\left(-i\omega + 1/\tau\right)^2 + v_F^2 q^2},
3382: \]
3383:
3384: \noindent
3385: which is similar to the notation $S$ used in Section~\ref{diagrams}
3386: [in fact, $C=S^*$, see Eq.~(\ref{sqrt})].
3387: Substituting Eq.~(\ref{eq:4.13}) into Eqs.~(\ref{K1}) -- (\ref{L0})
3388: and performing the angular integration we arrive to
3389:
3390: \begin{mathletters}
3391: \beqa
3392: &&K_1(\omega)= -{\Im} \int \frac{q dq}{4\pi} {\cal D}^R(\omega, q)
3393: \\
3394: &&
3395: \nonumber\\
3396: && \; \;
3397: \times\left\{ \frac{1}{v_F^2 q^2}
3398: \left( \frac{C - (-i\omega + 1/\tau)}{C-1/\tau} \right)^2
3399: + \frac{C- (-i\omega +1/\tau)}{C (C-1/\tau)^2} \right\},
3400: \nonumber\\
3401: &&
3402: \nonumber\\
3403: &&K_0(\omega)= {\Im} \int \frac{q dq}{4\pi}\ {\cal D}^R(\w, q)
3404: \\
3405: &&
3406: \nonumber\\
3407: && \quad
3408: \times\left\{ \frac{C-(-i\w + 1/\tau)}{C (C-1/\tau)^2} +
3409: \frac{(C-(-i\w + 1/\tau))^2}{C(C-1/\tau)} \frac{1}{v_F^2 q^2} \right\},
3410: \nonumber\\
3411: &&
3412: \nonumber\\
3413: &&\frac{L_0(\omega)}{v_F\tau} = - {\Im} \int \frac{q dq}{4\pi}
3414: {\cal D}^R(\w, q)
3415: \\
3416: &&
3417: \nonumber\\
3418: && \quad\quad\quad\quad
3419: \times\left\{ \frac{3}{2 \tau} \frac{v_F^2 q^2}{C^3 (C-1/\tau)^2} +
3420: \frac{v_F^2 q^2}{C^3} \frac{1/\tau^2}{(C-1/\tau)^3}
3421: \right\}.
3422: \nonumber
3423: \eeqa
3424: \label{fullkern}
3425: \end{mathletters}
3426:
3427: \noindent
3428: Together with the conductivity correction Eq.~(\ref{ccke}) the above
3429: expressions Eq.~(\ref{fullkern}) are identical to Eq.~(\ref{fock-int})
3430: obtained in Section~\ref{diagrams} by means of the standard perturbation
3431: theory. Thus the kinetic equation approach is completely equivalent to
3432: such diagrammatic calculation.
3433:
3434: Integration over the wave vector $q$ requires the knowledge of the
3435: interaction propagator. Substituting Eq.~(\ref{f-prop}) for the singlet
3436: channel and Eq.~(\ref{trip-prop}) for the triplet channel and performing
3437: the straightforward integration we arrive to the results in
3438: Section~\ref{results}.
3439:
3440:
3441:
3442:
3443: \subsection{Derivation of the kinetic equation}
3444: \label{derivation}
3445:
3446:
3447:
3448:
3449: In this section we derive the kinetic equation discussed in
3450: Section~\ref{kinetic}. For simplicity we show the derivation for the
3451: case of the singlet channel interaction Eq.~(\ref{Hrho}). The case
3452: of the triplet channel can be treated in the same manner with minor
3453: differences (introduction of extra spin indices) described in the end
3454: of this Section. To keep the discussion at the same level as in
3455: Section~\ref{diagrams}, we treat the Fermi liquid parameter $F^\rho$
3456: in Eq.~(\ref{Hrho}) as a constant, similar to our treatment of the
3457: triplet channel in Section ~\ref{triplet}.
3458:
3459:
3460:
3461: \subsubsection{Keldysh formalism}
3462:
3463:
3464:
3465: Here we summarize the results originally obtained by Keldysh
3466: \cite{Keldysh} that enable us to calculate correlation functions for
3467: any non-equilibrium distribution.
3468:
3469: Let us first consider a Green's function of
3470: electrons before disorder averaging.
3471: The electron-electron interaction
3472: is described by the Hamiltonian Eq.~(\ref{Hrho}). In the path-integral
3473: formulation it can be decoupled from fermion operators using an
3474: auxiliary bosonic field $\phi$. Then the Green's function can be written
3475: as
3476:
3477: \beq
3478: \label{fullgreen}
3479: \widehat G(x_1, x_2) = \int \left[ {\cal D} \phi \right]
3480: \widehat G(x_1, x_2 | \phi) e^{-i S_B[\phi]},
3481: \eeq
3482:
3483: \noindent
3484: with the action defined as
3485:
3486: \beq
3487: S_B[\phi] = \int\limits_{-\infty}^{\infty} dt d^2r
3488: \left\{\frac{1}{2} \phi^T V_0^{-1} \sigma_3 \phi \right\}
3489: + i \log Z[\phi] ,
3490: \label{sb}
3491: \eeq
3492:
3493: \noindent
3494: where $-V_0$ is the (bare; following Eq.~(\ref{Hrho})
3495: $V_0=V(q) +F_0^\rho/\nu$) electron-electron interaction propagator
3496: and $Z$ is the partition function,
3497:
3498: \beqa
3499: Z[\phi] = \< {\bf T_C } e^{-i S_F[ \phi, \psi]} \>
3500: \label{z}
3501: \eeqa
3502: \beqa
3503: S_{F}[ \phi, \psi] =
3504: \int\limits_{-\infty}^{\infty} dt d^2r \left\{
3505: \psi^\dagger \phi_{\alpha} \hat\gamma^{\alpha} \psi
3506: \right\},
3507: \label{sfu}
3508: \eeqa
3509:
3510: \noindent
3511: where $\hat\sigma_z={\rm diag}(-1,1)$ is the Pauli matrix in the Keldysh
3512: space.
3513:
3514: \begin{figure}
3515: \vglue 0cm
3516: \hspace{0.01\hsize}
3517: \epsfxsize=0.9\hsize
3518: \epsffile{f22.eps}
3519: \refstepcounter{figure} \label{contour}
3520: {\center{\small Fig.\ \ref{contour}.
3521: The Keldysh contour
3522: \par}}
3523: \end{figure}
3524:
3525: \noindent
3526: In the above expressions all the fields are defined on the Keldysh time
3527: contour shown in Fig.~\ref{contour}. In particular, the fermionic fields
3528: $\psi^\dagger$ and $\psi$ (as well as the bosonic field $\phi$) can be
3529: treated as doublets
3530:
3531: \beqa
3532: \psi = \left( \matrix{\psi_+ \cr \psi_-} \right),
3533: \eeqa
3534:
3535: \noindent
3536: where we adopt the notation that fields
3537: with a $-$ ($+$) subscript (also referred to by Greek letters in this
3538: Section) reside on the lower (upper) part of the
3539: contour on Fig.~(\ref{contour}).
3540: The time dependent fermionic operators $\psi$ are taken in the
3541: interaction represntation
3542: \[
3543: -i\partial_t \psi(t) = \left[\hat{H}_1(t); \psi (t) \right],
3544: \]
3545: where $\hat{H}_1$ is the one-electron Hamiltonian which includes
3546: the static disorder potential as well as external fields.
3547:
3548: Consequently, the Green's function in
3549: Eq.~(\ref{fullgreen}) is a $2\times2$ matrix. Time ordering along the
3550: contour is denoted in Eq.~(\ref{z}) by ${\bf T_C }$.
3551: Matrices $\hat\gamma^{\alpha}$ in Eq.~(\ref{sfu}) are defined as
3552:
3553: \[
3554: \hat\gamma^+ = \left(
3555: \matrix{
3556: -1 & 0 \cr
3557: 0 & 0
3558: }
3559: \right)
3560: \quad ;
3561: \quad\quad
3562: \hat\gamma^- = \left(
3563: \matrix{
3564: 0 & 0 \cr
3565: 0 & 1
3566: }
3567: \right)
3568: \]
3569:
3570: The Green's function $\widehat G(x_1, x_2 | \phi)$ in Eq.~(\ref{fullgreen})
3571: is given by
3572:
3573: \beq
3574: \widehat G(x_1, x_2 | \phi) = \frac{1}{Z[\phi]}
3575: \< T_C \psi^\dagger_{\alpha}(x_1) \psi_{\beta}(x_2)
3576: e^{-i S_F[ \phi, \psi]} \>.
3577: \label{fg2}
3578: \eeq
3579:
3580: \noindent
3581: Here, as well as in Eq.~(\ref{z}) the angular brackets
3582: $\langle \dots \rangle$ denote quantum-mechanical averaging.
3583: In this section we will use the short hand notation
3584:
3585: \[
3586: x_i \equiv (t_i, \vec r_i).
3587: \]
3588:
3589: The bosonic action Eq.~(\ref{sb}) can be treated in the
3590: saddle point approximation:
3591:
3592: \begin{mathletters}
3593: \beqa
3594: \< e^{-i S_B[\phi]}\> = e^{-i F[\phi]},
3595: \eeqa
3596: \beqa
3597: F[\phi] = F[\phi = 0] +
3598: \frac{1}{2}\phi^T \widehat\Pi \phi + {\cal O}(\phi^3),
3599: \label{fb}
3600: \eeqa
3601: \label{ba}
3602: \end{mathletters}
3603:
3604: \noindent
3605: where $\Pi$ is the electronic polarization operator, defined as
3606:
3607: \beq
3608: \Pi_{\alpha \beta}(x_1, x_2) = \left.
3609: \frac{\delta^2 F}{\delta\phi_{\alpha}(x_1)\delta\phi_{\beta}(x_2)}
3610: \right|_{\phi=0}.
3611: \label{pder}
3612: \eeq
3613:
3614: \noindent
3615: The quadratic expansion in Eq.~(\ref{fb}) is justified, provided that
3616: the fields $\phi$ are slowly changing on the scale much larger
3617: than $\lambda_F$.
3618:
3619: Let us now average the Green's function Eq.~(\ref{fullgreen}) over
3620: disorder:
3621:
3622: \beq
3623: \label{disorderedgreen}
3624: \< \widehat G(x_1, x_2) \>_{dis} =
3625: \int \left[ {\cal D} \phi \right]
3626: \< \widehat G(x_1, x_2 | \phi) \>_{dis} e^{-i \< S_B[\phi]\>_{dis}},
3627: %_{dis}
3628: \eeq
3629: where $\<\dots \>_{dis}$ hereafter denotes averaging over disorder.
3630:
3631: Here we average the electronic Green's function Eq.~(\ref{fg2})
3632: separately from the bosonic action Eq.~(\ref{ba}). This approximation
3633: means that we neglect correlations between mesoscopic
3634: fluctuations of the polarizability
3635: in Eq.~(\ref{fb}) and
3636: the fermionic operators in Eq.~(\ref{fg2}) (which describe the motion of
3637: conduction electrons).
3638: This is the same approximation we used in Section~\ref{diagrams}.
3639: It is justified by the well known fact that mesoscopic
3640: fluctuations are smaller than average quantities by a factor of the
3641: order $1/(E_F\tau)^2$.
3642:
3643: It is convenient \cite{LO77} to rotate the Keldysh basis as follows
3644:
3645: \beqa
3646: \widehat G \rightarrow \frac{1}{2}\hat\sigma_x
3647: \left(
3648: \matrix{
3649: 1 & -1 \cr
3650: 1 & 1
3651: }
3652: \right)
3653: \widehat G
3654: \left(
3655: \matrix{
3656: 1 & 1 \cr
3657: -1 & 1
3658: }
3659: \right).
3660: \eeqa
3661:
3662: \noindent
3663: In the new basis the Green's function Eq.~(\ref{fg2}) has the form
3664:
3665: \beq
3666: \widehat G(x_1, x_2 | \phi) =
3667: \left(
3668: \matrix{
3669: G^R(x_1, x_2 | \phi) & G^K(x_1, x_2 | \phi) \cr
3670: G^Z(x_1, x_2 | \phi) & G^A(x_1, x_2 | \phi)
3671: }
3672: \right).
3673: \label{nosense}
3674: \eeq
3675:
3676: \noindent
3677: After the averaging over the bosonic field and over the disorder
3678: according to Eq.~(\ref{disorderedgreen}) the entries in
3679: Eq.~(\ref{nosense}) acquires the following meaning/ where after
3680: integrating over the bosonic field $\phi$ the diagonal elements
3681: $G^{R(A)}$ become the retarded (advanced) Green's functions of the
3682: electron system
3683:
3684: \begin{eqnarray*}
3685: \< G^R(t_1, t_2) \> = -i\eta(t_1-t_2)\langle \psi (t_1) \psi^\dagger(t_2)+
3686: \psi^\dagger(t_2) \psi (t_1) \rangle,\\
3687: \< G^A(t_1, t_2) \> =
3688: i\eta(t_2-t_1)\langle \psi (t_1) \psi^\dagger(t_2)+
3689: \psi^\dagger(t_2) \psi (t_1) \rangle,
3690: \end{eqnarray*}
3691:
3692: \noindent
3693: where $\eta(t)$ is the Heaviside step function. The lower diagonal
3694: element vanishes due to the causality,
3695:
3696: \[
3697: \< G^Z(t_1, t_2) \>=0,
3698: \]
3699:
3700: \noindent
3701: even before the disorder averaging. Finally, the upper off-diagonal
3702: element (the so-called Keldysh Green's function) is related to the one
3703: particle density matrix
3704:
3705: \begin{eqnarray}
3706: \< G^K(t_1, t_2) \> = -i \langle \psi (t_1) \psi^\dagger(t_2)-
3707: \psi^\dagger(t_2) \psi (t_1) \rangle,
3708: \label{GK}
3709: \end{eqnarray}
3710:
3711: \noindent
3712: the quantum mechanical averaging is performed with an arbitrary
3713: distribution function to be found from the solution of the kinetic
3714: equation.
3715:
3716: The bosonic field in the rotated basis has the two components
3717:
3718: \beqa
3719: \phi_{1(2)} = \frac{1}{2} \left( \phi_+ \pm \phi_- \right)
3720: \eeqa
3721:
3722: \noindent
3723: which are described by the propagators
3724:
3725: \begin{mathletters}
3726: \label{Ds}
3727: \beqa
3728: &&
3729: \< \phi_1(t_1, \vec r_1) \phi_1(t_2, \vec r_2) \> =
3730: \frac{i}{2} {\cal D}^K(t_1,t_2; \vec r_1,\vec r_2),
3731: \\
3732: &&
3733: \nonumber\\
3734: &&
3735: \< \phi_1(t_1, \vec r_1) \phi_2(t_2, \vec r_2) \> =
3736: \frac{i}{2} {\cal D}^R(t_1,t_2; \vec r_1,\vec r_2),
3737: \\
3738: &&
3739: \nonumber\\
3740: &&
3741: \< \phi_2 (t_1, \vec r_1)\phi_1(t_2, \vec r_2) \> =
3742: \frac{i}{2} {\cal D}^A(t_1,t_2; \vec r_1,\vec r_2),
3743: \\
3744: &&
3745: \nonumber\\
3746: &&
3747: \< \phi_2 (t_1, \vec r_1)\phi_2(t_2, \vec r_2) \> = 0.
3748: \label{do}
3749: \eeqa
3750: \label{prp}
3751: \end{mathletters}
3752:
3753: \noindent
3754: The coupling Eq.~(\ref{sfu}) between the fermionic and bosonic fields
3755: in the rotated basis has the form:
3756:
3757: \beq
3758: \psi^\dagger \phi_{\alpha} \hat\gamma^{\alpha} \psi \rightarrow
3759: \psi^\dagger
3760: \left(
3761: \matrix{
3762: \phi_1 & \phi_2 \cr
3763: \phi_2 & \phi_1
3764: }
3765: \right) \psi.
3766: \eeq
3767:
3768: \noindent
3769: The propagators Eq.~(\ref{prp}) are solutions of the Dyson equations
3770:
3771: \beqa
3772: \hat{\cal D}(1,2) = \hat{\cal D}_0(1,2) + \int d 3 d4
3773: \hat{\cal D}_0(1,3)\hat{\Pi}(3,4)\hat{\cal D}(4,2)
3774: \nonumber
3775: \eeqa
3776: \beqa
3777: \hat{\cal D} =
3778: \left(
3779: \matrix{{\cal D}^R & {\cal D}^K \cr 0 & {\cal D}^A}
3780: \right),
3781: \quad
3782: \hat{\Pi} =
3783: \left(
3784: \matrix{\Pi^R & \Pi^K \cr 0 & \Pi^A}
3785: \right)
3786: \label{Dysonb}
3787: \eeqa
3788:
3789: \noindent
3790: and we introduced the short hand notation $(i) \equiv (t_i,\vec r_i)$.
3791: The bare interaction propagators are
3792:
3793: \beqa
3794: {\cal D}_0^R = {\cal D}_0^A && = - \left[V(\r_1-\r_2) +
3795: F_0^\rho \delta(\r_1-\r_2)
3796: \right] \delta(t_1-t_2),
3797: \nonumber \\
3798: &&
3799: \nonumber \\
3800: {\cal D}_0^K && = 0.
3801: \label{D00}
3802: \eeqa
3803:
3804: Any classical external field takes identical values on the two
3805: branches of the contour and, hence, in the rotated basis has only a
3806: diagonal component.
3807:
3808: The matrix Green's function (\ref{nosense}) satisfies the equation
3809:
3810: \beqa
3811: &&\Bigg\{
3812: i\partial_{t_1} + E_F
3813: - \frac{\left[-i\vec\nabla_{r_1} + \vec A_{ext}(x_1)\right]^2}{2m}
3814: - \hat{\phi}(x_1)
3815: \label{eq1}
3816: \\
3817: &&
3818: \nonumber\\
3819: &&
3820: \quad
3821: - U(\vec r_1) - \varphi_{ext}(x_1)
3822: \Bigg\} \;
3823: \widehat G(x_1, x_2 | \phi) = \hat{I}\delta(x_1-x_2)
3824: \nonumber
3825: \eeqa
3826:
3827: \noindent
3828: where $U(\vec r)$ is the potential due to the static disorder,
3829: $\vec A_{ext}(x_1)$ and $\varphi_{ext}(x_1)$ are the vector and
3830: scalar potential due to the external electric and magnetic fields.
3831:
3832: \beq
3833: e{\vec E}= \partial_t {\vec A}_{ext} -
3834: %\mbox{\boldmath $\nabla$}
3835: \vec\nabla
3836: \varphi_{ext}, \quad
3837: e{\vec B}=-\frac{1}{c}\vec\nabla \times {\vec A}_{ext}
3838: \label{EH}
3839: \eeq
3840:
3841: \noindent
3842: The equation (\ref{eq1}) is the basis for the further consideration.
3843: One can perform the disorder average in Eq.~(\ref{eq1}) in the leading
3844: in $1/(E_F\tau)$ approximation, which amounts to summation over all
3845: the non-intersecting impurity lines one obtains
3846:
3847: \beqa
3848: &&
3849: \Bigg\{
3850: i\partial_{t_1} + E_F
3851: - \frac{\left[-i\vec\nabla_{r_1} + \vec A_{ext}(x_1)\right]^2}{2m}
3852: - \hat{\phi}(x_1)
3853: \nonumber\\
3854: &&
3855: \nonumber\\
3856: &&\quad\quad
3857: - \varphi_{ext}(x_1)
3858: \Bigg\} \;
3859: \widehat G(x_1, x_2 | \phi) =
3860: \label{eq2}
3861: \\
3862: &&
3863: \nonumber\\
3864: &&\quad\quad\quad\quad
3865: \hat{I}\delta(x_1-x_2) +
3866: \int d x_3 \widehat\Sigma (x_1, x_3 | \phi) \widehat G(x_3, x_2 | \phi);
3867: \nonumber
3868: \eeqa
3869: \beqa
3870: \widehat\Sigma (x_1, x_2 | \phi) = \frac{\delta(r_1-r_2)}{2\pi\nu \tau}
3871: \widehat G(x_1, x_2 | \phi).
3872: \nonumber
3873: \eeqa
3874:
3875: \noindent
3876: The equation (\ref{eq2}) allows for semi-classical treatment
3877: introduced in Refs.~\onlinecite{Eilenberger,68}, and described in
3878: great details in Ref.~\onlinecite{LO77}.
3879: Since we have already averaged
3880: the equation of motion over disorder, the semi-classical approximation
3881: now amounts to averaging the Green's function $\widehat G(x_1, x_2 | \phi)$
3882: over the distance from the
3883: Fermi surface. This is done in two steps:
3884:
3885: \beqa
3886: \widehat G(t_1, t_2;\vec p; \vec R)=
3887: \int d^2r e^{i\vec P \cdot \vec r} \widehat G(x_1, x_2 | \phi),
3888: \eeqa
3889: \[
3890: \vec r = \vec r_1 - \vec r_2 ; \quad
3891: \vec R = \frac{1}{2}(\vec r_1 + \vec r_2);
3892: \]
3893: \[
3894: \vec P = \vec p - \frac{1}{2}
3895: \left[\vec A_{ext}\left(t_1, \vec R\right)+
3896: \vec A_{ext}\left(t_2, \vec R\right)\right];
3897: \]
3898: \beqa
3899: \label{intg}
3900: \hat g(t_1, t_2; \vec n, &&\vec r) =
3901: \\
3902: &&
3903: \nonumber\\
3904: &&
3905: \frac{i}{\pi}
3906: \int\limits_{-\infty}^{\infty} d\xi \widehat G\left(t_1,
3907: t_2; \vec n \left[p_F + \frac{\xi}{v_F}\right]; \vec r\right),
3908: \nonumber
3909: \eeqa
3910:
3911: \noindent
3912: Since we follow the avenue of Ref.~\onlinecite{LO77}, we will skip
3913: further intermediate steps, and use the semiclassical equation written
3914: in the next subsection.
3915:
3916:
3917:
3918: \subsubsection{Eilenberger equation}
3919:
3920:
3921:
3922: The dynamics of the electron matrix Green's function is then described
3923: by the Eilenberger equation\cite{Eilenberger}:
3924:
3925: \beq
3926: \left[
3927: {\tilde\partial_{t}} + v_F \vec n \tilde{\vec\nabla}
3928: + \vec \omega_c
3929: \left(\vec n \times
3930: \frac{\partial}{\partial \vec n} \right)
3931: \right] \hat{g} =
3932: \frac{\hat{g}\< \hat{g} \>_n - \< \hat{g} \>_n\hat{g}}{2\tau},
3933: \label{eil1}
3934: \eeq
3935:
3936: \noindent
3937: where angular averaging is defined as before
3938:
3939: \[
3940: \langle \dots \rangle_n =
3941: \int
3942: \frac{d\theta}{2\pi} \dots,
3943: \quad \vec n = (\cos\theta, \sin\theta) ,
3944: \]
3945:
3946: \noindent
3947: and the covariant derivatives in Eq.~(\ref{eil1}) are defined as
3948:
3949: \begin{mathletters}
3950: \beqa
3951: \tilde{\partial_t} \hat{g} =
3952: \partial_{t_1} \hat{g} + \partial_{t_2} \hat{g} +
3953: i\hat{\varphi}(t_1,r)\hat{g} -
3954: i\hat{g}\hat{\varphi}(t_2,r),
3955: \eeqa
3956: \beq
3957: \tilde{\vec\nabla}
3958: \hat{g} =
3959: \vec\nabla
3960: \hat{g} +
3961: i\hat{\vec A}(t_1, \vec r)\hat{g} -
3962: i\hat{g} \hat{\vec A}(t_2,\vec r).
3963: \eeq
3964: \label{covariant}
3965: \end{mathletters}
3966:
3967: \noindent
3968: Here $\hat{g}$ is a matrix in Keldysh space,
3969:
3970: \beq \hat{g}(t_1, t_2; \vec n, \vec r) =
3971: \left(
3972: \matrix{
3973: g^R & g^K \cr
3974: g^Z & g^A
3975: }
3976: \right),
3977: \eeq
3978:
3979: \noindent
3980: and we will suppress the coordinate and the time arguments unless
3981: otherwise is stated. A product of such matrices should be understood
3982: as a matrix product in Keldysh space and a convolution in time:
3983:
3984: \beqa
3985: \label{convolution}
3986: &&\big[\hat{g}(\vec n, \vec r)\hat{g}(\vec n_1, \vec r)\big]_{ij}
3987: \equiv
3988: \\
3989: &&
3990: \nonumber\\
3991: &&
3992: \quad
3993: \int\limits_{-\infty}^\infty
3994: d t_3 \sum_k\big[\hat{g}(t_1, t_3;\vec n, \vec r)\big]_{ik}
3995: \big[\hat{g}(t_3, t_2; \vec n_1, \vec r)\big]_{kj},
3996: \nonumber
3997: \eeqa
3998:
3999: \noindent
4000: and solutions of the homogeneous equation ~(\ref{eil1}) are subject
4001: to the constraints
4002:
4003: \beq
4004: \label{constraint}
4005: \hat{g}(\vec n, \vec r)\hat{g}(\vec n, \vec r)=
4006: {\hat I}^K, \quad
4007: \int\limits_{-\infty}^\infty dt \;
4008: {\bf Tr} \; \hat{g}(t,t;\vec n, \vec r)=0,
4009: \eeq
4010:
4011: \noindent
4012: where
4013:
4014: \[
4015: [ I^K ]_{ij} = \delta_{ij} \delta(t_1-t_2).
4016: \]
4017:
4018: \noindent
4019: The scalar and vector potentials in Eq.~(\ref{covariant}) have the
4020: following structure in the Keldysh space
4021:
4022: \beqa
4023: &&\hat{\vec A}(t,\vec r) =
4024: \left(
4025: \matrix{
4026: {\vec A}_{ext}(t,\vec r) & 0 \cr
4027: 0 & {\vec A}_{ext}(t,\vec r)
4028: }
4029: \right),
4030: \label{fields}
4031: \\
4032: &&
4033: \nonumber\\
4034: &&
4035: \hat{\varphi}(t,\vec r) =
4036: \left(
4037: \matrix{
4038: {\varphi}_{ext}(t,\vec r) +\phi_{1}(t,\vec r) & \phi_{2}(t,\vec r)\cr
4039: \phi_{2}(t,\vec r) & {\varphi}_{ext}(t,\vec r) + \phi_{1}(t,\vec r)
4040: }
4041: \right),
4042: \nonumber
4043: \eeqa
4044:
4045: \noindent
4046: where $\varphi_{ext}$ and ${\vec A}_{ext}$ are the external (classical)
4047: potentials due to the electric field ${\vec E}$,
4048:
4049: \beq
4050: e{\vec E}= \partial_t {\vec A}_{ext}
4051: -\vec\nabla
4052: \varphi_{ext}
4053: \label{E}
4054: \eeq
4055:
4056: \noindent
4057: acting on the electron system, and $\phi_{1,2}(t,\vec r)$ are the
4058: auxiliary fluctuating fields decoupling the interaction in the singlet
4059: channel. Because the singlet channel describes processes with small
4060: momentum transfers (smaller than $q*$, see Section~\ref{soft}), the
4061: fields $\phi_{1,2}(t,\vec r)$ vary slowly on the scale of the $1/q*$.
4062:
4063: The condition (\ref{do}) enforces causality of the physical response
4064: functions. It is worth noticing that the decoupling of interaction can
4065: be performed also using a fluctuating vector potential; our choice is
4066: strictly a matter of taste.
4067:
4068: In this formalism any observable quantity described by one electron
4069: operator ${\cal O}(\hat{p},\hat{r})$ is given by [see Eqs.~(\ref{GK})
4070: and (\ref{intg})]
4071:
4072: \beqa
4073: {\cal O}(t,{\vec r}) = -
4074: &&
4075: \nu \int \frac{d\theta}{2\pi}{\cal O}(p_F{\vec n},{\vec r})
4076: \label{O}
4077: \\
4078: &&
4079: \nonumber\\
4080: &&
4081: \times\lim_{t_1\to t} \left[
4082: \frac{\pi }{2}
4083: \langle g^K(t_1, t; \vec n, \vec r)\rangle_{\phi} +
4084: {\varphi}_{ext}(t,{\vec r})
4085: \right],
4086: \nonumber
4087: \eeqa
4088:
4089: \noindent
4090: where $\langle\dots \rangle_{\phi}$ stands for averaging over both
4091: auxiliary fields $\phi_{1,2}$ fluctuating according to
4092: Eqs.~(\ref{Ds}). The last term in brackets is a consequence of the
4093: ultraviolet anomaly, and its form is enforced by the requirement of
4094: the gauge invariance.
4095:
4096: Finally, the electronic polarization operators are determined [see
4097: Eqs.~(\ref{pder}) and (\ref{intg})] as variational derivatives of the
4098: solutions to the Eilenberger equation (\ref{eil1}):
4099:
4100: \beqa
4101: &&{\Pi}^R(1,2)= {\Pi}^A(2,1)
4102: \label{Pi}
4103: \\
4104: &&
4105: \nonumber\\
4106: &&
4107: \quad\quad\quad\quad
4108: =\nu \int \frac{d\theta}{2\pi}
4109: \left(\delta_{12} +
4110: \frac{\pi \langle \delta g^K(t_1, t_1; \vec n, \vec r_1)\rangle_{\phi}}
4111: {2\delta \phi_1(t_2,\r_2)}\right);
4112: \nonumber\\
4113: &&
4114: \nonumber\\
4115: &&
4116: {\Pi}^K(1,2)=\nu\pi
4117: \nonumber\\
4118: &&
4119: \nonumber\\
4120: &&\quad\times
4121: \int \frac{d\theta}{2\pi}
4122: \frac{ \langle \delta g^K(t_1, t_1; \vec n, \vec r_1)
4123: +\delta g^Z(t_1, t_1; \vec n, \vec r_1)
4124: \rangle_{\phi}}
4125: {2\delta \phi_2(t_2,\vec r_2)}.
4126: \nonumber
4127: \eeqa
4128:
4129:
4130:
4131:
4132: \subsubsection{Derivation of the kinetic equation}
4133:
4134:
4135:
4136:
4137: Our goal now is to obtain an equation for the Keldysh function
4138: averaged over the fluctuating fields, $\langle g^K(t_1, t_1; \vec n,
4139: \vec r_1)\rangle_{\phi}$. It is this quantity that determines physical
4140: observables, see Eq.~(\ref{O}). We will do this using the
4141: non-crossing approximation for bosonic propagators (i.e. the first
4142: loop approximation for the collision integral), see
4143: Fig.~\ref{15}. This approximation is justified provided that the
4144: resulting dynamics for the electrons (characterized by time
4145: $\tau_\epsilon$) is slow in comparison with motion of relevant bosonic
4146: mode, $T\tau_\epsilon \gg 1$.
4147:
4148: To do so, we notice that only two components of the matrix $\hat{g}$
4149: are independent, and the other two are fixed by the constraint
4150: (\ref{constraint}). For our purposes, we choose to fix the diagonal
4151: components
4152:
4153: \beq
4154: \label{normsolution}
4155: g^R=\sqrt{1-g^K g^Z};
4156: \quad g^A= -\sqrt{1-g^Z g^K},
4157: \eeq
4158:
4159: \noindent
4160: where the square root should be understood in operator sense: as a sum
4161: of its Taylor series, with all arising products hereafter being time
4162: convolutions, similar to Eq.~(\ref{convolution}). The two remaining
4163: independent components of the Eilenberger equation have the explicit
4164: form
4165:
4166: \begin{mathletters}
4167: \label{exact}
4168: \beqa
4169: &&\left[
4170: \tilde\partial_t + v_F \vec n
4171: \tilde{\vec\nabla}
4172: %\tilde{\mbox{\boldmath $\nabla$}}
4173: +\vec\omega_c
4174: %\mbox{ \boldmath $\w$}_c
4175: \left(\vec n \times
4176: \frac{\partial}{\partial \vec n} \right)
4177: \right] g^Z =
4178: \label{exactZ}
4179: \\
4180: &&
4181: \nonumber\\
4182: &&\;
4183: - i\big[\phi_1(t_1,\vec r)-\phi_1(t_2,\vec r) \big] g^Z
4184: - i\phi_2(t_1,\vec r) g^R + i g^A \phi_2(t_2,\vec r)
4185: \nonumber \\
4186: &&
4187: \nonumber \\
4188: &&\quad\quad
4189: +
4190: \frac{1}{2 \tau}
4191: \big[
4192: g^Z \< g^R \>_n - \<g^Z\>_n g^R
4193: + g^A \< g^Z \>_n - \< g^A \>_n g^Z
4194: \big],
4195: \nonumber
4196: \eeqa
4197: \beqa
4198: &&\left[
4199: {\tilde\partial_{t}} + v_F \vec n
4200: \tilde{\vec\nabla}
4201: %\tilde{\mbox{\boldmath $\nabla$}}
4202: +\vec\omega_c
4203: %\mbox{ \boldmath $\w$}_c
4204: \left(\vec n \times
4205: \frac{\partial}{\partial \vec n} \right)
4206: \right]g^K
4207: =
4208: \label{exactK}
4209: \\
4210: &&
4211: \nonumber \\
4212: &&\;
4213: - i\big[\phi_1(t_1,\vec r)-\phi_1(t_2,\vec r) \big] g^K
4214: - i\phi_2(t_1,\vec r) g^A + i g^R \phi_2(t_2,\vec r)
4215: \nonumber\\
4216: &&
4217: \nonumber \\
4218: &&\quad\quad
4219: +
4220: \frac{1}{2 \tau}
4221: \big[
4222: g^K \< g^A \>_n - \<g^K\>_n g^A
4223: + g^R \< g^K \>_n - \< g^R \>_n g^K
4224: \big],
4225: \nonumber
4226: \eeqa
4227: \end{mathletters}
4228:
4229: \noindent
4230: and we redefine the covariant derivatives
4231: Eq.~(\ref{covariant}) to include only the external scalar and vector
4232: potentials:
4233:
4234: \beqa
4235: &&\tilde{\partial_t} {g} = \partial_{t_1} {g} +
4236: \partial_{t_2} {g} + i\left[{\varphi}_{ext}(t_1,r)
4237: -{\varphi}_{ext}(t_2,r) \right]{g},
4238: \label{covariant2}
4239: \\
4240: &&
4241: \nonumber\\
4242: &&
4243: \tilde{\vec\nabla}
4244: %\tilde{\mbox{\boldmath $\nabla$}}
4245: {g} =
4246: \vec\nabla
4247: %\mbox{\boldmath $\nabla$}
4248: {g}
4249: + i\left[{\vec A}_{ext}(t_1,r) -{\vec A}_{ext}(t_2,r) \right] {g}.
4250: \nonumber
4251: \eeqa
4252:
4253: Now we are prepared to derive the collision integral. We notice that
4254: due to causality $\langle g_Z \rangle_{\phi} =0$ in all orders of the
4255: perturbation theory. We separate slow and fast degrees of freedom as
4256: follows
4257:
4258: \beqa
4259: \label{approx}
4260: g_K= \langle g_K
4261: \rangle_{\phi} +\delta g_K; \quad g_Z = \delta g_Z,
4262: \eeqa
4263:
4264: \noindent
4265: where $\delta g$ is the contribution fluctuating with the auxiliary
4266: fields and we calculate it to first order in $\phi$. In the same
4267: approximation Eq.~(\ref{normsolution}) becomes
4268:
4269: \beqa
4270: \label{anormsolution}
4271: &&g^R= \delta(t_1-t_2)- \frac{1}{2} g^K \delta g^Z;
4272: \nonumber\\
4273: &&
4274: \nonumber\\
4275: &&
4276: g^A= -\delta(t_1-t_2)+ \frac{1}{2} \delta g^Z g^K,
4277: \eeqa
4278:
4279: \noindent
4280: [expansion up to the second order in $\delta g^Z$ is unnecessary
4281: because terms of such kind vanish due to Eq.~(\ref{do})].
4282:
4283: We now substitute Eqs.~(\ref{approx}) and (\ref{anormsolution}) into
4284: Eqs.~(\ref{exact}) and obtain equations governing the behavior of the
4285: fluctuating parts
4286:
4287: \begin{mathletters}
4288: \label{first}
4289: \beqa
4290: \Bigg[
4291: {\tilde\partial_{t}} + v_F \vec n
4292: \tilde{\vec\nabla}
4293: %\tilde{\mbox{\boldmath $\nabla$}}
4294: + &&
4295: \vec\omega_c
4296: %\mbox{ \boldmath $\w$}_c
4297: \left(\vec n \times
4298: \frac{\partial}{\partial \vec n} \right)\Bigg]
4299: \delta g^Z - \frac{1}{ \tau}
4300: \left[ \delta g^Z - \<\delta g^Z\>_n \right]
4301: \nonumber\\
4302: &&
4303: \nonumber\\
4304: &&
4305: =- 2 i\phi_2(t_1,\r)\delta(t_1-t_2),
4306: \label{firstZ}
4307: \eeqa
4308: \beqa
4309: &&\left[
4310: {\tilde\partial_{t}} + v_F \vec n
4311: \tilde{\vec\nabla}
4312: %\tilde{\mbox{\boldmath $\nabla$}}
4313: +
4314: \vec\omega_c
4315: %\mbox{ \boldmath $\w$}_c
4316: \left(\vec n \times
4317: \frac{\partial}{\partial \vec n} \right) \right]\delta g^K +
4318: \frac{1}{\tau} \left[\delta g^K - \<\delta g^K\>_n \right]
4319: \nonumber\\
4320: &&
4321: \nonumber\\
4322: &&
4323: =
4324: 2 i\phi_2(t_1,\r)\delta(t_1-t_2)
4325: - i\left[\phi_1(t_1,\r)-\phi_1(t_2,\r) \right]
4326: \langle g^K\rangle_{\phi}
4327: \nonumber\\
4328: &&
4329: \nonumber\\
4330: &&
4331: + \frac{1}{4 \tau}
4332: \left[
4333: \langle g^K \rangle_{\phi}
4334: \< \delta g^Z \langle g^K \rangle_{\phi} \>_n
4335: - \langle \<g^K\>_n \rangle_{\phi}
4336: \delta g^Z \langle g^K \rangle_{\phi}
4337: \right]
4338: \label{firstK}
4339: \\
4340: &&
4341: \nonumber\\
4342: &&
4343: - \frac{1}{4 \tau}
4344: \left[
4345: \langle g^K \rangle_{\phi} \delta g^Z
4346: \langle \< g^K \>_n \rangle_{\phi}
4347: - \< \langle g^K \rangle_{\phi}
4348: \delta g^Z\>_n\langle g^K \rangle_{\phi}
4349: \right],
4350: \nonumber
4351: \eeqa
4352: \end{mathletters}
4353:
4354: Solutions to the equations (\ref{first}) should be substituted into
4355: the equation (\ref{exactK}) for the smooth part. Than the equation for
4356: the smooth part should be averaged over the fluctuating fields
4357: $\phi_{1,2}$ with the help of Eq.~(\ref{Ds}). As a result
4358:
4359: \beqa
4360: &&\left[
4361: {\tilde\partial_{t}} + v_F \vec n
4362: \tilde{\vec\nabla}
4363: %\tilde{\mbox{\boldmath $\nabla$}}
4364: +
4365: \vec\omega_c
4366: %\mbox{ \boldmath $\w$}_c
4367: \left(\vec n \times
4368: \frac{\partial}{\partial \vec n} \right)\right]
4369: \langle g^K \rangle_{\phi}
4370: \label{almostthere}
4371: \\
4372: &&
4373: \nonumber\\
4374: &&
4375: \quad
4376: = \St_{in}\left\{\langle g^K \rangle_{\phi} \right\}
4377: + \St_{el}\left\{\langle g^K \rangle_{\phi} \right\}.
4378: \nonumber
4379: \eeqa
4380:
4381: \noindent
4382: Here we separate the collision integrals into two contributions. The
4383: physical meaning of such separation will be discussed shortly. The
4384: first, inelastic part has the structure
4385:
4386: \beqa
4387: &&\St_{in}\left\{\langle g^K \rangle_{\phi} \right\}
4388: (t_1,t_2; {\vec n}, \vec r)
4389: \label{inelastic1}
4390: \\
4391: &&
4392: \nonumber\\
4393: &&
4394: \quad=- i\langle
4395: \left[\phi_1(t_1,\vec r)-\phi_1(t_2,\vec r) \right]
4396: \delta g^K(t_1,t_2; {\vec n}, \vec r)
4397: \rangle_{\phi}.
4398: \nonumber
4399: \eeqa
4400:
4401: \noindent
4402: The second, elastic contribution has the form
4403:
4404:
4405: \end{multicols}
4406: \widetext
4407:
4408:
4409: \beqa
4410: &&\St_{el}\left\{\langle g^K \rangle_{\phi}\right\}
4411: (t_1,t_2; \vec n; \vec r)
4412: = \frac{1}{\tau}\Big[
4413: \<\langle g^K (t_1,t_2; \vec n; \vec r)\rangle_{\phi} \>_n
4414: -\langle g^K(t_1,t_2; \vec n; \vec r) \rangle_{\phi} \Big]
4415: \label{elastic1}
4416: \\
4417: &&
4418: \nonumber\\
4419: &&
4420: \quad
4421: +
4422: \int d t_3 \int\frac{d\theta_1}{2\pi}
4423: \left[
4424: \<\langle g^K (t_1,t_3; \vec n_1, \vec r)\rangle_{\phi}
4425: F^A(t_3,t_2; \vec n_1, \vec n; \vec r)
4426: - \<\langle g^K (t_1,t_3; \vec n, \vec r)\rangle_{\phi}
4427: F^A(t_3,t_2; \vec n, \vec n_1; \vec r)
4428: \right]
4429: \nonumber\\
4430: &&
4431: \nonumber\\
4432: &&\quad+
4433: \int d t_3 \int\frac{d\theta_1}{2\pi}
4434: \left[
4435: F^R(t_1,t_3; \vec n, \vec n_1; \vec r)
4436: \<\langle g^K (t_3,t_2; \vec n_1, \vec r)\rangle_{\phi}
4437: - F^R(t_1,t_3; \vec n_1, \vec n; \vec r)
4438: \<\langle g^K (t_3,t_2; \vec n, \vec r)\rangle_{\phi}
4439: \right],
4440: \nonumber
4441: \eeqa
4442:
4443: \noindent
4444: where the first term is just the ordinary impurity scattering and the
4445: remaining terms characterize interaction effects. The kernels
4446: in Eq.~(\ref{elastic1}) are defined as
4447:
4448: \beqa
4449: F^R(t_1,t_2; \vec n, \vec n_1; \vec r)
4450: = \frac{1}{4\tau}\int d t_3
4451: \langle \delta g^K (t_1,t_3; \vec n, \vec r)
4452: \left[\delta g^Z (t_3,t_2; \vec n_1, \vec r)
4453: -\delta g^Z (t_3,t_2; \vec n, \vec r) \right]
4454: \rangle_{\phi}
4455: \nonumber\\
4456: \nonumber\\
4457: F^A(t_1,t_2; \vec n, \vec n_1; \vec r)
4458: =\frac{1}{4\tau} \int d t_3
4459: \langle\left[\delta g^Z (t_1,t_3; \vec n_1, \vec r)
4460: -\delta g^Z (t_1,t_3; \vec n, \vec r) \right]
4461: \delta g^K (t_3,t_2; \vec n, \vec r)
4462: \rangle_{\phi}
4463: \label{Fs}
4464: \eeqa
4465:
4466:
4467: \begin{multicols}{2}
4468:
4469:
4470: The equations (\ref{almostthere}), (\ref{inelastic1}), (\ref{elastic1}),
4471: (\ref{first}), (\ref{Ds}), and (\ref{Dysonb}) constitute a closed
4472: system of kinetic equations.
4473: Although sufficient for description of interaction effects in
4474: disorder systems, these equations are inconvenient for analytical
4475: calculations because the
4476: expressions for the collision integral are nonlocal in space and time.
4477: To simplify further calculations we will use the assumption that
4478: $\langle g^K (t_1,t_2;
4479: \nn, \r)\rangle_{\phi}$ is a smooth function so that a gradient
4480: expansion will be possible.
4481:
4482: Before embarking on such calculation we pause to discuss the
4483: physical distinction between the elastic (\ref{elastic1}) and
4484: inelastic (\ref{inelastic1}) collision terms. One immediately notices
4485: from Eq.~(\ref{elastic1}) that
4486:
4487: \beq
4488: \int d\theta\ \St_{el}(t_1,t_2; \vec n; \vec r)=0,
4489: \label{elastic2}
4490: \eeq
4491:
4492: \noindent
4493: for any $t_1$ and $t_2$. This indicates that this part of the
4494: collision integral preserves the number of particles on a given energy
4495: shell [see below for explicit connections between time representation
4496: and energy representation Eq.~(\ref{energy})].
4497:
4498: The inelastic term (\ref{inelastic1}) does not vanish after angular
4499: averaging. Therefore this part does promote electrons
4500: between energy shells. However, we notice that
4501:
4502: \beq
4503: \St_{in}\left\{\langle g^K \rangle_{\phi} \right\}
4504: (t_1,t_1; {\vec n}, \r)=0,
4505: \label{inelastic2}
4506: \eeq
4507:
4508: \noindent
4509: for any direction ${\vec n}$. Taking coinciding time arguments is
4510: equivalent to integrating over the whole energy spectrum [see
4511: Eq.~(\ref{energy})], so that not only the total number of particles is
4512: conserved, but the total number of particles moving along a given
4513: direction ${\vec n}$ is conserved (i.e. inelastic forward scattering).
4514:
4515: Let us now perform the actual calculation of the collision integrals. We
4516: solve Eq.~(\ref{firstZ}) and obtain
4517:
4518: \beqa
4519: &&\delta g^Z(t_1,t_2; \vec n,\vec r)=
4520: 2i\delta(t_1-t_2)
4521: \label{gzsol}
4522: \\
4523: &&
4524: \nonumber\\
4525: &&
4526: \quad\times
4527: \int d\r_1dt_3
4528: \phi_2(\r_1,t_3)\int \frac{d\theta'}{2\pi}
4529: D(t_3-t_1, \vec n', \vec n; \vec r_1, \vec r)
4530: \nonumber\\
4531: &&
4532: \nonumber\\
4533: &&
4534: D(t; \vec n, \vec n'; \vec r_1, \vec r_2)=
4535: \int \frac{d\omega d^2\q}{(2\pi)^3}
4536: e^{i{\vec q}(\vec r_1-\vec r_2)-i\omega t}
4537: D(\vec n, \vec n'; \omega, \vec q),
4538: \nonumber
4539: \eeqa
4540:
4541: \noindent
4542: where the diffuson propagator $D$ is defined in Eq.~(\ref{Diffuson}).
4543:
4544: To simplify the analytic solution of Eq.~(\ref{firstK}), we assume
4545: without loss of generality that $\langle g^K\rangle_{\phi}$
4546: varies slowly on the spatial scale $L_T = v_F {
4547: min}(1/T,\sqrt{\tau /T})$, and also a slow function of $t_1+t_2$ on the time
4548: scale $\simeq 1/T$. These assumptions are consistent with the first
4549: loop approximation we already invoked.
4550:
4551: In what follows we will keep only the zeroth and
4552: first angular harmonics (which is consistent with assumption about
4553: the spatial smoothness) in the direction dependence of the Keldysh
4554: function:
4555:
4556: \beqa
4557: \langle g\left(t_1, t_2; \vec n, \vec r\right) \rangle_\phi \approx &&
4558: \langle g\left(t_1, t_2; \vec n, \vec r\right) \rangle_n
4559: \nonumber\\
4560: &&
4561: \nonumber\\
4562: &&
4563: +2 \vec n
4564: \langle \vec n' g\left(t_1, t_2; \vec n', \vec r\right) \rangle_{n'}
4565: \label{angular}
4566: \eeqa
4567:
4568: \noindent
4569: This approximation does not affect results for any relevant
4570: quantities. From now on we will suppress the explicit sign of averaging
4571: over the fluctuating fields because we will not be dealing with
4572: non-averaged quantities anymore.
4573:
4574: We now substitute Eq.~(\ref{angular}) into the right-hand side of
4575: Eq.~(\ref{firstK}) and obtain
4576:
4577: \beqa
4578: \delta g^K (t_1, t_2; \vec n, \vec r) = &&
4579: \delta g^K_1 (t_1, t_2; \vec n, \vec r)
4580: \nonumber\\
4581: &&
4582: \nonumber\\
4583: &&
4584: +
4585: \delta g^K_2 (t_1, t_2; \vec n, \vec r)
4586: \label{gk12}
4587: \eeqa
4588:
4589: \noindent
4590: The first term in Eq.~(\ref{gk12}) is proportional to the field
4591: $\phi_1$ and gives contributions to both the elastic and the
4592: inelastic parts of the collision integral. To obtain non-vanishing
4593: contribution to the latter we have to do each one of the following:
4594: (i) take into account the first angular harmonic; (ii) perform the
4595: first order gradient expansion; (iii) expand up to the first order
4596: in external fields, ${\vec A}_{ext}$. The result is
4597:
4598:
4599: \end{multicols}
4600: \widetext
4601:
4602:
4603: \begin{mathletters}
4604: \label{gk1}
4605: \beqa
4606: \delta g^K_1 (t_1, t_2; \vec n, \vec r) &=&
4607: -i\int dt
4608: \left[\phi_1(\vec r_1,t_1-t) -\phi_1(\vec r_1,t_2-t)\right]
4609: \int \frac{d\vec n'}{2\pi}
4610: D(t, \vec n, \vec n'; \vec r, \vec r_1)
4611: \nonumber\\
4612: &&
4613: \nonumber\\
4614: &&
4615: \times \left.\Big\{
4616: \<g^K (t_1-t, t_2-t; \vec n_1, \vec r)\>_{\vec n_1}
4617: \right.
4618: \label{gk1a}
4619: \\
4620: &&
4621: \nonumber\\
4622: &&\quad\quad+
4623: 2 \vec n' \<\vec n_1g^K (t_1-t, t_2-t; \vec n_1, \vec r)\>_{\vec n_1}
4624: \label{gk1b}
4625: \\
4626: &&
4627: \nonumber\\
4628: &&\quad\quad+ \left(\vec r_1-\vec r\right)
4629: \tilde{\vec\nabla}
4630: %\tilde{\mbox{\boldmath $\nabla$}}
4631: \<g^K (t_1-t, t_2-t; \vec n_1, \vec r)\>_{\vec n_1}
4632: \Big\}
4633: \label{gk1c}
4634: \eeqa
4635: \end{mathletters}
4636:
4637: \noindent
4638: where the covariant derivative is defined in Eq.~(\ref{covariant2})
4639: and we neglected higher order derivatives of the external fields.
4640: Expansion in the time coordinate $t_1+t_2$ (using the covariant
4641: derivative $\tilde{\partial}_t$) is not necessary because it produces a
4642: negligible correction to the inelastic collision integral and does not
4643: affect the elastic one.
4644:
4645: The second term in the RHS of Eq.~(\ref{gk12}) is proportional to the
4646: field $\phi_2$, and according to Eqs.~(\ref{Fs}) and (\ref{do}) it
4647: does not contribute to the elastic collision integral. Therefore, it
4648: is sufficient to keep only the zeroth angular component and neglect
4649: gradient terms at all. This yields
4650:
4651: \beqa
4652: &&\delta g^K_2(t_1, t_2; \vec n, \vec r)
4653: =\int \frac{d\theta'}{2\pi}\frac{d\theta''}{2\pi}
4654: \int d\vec r_1 dt
4655: D(t, \vec n, \vec n'; \vec r, \vec r_1)
4656: \left. \Big\{
4657: 2 i\phi_2(\vec r_1,t_1-t) \delta(t_1-t_2)
4658: \right.
4659: \label{gk2}
4660: \\
4661: &&
4662: \nonumber\\
4663: &&
4664: \left.
4665: + \frac{i}{\tau}
4666: \langle g\left(t_1 - t, t_3; \vec n_1, \vec r\right) \rangle_{\vec n_1}
4667: \langle g\left(t_3, t_2 - t; \vec n_1, \vec r\right) \rangle_{\vec n_1}
4668: \left[
4669: \<D(t_4-t_3, \vec n'', \vec n_1; \vec r_2, \vec r_1)\>_{\vec n_1}-
4670: D(t_4-t_3, \vec n'', \vec n'; \vec r_2, \vec r_1)
4671: \right].
4672: \phi_2(\vec r_2,t_4)
4673: \right\}
4674: \nonumber
4675: \eeqa
4676:
4677:
4678: \begin{multicols}{2}
4679:
4680:
4681: As we already mentioned, $g(t_1,t_2)$ has a much faster dependence on
4682: the difference $t_1-t_2$ then on the sum $t_1+t_2$. Therefore it is more
4683: convenient to
4684: use a temporal transformation of the Green's function
4685:
4686: \beq
4687: g^K(t_1, t_2; \vec n, \vec r) = \int \frac{d\epsilon}{2\pi}
4688: g\left(\frac{t_1+t_2}{2}, \epsilon; \vec n, \vec r\right)
4689: e^{i\epsilon(t_2-t_1)},
4690: \label{energy}
4691: \eeq
4692:
4693: \noindent
4694: which defines the precise notion of energy $\epsilon$ in this
4695: context. We introduce the same transformation for the propagators of
4696: auxiliary fields (\ref{Ds})
4697:
4698: \begin{equation}
4699: {\cal D}(t_1,t_2) = \int \frac{d\w}{2\pi}
4700: {\cal D}\left( \frac{t_1+t_2}{2},\omega\right)e^{i\omega (t_2-t_1)}
4701: \label{dw}
4702: \end{equation}
4703:
4704: \noindent
4705: The transformed functions have the symmetry property (hereafter we
4706: omit the $K$ superscript for brevity since we are only dealing with
4707: the Keldysh function)
4708:
4709: \beqa
4710: g(t,\epsilon) &=& - g(t,-\epsilon);
4711: \label{symmetry}
4712: \\
4713: &&
4714: \nonumber\\
4715: {\cal D}^K\left(t,\omega;\vec r_1,\vec r_2\right)&=&
4716: {\cal D}^K\left(t,-\omega;\vec r_2,\vec r_1\right);
4717: \nonumber\\
4718: &&
4719: \nonumber\\
4720: {\cal D}^R\left(t,\omega;\vec r_1,\vec r_2\right)&=&
4721: {\cal D}^A\left(t,-\omega;\vec r_2,\vec r_1\right).
4722: \nonumber
4723: \eeqa
4724:
4725: Now, we are ready to obtain the explicit form of the collision
4726: integral. We start with the inelastic contribution and perform the
4727: following three steps: (1) substitute Eq.~(\ref{gk1a}) and (\ref{gk2})
4728: into Eq.~(\ref{inelastic}); (2) average over the fields $\phi_{1,2}$
4729: with the help of Eq.~(\ref{Ds}); (3) perform the temporal
4730: transformation (\ref{energy}) of the result. As a result we obtain
4731: with the help of Eqs.~(\ref{dw}) and (\ref{symmetry}) the following
4732: form of the collision integral:
4733:
4734:
4735: \end{multicols}
4736: \widetext
4737:
4738:
4739: \beqa
4740: &&\St_{in}\left\{g^K \right\}
4741: (t,\epsilon; \vec r)=
4742: - \frac{i}{2}
4743: \int d^2 r_1 \int \frac{d\omega}{2\pi}
4744: {\cal D}^K\left(t,\omega; \vec r, \vec r_1\right)
4745: \left[\langle D(\omega;\vec r,\vec r_1)\rangle +
4746: \langle D(-\omega;\vec r_1,\vec r)\rangle\right]
4747: \nonumber\\
4748: &&
4749: \nonumber\\
4750: &&
4751: \hspace*{5cm}
4752: \times
4753: \left[\langle g\left(t, \epsilon; \vec n, \vec r\right) \rangle_{\vec n}
4754: -
4755: \langle g\left(t, \epsilon -\omega ; \vec n, \vec r\right) \rangle_{\vec n}
4756: \right]
4757: \nonumber\\
4758: &&
4759: \nonumber\\
4760: &&
4761: \quad
4762: +\frac{i}{2\tau}
4763: \int d\vec r_1d\vec r_2 \int \frac{d\omega}{2\pi}
4764: \left[{\cal D}^R\left(t,\omega; \vec r_1, \vec r_2\right)-
4765: {\cal D}^A\left(t,\omega; \vec r_2, \vec r_1\right)\right]
4766: \left[\langle D(\omega;\vec r,\vec r_1)\rangle
4767: \langle D(-\omega;\vec r,\vec r_2)\rangle
4768: - \langle D(\omega;\vec r,\vec r_1) D(-\omega;\vec r,\vec r_2)\rangle
4769: \right]
4770: \nonumber\\
4771: &&
4772: \nonumber\\
4773: &&
4774: \hspace*{5cm}
4775: \times
4776: \langle g\left(t, \epsilon +\omega ; \vec n, \vec r\right) \rangle_{\vec n}
4777: \langle g\left(t, \epsilon; \vec n, \vec r\right) \rangle_{\vec n}
4778: \label{inelastic3}
4779: \eeqa
4780:
4781: \noindent
4782: where the angular averaging of the diffusons is defined after
4783: Eqs.~(\ref{eqs4.6}).
4784:
4785: Now, we have to express the bosonic propagator in terms of the
4786: fermionic polarization operators. The polarization operators are
4787: given by Eqs.~(\ref{Pi}), where we now substitute Eqs.~(\ref{gzsol}),
4788: (\ref{gk1a}), and (\ref{gk2}). After the temporal
4789: transformation (\ref{energy}) we find
4790:
4791: \begin{mathletters}
4792: \beqa
4793: \Pi^R(\omega;t, \vec r_1,\vec r_2)&=&\Pi^A(-\omega;t, \vec r_2,\vec r_1)
4794: \nonumber\\
4795: &&
4796: \nonumber\\
4797: &=&\nu
4798: \left[\delta(\vec r_1-\vec r_2) + \frac{i}{4}
4799: \langle D(\omega; \vec r_1, \vec r_2 )\rangle
4800: \int d \epsilon
4801: \left[\langle g\left(t, \epsilon; \vec n, \vec r\right) \rangle_{\vec n}
4802: -
4803: \langle g\left(t, \epsilon -\omega ; \vec n, \vec r\right) \rangle_{\vec n}
4804: \right] \right] \label{Pra}
4805: \\
4806: &&
4807: \nonumber\\
4808: &=&
4809: \nu \left[\delta(\vec r_1-\vec r_2)
4810: +i\omega \langle D(\omega; \vec r_1, \vec r_2 )\rangle\right]
4811: \label{Prb}
4812: \eeqa
4813: \end{mathletters}
4814: \beqa
4815: \Pi^K(t, \omega; \vec r_1,\vec r_2)&=&
4816: \frac{i\nu}{4\tau}
4817: \int d \vec r
4818: \Big[\langle D(\omega;\vec r,\vec r_1)\rangle
4819: \langle D(-\omega;\vec r,\vec r_2)\rangle
4820: - \langle D(\omega;\vec r,\vec r_1) D(-\omega;\vec r,\vec r_2)\rangle
4821: \Big]
4822: \label{Pk}
4823: \\
4824: &&
4825: \nonumber\\
4826: &\times&\int {d\epsilon} \left[
4827: \langle g\left(t, \epsilon +\omega ; \vec n, \vec r\right) \rangle_{\vec n}
4828: \langle g\left(t, \epsilon; \vec n, \vec r\right) \rangle_{\vec n}
4829: -4 \right]
4830: \nonumber
4831: \eeqa
4832:
4833: \noindent
4834: The last step in the calculation of the interaction propagators is to
4835: solve Eq.~(\ref{Dysonb}) with the polarization operators
4836: Eq.~(\ref{Pra}). This gives ${\cal D}^{R,A}$ in the form given by
4837: Eq.~(\ref{int-dyson}) and for the Keldysh component we obtain
4838:
4839: \begin{mathletters}
4840: \beq
4841: {\cal D}^K={\cal D}^{R}\Pi^K{\cal D}^{A}.
4842: \eeq
4843:
4844: \noindent
4845: Also we can relate the difference of the retarded and advanced
4846: propagators which enters the collision integral Eq.~(\ref{inelastic3})
4847: to the polarization operators:
4848:
4849: \beq
4850: {\cal D}^{R}-{\cal D}^{A}= {\cal D}^{R}
4851: \left[{\Pi}^{R}-{\Pi}^{A}\right]
4852: {\cal D}^{A}.
4853: \eeq
4854: \label{dsolution}
4855: \end{mathletters}
4856:
4857: To obtain the final form of the inelastic part of the collision
4858: integral Eqs.~(\ref{eq:4.8}) we need to
4859: substitute Eq.~(\ref{dsolution}) into Eq.~(\ref{inelastic3}), while
4860: using Eq.~(\ref{Pra}) for $\Pi^R-\Pi^A$. In addition, we note that
4861:
4862: \[
4863: \frac{2}{\tau}
4864: \Big[
4865: \langle D(\omega; q) D(-\omega;-q)\rangle
4866: - \langle D(\omega; q)\rangle\langle D(-\omega;-q)\rangle
4867: \Big] = \langle D(\omega; q)\rangle + \langle D(-\omega;-q)\rangle.
4868: \]
4869:
4870: \noindent
4871: Finally, we introduce the gauge invariant distribution function $f$ as
4872:
4873: \begin{equation}
4874: f(\epsilon,t; \vec n, \vec r)
4875: = \frac{1}{2} - \frac{1}{4}
4876: g(\epsilon + \varphi_{ext}(\vec r),t; \vec n, \vec r)
4877: \label{fg}
4878: \end{equation}
4879:
4880: \noindent
4881: and obtain Eqs.~(\ref{eq:4.8}).
4882:
4883: The calculation of the elastic part of the collision integral is
4884: completely analogous. We substitute Eqs.~(\ref{gk1a})--(\ref{gk1c})
4885: and Eq.~(\ref{gzsol}) into Eqs.~(\ref{Fs}) and average over
4886: fluctuating fields with the help of Eq.~(\ref{Ds}). After the temporal
4887: transformation (\ref{energy}) we find
4888:
4889: \begin{mathletters}
4890: \label{el4}
4891: \label{elastic4}
4892: \beqa
4893: F^R(\epsilon,t; \vec n_1,\vec n_2,\vec r)&=&
4894: F^A(\epsilon,t; \vec n_1,\vec n_2,\vec r)^*
4895: = \frac{i}{4\tau}
4896: \int \frac{d \omega}{2\pi}
4897: \int d\vec r_1 d\vec r_2
4898: {\cal D}^R(\omega, \vec r_1, \vec r_2)
4899: \nonumber\\
4900: &&
4901: \nonumber\\
4902: &\times&
4903: \int \frac{d\vec n_3}{2\pi}\frac{d\vec n_4}{2\pi}
4904: \left[D(\omega; \vec n_3, \vec n_2, \vec r_2, \vec r) -
4905: D(\omega; \vec n_3, \vec n_1, \vec r_2, \vec r)\right]
4906: D(\omega; \vec n_1, \vec n_4, \vec r, \vec r_1)
4907: \nonumber\\
4908: &&
4909: \nonumber\\
4910: &&
4911: \quad \times \Big\{
4912: \<g (t, \epsilon-\w; \vec n', \vec r)\>_{\vec n'}
4913: \label{el4a}
4914: \\
4915: &&
4916: \nonumber\\
4917: &&
4918: \quad +2 \vec n_4 \<\vec n' g (t, \epsilon-\w ; \vec n_1, \vec r)\>_{\vec n'}
4919: \label{el4b}
4920: \\
4921: &&
4922: \nonumber\\
4923: &&
4924: +\left(\vec r_1-\vec r\right)
4925: \left(
4926: \vec\nabla
4927: %\mbox{\boldmath $\nabla$}
4928: +\frac{\partial {\vec A}_{ext}}{\partial t}
4929: \frac{\partial}{\partial \epsilon}
4930: \right)
4931: \<g (t, \epsilon -\w; \vec n', \vec r)\>_{\vec n'}
4932: \Big\}
4933: \label{el4c}
4934: \eeqa
4935: \end{mathletters}
4936:
4937: \noindent
4938: Deriving Eq.~(\ref{elastic4}) we use the fact that $\int d\omega {\cal
4939: D}^R(\omega)D(\omega) = 0$.
4940:
4941: We substitute Eqs.~(\ref{el4}) and (\ref{angular}) into
4942: Eq.~(\ref{elastic2}). We expand the result into angular harmonics.
4943: The zeroth angular harmonic vanishes because of the conservation law
4944: (\ref{elastic2}), and the first harmonic gives (we write only
4945: interaction correction to the collision integral)
4946:
4947: \begin{mathletters}
4948: \label{el5}
4949: \beqa
4950: \St_{el}(t; \epsilon; \vec r)
4951: =\frac{2}{\tau}\int \frac{d\omega}{2\pi}
4952: &\Big[&
4953: n_\alpha K_1^{\alpha\beta}(\w)
4954: \langle n_\beta g(t; \epsilon, \vec r,\vec n)\rangle_n
4955: \langle g(t; \epsilon-\omega, \vec r,\vec n)\rangle_n
4956: \label{el5a}
4957: \\
4958: &&
4959: \nonumber\\
4960: &+&n_\alpha K_0^{\alpha\beta}(\w)
4961: \langle n_\beta g(t; \epsilon-\omega, \vec r,\vec n)\rangle_n
4962: \langle g(t; \epsilon, \vec r,\vec n)\rangle_n
4963: \label{el5b}
4964: \\
4965: &&
4966: \nonumber\\
4967: &+&
4968: \frac{n_\alpha L_0^{\alpha\beta}(\w)}{2}
4969: \langle g(t; \epsilon, \vec r,\vec n)\rangle_n
4970: \left(\nabla_\beta
4971: + \frac{\partial [A_{ext}]_\beta}{\partial t}
4972: \frac{\partial}{\partial \epsilon}
4973: \right)
4974: \langle g(t; \epsilon-\omega, \vec r,\vec n)\rangle_n
4975: \Big].
4976: \label{el5c}
4977: \eeqa
4978: \end{mathletters}
4979:
4980: \noindent
4981: Here the kernels $K$ and $L$ are given by Eq.~(\ref{eqs4.6}). Each
4982: labeled separately term in Eqs.~(\ref{el5}) corresponds to ones in
4983: Eqs.~(\ref{elastic4}) and in Eqs.~(\ref{gk1}) so the origin of terms
4984: can be easily traced.
4985:
4986:
4987:
4988: \begin{multicols}{2}
4989:
4990:
4991:
4992: Finally, we use the gauge invariant distribution function (\ref{fg})
4993: instead of $g$ and we arrive to Eqs.~(\ref{eq:4.5}).
4994:
4995: Closing this section, we remark that the above treatment can be easily
4996: generalized to include other channels as well as the higher angular
4997: harmonics of the Fermi liquid constant. For the latter task one has to
4998: introduce angle dependent auxiliary fields $\phi_{1,2}(\vec r,\vec n,t)$ and
4999: use $F_0^\rho \to F^\rho(\widehat{\vec n_1\vec n_2})$.
5000:
5001: The triplet channel requires introduction of the coupling of the form
5002: ${\vec h}_{1,2}(\vec r,t)
5003: \hat{\vec\sigma}$,
5004: %\hat{\mbox{\boldmath $\sigma$}}$,
5005: where
5006: $\hat{\sigma}^j,\ j=x,y,z$ are the Pauli matrices in the spin space,
5007: and ${\vec h}_{1,2}(\vec r,t)$ are the auxiliary fields. Accordingly, each
5008: bosonic propagator from Eq.~(\ref{Ds}) becomes a $3 \times 3$ matrix. The
5009: equation~(\ref{Dysonb}) retains the same form with the matrix multiplication
5010: in Keldysh and spin spaces implied. The equation (\ref{D00}) becomes
5011:
5012: \beq
5013: [{\cal D}_0^R]_{ij} = [{\cal D}_0^A]_{ij}
5014: =- F_0^\sigma \delta_{ij}\delta(\vec r_1-\vec r_2)\delta(t_1-t_2),
5015: \eeq
5016:
5017: \noindent
5018: and Eq.~(\ref{Pi}) is modified to
5019:
5020: \beqa
5021: &&{\Pi}_{ij}^R(1,2)= {\Pi}_{ji}^A(2,1)
5022: \label{Pi2}
5023: \\
5024: &&
5025: \nonumber\\
5026: &&
5027: \quad =\nu \int \frac{d\theta}{2\pi}
5028: \left(\delta_{12}\delta_{ji} +
5029: \frac{\pi \langle {\bf Tr}\sigma_i
5030: \delta g^K(t_1, t_1; \vec n, \vec r_1)\rangle_{\phi}}
5031: {4\delta h_1^j(t_2,\vec r_2)}\right);
5032: \nonumber\\
5033: &&
5034: \nonumber\\
5035: &&
5036: {\Pi}^K(1,2)= \pi\nu
5037: \nonumber\\
5038: &&
5039: \nonumber\\
5040: &&
5041: \int \frac{d\theta}{2\pi}
5042: \frac{ \langle {\bf Tr}\sigma_i
5043: \delta g^K(t_1, t_1; \vec n, \vec r_1)
5044: + {\bf Tr}\sigma_i\delta g^Z(t_1, t_1; \vec n, \vec r_1)
5045: \rangle_{\phi}}
5046: {4\delta h^j_2(t_2,\vec r_2)}
5047: \nonumber
5048: \eeqa
5049:
5050: \noindent
5051: where trace is performed in spin space.
5052:
5053: Further derivation consists of a repetition of the steps described
5054: in this section, and in the absence of the spin structure of the
5055: distribution function, $f_{ij}= \delta_{ij}f$, results in
5056: Eqs.~(\ref{r1}) and (\ref{r2}). The spin-orbit interaction or Zeeman
5057: splitting by external magnetic field slightly changes the results,
5058: but we will
5059: postpone the corresponding analysis until the future publication \cite{pre}.
5060:
5061: Finally, the Cooper channel interaction
5062: (\ref{106}) can be treated in the same manner by
5063: introducing auxiliary fields in the Gorkov-Nambu space. We will not
5064: discuss this question further in the present paper.
5065:
5066:
5067:
5068:
5069: \section{Conclusions}
5070: \label{projects}
5071:
5072:
5073: This paper is an attempt to consistently describe the effect of
5074: electron-electron interaction on longitudinal conductivity of disordered
5075: 2D electron gas at $T\ll E_F$. Our results are valid for an arbitrary
5076: relation between $T$ and $\hbar/\tau$ and are summarized in
5077: Section~\ref{results}. At low temperatures $T\tau\ll \hbar$ we reproduce the
5078: logarithmically divergent Altshuler-Aronov correction. At higher
5079: temperatures $T\tau > \hbar$, i.e. in the ballistic region, we found the
5080: linear temperature dependence in accord with Refs.~\onlinecite{gdl,rei}.
5081: However, even the sign of the slope of this dependence depends
5082: on the strength of electron-electron interaction in contradiction to
5083: the results of Refs.~\onlinecite{gdl,rei} (see Sections~\ref{qualitative}
5084: and \ref{single} for discussions of this discrepancy).
5085:
5086: We deliberately did not compare the theory with experimental data,
5087: postponing this comparison until the publication of theoretical
5088: results for Hall conductivity and magneto-resistance in the parallel
5089: field. For comparison with data obtained for Si-MOSFET samples the
5090: valley degeneracy should be taken into account (the degenaracy may
5091: increase the numerical factor in Eq.~(\ref{tc}) by as much as a factor
5092: of $5$ in the case of low intervalley scattering).
5093: We also relegate the
5094: corresponding discussion to a separate publication.
5095:
5096: Finally, we derived a kinetic equation to describe the effect of
5097: electron-electron interaction at arbitrary $T\tau$. The advantage of
5098: this approach is that it turns out to be more convenient for
5099: practical calculations of transport properties in magnetic field
5100: as well as thermal transport properties.
5101:
5102:
5103:
5104: \acknowledgements
5105:
5106:
5107: We are grateful to B.L. Altshuler,
5108: M.E. Gershenson, L.I. Glazman, A.I. Larkin, and M.Yu. Reizer
5109: for numerous discussions. I.A. acknowledges support of
5110: the Packard Foundation.
5111:
5112:
5113:
5114:
5115:
5116: \appendix
5117:
5118: \section*{}
5119:
5120:
5121:
5122: {
5123: \narrowtext
5124: \begin{figure}[ht]
5125: \vspace{0.5 cm}
5126: \epsfxsize=6 cm
5127: \centerline{\epsfbox{f23.eps}}
5128: \vspace{0.5cm}
5129: \caption{Integration contours for analytic continuation of
5130: Eqs.~(\protect\ref{aa1}).}
5131: \label{figA1}
5132: \end{figure}
5133: }
5134:
5135: In this Appendix we show in some detail the procedure of analytical
5136: continuation that leads to the expression for the interaction
5137: correction Eq.~(\ref{cc}) to the conductivity in terms of exact Green's
5138: function of noninteracting disordered system.
5139: The structure of the current correlator is
5140:
5141: \begin{mathletters}
5142: \label{aa1}
5143: \beqa
5144: &&\int\limits_{0}^{1/T} d\tau \langle {\bf T_\tau} \hat j_\alpha(\tau)
5145: \hat j_\beta(0) \rangle e^{i\Omega_n\tau} =
5146: \nonumber\\
5147: &&
5148: \
5149: - T\sum_{\epsilon_n}J_\alpha
5150: G(i\epsilon_n+ i\Omega_n)
5151: J_\beta G(i\epsilon_n)
5152: \label{aa1a} \\
5153: &&
5154: \
5155: - T\sum_{\epsilon_n}J_\alpha
5156: G(i\epsilon_n+ i\Omega_n)
5157: \Gamma_\beta (i\epsilon_n+ i\Omega_n, i\epsilon_n)
5158: G(i\epsilon_n),
5159: \label{aa1b}
5160: \eeqa
5161: \end{mathletters}
5162:
5163: \noindent
5164: where
5165: $\epsilon_n= \pi T(2n+1)$ is the fermionic Matsubara frequency,
5166: $G(i\epsilon_n)$ is the exact Green's function
5167: of the interacting system (diagrams $1$ and $2$ on Fig.~\ref{sr}
5168: are the first order correction to the Green's function),
5169: $\Gamma_\beta (i\epsilon_n+ i\Omega_n, i\epsilon_n)$ is the
5170: vertex function (not to be confused with disorder averaged
5171: interaction vertex $\Gamma$ from Sec.~\ref{diagrams}).
5172: Diagrams $3$ on Fig.~\ref{sr}
5173: are the first order correction to the vertex function.
5174: The current operator is defined in Eq.~(\ref{jop-def}).
5175: Note, that
5176: we omit the spatial coordinates and the integration
5177: whenever it should cause no confusion.
5178:
5179:
5180: We perform the analytic continuation in each term
5181: (\ref{aa1a}) and (\ref{aa1b}) separately. In Eq.~(\ref{aa1a})
5182: we use the standard procedure
5183: \beq
5184: T\sum\limits_{\epsilon_n} (\dots ) = \frac{1}{4 \pi i}
5185: \int\limits_{{\cal C}} d\epsilon \tanh\frac{\epsilon}{2T}(\dots ),
5186: \label{trick}
5187: \eeq
5188: where integration contour is shown on Fig.~\ref{figA1}. We deform
5189: this contour to form $C_1$, use the facts that
5190: $\tanh (\epsilon + i\Omega_n)/2T=\tanh \epsilon/2T$,
5191: and $G(\epsilon \pm i0)=G^{R(A)}(\epsilon)$ and we obtain
5192: \beqa
5193: &&M_1(i\Omega_n)=T\sum_{\epsilon_n}J_\alpha
5194: G(i\epsilon_n+ i\Omega_n) J_\beta G(i\epsilon_n)
5195: \nonumber\\
5196: &&=\int\frac{d\epsilon}{4 \pi i}
5197: \tanh\frac{\epsilon}{2T}
5198: \Big\{
5199: J_\alpha G(\epsilon + i\Omega_n) J_\beta
5200: \left[G^R(\epsilon ) - G^A(\epsilon ) \right]
5201: \nonumber\\
5202: &&+ J_\alpha \left[G^R(\epsilon ) - G^A(\epsilon ) \right]
5203: J_\beta G(\epsilon - i\Omega_n)
5204: \Big\}
5205: \label{aa2}
5206: \eeqa
5207: In the form (\ref{aa2}) frequency $\Omega$ is present only in
5208: functions which may have singularities only on the real axis, so that
5209: the required analytic continuation can be easily performed:
5210:
5211: \beqa
5212: &&M_1(\omega)=
5213: M_1(i\Omega_n \to \omega + i0)
5214: \nonumber\\
5215: &&=\int\frac{d\epsilon}{4 \pi i}
5216: \tanh\frac{\epsilon}{2T}
5217: \Big\{
5218: J_\alpha G^R(\epsilon + \omega) J_\beta
5219: \left[G^R(\epsilon ) - G^A(\epsilon ) \right]
5220: \nonumber\\
5221: &&\quad + J_\alpha \left[G^R(\epsilon ) - G^A(\epsilon ) \right]
5222: J_\beta G^A(\epsilon - \omega)
5223: \Big\}
5224: \label{aa3}
5225: \eeqa
5226: Thus one obtains for the quantity entering into conductivity
5227: (\ref{kubo})
5228: \beqa
5229: &&N_1=- \lim_{\omega \to 0}
5230: {\rm Im}
5231: \left(
5232: \frac{M_1(\omega)}{\omega}
5233: \right) \nonumber\\
5234: &&=
5235: {\rm Re}
5236: \int\frac{d\epsilon}{2 \pi}
5237: \tanh\frac{\epsilon}{2T}
5238: J_\alpha G^A(\epsilon )
5239: J_\beta \partial_\epsilon G^A(\epsilon)
5240: \nonumber\\
5241: &&
5242: + \int\frac{d\epsilon}{4 \pi}
5243: \left(\frac{d}{d \epsilon}\tanh\frac{\epsilon}{2T}
5244: \right)
5245: J_\alpha G^R(\epsilon ) J_\beta
5246: G^A(\epsilon )
5247: \label{aa4}
5248: \eeqa
5249: Equation (\ref{aa4}) can be further simplified for the calculation of
5250: the symmetric part of the conductivity
5251: \beqa
5252: &&N_1^{sym}=
5253: {\rm Re}
5254: \int\frac{d\epsilon}{8\pi}
5255: \left(\frac{d}{d \epsilon}\tanh\frac{\epsilon}{2T}
5256: \right)\label{aa5}\\
5257: &&\times
5258: \Big[
5259: - J_\alpha G^A(\epsilon )
5260: J_\beta G^A(\epsilon)
5261: + J_\alpha G^R(\epsilon ) J_\beta
5262: G^A(\epsilon )
5263: +(\alpha \leftrightarrow \beta)
5264: \Big]
5265: \nonumber
5266: \eeqa
5267:
5268: \end{multicols}
5269: \widetext
5270: Term (\ref{aa1b}) is considered analogously.
5271: We find similarly to Eq.~(\ref{aa2})
5272: \beqa
5273: &&M_2(i\Omega_n)=T\sum_{\epsilon_n}J_\alpha
5274: G(i\epsilon_n+ i\Omega_n)
5275: \Gamma_\beta (i\epsilon_n+ i\Omega_n, i\epsilon_n) G(i\epsilon_n)
5276: \nonumber\\
5277: &&=\int\frac{d\epsilon}{4 \pi i}
5278: \tanh\frac{\epsilon}{2T}
5279: \Big\{
5280: J_\alpha G(\epsilon + i\Omega_n)
5281: \left[\Gamma_\beta (\epsilon + i\Omega_n, \epsilon + i0)
5282: G^R(\epsilon ) -
5283: \Gamma_\beta (\epsilon + i\Omega_n, \epsilon - i0)
5284: G^A(\epsilon ) \right]
5285: \nonumber\\
5286: &&+ J_\alpha \left[G^R(\epsilon )
5287: \Gamma_\beta (\epsilon + i0, \epsilon - i\Omega_n)
5288: - G^A(\epsilon )
5289: \Gamma_\beta (\epsilon - i0, \epsilon - i\Omega_n)
5290: \right]
5291: G(\epsilon - i\Omega_n )
5292: \Big\}
5293: \label{aa6}
5294: \eeqa
5295: Using analytic properties of the Green's function and the vertex
5296: function, we perform the analytic continuation and obtain
5297: \beqa
5298: &&M_2(\omega)= M_2(i\Omega_n \to \omega+i0)
5299: \nonumber\\
5300: &&
5301: =
5302: \lim_{\delta_1 \to 0^+}\lim_{\delta_2 \to 0^+}
5303: \int\frac{d\epsilon}{4 \pi i}
5304: \tanh\frac{\epsilon}{2T}
5305: \Big\{
5306: J_\alpha G^R(\epsilon + \omega)
5307: \left[\Gamma_\beta (\epsilon + \omega + i\delta_1, \epsilon + i\delta_2)
5308: G^R(\epsilon ) -
5309: \Gamma_\beta (\epsilon + \omega + i0, \epsilon - i0)
5310: G^A(\epsilon ) \right]
5311: \nonumber\\
5312: &&+ J_\alpha \left[G^R(\epsilon )
5313: \Gamma_\beta (\epsilon + i0, \epsilon - \omega - i0)
5314: - G^A(\epsilon )
5315: \Gamma_\beta (\epsilon - i\delta_2, \epsilon - \omega - i\delta_1)
5316: \right]
5317: G^A(\epsilon - \omega )
5318: \Big\}.
5319: \label{aa7}
5320: \eeqa
5321: In the appropriate frequency limit, we find
5322: \beqa
5323: &&N_2=
5324: - \lim_{\omega \to 0}
5325: {\rm Im}
5326: \left(
5327: \frac{M_2(\omega)}{\omega}
5328: \right)
5329: =
5330: {\rm Re}
5331: \int\frac{d\epsilon}{2\pi }
5332: \tanh\frac{\epsilon}{2T}
5333: J_\alpha
5334: G^A(\epsilon)
5335: \left.
5336: \frac{\partial}{\partial \epsilon_1}
5337: \right|_{\epsilon_1=\epsilon }
5338: \Big[
5339: \lim_{\delta_1 \to 0^+}\lim_{\delta_2 \to 0^+}
5340: \Gamma_\beta (\epsilon - i\delta_2, \epsilon_1 - i\delta_1)
5341: G^A(\epsilon_1 )
5342: \Big]
5343: \nonumber\\
5344: &&+
5345: \int\frac{d\epsilon}{4 \pi}
5346: \left(\frac{d}{d \epsilon}\tanh\frac{\epsilon}{2T}
5347: \right)
5348: J_\alpha G^R(\epsilon )
5349: \Gamma_\beta (\epsilon + i0, \epsilon - i0)
5350: G^A(\epsilon )
5351: \label{aa8}
5352: \eeqa
5353:
5354: \begin{multicols}{2}
5355: Further calculation requires specification of the form of
5356: the self-energy and the vertex function. We have to find both
5357: in the first order in interaction propagator, and expand
5358: Eq.~(\ref{aa5}) up to the first order:
5359: \beqa
5360: &&\delta N_1^{sym}=
5361: {\rm Re}
5362: \int\frac{d\epsilon}{4\pi}
5363: \left(\frac{d}{d \epsilon}\tanh\frac{\epsilon}{2T}
5364: \right)\label{aa50}\\
5365: &&\times
5366: \Big[-
5367: J_\alpha G^A(\epsilon ) \Sigma^A(\epsilon ) G^A(\epsilon )
5368: J_\beta G^A(\epsilon)
5369: \nonumber\\
5370: &&
5371: + J_\alpha G^R(\epsilon ) J_\beta
5372: G^A(\epsilon )
5373: \Sigma^A(\epsilon ) G^A(\epsilon )
5374: +(\alpha \leftrightarrow \beta)
5375: \Big].
5376: \nonumber
5377: \eeqa
5378: For brevity we consider only the ``Fock'' contribution of
5379: Fig.~\ref{sr} b:
5380: \beq
5381: \Sigma(i\epsilon_n )_{12} = T\sum_{\Omega_m}
5382: {\cal D}_{12}(i\Omega_m)
5383: G_{12}(i\epsilon_n-i\Omega_m )
5384: \label{aa9}
5385: \eeq
5386: where ${\cal D}$ is the bosonic propagator, and we restored the
5387: notation for spatial coordinates.
5388: In the same order
5389: \beqa
5390: &&\left[\Gamma(i\epsilon_n, i\epsilon_m )_{\beta}\right]_{12}
5391: \label{aa10}\\
5392: =
5393: &&\quad
5394: T\sum_{\Omega_m}{\cal D}_{12}(\Omega_m)
5395: \left[G(i\epsilon_n- i\Omega_m ) J_\beta
5396: G(i\epsilon_m- i\Omega_m ) \right]_{12}.
5397: \nonumber
5398: \eeqa
5399:
5400:
5401: After analytic continuation similar to that in the derivation of
5402: Eq.~(\ref{aa3}) we find
5403: \beqa
5404: &&\Sigma^A_{12}(\epsilon) =
5405: - \int \frac{d \Omega}{2\pi} \coth \frac{\Omega}{2T}
5406: \Big[{\rm Im} {\cal D}_{12}^A(\Omega)
5407: \Big]G^A_{12}(\epsilon-\Omega)
5408: \label{aa11}\\
5409: &&+
5410: i\int \frac{d \Omega}{4\pi}
5411: \tanh \frac{\epsilon-\Omega}{2T}
5412: {\cal D}_{12}^A(\Omega)
5413: \Big[G^A_{12}(\epsilon-\Omega)
5414: - G^R_{12}(\epsilon-\Omega)
5415: \Big]
5416: \nonumber
5417: \eeqa
5418: and for the vertex function we have two cases:
5419:
5420: \end{multicols}
5421: \widetext
5422: \begin{mathletters}
5423: \label{aa12}
5424: \beqa
5425: \lim_{\delta_1 \to 0^+}\lim_{\delta_2 \to 0^+}
5426: \Gamma_\beta (\epsilon - i\delta_2, \epsilon_1 - i\delta_1)&&
5427: =
5428: - \int \frac{d \Omega}{2\pi} \coth \frac{\Omega}{2T}
5429: \Big[{\rm Im} {\cal D}_{12}^A(\Omega)
5430: \Big]\Big[G^A(\epsilon-\Omega)J_\beta G^A(\epsilon_1-\Omega)\Big]_{12}
5431: \nonumber\\
5432: &&
5433: +i\int \frac{d \Omega}{4\pi} \tanh \frac{\epsilon - \Omega}{2T}
5434: {\cal D}_{12}^A(\Omega)
5435: \Big[\Big(G^A(\epsilon -\Omega)
5436: - G^R(\epsilon -\Omega)\Big)J_\beta G^A(\epsilon_1-\Omega)\Big]_{12}
5437: \nonumber\\
5438: &&
5439: +i\int \frac{d \Omega}{4\pi} \tanh \frac{\epsilon_1 - \Omega}{2T}
5440: {\cal D}_{12}^A(\Omega)
5441: \Big[G^R(\epsilon -\Omega)J_\beta\Big(G^A(\epsilon_1-\Omega)
5442: - G^R(\epsilon_1-\Omega)\Big) \Big]_{12}
5443: \eeqa
5444: \beqa
5445: \Gamma_\beta (\epsilon + i0, \epsilon - i0) &&
5446: =
5447: - \int \frac{d \Omega}{2\pi} \coth \frac{\Omega}{2T}
5448: \Big[{\rm Im} {\cal D}_{12}^A(\Omega)
5449: \Big]\Big[G^R(\epsilon-\Omega)J_\beta G^A(\epsilon-\Omega)\Big]_{12}
5450: \nonumber\\
5451: &&
5452: +i\int \frac{d \Omega}{4\pi} \tanh \frac{\epsilon - \Omega}{2T}
5453: {\cal D}_{12}^R(\Omega)
5454: \Big[\Big(G^A(\epsilon -\Omega)
5455: - G^R(\epsilon -\Omega)\Big)J_\beta G^A(\epsilon-\Omega)\Big]_{12}
5456: \nonumber\\
5457: &&
5458: +i\int \frac{d \Omega}{4\pi} \tanh \frac{\epsilon - \Omega}{2T}
5459: {\cal D}_{12}^A(\Omega)
5460: \Big[G^R(\epsilon -\Omega)J_\beta\Big(G^A(\epsilon -\Omega)
5461: - G^R(\epsilon -\Omega)\Big) \Big]_{12}
5462: \eeqa
5463: \end{mathletters}
5464:
5465:
5466:
5467: We now substitute Eq.~(\ref{aa11}) into Eq.~(\ref{aa50}). We use the
5468: fact that the combination containing only retarded or only advanced
5469: Green's functions vanish upon the disorder averaging. Moreover, the
5470: average of the combinations like $G(\epsilon)G(\epsilon-\Omega_1)\dots
5471: G(\epsilon-\Omega_N)$ does not depend on the energy $\epsilon$, which
5472: enable us to perform the integration over $\epsilon$ using
5473:
5474: \[
5475: \int d\epsilon \tanh \frac{\epsilon-\Omega}{2T}
5476: \frac{d}{d\epsilon}\tanh \frac{\epsilon}{2T}=
5477: - 2 \frac{d}{d\Omega}\left(\Omega \coth \frac{\Omega}{2T}\right).
5478: \]
5479: We find using $D^A(\Omega) = [D^A(-\Omega)]^*$
5480: \begin{mathletters}
5481: \label{aa500}
5482: \beqa
5483: &&\delta N_1^{sym}= {\rm Im}
5484: \int\frac{d\Omega}{8\pi^2}
5485: \left[\frac{d}{d\Omega}\left(\Omega \coth \frac{\Omega}{2T}\right)
5486: \right]{\cal D}_{12}^A(\Omega)
5487: \label{aa500a}\\
5488: &&
5489: \times
5490: \Big[
5491: J_\alpha G^A(\epsilon )
5492: G^R_{12} (\epsilon -\Omega) G^A(\epsilon )
5493: J_\beta G^A(\epsilon) -
5494: J_\alpha G^A(\epsilon )
5495: G^R_{12} (\epsilon -\Omega) G^A(\epsilon )
5496: J_\beta G^R(\epsilon)
5497: - J_\alpha G^R(\epsilon )
5498: G^R_{12} (\epsilon -\Omega) G^R(\epsilon )
5499: J_\beta G^A(\epsilon)
5500: \Big]
5501: \nonumber\\
5502: &&
5503: -
5504: \int\frac{d\Omega}{4\pi^2}
5505: \Big[
5506: \frac{\Omega}{2T\sinh^2\frac{\Omega}{2T}}
5507: {\rm Im} {\cal D}_{12}^A(\Omega)
5508: \Big] {\rm Re}
5509: \Big[ J_\alpha G^A(\epsilon )
5510: G^A_{12} (\epsilon -\Omega) G^A(\epsilon )
5511: J_\beta G^R(\epsilon)\Big] + (\alpha \leftrightarrow \beta).
5512: \label{aa500b}
5513: \eeqa
5514: \end{mathletters}
5515:
5516: The same manipulations are performed with substituition of
5517: Eqs.~(\ref{aa12}) into Eq.~(\ref{aa8}). One finds for the symmetrized
5518: part
5519:
5520: \begin{mathletters}
5521: \label{aa800}
5522: \beqa
5523: &&\delta N_2^{sym}= {\rm Im}
5524: \int\frac{d\Omega}{8\pi^2}
5525: \left[\frac{d}{d\Omega}\left(\Omega \coth \frac{\Omega}{2T}\right)
5526: \right]{\cal D}_{12}^A(\Omega)
5527: \label{aa800a}\\
5528: &&
5529: \times
5530: \Bigg\{
5531: \Big[ G^A(\epsilon ) J_\alpha G^A(\epsilon ) \Big]_{12}
5532: \Big[ G^R (\epsilon -\Omega)
5533: J_\beta G^R(\epsilon- \Omega ) \Big]_{21}
5534: - 2 \Big[ G^R(\epsilon ) J_\alpha G^A(\epsilon ) \Big]_{12}
5535: \Big[ G^R (\epsilon -\Omega)
5536: J_\beta G^R(\epsilon- \Omega ) \Big]_{21}
5537: \Bigg\}
5538: \nonumber\\
5539: &&
5540: -
5541: \int\frac{d\Omega}{8\pi^2}
5542: \Big[
5543: \frac{\Omega}{2T\sinh^2\frac{\Omega}{2T}}
5544: {\rm Im} {\cal D}_{12}^A(\Omega)
5545: \Big]
5546: \Big\{ \Big[ G^R(\epsilon ) J_\alpha G^A(\epsilon ) \Big]_{12}
5547: \Big[ G^A (\epsilon -\Omega)
5548: J_\beta G^R(\epsilon- \Omega ) \Big]_{21}\Big\}
5549: + (\alpha \leftrightarrow \beta).
5550: \label{aa800b}
5551: \eeqa
5552: \end{mathletters}
5553:
5554:
5555: Total correction to the conductivity is just $N_1+N_2$.
5556: In the Hartree-Fock approximation $D^A=-V(q)$ and we obtain
5557: Eq.~(\ref{cc}). In the case for the stronger interaction
5558: terms (\ref{aa500a}) and (\ref{aa800a}) are added to produce
5559: Eq.~(\ref{cci}) and terms (\ref{aa500b}), (\ref{aa800b})
5560: give rise to the inelastic or so-called dephasing term:
5561: \beqa
5562: \delta\sigma^{deph}_{\alpha\beta}
5563: &=& -
5564: \int\frac{d\Omega}{8\pi^2}
5565: \Big[
5566: \frac{\Omega}{2T\sinh^2\frac{\Omega}{2T}}
5567: {\rm Im} {\cal D}_{12}^A(\Omega)
5568: \Big] {\rm Re}
5569: \Bigg\{
5570: \Big[ 2 J_\alpha G^A(\epsilon )
5571: G^A_{12} (\epsilon -\Omega) G^A(\epsilon )
5572: J_\beta G^R(\epsilon)\Big]
5573: \nonumber\\
5574: &+&
5575: \Big[ G^R(\epsilon ) J_\alpha G^A(\epsilon ) \Big]_{12}
5576: \Big[ G^A (\epsilon -\Omega)
5577: J_\beta G^R(\epsilon- \Omega ) \Big]_{21}
5578: \Bigg\}
5579: +
5580: (\alpha \leftrightarrow \beta).
5581: \label{adeph}
5582: \eeqa
5583:
5584: \begin{multicols}{2}
5585:
5586:
5587: In our leading approximation in $1/E_F\tau$ this term vanishes,
5588: see Fig.~\ref{vanishing}. The role of this term in the temperature
5589: dependence of weak localization correction is discussed in detail in
5590: Ref.~\onlinecite{aag}.
5591: {
5592: \narrowtext
5593: \begin{figure}[ht]
5594: \vspace{0.2 cm}
5595: \epsfxsize=7 cm
5596: \centerline{\epsfbox{f24.eps}}
5597: \vspace{0.2cm}
5598: \caption{Cancellation of inelastic term (\protect\ref{adeph}) in
5599: the leading ladder approximation.}
5600: \label{vanishing}
5601: \end{figure}
5602: }
5603:
5604:
5605: \begin{references}
5606:
5607: \bibitem{aar} B.L. Altshuler and A.G. Aronov in
5608: {\em Electron-Electron Interactions in Disordered Systems},
5609: eds. A.L. Efros, M. Pollak (North-Holland, Amsterdam, 1985).
5610: \bibitem{fin} A.M. Finkelstein, Zh. Eksp. Teor. Fiz. {\bf 84}, 168 (1983)
5611: [Sov. Phys. JETP {\bf 57}, 97 (1983)]; Z. Phys. B {\bf 56}, 189 (1984).
5612: \bibitem{Lee} C. Castellani, C. Di Castro, P. A. Lee, and M. Ma,
5613: Phys. Rev. B {\bf 30}, 527 (1984); {\it ibid} {\bf 30}, 1596 (1984);
5614: C. Castellani, C. Di Castro, P.A. Lee, {\it ibid} {\bf 57}, R9381 (1998).
5615: \bibitem{stern}F. Stern, Phys. Rev. Lett. {\bf 44}, 1469 (1980).
5616: \bibitem{dstern} F. Stern and S. Das Sarma, Solid State Electron.,
5617: {\bf 28}, 158 (1985).
5618: \bibitem{gdl} A. Gold and
5619: V.T. Dolgopolov, Phys. Rev. B {\bf 33}, 1076 (1986).
5620: \bibitem{das} S. Das Sarma and E.H. Hwang, Phys. Rev. Lett.
5621: {\bf 83}, 164 (1999).
5622: \bibitem{aae} B.L. Altshuler, A.G. Aronov,
5623: M.E. Gershenson, and Yu.V. Sharvin,
5624: Sov. Sci. A. Phys. {\bf 9}, 223 (1987).
5625: \bibitem{Bergman}G. Bergman, Physics Reports, {\bf 107}, 1 (1984).
5626: \bibitem{cwh} K.M. Cham and R.G. Wheeler, Phys. Rev. Lett.
5627: {\bf 44}, 1472 (1980).
5628: \bibitem{new} For a review see B.L. Altshuler, D.L. Maslov, and
5629: V.M. Pudalov, Physica E, {\bf 9} (2), 209 (2001).
5630: \bibitem{nfl} For a recent review and references see
5631: E. Abrahams, S. V. Kravchenko, and M. P. Sarachik,
5632: Rev. Mod. Phys. {\bf 73}, 251 (2001).
5633: \bibitem{footnote1} Logarithmic renormalization of $F_0^\sigma$,
5634: see Ref. \onlinecite{fin,Lee} will not be relevant for the further discussion.
5635: \bibitem{pre} G. Zala, B.N. Narozhny, and I.L. Aleiner (in
5636: preparation).
5637: \bibitem{rag} A.M. Rudin, I.L. Aleiner, and L.I. Glazman,
5638: Phys. Rev. B, {\bf 55}, 9322 (1997).
5639: \bibitem{aag} I.L. Aleiner, B.L. Altshuler, and M.E. Gershenson,
5640: Waves Random Media {\bf 9}, 201 (1999).
5641: \bibitem{kit} C. Kittel, {\it ``Quantum Theory of Solids''}
5642: (Wiley, New York, 1963).
5643: \bibitem{llq} L.D. Landau, I.M. Lifshits, {\it ``Quantum Mechanics''}
5644: (Pergamon, New York, 1977).
5645: \bibitem{gri} I.S. Gradstein I.M. Ryzhik, {\it Table of Integrals,
5646: Series, and Products} (Academic Press, San Diego, 1994).
5647: \bibitem{gla} K.A. Matveev, D. Yue, and L. I. Glazman,
5648: Phys. Rev. Lett. {\bf 71}, 3351 (1993); D. Yue, L.I. Glazman, and
5649: K.A. Matveev, Phys. Rev. B {\bf 49}, 1966 (1994);
5650: \bibitem{Fermiliquid} L.D. Landau, Zh. Eksp. Teor. Fiz. {\bf 30}, 1058 (1956);
5651: {\it ibid} {\bf 32}, 59 (1957).
5652: \bibitem{AGD} A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski,
5653: {\it ``Methods of quantum field theory in statistical physics''}
5654: (Prentice-Hall, Englewood Cliffs, N.J., 1963).
5655: \bibitem{Wignercrystal} B. Tanatar and D. M. Ceperley,
5656: Phys. Rev. {\bf B} 39, 5005 (1989)
5657: \bibitem{vsi} This statement contradicts conclusions of
5658: Q. Si, C.M. Varma, Phys. Rev. Lett. {\bf 81}, 4951 (1998).
5659: We believe that this paper mishandles the Fermi liquid renormalizations,
5660: which amounts to omission of $F_0^\rho$ in the numerator of
5661: Eq.~(\ref{f-prop}), see also comment by
5662: P. Schwab and C. Castellani, Phys. Rev. Lett. {\bf 84}, 4779 (2000).
5663: \bibitem{bkm} D. Belitz, T.R. Kirkpatrick, and T. Vojta,
5664: Phys. Rev. B {\bf 55}, 9452 (1997); D. Belitz, T.R. Kirkpatrick,
5665: A.J. Millis, and T. Vojta, Phys. Rev. B {\bf 58}, 14155 (1998).
5666: \bibitem{lar} We are grateful to A.I. Larkin for
5667: calling our attention to the momentum dependence of $F^\sigma_0$.
5668: \bibitem{rei} M.Yu. Reizer, Phys. Rev. B {\bf 57}, 12338 (1998).
5669: \bibitem{Keldysh} L.V. Keldysh, Zh. Eksp. Teor. Fiz. {\bf 47}, 1945 (1964)
5670: [Sov. Phys. JETP {\bf 20}, 1018 (1964)].
5671: \bibitem{Eilenberger}G. Eilenberger, Z. Phys. Bd. {\bf 214}, 195
5672: (1968).
5673: \bibitem{68} A.I. Larkin and Y.N. Ovchinnikov,
5674: Zh. Eksp. Teor. Fiz. {\bf 55}, 2262 (1968)
5675: [Sov. Phys. JETP {\bf 28}, 1200 (1969)].
5676: \bibitem{LO77} A.I. Larkin and Y.N. Ovchinnikov,
5677: Zh. Eksp. Teor. Fiz. {\bf 73}, 299 (1977)
5678: [Sov. Phys. JETP {\bf 46}, 155 (1977)].
5679: \end{references}
5680:
5681:
5682: \end{multicols}
5683:
5684: \end{document}
5685: