cond-mat0505595/ER7.tex
1: %% This document created by Scientific Word (R) Version 3.5
2: 
3: \documentclass
4: [prb,10pt,a4paper,twocolumn,tightenlines,showpacs,preprint]{revtex4}%
5: \usepackage{amsmath}
6: \usepackage{graphicx}
7: \usepackage{amsfonts}
8: \usepackage{amssymb}%
9: \setcounter{MaxMatrixCols}{30}
10: %TCIDATA{OutputFilter=latex2.dll}
11: %TCIDATA{Version=4.10.0.2363}
12: %TCIDATA{CSTFile=revtex4.cst}
13: %TCIDATA{Created=Thursday, July 10, 2003 06:47:20}
14: %TCIDATA{LastRevised=Tuesday, May 17, 2005 14:35:03}
15: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
16: %TCIDATA{<META NAME="DocumentShell" CONTENT="Articles\SW\REVTeX 4">}
17: %TCIDATA{Language=American English}
18: \newtheorem{theorem}{Theorem}
19: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
20: \newtheorem{algorithm}[theorem]{Algorithm}
21: \newtheorem{axiom}[theorem]{Axiom}
22: \newtheorem{claim}[theorem]{Claim}
23: \newtheorem{conclusion}[theorem]{Conclusion}
24: \newtheorem{condition}[theorem]{Condition}
25: \newtheorem{conjecture}[theorem]{Conjecture}
26: \newtheorem{corollary}[theorem]{Corollary}
27: \newtheorem{criterion}[theorem]{Criterion}
28: \newtheorem{definition}[theorem]{Definition}
29: \newtheorem{example}[theorem]{Example}
30: \newtheorem{exercise}[theorem]{Exercise}
31: \newtheorem{lemma}[theorem]{Lemma}
32: \newtheorem{notation}[theorem]{Notation}
33: \newtheorem{problem}[theorem]{Problem}
34: \newtheorem{proposition}[theorem]{Proposition}
35: \newtheorem{remark}[theorem]{Remark}
36: \newtheorem{solution}[theorem]{Solution}
37: \newtheorem{summary}[theorem]{Summary}
38: \newenvironment{proof}[1][Proof]{\noindent\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
39: \begin{document}
40: \title{Non-ohmicity and energy relaxation in diffusive 2D metals }
41: \author{Roy Ceder, Oded Agam, and Zvi Ovadyahu}
42: \affiliation{Racah Institute of Physics, The Hebrew University, Jerusalem 91904, Israel}
43: 
44: \begin{abstract}
45: 
46: We analyze current-voltage characteristics taken on Au-doped indium-oxide films.
47: By fitting a scaling function to the data, we extract the electron-phonon
48: scattering rate as function of temperature, which yields a quadratic dependence
49: of the electron-phonon scattering rate on temperature from 1K down to 0.28K. The
50: origin of this enhanced electron-phonon scattering rate is ascribed to the
51: mechanism proposed by Sergeev and Mitin.
52: \end{abstract}
53: \pacs{ 72.15.Lh,72.10.-d, 73.61.-r, 63.20.Kr, 72.10.Di } 
54: \maketitle
55: 
56: 
57: \section{Introduction}
58: 
59: Energy relaxation processes play an important role in the low temperature
60: transport properties of diffusive metals, alloys and semiconductors. In
61: particular, they determine the maximum electric field $F$ allowed if the
62: system is to be measured under near-equilibrium conditions at a given
63: temperature. The condition for that may be expressed as $eFL_{\epsilon}\ll
64: k_{B}T$ where the energy relaxation length $L_{\epsilon}$ is the length over
65: which the electron diffuses under $F$ before the energy gained from it is
66: dissipated into the thermal bath. In particular, this issue is relevant for
67: all aspects of quantum transport such as corrections to the conductivity due
68: to interference and electron-electron interactions. For a system of size $L\gg
69: L_{\epsilon}$, energy relaxation processes are usually controlled by
70: electron-phonon inelastic scattering, and in the following we shall assume
71: that $L_{\epsilon}$ is dominated by the electron-phonon diffusion length
72: $L_{ep}$.
73: 
74: In clean samples the electron-phonon scattering mechanism is well understood,
75: and the scattering rate, $\tau_{ep}^{-1}$, is known to be proportional to
76: $T^{3}$ (where $T$ denotes the temperature). In dirty systems however, where
77: the elastic mean free path of the electron is smaller than the phonon thermal
78: length, the situation is more complicated. Schmid \cite{Schmid73} showed that
79: in this case $\tau_{ep}^{-1}$ is suppressed and becomes proportional to
80: $T^{4}$, in accordance with Pipard's ineffectiveness condition
81: \cite{Pippard55}. His model assumed that the impurities are anchored to the
82: lattice, and the scattering rate was calculated by moving into a reference
83: frame which follows the lattice vibrations. Riezer and Sergeev
84: obtained the same result using the laboratory frame of reference
85: \cite{Reizer86}. A scattering rate proportional to $T^{4}$ has indeed been
86: observed in a number of experiments \cite{Gershenson01,Karvonen04, Kivinen04}.
87: However, a $T^{3}$ law was frequently observed $\ $ even in systems that were
88: presumably well into the dirty limit regime \cite{Roukes85, Wennberg86,
89: Eshternach92, Wellstood94}. Moreover, quite a few observations of a $T^{2}$
90: scattering law were reported in other experiments \cite{Bergmann90,
91: DiTusa92,Watson95, Wu98}. The latter experimental results triggered further
92: studies with the aim of understanding better the electron-phonon scattering
93: mechanism in disordered metals. In particular, to obtain an electron-phonon
94: scattering rate proportional to $T^{2}$ (rather than the "ineffective" $T^{4}$
95: law), Belitz and Wybourne \cite{Belitz95} assumed a strong phonon damping,
96: while Jan Wu and Wei \cite{Jan05} included effects associated with the
97: discrete lattice structure. Sergeev and Mitin \cite{Sergeev00} obtained the
98: $T^{2}$ behavior from a model were impurities are assumed to be fixed, namely,
99: impurities which do not follow the lattice vibrations. They argued that heavy
100: impurities or boundaries which move differently from the host lattice produce
101: the same effect.
102: 
103: \begin{figure*}
104: \includegraphics[width=79mm]{fig_1c.ps}
105: \includegraphics[width=8.1cm]{fig_2c.ps}
106: \caption{ The non-ohmic characteristic of 200\AA  thick In$_{2}$O$_{3-x}$:Au sample with
107: length L=3500$\mu m$ and width W=1mm. Left panel: The differential resistance
108: (defined in Eq.~(\ref{scaling-function}) as function of the voltage. Right
109: panel: The same data plotted as function of $F(=V/L)$ normalized by $T^{2}$.
110: Note the near perfect data collapse as well as the fit to formula
111: (\ref{R-scaling-1}) represented by the continuous line.}%
112: \end{figure*}
113: 
114: In this paper we analyze the non-ohmic characteristic of thin films of
115: In$_{2}$O$_{3-x}$:Au (crystalline indium-oxide doped with 2\% gold), that were
116: characterized and measured as described elsewhere \cite{Zvi_1}. We show that
117: in this system $\tau_{ep}^{-1}\propto T^{2}$, and interpret this behavior as a
118: manifestation of the Sergeev Mitin mechanism where the Au atoms play the role
119: of the \textquotedblleft immobile impurities\textquotedblright.  The sample
120: resistance is used as a thermometer of the electron temperature. The latter is determined
121:  by the energy balance between the Joule heating, due to the presence of electric field, 
122: and the heat transfer to the lattice phonons governed by $\tau_{ep}^{-1}.$
123: 
124: \bigskip The data we shall analyze are dynamic resistance $R$ versus voltage
125: traces for a typical film at $T\leq1K$ shown in Fig.~1 (left panel). As
126: expected, the deviations from ohm's law become more pronounced as $T$
127: decreases. In the right panel of Fig.~1 it is shown that these curves taken at
128: different temperatures can be collapsed onto a common function:
129: \begin{equation}
130: \frac{\Delta R}{R}\equiv\frac{R(F,T)-R(0,T)}{R(0,T)}=\Delta\mathcal{R}%
131: _{p}\left(  \frac{F}{T^{\frac{p}{2}+1}}\right)  . \label{scaling-function}%
132: \end{equation}
133: with $p=2$. The scaling of the electric field as a power of the temperature,
134: $F\sim T^{\frac{p}{2}+1}$, and its relation to the electron-phonon time
135: $\tau_{ep}^{-1}(T)\propto T^{p}$, has been already recognized by Anderson,
136: Abrahams and Ramakrishnan \cite{Anderson79}, and later by Arai \cite{Arai83}.
137: In the next section we shall calculate the function $\Delta\mathcal{R}_{p}$,
138:  and clarify its relation to the electron-phonon relaxation time. We shall 
139: consider the dependence on the sample length and
140: compare with further experimental results in section 3.
141: 
142: \section{Theory}
143: 
144: \bigskip The essence of the picture developed below is the assignment of an
145: effective temperature for a given field and bath-temperature. The electrons
146: are accelerated by the electric field, and the collisions with other electrons
147: and phonons result in a local equilibrium distribution characterized by an
148: effective temperature $T_{eff}$. The latter differs from the bath temperature
149: $T$, and depends on the electric field. We shall make the association
150: $R(F,T)\simeq R(0,T_{eff})$. Hence knowing $T_{eff}(F,T)$ and the form of the
151: near-equilibrium $R(0,T)$ yields the desired $R(F,T)$ from which we deduce the
152: scaling function (\ref{scaling-function}).
153: 
154: We begin by considering the Boltzmann equation for the electron distribution
155: function $f$:
156: \begin{equation}
157: \frac{\partial f}{\partial t}+\vec{v}\cdot\frac{\partial f}{\partial\vec{r}%
158: }+e\vec{F}\cdot\frac{\partial f}{\partial\vec{p}}=I[f]. \label{boltzmann}%
159: \end{equation}
160: Here $\vec{r}$ is the position, $\vec{v}$ is the velocity, $\vec{p}$ denotes
161: the momentum, $\vec{F}$ the electric field, and $I[f]=I_{im}[f]+I_{ee}%
162: [f]+I_{ep}[f,N]$ represents the collision integrals due to impurity
163: scattering, electron-electron, and electron-phonon interactions. The latter
164: depends also on the phonon distribution function, assumed to be the
165: equilibrium distribution function denoted by $N$.
166: 
167: Following Nagaev\cite{Nag_1} and Kozub and Rudin\cite{Kozub}, we look for a
168: steady state solution of the form:%
169: \begin{equation}
170: f=f(\vec{r},\hat{n},\epsilon-e\vec{F}\cdot\vec{r}) \label{f-form}%
171: \end{equation}
172: where $\hat{n}$ denotes a unit vector in the direction of the momentum, and
173: $\epsilon=\epsilon(p)$ is the energy assumed to depend on the absolute value
174: of the momentum $p=|\vec{p}|$. Substituting (\ref{f-form}) in (\ref{boltzmann}%
175: ) leads to:
176: \begin{equation}
177: \vec{v}\cdot\frac{\partial f}{\partial\vec{r}}+\sum_{ij}\frac{eF_{i}}%
178: {p}\left(  \delta_{ij}-n_{i}n_{j}\right)  \frac{\partial f}{\partial n_{j}%
179: }=I[f]. \label{boltzmann1}%
180: \end{equation}
181: Next, we define the symmetric and anti-symmetric parts of the distribution
182: function with respect to the momentum direction $f^{\pm}=\left(  f(\hat{n})\pm
183: f(-\hat{n})\right)  /2$, and construct two new equations from
184: (\ref{boltzmann1}) associated with the addition and subtraction of Boltzmann
185: equations for $f(\hat{n})$ and $f(-\hat{n})$. Then assuming that momentum
186: relaxation is dominated by scattering from impurities, and that
187: electron-electron and electron-phonon interactions are essentially independent
188: of the momentum direction (i.e., $I_{ee}[f]\simeq I_{ee}[f^{+}]$, and
189: $I_{ep}[f,N]\simeq I_{ep}[f^{+},N]$), we obtain:
190: \begin{equation}
191: \vec{v}\frac{\partial f^{-}}{\partial\vec{r}}+\sum_{ij}\frac{eF_{i}}{p}\left(
192: \delta_{ij}-n_{i}n_{j}\right)  \frac{\partial f^{-}}{\partial n_{j}}=\bar
193: {I}[f^+], \label{boltzmann+}%
194: \end{equation}
195: where $\bar{I}[f^+]\simeq I_{ee}[f^{+}]+I_{ep}[f^{+},N]$, and
196: \begin{equation}
197: \vec{v}\frac{\partial f^{+}}{\partial \vec{r}}=I_{im}[f^{-}]. \label{boltzmann-}%
198: \end{equation}
199: 
200: 
201: In the simplest approximation, the impurity collision term takes the form
202: $I_{im}[f^{-}]=-f^{-}/\tau$ where $\tau$ is the elastic mean free time. Thus
203: using (\ref{boltzmann-}), the antisymmetric part of the distribution function
204: can be expressed in terms of the symmetric part and substituted into
205: (\ref{boltzmann+}). The resulting equation is now averaged over the momentum
206: directions to give
207: \begin{equation}
208: -D\nabla^{2}f^{+}=I_{ee}[f^{+}]+I_{ep}[f^{+},N], \label{diffuion-boltzmann}%
209: \end{equation}
210: where $D=\tau v_{F}^{2}/3$ is the diffusion constant ($v_{F}$ is the Fermi
211: velocity). Here and henceforth we neglect the energy dependence of the
212: diffusion constant.
213: 
214: We wish to solve Eq.~(\ref{diffuion-boltzmann}) for homogeneous quasi-two
215: dimensional samples of rectangular shape with voltage contacts located at
216: $x=\pm L/2$. Thus $f^{+}$ is independent of the transverse coordinate, and the
217: boundary conditions assuming ideal contacts are:
218: \begin{equation}
219: \left.  f^{+}\right\vert _{x=\pm L/2}=n_{F}\left(  \frac{\epsilon-\epsilon
220: _{F}\mp\frac{eV}{2}}{k_{B}T}\right)  , \label{f-boundary-conditions}%
221: \end{equation}
222: where $n_{F}(x)=(1+\exp(x))^{-1}$ is the Fermi distribution function,
223: $\epsilon$ is the electron energy, $\epsilon_{F}$ is the Fermi energy, $V=FL$
224: is the voltage drop across the sample, $k_{B}$ is Boltzmann constant, and $T$
225: is the bath temperature.
226: 
227: Equation (\ref{diffuion-boltzmann}) is a nonlinear equation for the electron
228: distribution function. To make progress, we shall assume that the
229: electron-electron diffusion length $L_{ee}$ is much smaller than the system
230: size, $L$, and the energy relaxation length $L_{ep}$. This should secure an
231: effective local thermalization due to the large number of collisions an
232: electron experiences. Looking then for a solution which describes local
233: equilibrium of the electrons:
234: \begin{equation}
235: f^{+}=n_{F}\left(  \frac{\epsilon-\epsilon_{F}-eFx}{k_{B}T_{loc}(x)}\right)  ,
236: \label{anstz}%
237: \end{equation}
238: where $T_{loc}(x)$ is a local temperature of the electrons. At the contacts
239: the local temperature by assumption coincides with the bath temperature:
240: \begin{equation}
241: T_{loc}\left(  \pm\frac{L}{2}\right)  =T, \label{boundaryT}%
242: \end{equation}
243: so that the solution (\ref{anstz}) with the boundary conditions
244: (\ref{boundaryT}) satisfies the requirement (\ref{f-boundary-conditions}).
245: Under the assumption of local equilibrium the electron-electron collision term
246: vanishes and equation (\ref{diffuion-boltzmann}) reduces to
247: \begin{equation}
248: -D\frac{\partial^{2}f^{+}}{\partial x^{2}}=I_{ep}[f^{+},N].
249: \end{equation}
250: Finally, to extract the local temperature behavior we multiply this equation
251: by $\epsilon$ and integrate over the energy. The resulting equation is:
252: \begin{equation}
253: D\left[  \frac{\pi^{2}k_{B}^{2}}{6}\frac{\partial T_{loc}^{2}(x)}{\partial
254: x^{2}}+(eF)^{2}\right]  =-\int  d\epsilon \epsilon I_{ep}[f^{+},N].
255: \label{local-temperature}%
256: \end{equation}
257: To further simplify (\ref{local-temperature}), we take the electron-phonon 
258: collision integral to have
259: the form:
260: \begin{align}
261: I_{ep}  &  =\int d\omega K(\omega)\left[  -\left(  1-f^{+}(\epsilon
262: +\omega)\right)  f^{+}(\epsilon)N\left(  \frac{\omega}{k_{B}T}\right)  \right.
263: \nonumber\\
264: &  +\left(  1-f^{+}(\epsilon)\right)  f^{+}(\epsilon+\omega)\left(  N\left(
265: \frac{\omega}{k_{B}T}\right)  +1\right) \nonumber\\
266: &  -\left(  1-f^{+}(\epsilon-\omega)\right)  f^{+}(\epsilon)\left(  N\left(
267: \frac{\omega}{k_{B}T}\right)  +1\right) \nonumber\\
268: &  \left.  +\left(  1-f^{+}(\epsilon)\right)  f^{+}(\epsilon-\omega)N\left(
269: \frac{\omega}{k_{B}T}\right)  \right]  , ~~~~~~~~ \label{ep-collision}%
270: \end{align}
271: where $K(\omega)=\alpha\omega^{p-1}$, is a kernel depending on the nature of
272: the collision between the electrons and phonons, while $N(x)=(\exp(x)-1)^{-1}$
273: is the equilibrium phonon  distribution function. We substitute this expression
274: into (\ref{local-temperature}) with the approximate local equilibrium form of
275: the electron distribution function (\ref{anstz}), and integrate over $\epsilon$
276: and $\omega$. The resulting equation is an equation for the local temperature
277: \begin{equation}
278: \frac{\pi^{2}k_{B}^{2}}{6}\frac{\partial T_{loc}^{2}(x)}{\partial x^{2}%
279: }+(eF)^{2}=\eta k_{B}^{p+2}\left[  T_{loc}^{p+2}(x)-T^{p+2}\right]  ,
280: \label{local-temp-eq}%
281: \end{equation}
282: where
283: \begin{equation}
284: \eta=(1-2^{-(p+1)})(p+1)!\zeta(p+2)\frac{\alpha}{D}.
285: \end{equation}
286: 
287: 
288: In understanding the form of the solution of equation~(\ref{local-temp-eq}),
289: it is instructive to identify first the relevant length scale for this
290: equation. To this end one may linearize the equation by substituting
291: $T_{loc}^{2}(x)=T^{2}+\delta T^{2}(x)$ and expanding to linear order in
292: $\delta T^{2}(x)$. The solution of the resulting equation with the boundary
293: conditions (\ref{boundaryT}) is
294: \begin{equation}
295: T_{loc}^{2}(x)=T^{2}+\frac{6(eFL_{ep})^{2}}{k_{B}^{2}\pi^{2}}\left(
296: 1-\frac{\cosh\left(  \frac{x}{L_{ep}}\right)  }{\cosh\left(  \frac{L}{2L_{ep}%
297: }\right)  }\right)  , \label{linear-solution}%
298: \end{equation}
299: where
300: \begin{equation}
301: L_{ep}=\frac{\pi\left(  k_{B}T\right)  ^{-\frac{p}{2}}}{\sqrt{(3(p+2)\eta}}
302: \label{ep-length}%
303: \end{equation}
304: is essentially the electron-phonon length at equilibrium. From
305: (\ref{linear-solution}) one can see that $L_{ep}$ sets the distance from the
306: contacts over which the temperature profile reaches a saturated value. Thus
307: long sample satisfies $L\gg L_{ep}$, and these may be considered to have an
308: essentially space independent local temperature which we shall refer to as the
309: effective temperature.
310: 
311: The linearized solution for long samples (\ref{linear-solution}) is strictly
312: justified when the electric field is weak, i.e. $eFL_{ep}<k_{B}T$. We shall be
313: interested in the limit of strong fields where $L_{ep}$ will presumably be
314: smaller than its equilibrium value, the r.h.s of (\ref{ep-length}). Provided
315: $L\gg L_{ep}(T,F)$ it makes sense to assume an electron temperature, which is
316: essentially constant throughout the sample. We then neglect the space
317: dependent term in equation~(\ref{local-temp-eq}), and the effective
318: temperature of the electrons, at any field strength, is readily deduced to be:
319: \begin{equation}
320: T_{eff}\simeq\left[  T^{p+2}+\frac{(eF)^{2}}{\eta k_{B}^{p+2}}\right]
321: ^{\frac{1}{p+2}}.
322: \end{equation}
323: 
324: 
325: As mentioned earlier, once $T_{eff}$ is known, to find the scaling function
326: $\Delta\mathcal{R}_{p}$ requires only the temperature dependence of the
327: resistance, $R(T)$, since $R(F,T)\simeq R(0,T_{eff})$. At low temperatures,
328: the temperature dependent terms of the resistance are the weak localization
329: \cite{Brg_1} and the Altshuler-Aronov corrections \cite{alt_1}. For thin
330: films, both of these have a logarithmic behavior, thus
331: \begin{equation}
332: R(0,T)\simeq R_{D}(1-\gamma\ln(T)), \label{equilibrium-resistance}%
333: \end{equation}
334: where $\gamma$ is a small dimensionless constant, depending on the nature of
335: the electron-electron interactions, the spin-orbit coupling, and the ratio of
336: the quantum unit resistance to the Drude resistance of the sample, $R_{D}$.
337: Substituting $R(F,T)\simeq R(0,T_{eff})$ and (\ref{equilibrium-resistance}) in
338: the definition of the scaling function (\ref{scaling-function}), and expanding
339: to the leading order in $\gamma$ we obtain:
340: \begin{equation}
341: \Delta\mathcal{R}_{p}\left(  \frac{F}{T^{\frac{p}{2}+1}}\right)  \simeq
342: -\frac{\gamma}{p+2}\ln\left[  1+\frac{(eF)^{2}}{\eta\left(  k_{B}T\right)
343: ^{p+2}}\right]  , \label{R-scaling-1}%
344: \end{equation}
345: Note that this has the scaling form as in Eq.~(\ref{scaling-function}) above.
346: 
347: To anticipate the discussion in the next section, it is instructive to
348: consider the case of short samples $L_{ep}\gg L$. Here one may neglect the
349: contribution of the electron-phonon collision term in Eq.~(\ref{local-temperature}) 
350: and readily obtain the solution for
351: $T_{loc}(x)$, whose space dependence, in general, cannot be ignored:
352: \begin{equation}
353: T_{loc}^{2}(x)=T^{2}+\frac{3(eF)^{2}}{k_{B}^{2}\pi^{2}}\left(  \frac{L^{2}}%
354: {4}-x^{2}\right)  . \label{local_T}%
355: \end{equation}
356: At the temperature range where the dephasing length is much smaller than the
357: system size one may view the sample as a set of classical resistors connected in
358: series. Therefore the total resistance can be approximated by the sum:
359: $R(F,T)\simeq\sum_{j}R_{j}$, where $R_{j}=R\left(  T_{loc}(x_{j})\right)  $ is
360: the resistance of the j-th segment (of size of the dephasing length), centered at the point
361: $x_{j}$. Thus, assuming homogeneous sample, the experimentally measured
362: resistivity is essentially an average over the position. From this average one
363: immediately obtains the scaling function of short samples:
364: \begin{equation}
365: \Delta\mathcal{R}_{0}\left(  \frac{V}{T}\right)  \simeq-\gamma\left[
366: \chi\tanh^{-1}\left(  \frac{1}{\chi}\right)  -1\right]  ,
367: \label{short-scaling}%
368: \end{equation}
369: where
370: \begin{equation}
371: \chi=\sqrt{1+\frac{4\pi^{2}}{3}\left(  \frac{k_{B}T}{eV}\right)  ^{2}},
372: \end{equation}
373: and $V=FL$ is the voltage drop along the sample.
374: 
375: \section{Data analysis and discussion}
376: 
377: Comparing the resistance curves shown in Fig.~1 with Eq.~(\ref{R-scaling-1}),
378: one notes that a scaling form that leads to the data collapse occurs for
379: $p=2$. This means that the energy relaxation time is quadratic with
380: temperature: $\tau_{ep}^{-1}\propto\frac{D}{L_{ep}^{2}}\propto T^{2}$, and
381: therefore the electron-phonon length (\ref{ep-length}) is inversely
382: proportional to the temperature.
383: 
384: \begin{figure}[ptb]
385: \includegraphics[width=8cm]{fig_3c.ps}
386: \caption{Resistance versus temperature for the sample of Fig.~1 (open
387: squares) and for a In$_{2}$O$_{3-x}$:Au sample with length L=80$\mu m$ and
388: width W=500$\mu m$ (full circles). The dashed lines are the logarithmic slopes
389: used in defining the value of $\gamma$ (see Eq.~(\ref{equilibrium-resistance})).}%
390: \end{figure}
391: 
392: The procedure of extracting the detailed form of $\tau_{ep}^{-1}$ or $L_{ep}$
393: from the experimental results is as follows. First, the value of $\eta$ is
394: obtained by fitting the $R(F,T)$ data of Fig.~1 to Eq.~(\ref{R-scaling-1}),
395:  as shown in the right panel of this figure. Note that the latter needs the
396: parameter $\gamma$ as input. This is defined by equation
397: (\ref{equilibrium-resistance}) above and is thus obtained from the
398: near-equilibrium $R(T)$ measurement performed on the same sample. Such $R(T)$
399: data and their associated $\gamma$ are shown in Fig.~2 for two samples. These
400: are made from the same batch of a Au-doped $In_{2}O_{3-x}$ film, they only
401: differ in their lateral dimensions. The first is the 3500$\mu$m sample of
402: Fig.~1. The second sample is $80\mu m$ long. The sheet resistances of the two
403: are within 1\% of each other, yet their logarithmic slopes are somewhat
404: different. Also, both samples show a systematic deviation from the theoretical
405: $\ln(T)$ dependence, a feature that seem to occur in some other 2D systems
406: \cite{Mueller94}. Since we are mainly interested in the restricted temperature
407: range $0.28-1K$, this feature may be ignored, and $\gamma$ is defined by
408: fitting $R(T)$ to a simple $\ln(T)$ over the relevant range as shown in the
409: figure. The fits yield $\gamma\approxeq0.0098$ and $\gamma\approxeq0.0081$ for
410: the long and short samples respectively, and these values are used in the
411: subsequent analysis below. 
412: \begin{figure}[ptb]
413: \includegraphics[ width=8cm]{fig_4c.ps}
414: \caption{Dependence of the differential resistance on voltage at several
415: temperatures. In$_{2}$O$_{3-x}$:Au sample with length L=80$\mu m$ and width
416: W=500$\mu m$. The noisier data here (as compared with those of Fig.~1), is
417: mainly due to the much smaller sample size.}%
418: \end{figure}
419: 
420: \begin{figure*}
421: \includegraphics[width=8cm]{fig_5c.ps}
422: \includegraphics[width=8.2cm]{fig_6c.ps}
423: \caption{Left panel: The differential resistance versus the voltage $V$
424: normalized by the temperature $T$ using the data of Fig.~3. The full line is a
425: best fit to the data using Eq.~(\ref{short-scaling}) . Note the data collapse
426: for the three lowest temperatures and the deviations from the short-sample
427: behavior at $T=837mK$.Right panel: Data from the same figure (Fig.~3) plotted
428: as function of $F(=V/L)$ normalized by $T^{2}$. The full line is a best fit to
429: the data using Eq.~(\ref{R-scaling-1}). Note the data collapse for the two
430: highest temperatures and the deviations from the long-sample behavior at
431: $T=279mK$.}%
432: \end{figure*}
433: 
434: An excellent fit to the data in the right panel of Fig.~1 can be
435: obtained using Eq.~(\ref{R-scaling-1}) with $\eta = 3.7\cdot
436: 10^{55}\frac{1}{Joule^{2}m^{2}}$, $\gamma= 0.0098$ and $p=2$. The use
437: of this formula, appropriate for the $L\gg L_{ep}$ limit, is justified for
438: this sample as can be seen by estimating $L_{ep}$. Inserting the above value
439: of $\eta$ in equation (\ref{ep-length}), gives $L_{ep}\simeq$ $20\mu m$ at
440: $1K$ and only $L_{ep}\simeq60\mu m$ at $T\simeq 0.28K$. Thus $L_{ep}$ is much
441: smaller than $L$ down to the lowest temperature we are dealing with here.
442: 
443: 
444: For the 80$\mu m$ sample, on the other hand, a crossover to the short-sample
445: regime is realized in the temperature interval covered in our experiments. The
446: crossover can be seen by studying the data depicted in Figs. 3, and 4. Fig. 3
447: shows the raw $R(V)$ data measured at different temperatures. Fig.~4 show
448: these data plotted according to the short-sample formula
449: (Eq.~(\ref{short-scaling})) in the left panel,  and  according to the long-sample scheme
450: (Eq.~(\ref{R-scaling-1}) with $p=2$) in the right panel. The crossover temperature is the
451: temperature below which the $R(F,T)$ data can be scaled as a function of
452: $V/T$, and above which it scales as $F/T^{2}$. Comparing between the left and
453: the right panels of Fig.~4 it is evident that this temperature is
454: approximately $0.5K$. Thus, with this sample, the consistency of our 
455: approach can be tested in the two limits. As shown in the left panel of Fig.~4,
456: a good fit to the $\Delta R/R(0)$ data is obtained using
457: Eq.~(\ref{short-scaling}), which involves just the parameter $\gamma$. Note
458: that the best fit $\gamma$ is quite close to the $\gamma$ that one gets from
459: the logarithmic fit to the $R(T)$ data of this sample (Fig.~2). The other
460: limit, which conforms to Eq.~(\ref{R-scaling-1}), also yields a reasonable
461: agreement, with the \textit{same} $\gamma$ as used above, and with
462: $\eta=2.2\cdot10^{55}\frac{1}{Joule^{2}m^{2}}$ (see right panel of Fig.~4).
463: The quality of the fit here is less good than in the 3500$\mu m$ sample,
464: perhaps due to the fact that even at the highest temperature used the sample
465: is not really in the long-sample limit.
466: 
467: 
468: 
469: Finally, using the value of $\eta$ for this sample in Eq.~(\ref{ep-length}),
470: one gets $L_{ep}\simeq$ $34\mu m$ at $1K$ and$\simeq100\mu m$ at
471: $T\approxeq0.28K$ . Since it is plausible to expect that the crossover from
472: the short-sample to long-sample regime should occur when $L_{ep}$ becomes
473: comparable with $L/2$, these numbers are consistent with our picture.
474: 
475: We turn now to discuss the physics that underlies the $\tau_{ep}^{-1}\propto
476: T^{2}$ law for the electron-phonon scattering rate suggested by our analysis.
477: It is important to note that the enhanced electron-phonon inelastic scattering
478: resulted from the inclusions of Au atoms in the indium-oxide matrix: By
479: comparison, an undoped In$_{2}$O$_{3-x}$ sample showed $\tau_{ep}^{-1}$ that,
480: at $T\approx0.5K,$ was more than \textit{three orders of magnitude smaller}
481: \cite{Zvi_1}. Since these 'pure' and the Au-doped samples had otherwise quite
482: similar parameters (their $R_{\square},$ and diffusion constant were the same
483: to within 30\% ), the non-trivial role of the gold in enhancing $\tau
484: _{ep}^{-1}$\ must be considered. It seems likely that this is a manifestation
485: of the Sergeev-Mitin mechanism for electron-phonon scattering in disordered
486: metals. The gold impurities in our samples are heavier than the host atoms,
487: and being inert they are also loosely attached to the indium-oxide lattice.
488: These factors limit their ability to follow the lattice movement, and thus the
489: main assumption of the Sergeev-Mitin mechanism
490: is fulfilled. At the same time, the Au atoms are active as local soft-modes
491: which could be very effective in dephasing the electrons\cite{Imry}.
492: However, being weakly coupled to the lattice, they cannot efficiently dissipate the
493: energy, gained by inelastic collisions with the electrons,
494: to the bath.  Therefore the Au inclusions contribute to dephasing 
495: much more than to energy relaxation. Note indeed that the phase coherence length in these 
496: samples  is dominated by the interaction of the electrons with these local modes\cite{Zvi_1}
497: and it is $\approx0.4\mu m$ at $T=0.3K$ as compared with $L_{ep}\approx$ $100\mu m$.
498: That the dephasing rate exceeds the energy relaxation rate by many orders of
499: magnitude is quite a general property of low temperature transport, which
500: follows from the different$\ $temperature dependencies of energy relaxation
501: processes on one hand and dephasing on the other hand.
502: 
503: 
504: To summarize, we have employed a scaling analysis of non-ohmic resistance
505: curves in order to extract the electron-phonon scattering rate of metallic
506: films. The method makes use of the temperature dependence of the resistivity
507: therefore it is best suited for those cases where the resistance can be used
508: as a sensitive thermometer. Our analysis, which assumes quasi-two dimensional
509: samples may be easily extended to other dimensionalities. In these cases the
510: temperature dependence of the resistance, at sufficiently low temperatures, is
511: dominated by a power low behavior, $R(T)=R_{D}(1-\gamma T^{\nu})$. From here
512: it follows that the scaling function (\ref{scaling-function}) of long samples
513: satisfies the relation:
514: \[
515: T^{-\nu}\Delta\mathcal{R}=\gamma\left[  1-\left(  1+\frac{(eF)^{2}}{\eta
516: T^{p+2}}\right)  ^{\frac{\nu}{p+2}}\right]  .
517: \]
518: Having the equilibrium parameters ($\gamma$ and $\nu$), the experimental data
519: of $R(T,F)$ can be fitted to the above form and both $\eta$ and $p$ can be
520: extracted. The electron phonon length is then deduced from Eq.~(\ref{ep-length}).
521: 
522: We reiterate that in order to apply the scaling approach, the following
523: conditions should be satisfied: (a) The heat transfer from the electrons to
524: the bath is dominated by the electron-phonon collisions; (b) the
525: electron-electron diffusion length should be much smaller than the energy
526: relaxation length.
527: 
528: On the other hand, the scaling approach is insensitive to the inclusion of
529: other ingredients such as two level systems and Kondo impurities as long as
530: they do not serve as additional channels for heat conduction to the bath.
531: Furthermore, for long samples the quality of the contacts is of minor
532: importance, since the amount of heat transferred by the electrons through the
533: contacts is anyhow negligible. For short samples, however, our analysis
534: assumes that the electrons near the contacts are at the bath temperature. This
535: means that the contacts are ideal heat sinks, a caveat that should be borne in
536: mind when using contacts made of a superconducting material. \bigskip\ 
537: 
538: The authors gratefully acknowledge discussions with I.~Aleiner and Y.~Imry.
539: This work has been supported in part by the Israel Science Foundation (ISF), 
540: and by the German-Israel Foundation (GIF).   
541: 
542: \begin{thebibliography}{99}                                                                                               %
543: 
544: 
545: \bibitem {Schmid73}A.~Schmid, Z. Phys. \textbf{259}, 421 (1973).
546: 
547: \bibitem {Pippard55}A.~B.~Pippard, Philos. Mag. \textbf{46}, 1104, (1955).
548: 
549: \bibitem {Reizer86}M.Yu.~Reizer and A.V.~Sergeev, Zh. Eksp. Teor. Fiz.
550: \textbf{90}, 1056 (1986) [Sov. Phys. JETP \textbf{63}, 616 (1986).
551: 
552: \bibitem {Gershenson01}M.~E.~Gershenson, D.~Gong, T.~Sato, B.~S.~Karasik, and
553: A.~V.~Sergeev, App. Phys. Lett. \textbf{79}, 2049 (2001).
554: 
555: \bibitem {Karvonen04}J.~T.~Karvonen, L.~J.~Taskinen, and I.~J.~Maasilta, Phys.
556: Stat. S0l \textbf{1}, 2799 (2004).
557: 
558: \bibitem {Kivinen04}P. Kivinen, M. Prunnila, A. Savin, P. Torma, J. Pekola,
559: and J. Ahopelto, Phys. Stat. Sol. \textbf{1} 2848 (2004).
560: 
561: \bibitem {Roukes85}M.~L.~Roukes, M.~R. Freeman, R.~S.~Germain,
562: R.~C.~Richardson, and M.~B.~Ketchen, Phys. Rev. Lett. \textbf{55}, 422 (1985).
563: 
564: \bibitem {Wennberg86}A.~K.~Wennberg, S.~N.~Ytterboe, C.~M.~Gould,
565: H.~M.~Bozler, J.Klem, and H.~Morkoc, Phys. Rev. B \textbf{34}, 4409 (1986).
566: 
567: \bibitem {Eshternach92}P.~M.~Eshternach, M.~R.~Thoman, C.~M.~Gould, and
568: H.~M.~Bozler, Phys. Rev. B \textbf{46}, 10339 (1992).
569: 
570: \bibitem {Wellstood94}F.~C.~Wellstood, C.~Urbina, and J.~Clarke, Phys. Rev. B
571: \textbf{49}, 5942 (1994).
572: 
573: \bibitem {Bergmann90}G.~Bergmann, W.~Wei, Y.~Zou, and R.~M.~Mueller, Phys.
574: Rev. B \textbf{41}, 7386 (1990);
575: 
576: \bibitem {DiTusa92}J.~F.~DiTusa, K.~Lin, M.~Park, M.~S.~Isaacson, and
577: J.~M.~Parpia, Phys. Rev. Lett. \textbf{68}, 1156 (1992);
578: 
579: \bibitem {Watson95}P.~W.~Watson III and D.~G.~Naugle, Phys. Rev. B
580: \textbf{51}, 685 (1995).
581: 
582: \bibitem {Wu98}C.~Y.~Wu, W.~B.~Jian, and J.~J.~Lin, Phys. Rev. B \textbf{57},
583: 11232 (1998).
584: 
585: \bibitem {Belitz95}D.~Belitz and M.~N.~Wybourne, Phys. Rev. B \textbf{51}, 689 (1995).
586: 
587: \bibitem {Jan05}W.~Jan, G.~Y.~Wu, and H,-S. Wei, Physica Scripta \textbf{71},
588: 552 (2005).
589: 
590: \bibitem {Sergeev00}A.~Sergeev and V.~Mitin, Phys. Rev. B \textbf{61}, 6041 (2000).
591: 
592: \bibitem {Zvi_1}Z.\ Ovadyahu, Phys. Rev. \textbf{B63}, 2054 (2000).
593: 
594: \bibitem {Anderson79}P.~W.~Anderson, E. Abrahams, and T.~V.~Ramakrishnan,
595: Phys. Rev. Lett. \textbf{43}, 718 (1979).
596: 
597: \bibitem {Arai83}M.~R.~Arai, Appl. Phys. Lett. \textbf{42}, 15 (1983).
598: 
599: \bibitem {Nag_1}K.\ E.\ Nagaev, Phys. Lett. A, \textbf{169},103(1992);
600: K.\ E.\ Nagaev, Phys. Rev. B,\textbf{52}, 4740, (1995).
601: 
602: \bibitem {Kozub}V. I. Kozub and A. M. Rudin, Phys. Rev. \textbf{B52}, 7853 (1995).
603: 
604: \bibitem {Brg_1}G.\ Bergmann, Phys. Rep., \textbf{107}, 1, (1984).
605: 
606: \bibitem {alt_1}B.\ L.\ Altshuler and A.\ G.\ Aronov ,in
607: \textit{Electron-Electron Interaction in Disordered Systems},edited by
608: A.\ L.\ Efros and M.\ Pollak (North-Holland, Amsterdam 1985).
609: 
610: \bibitem {Mueller94}R.~M.~Mueller, R.~Stasch, and G.~Bergmann, Solid Stat.
611: Comm. \textbf{91}, 255 (1994).
612: 
613: \bibitem {Imry}Y. Imry, Z. Ovadyahu, and A. Schiller, Euresco Conference on
614: Fundamental Problems of Mesoscopic Physics, Granada, September (2003), cond-mat/0312135
615: \end{thebibliography}
616: 
617: 
618: \end{document}