1:
2: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
3: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
4: \documentclass[aps,prl,twocolumn,groupedaddress,dvips]{revtex4}
5: \usepackage[dvips]{graphicx}
6: % You should use BibTeX and apsrev.bst for references
7: % Choosing a journal automatically selects the correct APS
8: % BibTeX style file (bst file), so only uncomment the line
9: % below if necessary.
10: \bibliographystyle{apsrev}
11: \begin{document}
12: % Use the \preprint command to place your local institutional report
13: % number in the upper righthand corner of the title page in preprint mode.
14: % Multiple \preprint commands are allowed.
15: % Use the 'preprintnumbers' class option to override journal defaults
16: % to display numbers if necessary
17: \preprint{}
18: %Title of paper
19: \title{Saturation of Spin-Polarized Current in Nanometer Scale Aluminum Grains}
20:
21:
22: % repeat the \author .. \affiliation etc. as needed
23: % \email, \thanks, \homepage, \altaffiliation all apply to the current
24: % author. Explanatory text should go in the []'s, actual e-mail
25: % address or url should go in the {}'s for \email and \homepage.
26: % Please use the appropriate macro foreach each type of information
27: % \affiliation command applies to all authors since the last
28: % \affiliation command. The \affiliation command should follow the
29: % other information
30: % \affiliation can be followed by \email, \homepage, \thanks as well.
31: %\email[]{dragomir.davidovic@physics.gatech.edu}
32: %\homepage[]{Your web page}
33: %\thanks{}
34: %\altaffiliation{}
35: \author{Y. G. Wei, C. E. Malec, D. Davidovi\'c }
36: \affiliation{School of Physics, Georgia Institute of Technology,
37: Atlanta, GA 30332}
38: %Collaboration name if desired (requires use of superscriptaddress
39: %option in \documentclass). \noaffiliation is required (may also be
40: %used with the \author command).
41: %\collaboration can be followed by \email, \homepage, \thanks as well.
42: %\collaboration{} FG^{as}
43: %\noaffiliation
44: \date{\today}
45: \begin{abstract}
46: We describe measurements of spin-polarized tunnelling via discrete
47: energy levels of single Aluminum grains. In high resistance
48: samples ($\sim G\Omega$), the spin-polarized tunnelling current
49: rapidly saturates as a function of the bias voltage. This
50: indicates that spin-polarized current is carried only via the
51: ground state and the few lowest in energy excited states of the
52: grain. At the saturation voltage, the spin-relaxation rate
53: $T_1^{-1}$ of the highest excited states is comparable to the
54: electron tunnelling rate: $T_1^{-1}\approx 1.5\cdot 10^6 s^{-1}$
55: and $10^7s^{-1}$ in two samples. The ratio of $T_1^{-1}$ to the
56: electron-phonon relaxation rate is in agreement with the
57: Elliot-Yafet scaling, an evidence that spin-relaxation in Al
58: grains is governed by the spin-orbit interaction.
59:
60: \end{abstract}
61:
62: % insert suggested PACS numbers in braces on next line
63: \pacs{73.21.La,72.25.Hg,72.25.Rb,73.23.Hk}
64: % insert suggested keywords - APS authors don't need to do this
65: %\keywords{}
66: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
67: \maketitle
68:
69: \section{introduction}
70:
71: Electron tunnelling through single nanometer scale metallic grains
72: at low temperatures can display a discrete energy level
73: spectrum.~\cite{ralph} Tunnelling spectroscopy of the energy
74: spectra have led to numerous discoveries, including Fermi-Liquid
75: coupling constants between quasiparticles,~\cite{agam} spin-orbit
76: interactions,~\cite{davidovic,petta} and superconducting
77: correlations in zero-dimensional systems.~\cite{black} Some
78: information regarding the spin of an electron occupying a discrete
79: level can be obtained using spin-unpolarized tunnelling, such as
80: spin-multiplicity and electron g-factors.~\cite{ralph}
81:
82: In this letter we report on spin-polarized tunnelling via discrete
83: energy levels of single aluminum grains. Spin-polarized electron
84: transport permits studies of spin relaxation and spin
85: dephasing.~\cite{johnson,jedema1} By comparison, spin-unpolarized
86: spectroscopy is suitable for the studies of energy relaxation in
87: the grains.~\cite{agam,ralph} Since spin-relaxation times are
88: generally many orders of magnitude longer than energy relaxation
89: times, spin-unpolarized spectroscopy is not an easy tool to study
90: spin-relaxation in the grains and spin-polarized tunnelling is
91: needed. We find that some electron spin-relaxation times in Al
92: grains are exceptionally long compared to bulk Al with similar
93: disorder, on the order of $\mu s$.
94:
95: Spin-polarized transport via metallic grains has recently
96: generated a lot of theoretical
97: interest.~\cite{braun,weymann,vandermolen,wetzels,cottet1} In
98: addition, there is a major effort to study nano-spintronics using
99: carbon-nanotubes; see Ref.~\cite{cottet} and references therein.
100: Spin-coherent electron tunnelling via nanometer scale normal
101: metallic grains has been confirmed in arrays~\cite{zhang1,ernult}
102: and in single grains.~\cite{bernard} However, the electron
103: spin-relaxation time $T_1$ in a metallic grain has not been
104: reported yet.
105:
106: \section{Sample fabrication}
107:
108: Our samples are prepared by electron beam lithography and shadow
109: evaporation, similar to the technique described
110: previously.~\cite{davidovic} First we define a resist bridge
111: placed 250 nm above the Si wafer; this bridge acts as a mask. Next
112: (Fig.~\ref{fig1}-A), we deposit 11 nm permalloy (Py =
113: Ni$_{0.8}$Fe$_{0.2}$) onto oxidized silicon substrate at $4\cdot
114: 10^{-7}$ Torr base pressure, measured near the gate valve, along
115: the direction indicated by the arrow. Then we rotate the sample by
116: 36 degrees without breaking the vacuum and deposit 1.2 nm of
117: Al$_2$O$_3$ by reactive evaporation of Al,~\cite{davidovic} at a
118: rate of 0.35 nm/s, at an oxygen pressure of $2.5\cdot 10^{-5}$
119: Torr. Now, oxygen flow is shut down. When pressure decreases to
120: the $10^{-7}$ Torr range, we deposit a 0.6 nm thick film of Al, as
121: shown in Fig.~\ref{fig1}-B. Al forms isolated grains with a
122: typical diameter of 5 nm. The grains are displayed by the scanning
123: electron microscope (SEM) image in Fig.~\ref{fig1}-D. Finally we
124: deposit another 1.2 nm layer of Al$_2$O$_3$ by the reactive
125: evaporation and top it of by an 11 nm thick film of Py
126: (Fig.~\ref{fig1}-C).
127: \begin{figure}%[p]
128: \includegraphics[width=0.45\textwidth]{fig1.eps}
129: \caption{A, B, C: Sample fabrication steps. D. Image of Al grains.
130: E, F: Image of a typical sample. G, H: I-V curves at the base
131: temperature.~\label{fig1}}
132: \end{figure}
133: We make many samples on the same silicon wafer, and vary the
134: overlap from 0 to 50 nm and select the devices with the highest
135: resistance, as they have the smallest overlap. Figs.~\ref{fig1}-E
136: and F show SEM images of a typical device.
137:
138: \section{Discrete energy levels}
139:
140: Transport properties of the samples at low temperatures were
141: measured using an Ithaco current amplifier. The samples were
142: cooled down to $\approx 0.035K$ base temperature. The sample leads
143: were cryogenically filtered to reduce the electron temperature
144: down to $\approx 0.1K$.
145:
146: The majority of samples ($>$80\%) exhibit Coulomb Blockade at low
147: temperature. About 150 samples were measured at 4.2K and 16
148: samples at $0.035 K$. In this paper we describe two samples. The
149: I-V curve of two samples are shown in Fig.~\ref{fig1}-G and H. The
150: tunnelling current increases in discrete steps as a function of
151: bias voltage, corresponding to discrete electron-in-a-box energy
152: levels of the grain.
153:
154:
155: In sample 1, the average electron-in-a-box level spacing caused by
156: electron geometric confinement is $\delta\approx 0.8meV$, which
157: corresponds to diameter of $D\approx 6nm$ assuming a spherical Al
158: grain. The average current step is ${\overline I}\approx 0.47pA$.
159: We make a connection with the tunnelling rates from the leads to
160: the grain and the measured current response. The tunnel junctions
161: are highly asymmetric, and therefore one of the tunnelling rates
162: is much smaller than the other, and thus rate limiting. Throughout
163: this paper, we choose the rate limiting step to be across the left
164: junction, corresponding to the tunnelling rate $\Gamma_L$.
165: Therefore, our measured current corresponds to the average
166: tunnelling-in rate of $\overline{\Gamma}_L={\overline
167: I}/2|e|\approx 1.5\cdot 10^6 s^{-1}$. Similarly, in sample 2,
168: $\delta\approx 2.7 meV$, $D\approx 4nm$, and
169: ${\overline\Gamma_L}\approx 9.6\cdot 10^6 s^{-1}$.
170:
171:
172: The
173: spin-conserving energy-relaxation in Al grains takes place by
174: phonon emission with the relaxation rate~\cite{agam}
175: \begin{equation}
176: {\tau_{e-ph}^{-1}(\omega)}=\left(\frac{2}{3}E_F\right)^2\frac{\omega^3\tau_e\delta}{2\rho\hbar^5
177: v_S^5}, \label{eph}
178: \end{equation}
179: where $E_F=11.7 eV$ is the Fermi energy, $\omega$ is the energy
180: difference between the initial and the final state, $\rho=2.7
181: g/cm^3$ is the ion-mass density, and $v_s=6420 m/s$ is the sound
182: velocity. We obtain $\tau_{e-ph}^{-1}(\delta)\approx 1.6\cdot 10^9
183: s^{-1}$ and $4.1\cdot 10^{10} s^{-1}$ in samples 1 and 2,
184: respectively. Sample 2 has significantly larger relaxation rate
185: because of the larger level spacing. Since the tunnelling rates in
186: our samples are $\sim 10^6s^{-1}$, if the grain is excited by
187: electron tunnelling in and out, it will instantly relax to the
188: lowest energy state accessible by spin-conserving transitions.
189:
190: As shown by Fig.~\ref{fig2}, the energy levels exhibit Zeeman
191: splitting as a function of an applied magnetic field. In sample 1,
192: the I-V curve probes the same energy spectrum at negative and
193: positive bias voltage. This is evident from the equivalence of the
194: magnetic field dependencies at negative and positive bias. The
195: lowest tunnelling threshold is two fold degenerate at zero
196: magnetic field, showing that $N_0$, the number of electrons on the
197: grain before tunnelling in, is even. The conductance peaks are
198: similar in magnitude at negative bias, because the first
199: tunnelling step, in which an electron tunnels in to the grain
200: through the higher resistance junction, is rate limiting. At
201: positive bias, the first conductance peak is much larger than the
202: subsequent conductance peaks, because the first tunnelling step
203: takes place via the lower resistance junction, and the rates are
204: limited by the electron discharge process across the high
205: resistance junction.
206:
207: In sample 1, the first two peaks split corresponding to g-factors:
208: $g=1.83\pm 0.05$ and $1.95\pm 0.05$. Slight reduction of the
209: g-factors from $2$ indicates spin-orbit interaction in
210: Al.~\cite{ralph} The avoided level crossings clearly are resolved
211: in Fig.~\ref{fig2}, near points $(-11.5mV,5T)$ and
212: $(-13mV,11.5T)$. The corresponding avoided crossings at positive
213: bias are located near $(13.5mV,5T)$ and $(15.5mV,11.5T)$,
214: respectively. In the regime where g factors are slightly reduced,
215: the spin-orbit scattering rate ($\tau_{SO}^{-1}$) can be obtained
216: from the avoided crossing energies $\Delta_{SO}\approx
217: 0.1meV$.~\cite{adam} Theory predicts that
218: $\tau_{SO}\approx\hbar\delta/\pi\Delta_{SO}^2$,~\cite{adam} within
219: a factor of two. Thus, we obtain $\tau_{SO}^{-1}\approx 5.5\cdot
220: 10^{10} s^{-1}$. By the Elliot-Yafet relation,~\cite{yafet}
221: $\tau_{SO}^{-1}$ is related to the elastic scattering rate
222: $\tau_e^{-1}$: $\tau_{SO}^{-1}=\alpha \tau_e^{-1}$. Assuming
223: ballistic grain, $\tau_{e}^{-1}\approx v_F/D=3.4\cdot 10^{14}
224: sec^{-1}$. We obtain $\alpha\approx 1.6\cdot 10^{-4}$, in
225: excellent agreement with $\alpha\approx 10^{-4}$ in Al thin
226: films.~\cite{jedema3}
227:
228:
229: \begin{figure}%[p]
230: \includegraphics[width=0.45\textwidth]{fig2.eps} \caption{A, B: Differential conductrance (gray) versus bias
231: voltage and the applied magnetic field in sample 1 at the base
232: temperature.~\label{fig2}}
233: \end{figure}
234:
235: \section{Spin-polarized tunnelling}
236:
237: Now we discuss magnetoresistance from the spin-polarized
238: tunnelling. In the magnetic field range of $\pm 50 mT$,
239: approximately 90\% of the samples do not display any of the
240: tunnelling magnetoresistance effect (TMR) . By contrast, we tested
241: about 10 tunnelling junctions without the embedded grains and with
242: similar resistance (empty junctions) at 4.2 K. All of the empty
243: junctions exhibit a significant TMR in this field range,
244: comparable to $10\%$. Approximately one half of the empty
245: junctions display a simple spin-valve effect. So, the absence of
246: TMR for electron tunnelling via grains shows that the
247: spin-dephasing rate $T_2^{-1}$ in 90\% of the samples must be much
248: larger than the tunnelling rate.
249:
250: Nevertheless, approximately 10\% of the samples with embedded
251: grains display significant TMR, so the dephasing must be weak,
252: e.g. $T_2^{-1}$ must be smaller than or comparable to the
253: tunnelling rate in these samples. $T_2$ variation among different
254: samples could be explained by magnetic defects, such as
255: paramagnetic impurities from the Py layer. Paramagnetic impurities
256: are common sources of dephasing.~\cite{bergman} The defects would
257: be located on the grain surface, since bulk Al does not support
258: paramagnetism. Since the number of atoms on the surface is
259: relatively small ($\sim$1000), we could occasionally obtain a
260: sample free of impurities. More insight into the nature of $T_2$
261: in this device will require a more in depth theoretical study.
262:
263: A majority of the samples with nonzero TMR show positive TMR near
264: the Coulomb-Blockade conduction threshold; only about 30\% of the
265: samples show negative TMR. The sign of TMR in quantum dots is
266: determined by the interplay between charging effects and
267: spin-accumulation.~\cite{barnas,ernult} For any given sample, the
268: data in this paper correspond to the voltage range within the
269: first step of the Coulomb staircase. In this range the sign of TMR
270: is found to be constant as expected.
271:
272: TMR in our devices usually does not display a simple spin-valve
273: effect. We believe this is because there are spin-dependent
274: interactions inside the grain that induce a complicated TMR even
275: when the magnetic transitions in the drain and source leads are
276: sharp and as expected. For example, a rotation of stray magnetic
277: field acting on the grain will alter the direction of the
278: spin-quantization axis in the grain, thereby changing the
279: conductance~\cite{braun}. A rotation or a switch of a remote
280: domain can change the tunnelling current through the grain via the
281: magnetic field generated by the domain. Similarly, the orientation
282: of the nuclear spin in the grain can change the quantization axes
283: via the hyperfine interaction.
284: \begin{figure}%[p]
285: \includegraphics[width=0.45\textwidth]{fig3.eps} \caption{A-F: Spin-valve effect in current versus
286: applied magnetic field in two samples at the base temperature. The
287: current magnitude is reduced in the antiparallel
288: state.~\label{fig3}}
289: \end{figure}
290:
291: We select only those samples that display a simple spin-valve TMR
292: effect, which is shown in Figs.~\ref{fig3}. Fig.~\ref{fig3}-A is
293: the TMR of sample 1 at a bias voltage corresponding to the second
294: current plateau. TMR is barely resolved in this case, since the
295: current changes by only about 40fA. We do not have good data to
296: display TMR at the first current plateau. By comparison,
297: Figs.~\ref{fig3}-B and C display TMR at bias voltage where the
298: number of electron-in-a-box levels energetically available for
299: tunnelling-in are approximately 19 and 48, respectively. To
300: facilitate comparisons, the current intervals on the vertical axes
301: in Figs.~\ref{fig3} A-C and D-F have equal lengths.
302:
303: The main observation in this letter is that $\Delta
304: I=I_{\uparrow\uparrow}-I_{\uparrow\downarrow}$ is nearly constant
305: with current above a certain current. There is hardly any increase
306: in $\Delta I$ between Figs.~\ref{fig3} B and C and between
307: Figs.~\ref{fig3} E and F. This behavior is shown in more detail
308: Fig.~\ref{fig4}-A and B, which displays $\Delta I$ versus bias
309: voltage. $\Delta I$ versus negative bias voltage in sample 1 is
310: fully saturated at the third current plateau; at the second
311: current plateau, $\Delta I$ is already at one half of the
312: saturation value. Similarly, in sample 2 $\Delta I$ reaches
313: saturation at the second current plateau. Our samples should be
314: contrasted with ordinary ferromagnetic tunnelling junctions, where
315: $\Delta I$ is proportional to the current over a significantly
316: wider range of bias voltage~\cite{moodera1,zhang2}.
317:
318: \section{Interpretation of the results}
319:
320: In coulomb-blockade samples containing magnetic leads, the
321: electrochemical potential difference between the island and leads
322: can jump when the magnetization in one of the leads changes
323: direction~\cite{vandermolen}. This can lead to a sudden shift in
324: energy levels, producing a jump in current that is constant as a
325: function of bias voltage. The shift in energy levels is seen as a
326: discontinuity near zero magnetic field in Fig. 2, and is $\sim 0.1
327: mV$.
328:
329: To show that the electrocemical shift is not responsible for the
330: saturation of the spin-polarized current with voltage in our
331: sample, we performed other measurements by sweeping the magnetic
332: field both on and between the current plateaus, coming up with
333: similar values for the electrochemical shift. The shift is lower
334: than the average level spacing of 0.8 meV and 2.7 meV for sample 1
335: and sample 2 respectively. Therefore, since we measured
336: magnetoresistance in the middle of the current plateau, the
337: threshold voltage shift should not effect our measurements of the
338: saturation in $\Delta I$.
339:
340: %This sets up the long argument
341: To explain $I_{\uparrow\uparrow}-I_{\uparrow\downarrow}=const$, we
342: must discuss the relative magnitudes of three rates:
343: $\tau_{e-ph}^{-1}$, the rate of energy relaxation from excited to
344: lower energy states by spin-conserving phonon emission;
345: $\Gamma_L$, the rate electrons tunnel into the grain; and
346: $T_1^{-1}$, the rate of transitions between levels that result in
347: an electron flipping its spin orientation. $\tau_{e-ph}^{-1}$ is
348: obtained theoretically, the measured I-V spectrum fixes the
349: tunnelling rate, and $T_1^{-1}$ is obtained from the saturation in
350: $I_{\uparrow\uparrow}-I_{\uparrow\downarrow}$ with bias voltage.
351:
352: Finally we must deduce the relative magnitude of $T_1^{-1}$. The
353: rate of spin-flip transitions is expected to be significantly
354: smaller than $\tau_{e-ph}^{-1}$~\cite{yafet}. In this case the
355: ground state would not necessarily be accessible by energy
356: relaxation. The grain could remain in an excited, spin-polarized
357: state, as sketched in Fig.~\ref{fig4}-C. These spin-polarized
358: excited states are responsible for spin accumulation in the
359: antiparallel magnetic configuration of the leads. If the
360: relaxation rates for the spin-flip transitions are much smaller
361: than the tunnelling rate, then various spin-polarized states would
362: have similar probabilities, which are determined by the tunnelling
363: rates. In the antiparallel configuration of the leads, the
364: probabilities of the excitations with spin up would be enhanced by
365: $1+ P$ and probabilities of the excitations with spin down would
366: be suppressed by $1-P$, where $P$ is the spin-polarization in the
367: leads. In the parallel configurations, the probabilities of the
368: excitations with spin up and spin down are the same. In this
369: regime, $I_{\uparrow\uparrow}-I_{\uparrow\downarrow}$ is
370: proportional to the current, similar to the usual ferromagnetic
371: tunnelling junctions.
372:
373: It is reasonable to expect that the spin-flip rate
374: $T_1^{-1}(\omega)$ increases rapidly with energy difference
375: $\omega$ between the initial and the final state~\footnote{In bulk
376: metals the spin-orbit scattering rate increases rapidly with
377: electron excitation energy}. If $T_1^{-1}(\omega)$ exceeds the
378: tunnelling rate above some $\omega$, then the excitations with
379: energy $>\omega$ will occur with a reduced probability in the
380: ensemble of states generated by tunnelling in and out. Thus
381: $\Delta I$ is limited by tunnelling via the ground state and those
382: low lying spin-polarized states where $T_1^{-1}(\omega)<\Gamma_L$.
383: $\Delta I$ versus bias voltage approaches saturation approximately
384: when $T_1^{-1}(\omega)=\Gamma_L$, where $\omega$ is the highest
385: excitation energy in the ensemble of spin-polarized states
386: generated by tunnelling in and out: $\omega\approx\delta
387: \frac{I}{|e|\Gamma_L}$. This is how we determine the
388: spin-relaxation time $T_1(\omega)$ at an energy $\omega$ in a
389: given sample.
390:
391: \begin{figure}%[p]
392: \includegraphics[width=0.45\textwidth]{fig4.eps}
393: \caption{A and B: $\Delta
394: I=|I_{\uparrow\uparrow}-I_{\uparrow\downarrow}|$ versus bias
395: voltage in samples 1 and 2, respectively, at the base temperature.
396: The numbers near the circles indicate how many doubly degenerate
397: electron-in-a-box levels are available for tunnelling in. C:
398: Possible spin-polarized electron configurations caused by electron
399: tunnelling in and out, before an electron tunnels in, at the
400: second current plateau, for $N_0$ even.~\label{fig4}}
401: \end{figure}
402:
403: In sample 1, $\Delta I$ is at 50\% of the saturation value at the
404: second current plateau, and $\Delta I$ is saturated at the third
405: current plateau. At the second current plateau, the
406: spin-relaxation rate of the highest energy excited state generated
407: by tunnelling must be close to the tunnelling rate. Since the
408: spin relaxation is very rapid in configurations more than
409: $3\delta$ above the ground state, and $N_0$ is even as noted
410: above, the grain spends most of the time among the five
411: configurations shown in Fig.~\ref{fig4}-C: $N_0$, $N_0^+$,
412: $N_0^-$, $N_0^{++}$, and $N_0^{--}$. The highest energy
413: spin-polarized states are $N_0^{++}$ and $N_0^{--}$. Thus,
414: $T_1^{-1}(3\delta)\approx \Gamma_L=1.5\cdot 10^6 s^{-1}$. In
415: sample 2, this analysis leads to $T_1^{-1}(2\delta)\approx 10^7
416: s^{-1}$.
417:
418: Now we discuss the origin of spin relaxation and its rapid
419: enhancement with the energy difference. Note that the rate of
420: spin-conserving transitions in Eq.~\ref{eph} increases as
421: $\omega^3$. We suggest that the electron-phonon transition rates
422: without and with spin flip scale by the Elliot-Yafet relation:
423: $T_1^{-1}(\omega)=\alpha'\tau_{e-ph}^{-1}(\omega)$. This scaling
424: would certainly explain the rapid increase in spin-relaxation rate
425: with excitation energy. In metallic films, it is well established
426: that the Elliot-Yafet scaling applies for both elastic and
427: inelastic scattering processes, with $\alpha\approx
428: \alpha'$.~\cite{jedema3}
429:
430:
431: In sample 1, Eq.~\ref{eph} leads to
432: $\tau_{e-ph}^{-1}(3\delta)\approx 4\cdot 10^{10} s^{-1}$. Since
433: $T_1^{-1}(3\delta)\approx 1.5\cdot 10^6 s^{-1}$, we obtain
434: $\alpha'\approx 0.4\cdot 10^{-4}$. Similarly, in sample 2,
435: $\tau_{e-ph}^{-1}(2\delta)\approx 3.3\cdot 10^{11} s^{-1}$ and we
436: obtain $\alpha'\approx 0.3\cdot 10^{-4}$. $\alpha'$ agrees with
437: $\alpha\approx 1.5\cdot 10^{-4}$ obtained earlier, within an order
438: of magnitude. So the ratio of $\tau_{e-ph}$ and $T_1$ is in
439: agreement with the Elliot-Yafet scaling. This is an evidence that
440: the spin-flip transitions in Al grains are driven by the
441: spin-orbit interaction. By this relaxation mechanism, the spin of
442: an electron on the grain is coupled to the phonon continuum via
443: the spin-orbit interaction. An electron in an excited
444: spin-polarized state relaxes by an emission of a phonon, which has
445: an angular momentum equal to the difference between the initial
446: and final electron spin.
447:
448: \section{Conclusion}
449:
450: In summary, we have observed spin-coherent electron transport via
451: discrete energy levels of single Al grains. Spin polarized current
452: saturates quickly as a function of bias voltage, which
453: demonstrates that the ground state and the lowest excited states
454: carry spin polarized current. Higher excited states have a
455: relaxation time shorter than the tunnelling time and they do not
456: carry spin-polarized current. The spin-relaxation time of the
457: low-lying excited states is $T_1\approx 0.7\mu s$ and $0.1\mu s $
458: in two samples. Finally, the ratio of the spin-flip transition
459: rate and the electron-phonon relaxation rate is in quantitative
460: agreement with the Elliot-Yafet scaling ratio, an evidence that
461: the spin-relaxation transitions are driven by the spin-orbit
462: interaction.
463:
464: This work was performed in part at the Georgia-Tech electron
465: microscopy facility. We thank Matthias Braun and Markus Kindermann
466: for consultation. This research is supported by the DOE grant
467: DE-FG02-06ER46281 and David and Lucile Packard Foundation grant
468: 2000-13874.
469:
470: \bibliography{career1}
471:
472: \end{document}
473: