cond-mat0405528/rmp.tex
1: %documentstyle[rmp,aps,floats,psfig,preprint]{revtex} 
2: \documentclass[rmp,aps,floatfix,twocolumn,unsortedaddress]{revtex4}
3: %\documentclass[rmp,aps,floatfix,preprint,unsortedaddress]{revtex4}
4: \usepackage{epsfig}
5: \begin{document}
6: 
7: 
8: %\twocolumn[
9: %\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
10: 
11: \title{Spintronics: Fundamentals and applications}
12: 
13: \author{Igor \v{Z}uti\'{c}\footnote{Electronic address: igorz@physics.umd.edu.
14: Present address: Center for Computational Materials Science,
15: Naval Research Laboratory, Washington, D.C. 20375, USA}}
16: \affiliation{Condensed Matter Theory Center,
17: Department of Physics, University of Maryland at College
18: Park, College Park, Maryland 20742-4111, USA}
19: 
20: \author{Jaroslav Fabian\footnote{Electronic address: 
21: jaroslav.fabian@uni.graz.at}}
22: \affiliation{Institute for Theoretical Physics, Karl-Franzens University,
23: Universit\"atsplatz 5, 8010 Graz, Austria }
24: 
25: \author{S. Das Sarma}
26: \affiliation{Condensed Matter Theory Center,
27: Department of Physics, University of Maryland at College
28: Park, College Park, Maryland 20742-4111, USA}
29: 
30: 
31: %\maketitle
32: \begin{abstract}
33: {Spintronics, or spin electronics, involves the study of active control
34: and manipulation of spin degrees of freedom in solid-state systems.
35: This article reviews
36: the current status of this subject, including both recent advances
37: and well-established results.
38: The primary focus is on the basic physical principles underlying the
39: generation of carrier spin polarization, spin dynamics, and spin-polarized
40: transport in semiconductors and metals.
41: Spin transport differs from charge transport in that spin is
42: a nonconserved quantity in solids due to spin-orbit and hyperfine coupling.
43: The authors discuss in detail spin decoherence mechanisms in metals and 
44: semiconductors. Various theories of spin injection
45: and spin-polarized transport are applied to hybrid structures relevant to
46: spin-based devices and fundamental studies of materials properties.
47: Experimental work is reviewed with the emphasis on projected applications, 
48: in which
49: external electric and magnetic fields and illumination by light
50: will be used to control spin and charge dynamics to create new functionalities 
51: not feasible or ineffective with conventional electronics.}
52: \end{abstract}
53: \maketitle
54: 
55: \setcounter{tocdepth}{4}
56: \tableofcontents
57: 
58: 
59: \section{\label{sec:I} Introduction}
60: %===========================================
61: 
62: 
63: \subsection{\label{sec:IA} Overview}
64: %-------------------------------------------
65: 
66: Spintronics is a multidisciplinary field
67: whose central theme is the
68: active manipulation of spin degrees of freedom in 
69: solid-state systems.\footnote{While there are proposals for spintronic
70: devices based on deoxyribonucleic acid (DNA) molecules
71: \cite{Zwolak2002:APL}, the whole device, which includes 
72: electrodes, voltage/current source, etc., is still a solid-state system.}
73: In this article the term spin stands for either the spin of a single
74: electron $\bf s$, which can be detected by its magnetic moment 
75: $-g \mu_B {\bf s}$
76: ($\mu_B$ is the Bohr magneton and $g$ is the electron $g$ factor, in a
77: solid generally different from the free electron value of 
78: $g_0=2.0023$),
79: or the average spin of an ensemble of electrons, manifested by magnetization.
80: The control of spin is then a control of either the population and
81: the phase of the spin of an ensemble of particles, or a coherent
82: spin manipulation of a single or a few-spin system.
83: The goal of spintronics is to understand the interaction
84: between the particle spin and its solid-state environments
85: and to make useful devices using the acquired knowledge. 
86: Fundamental studies of spintronics include investigations of spin 
87: transport in electronic materials, as well as 
88: understanding spin dynamics and spin relaxation. 
89: Typical questions that are posed are 
90: (a) what is an effective way to polarize
91: a spin system?
92: (b) how long is the system able to remember its spin
93: orientation? (c) how 
94: can spin be detected? 
95: 
96: Generation of spin polarization usually means creating a nonequilibrium
97: spin population. This can be achieved in several ways. While
98: traditionally spin has been oriented using optical techniques in which 
99: circularly polarized photons transfer their angular momenta
100: to electrons, for device applications electrical spin injection
101: is more desirable. In electrical spin injection a 
102: magnetic electrode is connected
103: to the sample. When the current drives spin-polarized electrons
104: from the electrode to the sample, nonequilibrium spin accumulates
105: there. The rate of spin accumulation depends
106: on spin relaxation, 
107: the
108: process of bringing the accumulated spin population 
109: back to equilibrium. There are several relevant mechanisms
110: of spin relaxation, most involving spin-orbit coupling to  provide
111: the spin-dependent potential, in combination with momentum
112: scattering providing a randomizing force. Typical time scales
113: for spin relaxation in electronic systems are 
114: measured in 
115: nanoseconds, 
116: while the range is from pico to microseconds.
117: Spin detection, also 
118: part of a generic spintronic scheme, typically
119: relies on sensing the changes in the signals caused by the
120: presence of nonequilibrium spin in the system. The common
121: goal in many spintronic devices is to maximize the 
122: spin detection sensitivity  to the point it detects 
123: not the spin itself, but changes in the spin states.  
124: 
125: Let us illustrate the generic spintronic scheme on a prototypical
126: device, the Datta-Das spin field effect
127: transistor (SFET) \cite{Datta1990:APL}, depicted in Fig.~\ref{fig:DD}. 
128: The scheme shows the structure of the usual FET, 
129: with a drain, a source, a narrow channel, and a gate for controlling the 
130: current.
131: The gate either allows 
132: the current to flow (ON) or does not (OFF). The
133: spin transistor is similar in that the result is also a
134: control of the charge current through the narrow channel. 
135: The difference, however, is in the physical realization
136: of the current control. In the Datta-Das SFET the source
137: and the drain are ferromagnets
138: acting as the injector and detector of the electron spin. 
139: The drain injects electrons with spins parallel to the
140: transport direction. The electrons are transported ballistically
141: through the channel. When they arrive at the drain, their spin is 
142: detected. In a simplified picture, the electron can enter 
143: he drain (ON) if its spin 
144: points in the same direction as the spin of the drain. Otherwise
145: it is scattered away (OFF). The role of the gate is to generate an
146: effective magnetic field 
147: (in the direction of $\bf\Omega$ in Fig.~\ref{fig:DD}), 
148: arising from the 
149: spin-orbit coupling in the substrate material,
150: from the confinement geometry of the transport channel, and the
151: electrostatic potential of the gate.
152: This 
153: effective magnetic field causes
154: the electron spins to precess. By modifying the voltage, one can cause
155: the precession to lead to either parallel or antiparallel 
156: (or anything between) electron spin at the drain, effectively
157: controlling the current. 
158: 
159: \begin{figure}
160: \centerline{\psfig{file=zutic_fig01.eps,width=1\linewidth,angle=0}}
161: \caption{Scheme of the Datta-Das spin field-effect transistor (SFET). 
162: The source (spin injector) and the drain (spin detector) are ferromagnetic 
163: metals or semiconductors,
164: with parallel magnetic moments. 
165: The injected spin-polarized electrons with wave vector ${\bf k}$ 
166: move ballistically
167: along a quasi-one-dimensional channel formed by, for example, 
168: an InGaAs/InAlAs heterojunction in a plane normal to ${\bf n}$. Electron spins
169: precess about the precession vector ${\bf \Omega}$, which arises from  
170: spin-orbit coupling and which is defined by the structure and 
171: the materials properties of the channel. The magnitude
172: of ${\bf \Omega}$ is tunable by the gate voltage $V_G$ 
173: at the top of the channel.  The current is large if the electron 
174: spin at the drain points in the initial direction (top row), for
175: example, if the precession period is much larger than the 
176: time of flight, and small if the direction is reversed (bottom).
177: }
178: \label{fig:DD}
179: \end{figure}
180: 
181: Even though the name {\it spintronics} is rather novel,\footnote{The term
182: was coined by S. A. Wolf in 1996, as a name for a 
183: DARPA initiative for novel magnetic materials and devices.} 
184: contemporary research in spintronics relies closely on a long
185: tradition of results obtained in diverse areas of physics (for
186: example, magnetism, semiconductor physics, superconductivity, optics,
187: and mesoscopic physics) and establishes new connections between its
188: different subfields \cite{Rashba2002:JS,Zutic2002:JS}.
189: We review here both well-established results and the physical
190: principles relevant to the present and future applications. 
191: Our strategy is to give a comprehensive view of what
192: has been accomplished, focusing in detail on a few
193: selected topics that we believe are representative for
194: the broader subject within which they appear. For example, 
195: while discussing the generation of spin polarization, we survey
196: many experimental and theoretical studies of both optical orientation
197: and electrical spin injection and present a detailed and
198: self-contained formalism of electrical spin injection. Similarly,
199: when we discuss spin relaxation, we give a catalog 
200: of important work, while studying spin relaxation in 
201: the cases of Al and GaAs as representative of the whole field.
202: Finally, in the section on spin devices we give detailed
203: physical principles of several selected devices, such as, 
204: for example, the above-mentioned Datta-Das SFET.
205: 
206: There have been many other reviews written on spintronics, 
207: most focusing on a particular aspect of the field. We divide
208: them here, for an easier orientation, into two groups,
209: those that cover the emerging applications\footnote{Reviews on
210: emerging application include those of 
211: \cite{Wolf2000:IEEE,
212: DasSarma2000:SM,DasSarma2000:IEEE,DasSarma2000:DRC,%
213: DasSarma2001:SSC,DasSarma2001:AS,Zutic2002:JS,Rashba2002:JS,Wolf2001:S,%
214: Oestreich2002:SST,Zutic2002:PROC}.} 
215: and those covering already well-established schemes 
216: and materials\footnote{Established schemes and materials are
217: reviewed by 
218: \cite{Prinz1995:PT,Prinz1998:S,Gregg1997:JMMM,Bass1999:JMMM,%
219: Ansermet1998:JPCM,Gijs1997:AP,Tedrow1994:PR,Daughton1999:JPDAP,Stiles2003:P}.}
220: The latter group, often described as {\it magnetoelectronics} 
221: typically covers paramagnetic and ferromagnetic metals and insulators,
222: which utilize magnetoresistive effects, 
223: realized, for example, as
224: magnetic read heads in computer hard drives, nonvolatile magnetic
225: random access memory (MRAM), and circuit isolators
226: \cite{Wang2002:JAP}. These more established aspects of spintronics
227: have been also addressed in several books\footnote{See, for example, the
228: books of 
229: \cite{Hartmann:2000,Parkin:2002,Shinjo:2002,Maekawa:2002,Hirota:2002,%
230: Levy:2002,Chtchelkanova:2003,Ziese:2001}}
231: and will be discussed in another review,\footnote{In preparation
232: by S. S. P. Parkin for Review Modern Physics.} complementary to
233: ours.
234: 
235: Spintronics also benefits from a large class of emerging materials, such as
236: ferromagnetic semiconductors \cite{Ohno1998:S,Pearton2003:JAP},
237: organic semiconductors \cite{Dediu2002:SSC}, organic ferromagnets 
238: \cite{Pejakovic2002:PRL,Epstein2003:MRS},
239: high temperature superconductors \cite{Goldman1999:JMMM},
240: and carbon nanotubes \cite{Tsukagoshi1999:N,Zhao2002:APL},
241: which can bring novel functionalities to the traditional devices.
242: There is a continuing need for fundamental studies before the
243: potential of spintronic applications is fully realized.
244: 
245: After an overview, Sec.~\ref{sec:I} covers some basic historical and 
246: background material, part of which has already  been extensively covered 
247: in the context of magnetoelectronics
248: and will not be discussed further in this review. Techniques 
249: for generating spin polarization, focusing on optical spin orientation and
250: electrical spin injection, are described in Sec.~\ref{sec:II}. 
251: The underlying mechanisms responsible for the
252: loss of spin orientation and coherence, which impose fundamental limits on 
253: the length  and time scales in spintronic devices, are addressed in 
254: Sec.~\ref{sec:III}.
255: Spintronic applications and devices, with the emphasis on 
256: those based on semiconductors, are discussed in Sec.~\ref{sec:IV}.
257: The review concludes with a look at future prospects
258: in Sec.~\ref{sec:V}
259: and with the table (Tab. II) listing the most common abbreviations 
260: used in the text. 
261: 
262: 
263: \subsection{\label{sec:IB} History and background}
264: %-----------------------------------------------------------
265: 
266: \subsubsection{\label{sec:IB1} 
267: Spin-polarized transport and magnetoresistive effects}
268: 
269: In a pioneering work, \textcite{Mott1936:PRCa,Mott1936:PRCb} provided
270: a basis for our understanding of spin-polarized transport.  Mott
271: sought an explanation for an unusual behavior of resistance in
272: ferromagnetic metals. He realized that at sufficiently low
273: temperatures, where magnon scattering becomes vanishingly small,
274: electrons of majority and minority spin, with magnetic moment parallel
275: and antiparallel to the magnetization of a ferromagnet, respectively,
276: do not mix in the scattering processes. The conductivity can then be
277: expressed as the sum of two independent and unequal parts for two
278: different spin projections--the current in ferromagnets is
279: spin polarized.  This is also known as the two-current model and has
280: been extended by \textcite{Campbell1967:PM,Fert1968:PRL}. It 
281: continues, in its modifications, to provide an explanation for various
282: magnetoresistive phenomena \cite{Valet1993:PRB}.
283: 
284: Tunneling measurements played a key role in early experimental work on
285: spin-polarized transport. Studying N/F/N junctions, where N was a
286: nonmagnetic\footnote{Unless explicitly specified, we shall use the
287: terms ``nonmagnetic'' and ``paramagnetic'' interchangeably, i.e.,
288: assume that they both refer to a material with no long-range
289: ferromagnetic order and with Zeeman-split carrier spin subbands in
290: an applied magnetic field.} metal and F was an Eu-based ferromagnetic
291: semiconductor \cite{Kasuya1968:RMP,Nagaev:1983}, revealed that I-V
292: curves could be modified by an applied magnetic field
293: \cite{Esaki1967:PRL} and show potential for developing a solid-state
294: spin-filter. When unpolarized current is passed across a 
295: ferromagnetic semiconductor, the current becomes spin-polarized
296: \cite{Moodera1988:PRL,Hao1990:PRB}.
297: 
298: \begin{figure}
299: \centerline{\psfig{file=zutic_fig02.eps,width=\linewidth,angle=0}}
300: \caption{Schematic illustration of electron tunneling in
301: ferromagnet/insulator/ferromagnet (F/I/F) tunnel junctions: 
302: (a) Parallel and (b) antiparallel orientation of magnetizations 
303: with the corresponding spin-resolved density of the d states
304: in ferromagnetic metals that have exchange spin splitting $\Delta_{ex}$.
305: Arrows in the two ferromagnetic regions are determined by the
306: majority-spin subband. Dashed lines depict spin-conserved tunneling.
307: }
308: \label{intro:1}
309: \end{figure}
310: 
311: A series of experiments
312: \cite{Tedrow1971:PRLa,Tedrow1973:PRB,Tedrow1994:PR} in
313: ferromagnet/insulator/superconductor (F/I/S) junctions has unambiguously
314: proved that the tunneling current remains spin-polarized even outside
315: of the ferromagnetic region.\footnote{It has been shown
316: that electrons photoemitted from ferromagnetic gadolinium remain 
317: spin polarized \cite{Busch1969:PRL}.}
318: The Zeeman-split quasiparticle density of
319: states in a superconductor \cite{Tedrow1970:PRL,Fulde1973:AP}
320: was used as a detector of spin polarization of conduction electrons in
321: various magnetic materials.  
322: \textcite{Julliere1975:PL} measured  
323: tunneling conductance of F/I/F junctions, where I was 
324: an amorphous Ge. 
325: By adopting the
326: \textcite{Tedrow1971:PRLa,Tedrow1973:PRB} analysis 
327: of the tunneling
328: conductance from F/I/S to the F/I/F junctions,  
329: \textcite{Julliere1975:PL} formulated a model for a change of
330: conductance between the parallel ($\uparrow \uparrow$) and
331: antiparallel ($\uparrow \downarrow$) magnetization in the two
332: ferromagnetic regions F1 and F2, as depicted in Fig.~\ref{intro:1}.
333: The corresponding tunneling magnetoresistance\footnote{Starting with
334: \textcite{Julliere1975:PL} an equivalent expression $(G_{\uparrow
335: \uparrow}-G_{\uparrow \downarrow})/G_{\uparrow \uparrow}$ has also also
336: used by different authors and often referred to as junction
337: magnetoresistance (JMR) \cite{Moodera1999:JMMM}.}  
338: (TMR) in an 
339: F/I/F 
340: magnetic tunnel   
341: junction (MTJ) is defined as 
342: \begin{equation}
343: TMR= \frac{\Delta R}{R_{\uparrow \uparrow}}
344: =\frac{R_{\uparrow \downarrow}-R_{\uparrow \uparrow}}{R_{\uparrow \uparrow}}  
345: =\frac{G_{\uparrow \uparrow}-G_{\uparrow \downarrow}}{G_{\uparrow \downarrow}}, 
346: \label{eq:tmr}
347: \end{equation}
348: where conductance G and resistance R=1/G are labeled by the relative
349: orientations of the magnetizations in F1 and F2 (it is possible
350: to change the relative orientations, between $\uparrow \uparrow$ and
351: $\uparrow \downarrow$, even at small applied magnetic fields $\sim$ 10
352: G). 
353: TMR is a particular manifestation of a magnetoresistance (MR) that 
354: yields a change of electrical resistance in the presence of an  
355: external magnetic field.\footnote{The concept of TMR was proposed
356: independently by R. C. Barker in 1975 [see \textcite{Meservey1983:JMMM}] 
357: and by \textcite{Slonczewski1976:IBM}, 
358: who envisioned its use for 
359: magnetic bubble memory \cite{Parkin:2002}.}
360: Historically, the anisotropic
361: MR in bulk ferromagnets such as Fe and Ni was discovered
362: first, dating back the to experiments of Lord Kelvin
363: \cite{Thomson1857:PRSL}.  Due to spin-orbit interaction, electrical
364: resistivity changes with the relative direction of the charge current
365: (for example, parallel or perpendicular) with respect to the direction
366: of magnetization.
367: 
368: Within Julli{\`{e}}re's model, which assumes constant tunneling matrix elements
369: and that electrons tunnel without spin flip, Eq.~(\ref{eq:tmr}) yields
370: \begin{equation}
371: TMR=\frac{2P_1 P_2}{1-P_1 P_2},
372: \label{eq:julliere}
373: \end{equation}
374: where the polarization $P_i=({\cal N}_{M i} - {\cal N}_{m i})/
375: ({\cal N}_{M i} + {\cal N}_{m i})$
376: is expressed in terms of
377: the spin-resolved density of states ${\cal N}_{M i}$ and 
378: ${\cal N}_{m i}$, for majority and minority spin in F$_i$, respectively. 
379: Conductance in Eq.~(\ref{eq:tmr}) can then 
380: be expressed as
381: \cite{Maekawa1982:IEEE} $G_{\uparrow \uparrow} \sim {\cal N}_{M 1}
382: {\cal N}_{M 2}+ {\cal N}_{m 1} {\cal N}_{m 2}$ and $G_{\uparrow
383: \downarrow} \sim {\cal N}_{M 1} {\cal N}_{m 2}+ {\cal N}_{m 1}
384: {\cal N}_{M 2}$ to give Eq.~(\ref{eq:julliere}).\footnote{In
385: ~\ref{sec:IV} we address some limitations of the Julli{\`{e}}re's
386: model and its potential ambiguities to identify precisely which spin
387: polarization is actually measured.}  While the early results of
388: \textcite{Julliere1975:PL} were not confirmed, TMR at 4.2 K was
389: observed using NiO as a tunnel barrier by \textcite{Maekawa1982:IEEE}.
390: 
391: The prediction of Julli{\`{e}}re's model illustrates the spin-valve effect:
392: the resistance of a device can be changed by manipulating the relative
393: orientation of the magnetizations {\bf M$_1$} and {\bf M$_2$}, in
394: F1 and F2, respectively.  
395: Such  orientation 
396: can be preserved even in the absence of a power
397: supply and the spin-valve effect,\footnote{The term was coined by
398: \textcite{Dieny1991:PRB} in the context of GMR,
399: by invoking an analogy with the physics of the TMR.}  later
400: discovered in multilayer structures displaying the giant
401: magnetoresistance (GMR) effect,\footnote{The term ``giant'' reflected the
402: magnitude of the effect 
403: (more than $\sim 10$ \%), as compared to the better known anisotropic
404: magnetoresistance ($\sim 1$ \%).} \cite{Baibich1988:PRL,Binasch1989:PRB} can be
405: used for nonvolatile memory applications
406: \cite{Hartmann:2000,Parkin:2002,Hirota:2002}. GMR structures are often
407: classified according whether the current flows parallel (CIP) or
408: perpendicular (CPP) to the interfaces between the different layers,
409: as depicted in Fig.~\ref{gmr:1}.  Most of the GMR applications use the
410: CIP geometry, while the CPP version, first realized by
411: \cite{Pratt1991:PRL}, is easier to analyze theoretically
412: \cite{Gijs1997:AP,Levy:2002} and relates to the physics of the TMR
413: effect \cite{Mathon1997:PRB}. The size of magnetoresistance in the GMR
414: structures can be expressed analogously to  Eq.~(\ref{eq:tmr}), where
415: parallel and antiparallel orientations of the magnetizations in the
416: two ferromagnetic regions are often denoted by ``P'' and ``AP,''
417: respectively (instead of $\uparrow \uparrow$ and $\uparrow
418: \downarrow$). 
419: Realization of a large room temperature GMR 
420: \cite{Parkin1991:PRL,Parkin1991:APL} enabled a quick transition from basic 
421: physics to commercial applications in magnetic recording \cite{Parkin2003:PIEEE}.
422: 
423: One of the keys to the success of the
424: MR-based-applications is their ability to 
425: control\footnote{For example, with small magnetic field \cite{Parkin:2002} 
426: or at high
427: switching speeds \cite{Schumacher2003:PRLa,Schumacher2003:PRLb}.} the
428: relative orientation of {\bf M$_1$} and {\bf M$_2$}. An interesting
429: realization of such control was proposed independently by 
430: \textcite{Berger1996:PRB} and \textcite{Slonczewski1996:JMMM}. 
431: While in GMR or TMR structures the relative orientation of magnetizations
432: will affect the flow of spin-polarized current, they predicted
433: a reverse effect. The flow of spin-polarized current can transfer angular
434: momentum from carriers to ferromagnet and alter the orientation 
435: of the corresponding magnetization, even in the absence of an applied
436: magnetic field.
437: This phenomenon, known as spin-transfer
438: torque, has since been extensively studied both theoretically and 
439: experimentally
440: \cite{Bazily1998:PRB,Tsoi1998:PRLa,Myers1999:S,Sun2000:PRB,%Tsoi1998:PRLb
441: Stiles2002:PRB,Wanital2000:PRB}
442: and current-induced magnetization reversal was demonstrated at room 
443: temperature \cite{Katine2000:PRL}. 
444: It was also shown that the magnetic field generated by passing the
445: current through a CPP GMR device could produce room temperature
446: magnetization reversal \cite{Bussman1999:APL}.
447: \begin{figure}
448: \centerline{\psfig{file=zutic_fig03.eps,width=0.33\linewidth,angle=-90}}
449: \caption{Schematic illustration of (a) the current in plane (CIP)
450: (b) the current perpendicular to the plane (CPP) giant magnetoresistance
451: geometry.} 
452: \label{gmr:1}
453: \end{figure}
454: In the context of ferromagnetic semiconductors additional control of
455: magnetization was demonstrated optically (by shining light)
456: \cite{Koshihara1997:PRL,Oiwa2002:PRL,Boukari2002:PRL} and electrically
457: (by applying gate voltage)
458: \cite{Ohno2000:N,Park2002:S,Boukari2002:PRL} to perform switching
459: between the ferromagnetic and paramagnetic states.
460: 
461: 
462: Julli{\`{e}}re's model also justifies the  
463: continued quest for highly spin-polarized materials -- they
464: would provide large magnetoresistive effects, desirable for device 
465: applications. In an extreme case, spins would be completely polarized even in
466: the absence of magnetic field. Numerical support for the existence of such
467: materials--the so called half-metallic ferromagnets\footnote{Near the Fermi
468: level they behave as metals only for one spin, the density of states 
469: vanishes completely for the other spin.} was provided 
470: by \textcite{deGroot1983:PRL},
471: and these materials were reviewed by \textcite{Pickett2001:PT}.
472: In addition to ferromagnets, such as CrO$_2$ \cite{Soulen1998:S,Parker2002:PRL}
473: and manganite perovskites \cite{Park1998:N}, there is evidence for high
474: spin polarization in III-V ferromagnetic semiconductors like 
475: (Ga,Mn)As \cite{Braden2003:P,Panguluri2003:P}.
476: The challenge remains
477: to preserve such spin polarization above room temperature and in junctions
478: with other materials, since the surface (interface) and bulk magnetic
479: properties can be significantly different 
480: \cite{Falicov1990:JMR,Fisher1967:PRL,Mills1971:PRB}. 
481: 
482: While many existing spintronic applications
483: \cite{Hartmann:2000,Hirota:2002} are based on the GMR effects, the
484: discovery of large room-temperature TMR
485: \cite{Moodera1995:PRL,Miyazaki1995:JMMM} has renewed interest in the
486: study of magnetic tunnel junctions 
487: which are now the basis for the
488: several MRAM prototypes\footnote{Realization of the early MRAM
489: proposals used the effect of anisotropic magnetoresistance
490: \cite{Pohm1987:IEEETM,Pohm1988:IEEETM}.} 
491: \cite{Parkin1999:JAP,Tehrani2000:IEEE}. 
492: Future generations of magnetic read heads are expected to  
493: use MTJ's instead of CIP GMR.
494: To improve the switching
495: performance of related devices it is important to reduce the junction
496: resistance, which determines the RC time constant of the MTJ cell.
497: Consequently, semiconductors, which would provide a lower tunneling
498: barrier than the usually employed oxides, are being investigated both as the
499: non-ferromagnetic region in MTJ's and as the basis for an
500: all-semiconductor junction that would demonstrate large TMR at low
501: temperatures \cite{Tanaka2002:SST,Tanaka2001:PRL}. Another
502: desirable property of semiconductors has been demonstrated by
503: the extraordinary large room-temperature MR in hybrid
504: structures with metals reaching 750 000\% at a magnetic field of 4 T 
505: \cite{Solin2000:S} which could lead to improved magnetic
506: read heads \cite{Solin2002:APL,Moussa2003:JAP}. MR effects of
507: similar magnitude have also been found in hybrid metal/semiconductor
508: granular films \cite{Akinaga2002:SST}. Another 
509: approach to obtaining large room-temperature magnetoresistance 
510: ($>100$\% at $B\sim 100$ G)
511: is to fabricate ferromagnetic regions separated by 
512: a nanosize contact. For simplicity, such a structure could be thought of 
513: as the limiting case of the CPP GMR scheme in Fig.~\ref{gmr:1}(b).
514: This behavior, also known as ballistic magnetoresistance, has already 
515: been studied in a large number of materials and geometries 
516: \cite{Garcia1999:PRL,Tatara1999:PRL,Imamura2000:PRL,Chung2002:PRL,%
517: Bruno1999:PRL,Versluijs2002:PRL}.
518: 
519: 
520: \subsubsection{\label{sec:IB2} 
521: Spin injection and optical orientation}
522: 
523: Many  materials in their ferromagnetic state can have a substantial degree of 
524: {\it equilibrium} carrier spin polarization. However, as illustrated
525: in Fig.~\ref{fig:DD}, this alone is 
526: usually  not sufficient for spintronic applications, which typically 
527: require current 
528: flow and/or manipulation of the {\it nonequilibrium} spin 
529: (polarization).\footnote{Important exceptions are tunneling devices operating 
530: at low bias
531: and near {\it equilibrium} spin. Equilibrium polarization 
532: and the current flow can be potentially realized, for example,
533: in spin-triplet superconductors and thin-film ferromagnets 
534: \cite{Konig2001:PRL},
535: accompanied by dissipationless spin currents. 
536: Using an analogy with the quantum Hall effect, it has been suggested that the 
537: spin-orbit interaction could lead to dissipationless spin currents in
538: hole-doped semiconductors \cite{Murakami2003:P}. 
539: \textcite{Rashba2003:PRB} has pointed out that similar dissipationless spin
540: currents in thermodynamic equilibrium, due to spin-orbit interaction,
541: are not transport currents which could be employed for transporting spins
542: and spin injection.
543: It is also 
544: instructive to compare several earlier proposals that use spin-orbit coupling 
545: to generate spin currents, discussed in Sec.~\ref{sec:IIA}.
546: } 
547: The importance of generating {\it nonequilibrium} spin is not limited to 
548: device applications; it can also be used as a sensitive spectroscopic
549: tool to study a wide variety of fundamental properties ranging from spin-orbit 
550: and hyperfine interactions \cite{Meier:1984} to the pairing symmetry of high 
551: temperature superconductors 
552: \cite{Tsuei2000:RMP,Vasko1997:PRL,Wei1999:JAP,Ngai2003:P} 
553: and the
554: creation of  spin-polarized beams to measure parity violation in high energy 
555: physics \cite{Pierce:1984}.
556: 
557: Nonequilibrium spin is the result of some source
558: of pumping arising from transport, optical, or resonance methods. 
559: Once the pumping is turned off the spin will return to its equilibrium value.
560: While for most applications it is desirable to 
561: have long spin relaxation times, it has been demonstrated that short spin 
562: relaxation times are useful in the implementation of fast switching 
563: \cite{Nishikawa1995:APL}. 
564: 
565: Electrical spin injection, an example of a transport method for generating 
566: nonequilibrium spin, has already been realized
567: experimentally by
568: \textcite{Clark1963:PRL}, 
569: who drove a direct current
570: through a sample of InSb in the presence of
571: constant applied magnetic filed.
572: The principle was based on the Feher effect,\footnote{The importance and 
573: possible applications
574: of the Feher effect \cite{Feher1959:PRL} to polarize electrons was discussed
575: by \cite{DasSarma2000:IEEE,Suhl2002:P}.}
576: in which the hyperfine coupling between the electron and nuclear spins, 
577: together with
578: different temperatures representing electron velocity  and  
579: electron spin populations, is responsible for the dynamical
580: nuclear polarization \cite{Slichter:1989}.\footnote{Such an effect can be 
581: thought of as a generalization of the 
582: Overhauser effect \cite{Overhauser1953b:PR} in which the use of 
583: a resonant microwave excitation causes 
584: the spin relaxation of the nonequilibrium  
585: electron 
586: population through hyperfine coupling to lead to the spin polarization of 
587: nuclei.
588: \textcite{Feher1959:PRL} suggested several other methods, instead 
589: of microwave excitation, that could 
590: produce a  
591: nonequilibrium electron population and yield a dynamical polarization of 
592: nuclei [see also
593: \textcite{Weger1963:PR}].} 
594: Motivated by the work of \textcite{Clark1963:PRL},
595: \textcite{Tedrow1971:PRLa,Tedrow1973:PRB}, and the principle of
596: optical orientation \cite{Meier:1984},
597: \textcite{Aronov1976:JETPL,Aronov1976:SPJETP} 
598: and \textcite{Aronov1976:SPS} established several key concepts
599: in electrical spin injection from ferromagnets into metals,
600: semiconductors\footnote{In an earlier
601: work spin injection of minority carriers was proposed in a 
602: ferromagnet/insulator/$p$-type semiconductor structure. Measuring polarization
603: of electroluminescence was suggested as a technique for detecting injection
604: of polarized carriers in a semiconductor \cite{Scifres1973:SSC}.}
605: and superconductors. 
606: When a charge current flowed across 
607: the F/N junction (Fig.~\ref{intro:2}) 
608: \textcite{Aronov1976:JETPL} predicted 
609: that spin-polarized carriers in a ferromagnet would
610: contribute to the net current of magnetization entering the nonmagnetic
611: region and would lead to nonequilibrium magnetization $\delta M$, 
612: depicted in Fig.~\ref{intro:2}(b), 
613: with the spatial extent
614: given by the spin diffusion length
615: \cite{Aronov1976:SPS,Aronov1976:JETPL}.\footnote{Supporting the findings of 
616: \textcite{Clark1963:PRL}, Aronov calculated
617: that the electrical spin injection would polarize nuclei and lead to a 
618: measurable effect in the electron spin resonance (ESR). Several decades
619: later related experiments on spin injection are also examining other
620: implications of dynamical nuclear polarization 
621: \cite{Johnson2000:APL,Strand2003:PRL}.} 
622: Such $\delta M$, which is also
623: equivalent to a {\it nonequilibrium} spin accumulation, was first measured 
624: in metals by \textcite{Johnson1985:PRL,Johnson1988:PRBb}.
625: In the steady state $\delta M$ is realized as the balance between spins 
626: added by 
627: the magnetization current and spins removed by spin 
628: relaxation.\footnote{The spin diffusion length 
629: is an important quantity
630: for CPP GMR. The thickness of the N region in Fig.~\ref{gmr:1}
631: should not exceed the spin diffusion length, otherwise the 
632: information on the orientation of the magnetization in F1 will
633: not be transferred to the F2 region.}
634: \begin{figure}
635: \centerline{\psfig{file=zutic_fig04.eps,width=\linewidth,angle=0}}
636: \caption{Pedagogical illustration of the concept of electrical
637: spin injection from a ferromagnet (F) into a normal metal (N).
638: Electrons flow from F to N:
639: (a) schematic device geometry; (b) magnetization M as a function of
640: position. Nonequilibrium magnetization $\delta M$ (spin accumulation)
641: is injected into a normal metal;
642: (c) contribution of 
643: different spin-resolved densities of states to 
644: charge and spin transport across the F/N interface.
645: Unequal filled levels in the density of states depict spin-resolved
646: electrochemical potentials different from the equilibrium value
647: $\mu_0$.}
648: \label{intro:2}
649: \end{figure}
650: 
651: Generation of nonequilibrium spin polarization and spin accumulation
652: is also possible by optical methods known as optical orientation or
653: optical pumping. In optical orientation, the angular momentum of absorbed 
654: circularly polarized 
655: light is transferred to the medium. Electron orbital momenta are directly 
656: oriented by light and through spin-orbit interaction electron spins 
657: become polarized. In ~\ref{sec:IIB} we focus on the optical orientation in 
658: semiconductors, a well-established technique \cite{Meier:1984}.
659: In a pioneering work \textcite{Lampel1968:PRL} demonstrated 
660: that spins in silicon can be optically oriented (polarized). 
661: This technique is derived from optical pumping proposed by 
662: \textcite{Kastler1950:JDP} in which optical irradiation changes the relative
663: populations within the  Zeeman and the hyperfine levels of the ground states 
664: of atoms.
665: While there are similarities with previous studies of free atoms 
666: \cite{Cohen-Tannoudji:1966,Happer1972:RMP},
667: optical orientation in semiconductors has important differences related
668: to the strong coupling between the electron and nuclear
669: spin and macroscopic number of particles 
670: \cite{Paget1977:PRB,Hermann1985:APF,Meier:1984}.  
671: Polarized nuclei can exert large magnetic 
672: fields ($\sim 5$ T) 
673:  on electrons. 
674: In bulk III-V semiconductors, such as GaAs, optical orientation
675: can lead to 50\% polarization of electron density which could be further 
676: enhanced 
677: in quantum structures of reduced dimensionality or by applying a strain.
678: A simple reversal in the polarization of the illuminating light (from
679: positive to negative helicity) also reverses the sign of the electron density
680: polarization. Combining these properties of optical orientation with the 
681: semiconductors tailored to have a negative electron affinity allows 
682: photoemission of spin-polarized electrons to be used as a powerful detection 
683: technique in high-energy physics and for investigating surface 
684: magnetism \cite{Pierce:1984}.
685: 
686: 
687: \section{\label{sec:II} Generation of spin polarization}
688: %====================================================================
689: 
690: \subsection{\label{sec:IIA} Introduction}
691: %--------------------------------------------------------------------
692: 
693: Transport, optical, and resonance methods (as well as their combination) 
694:  have all been used to create nonequilibrium spin. 
695: After introducing the concept of spin polarization in solid-state systems
696: we give a pedagogical picture of electrical spin injection and
697: detection of polarized carriers. While electrical spin injection
698: and optical orientation will be discussed in more detail later in
699: this section, we also survey here several other techniques for polarizing
700: carriers. 
701: 
702: Spin polarization not only of electrons, but also of holes, nuclei, 
703: and excitations can be defined as 
704: \begin{equation}
705: P_X=X_s/X,
706: \label{eq:polar}
707: \end{equation}                                                                     
708: the ratio of the difference 
709: $X_s=X_\lambda-X_{-\lambda}$ and the sum $X=X_\lambda+X_{-\lambda}$, 
710: of the spin-resolved   
711: $\lambda$ components for a particular quantity $X$. 
712: To avoid ambiguity as to what precisely is meant by spin polarization 
713: both the choice of the spin-resolved components and the relevant
714: physical quantity $X$ need to be specified.
715: Conventionally, $\lambda$ is taken to be $\uparrow$ or $+$
716: (numerical value +1) for spin up, $\downarrow$ or $-$ (numerical
717: value -1) for spin down, with respect to the chosen axis of 
718: quantization.\footnote{For example, 
719: along the spin angular momentum, applied magnetic field, magnetization, or 
720: direction of light propagation.}
721: In ferromagnetic metals it is customary to refer to $\uparrow$ ($\downarrow$) 
722: as carriers with magnetic moment parallel (antiparallel) to the magnetization 
723: or, 
724: equivalently, as carriers with majority (minority) spin \cite{Tedrow1973:PRB}. 
725: In semiconductors the terms majority and minority usually refer to relative 
726: populations of the carriers while $\uparrow$ or $+$ and $\downarrow$ or $-$
727: correspond to the quantum numbers $m_j$ with respect to the $z$-axis taken 
728: along the direction of the light propagation or along the applied magnetic 
729: field \cite{Meier:1984,Jonker2003:P}.
730: It is important to emphasize that both the magnitude and the sign of
731: the spin polarization in Eq.~(\ref{eq:polar}) depends of the choice of 
732: $X$, relevant to the detection technique employed, say optical vs. transport 
733: and bulk vs. surface measurements \cite{Mazin1999:PRL,Jonker2003:P}.
734: Even 
735: in the same homogeneous material the measured $P_X$ can vary for
736: different $X$, and it is crucial to identify which physical 
737: quantity---charge current, carrier density, conductivity, 
738: or the density of states---is being measured experimentally.
739: 
740: \begin{figure}
741: \centerline{\psfig{file=zutic_fig05.eps,width=\linewidth,angle=0}} 
742: \caption{Spin injection, spin accumulation, and spin detection:
743: (a) two idealized completely polarized ferromagnets F1 and F2
744: (the spin down density of states ${\cal N}_\downarrow$ is zero
745: at the energy of electrochemical potential $E=\mu_0$) 
746: with parallel
747: magnetizations are separated by the nonmagnetic region N;
748: (b) density-of-states diagrams for spin injection from F1 into N,
749: accompanied by the spin accumulation--generation of the nonequilibrium
750: magnetization $\delta M$. At F2
751: in the limit of low impedance ($Z$=0) 
752: spin is detected
753: by measuring the spin-polarized
754: current across the N/F2 interface. In the limit of high impedance 
755: ($Z=\infty$) spin is detected
756: by measuring the voltage $V_s\sim \delta M$ developed across the N/F2 
757: interface;
758: (c) spin accumulation in a device in which  
759: a superconductor (with the superconducting gap $\Delta$) is
760: occupying the region between F1 and F2.}
761: \label{inj:1}
762: \end{figure}                                                                                       
763: The spin polarization of electrical current or carrier density, 
764: generated in a nonmagnetic region,
765: is typically used to describe the efficiency of electrical spin injection.
766: \textcite{Silsbee1980:BMR}  suggested that the nonequilibrium 
767: density polarization in the N region, or equivalently the
768: nonequilibrium magnetization, acts as the source of spin electromotive force 
769: (EMF) and produces a measurable ``spin-coupled'' 
770: voltage $V_s \propto \delta M$. Using this concept, also referred to
771: as  {\it spin-charge coupling},
772: \textcite{Silsbee1980:BMR} proposed a detection technique  
773: consisting of two ferromagnets F1 and F2 (see Fig.~\ref{inj:1})
774: separated by a nonmagnetic region.\footnote{A similar geometry was also 
775: proposed independently by \textcite{deGroot1983:PA}, where F1 and F2 were two
776: half-metallic ferromagnets with the goal of implementing  spin-based
777: devices to amplify and/or switch current.} 
778: F1 serves as the spin injector (spin aligner)
779: and F2 as the spin detector. This could be called the polarizer-analyzer 
780: method, 
781: the optical counterpart of the transmission of light through 
782: two optical linear polarizers.
783: From  Fig.~\ref{inj:1} it follows that
784: the reversal of the magnetization
785: direction in one of the ferromagnets would lead either to 
786: $V_s \rightarrow -V_s$,
787: in an open circuit (in the limit of large impedance Z), or to the reversal of
788: charge current $j \rightarrow -j$, in a short circuit (at small Z), 
789: a consequence
790: of Silsbee-Johnson spin-charge coupling 
791: \cite{Silsbee1980:BMR,Johnson1987:PRB,Johnson1988:PRBa}.
792: Correspondingly, as discussed in the following sections,
793: the spin injection could be detected through the spin accumulation signal 
794: either as a voltage or the resistance change when the magnetizations 
795: in F1 and F2 are changed from parallel to antiparallel alignment. 
796: 
797: Since the experiments demonstrating
798: the spin accumulation of conduction electrons
799: in metals \cite{Johnson1985:PRL}, spin injection has been realized
800: in a wide range of materials. While in Sec.~\ref{sec:IIC} we focus on related
801: theoretical work motivated by potential applications, experiments on
802: spin injection have also stimulated proposals for examining the fundamental
803: properties of electronic systems.\footnote{For example, studies          
804: probing the  spin-charge separation
805: in the non-Fermi liquids have been proposed by
806: \cite{Kivelson1990:PRB,Zhao1995:PRB,Si1997:PRL,Si1998:PRL,Balents2000:PRL,Balents2001:PRB}. 
807: Spin and charge are carried by
808: separate excitations and can lead to spatially separated spin and 
809: charge currents \cite{Kivelson1990:PRB}.}
810: 
811: The generation of nonequilibrium spin polarization has a long
812: tradition in magnetic resonance methods
813: \cite{Abragam:1961,Slichter:1989}. However, transport methods
814: to generate carrier spin polarization are not limited to electrical
815: spin injection. For example, they also include scattering of unpolarized
816: electrons in the presence of spin-orbit coupling 
817: \cite{Mott:1965,Kessler:1976} and in materials that lack the
818: inversion symmetry \cite{Levitov1985:SPJETP}, adiabatic 
819: \cite{Mucciolo2002:PRL,Sharma2003:PRB,Watson2003:PRL} and nonadiabatic  
820: quantum spin pumping \cite{Zheng2002:P} [for an instructive description
821: of parametric pumping see \cite{Brouwer1998:PRB}],
822: and the proximity effects \cite{Ciuti2002:PRL}.  
823: 
824: It would be interesting to know what the limits are on
825: the magnitude of various spin polarizations. Could we have a completely
826: polarized current [$P_j\rightarrow \infty$, see Eq.~(\ref{eq:polar})],
827: with only a spin current ($j_\uparrow-j_\downarrow$) and no charge current
828: ($j_\uparrow+j_\downarrow=0$)?
829: While it is tempting to
830: recall the Stern-Gerlach experiment and try to set up magnetic
831: drift through  inhomogeneous magnets \cite{Kessler:1976}, 
832: this would most likely work
833: only as a transient effect \cite{Fabian2002b:PRB}. It was proposed
834: already by \textcite{Dyakonov1971:PL,Dyakonov1971:JETPLa} that a
835: transverse spin current (and transverse spin polarization in a
836: closed sample) would form as a result of spin-orbit coupling-induced
837: skew scattering in the presence of a longitudinal electric field.
838: This interesting effect, also called the {\it spin Hall effect} 
839: \cite{Hirsch1999:PRL,Zhang2000:PRL},
840: has yet to be demonstrated. An alternative scheme for producing pure
841: spin currents was proposed by \textcite{Bhat2000:PRL},
842: motivated by the experimental demonstration of
843: phase coherent control of charge currents 
844: \cite{Atanasov1996:PRL,Hache1997:PRL}
845: and carrier population \cite{Fraser1999:PRL}. 
846: A quantum-mechanical
847: interference between one- and two-photon absorptions of orthogonal
848: linear polarizations creates an opposite ballistic flow of spin up and
849: spin down electrons in a semiconductor. Only a spin
850: current can flow, without a charge current, as
851: demonstrated by \textcite{Stevens2003:PRL} and \textcite{Hubner2003:PRL}, 
852: who were able to achieve coherent 
853: control of the spin current direction and magnitude by the
854: polarization and the relative phase of two exciting laser light fields.
855: 
856: Charge current also can be driven by circularly polarized
857: light \cite{Ivchenko:1997}. 
858: Using the principles of optical orientation (see Sec.~\ref{sec:IB2}
859: and 
860: further discussion in Sec.~\ref{sec:IIB}) in semiconductors of reduced 
861: dimensionality or lower symmetry, 
862: both  the direction and  the magnitude 
863: of a generated charge current 
864: can be controlled by
865: circular polarization of the light. 
866: This is called the circular photo-voltaic effect \cite{Ganichev2003:JPCM},
867: which can be viewed as a transfer of the angular momentum of photons to 
868: directed motion of electrons. This could also be called a spin
869: corkscrew effect, since a nice mechanical analog is a corkscrew
870: whose rotation generates linear directed motion. A related effect,
871: in which the photocurrent is driven, is called the spin-galvanic effect
872: \cite{Ganichev2003:JPCM}. The current here is causes by the difference
873: in spin-flip scattering rates for electrons with different spin states
874: in some systems with broken inversion symmetry.
875: A comprehensive survey of the related effects 
876: from the circular photo-galvanic effect \cite{Asnin1979:SSC}
877: to recent demonstrations in semiconductor quantum wells 
878: \cite{Ganichev2001:PRL,Ganichev2002:PRL,Ganichev2002:N,Ganichev2003:PRB}]
879: is given by \textcite{Ganichev2003:JPCM}.
880: 
881: There is a wide range of recent theoretical proposals for devices
882: that would give rise to a spin electromotive force 
883: \cite{Long2002:APL,Zutic2001:APL,Zutic2001:PRB,Governale2003:PRB,%
884: Brataas2002:PRB,Malshukov2003:PRB,Ting2003:APL}, 
885: often
886: referred to as spin(-polarized) pumps, cells, or batteries. However,
887: even when it is feasible to generate pure spin current, this does 
888: not directly imply that 
889: it would be dissipationless. 
890: In the context of superconductors, it has been 
891: shown that Joule heating can arise from pure spin current 
892: flowing through a Josephson junction \cite{Takahashi2001:JAP}.
893: 
894: \subsection{\label{sec:IIB} Optical spin orientation}
895: %-----------------------------------------------------------------------------
896: 
897: 
898: \begin{figure}
899: \centerline{\psfig{file=zutic_fig06.eps,width=\linewidth,angle=0}}
900: \caption{(a) Interband transitions in GaAs: 
901: (a) schematic band structure of GaAs near the center 
902: of the Brillouin zone ($\Gamma$ point), where E$_g$ 
903: is the band gap and $\Delta_{so}$ is
904: the spin-orbit splitting; 
905: CB, conduction band; HH, valence heavy hole; LH, light hole; SO,
906: spin-orbit-split-off 
907: subbands; $\Gamma_{6,7,8}$ are the corresponding symmetries at the $k=0$ 
908: point, more precisely, the irreducible 
909: representations of the tetrahedron group $T_d$ \cite{Ivchenko:1997};
910: (b) selection rules for interband transitions between 
911: $m_j$ sublevels 
912: for circularly polarized light $\sigma^+$ 
913: and $\sigma^-$ (positive and negative helicity). The circled numbers
914: denote the relative transition intensities that apply for both
915: excitations (depicted by the arrows) and radiative recombinations.} 
916: \label{oo:1}
917: \end{figure}
918: 
919: In a semiconductor the photoexcited spin-polarized electrons and holes 
920: exist for the time $\tau$ before they recombine. If a fraction of the 
921: carriers' initial 
922: orientation survives longer than the recombination time, that is, 
923: if $\tau < \tau_s$, \footnote{In Si this condition is not
924: fulfilled. Instead of measuring the luminescence polarization, 
925: \textcite{Lampel1968:PRL} has used NMR to detect optical spin orientation.}
926: where $\tau_s$ is the spin relaxation time (see Sec.~\ref{sec:III}), 
927: the luminescence  (recombination radiation) will be partially polarized.  
928: By measuring the circular polarization of the luminescence it is 
929: possible to study the spin dynamics of the nonequilibrium carriers 
930: in semiconductors \cite{Oestreich2002:SST} and to extract such 
931: useful quantities as the spin orientation, the recombination time, or  
932: the spin relaxation time of the carriers
933: \cite{Parsons1969:PRL,Garbuzov1971:ZhETF,Ekimov1970:ZhETF,Meier:1984}.
934: 
935: We illustrate the basic principles of optical orientation
936: by the example of GaAs which is representative of a large
937: class of III-V and II-VI zincblende semiconductors. The
938: band structure is depicted in Fig.~\ref{oo:1}(a).
939: The band gap is $E_g=1.52$ eV at $T=0$  $K$, while the spin 
940: split-off band is separated from the light and 
941: heavy hole bands by $\Delta_{so}=0.34$ eV. 
942: We denote the Bloch states according to the total angular
943: momentum $J$ and its projection onto the positive
944: $z$ axis $m_j$: $|J,m_j \rangle$. Expressing the wave
945: functions with the symmetry of $s$, $p_x$, $p_y$, and $p_z$
946: orbitals as $|S\rangle $, $|X\rangle $, $|Y\rangle$, and $|Z\rangle$, 
947: respectively, the band wave functions can be written as listed 
948: in Table~\ref{tab:1} 
949: \cite{Pierce1976:PRB} [with minor typos removed, 
950: see also \cite{Kittel:1963}]. 
951: 
952: To obtain the excitation (or recombination) probabilities  
953: consider photons arriving in the $z$ direction. 
954: Let $\sigma^{\pm}$ represent the helicity of the exciting light. 
955: When we represent the dipole operator corresponding to the 
956: $\sigma^{\pm}$ optical transitions as \footnote{For an outgoing light 
957: in the $-z$ direction the helicities are reversed.}
958: $\propto (X\pm iY) \propto Y^{\pm 1}_1$, where $Y^m_l$  is the spherical
959: harmonic, it follows from Table~\ref{tab:1} that  
960: \begin{equation}
961: \label{eq:three}
962: \frac{|\langle 1/2,-1/2 | Y^1_1 |3/2, -3/2 \rangle|^2}
963: {|\langle 1/2,1/2 | Y^1_1 |3/2, -1/2 \rangle|^2}=3
964: \end{equation}
965: for the relative intensity of the $\sigma^+$ transition
966: between the heavy ($|m_j=3/2|$) and the light ($|m_j=1/2|$) hole subbands 
967: and the conduction band.
968: Other transitions are analogous. The relative transition rates
969: are indicated  in Fig.~\ref{oo:1}(b). The same selection rules apply
970: to the optical orientation of shallow impurities 
971: \cite{Parsons1969:PRL,Ekimov1970:ZhETF}. 
972: 
973: 
974: \begin{table}
975: \begin{tabular}{lll}
976: \hline
977: \hline
978: & &  \\
979: Symmetry   &  $|J,m_j \rangle$ & Wave function  \\
980: & &  \\
981: \hline 
982: $\Gamma_6$ & $|1/2,1/2 \rangle$ &  $|S \! \uparrow  \rangle$  \\
983:  & $|1/2,-1/2 \rangle$ &  $|S \! \downarrow  \rangle$  \\
984: \hline
985: $\Gamma_7$ & $|1/2,1/2 \rangle$ &  
986: $|-(1/3)^{1/2}[\, (X+iY) \! \downarrow  - Z \! \uparrow] \, \rangle$  \\
987:  & $|1/2,-1/2 \rangle$ &  
988: $|(1/3)^{1/2}[\, (X-iY) \! \uparrow  + Z \! \downarrow]  \, \rangle$  \\
989: \hline
990: $\Gamma_8$ & $|3/2,3/2 \rangle$ &  
991: $|(1/2)^{1/2}(X+iY) \! \uparrow \rangle$  \\
992:  & $|3/2,1/2 \rangle$ &  
993: $|(1/6)^{1/2}[\, (X+iY) \! \downarrow  + 2 Z \! \uparrow] \, \rangle$  \\
994:  & $|3/2,-1/2 \rangle$ &  
995: $|-(1/6)^{1/2}[\, (X-iY) \! \uparrow  - 2 Z \! \downarrow] \, \rangle$  \\
996:  & $|3/2,-3/2 \rangle$ &  
997: $|(1/2)^{1/2}(X-iY) \! \downarrow \rangle$  \\
998: \hline
999: \hline
1000: \end{tabular}
1001: \caption{Angular and spin part of the wave function at $\Gamma$.}
1002: \label{tab:1}
1003: \end{table}
1004:   
1005: The spin polarization of the excited electrons\footnote{Although holes
1006: are initially polarized too, they lose spin orientation very
1007: fast, on the time scale of the momentum relaxation time 
1008: (see Sec.~\ref{sec:IIID1})
1009: However, it was suggested that manipulating hole spin by short electric
1010: field pulses, between momentum scattering events, could be useful for
1011: ultrafast spintronics applications \cite{Dargys2002:PRB}.
1012: }.
1013: depends on the photon energy $\hbar\omega$. For $\hbar\omega$ between $E_g$ 
1014: and $E_g+\Delta_{so}$, only the light and heavy hole subbands contribute. 
1015: Denoting by $n_{+}$ and $n_{-}$ the density of electrons polarized 
1016: parallel ($m_j=1/2$) and
1017: antiparallel ($m_j=-1/2$) to the direction of light propagation, we
1018: define the 
1019: spin polarization as (see Sec.~\ref{sec:IIA})
1020: \begin{equation} \label{eq:pn}
1021: P_n=(n_+-n_-)/(n_++n_-).
1022: \end{equation}
1023: For our example of the zincblende structure, 
1024: \begin{equation}
1025: P_n=(1-3)/(3+1)=-1/2
1026: \end{equation}
1027: is the spin polarization at the moment of photoexcitation. The spin is 
1028: oriented against the direction of light propagation, since there are more 
1029: transitions from the heavy hole than from the light hole subbands. 
1030: The circular polarization of the luminescence is defined as
1031: \begin{equation}
1032: \label{eq:circ}
1033: P_{\rm circ}=(I^+ -I^-)/(I^+ + I^-),
1034: \end{equation}
1035: where $I^\pm$ is the radiation intensity for the helicity $\sigma^{\pm}$.
1036: The polarization of the $\sigma^+$ photoluminescence is then
1037: \begin{equation}
1038: P_{\rm circ}=\frac{(n_+ + 3n_-)-(3n_+ + n_-)}{(n_+ + 3n_-)+(3n_+ + n_-)}
1039:              =-\frac{P_n}{2}=\frac{1}{4}.
1040: \label{eq:1/4}       
1041: \end{equation}
1042: 
1043: If the excitation involves transitions from the spin split-off band, 
1044: that is, if  $\hbar \omega \gg E_g + \Delta_{so}$, the electrons
1045: will not be spin polarized ($P_n=P_{\rm circ}=0$), underlining the 
1046: vital role of spin-orbit coupling for spin orientation.  On the
1047: other hand, Fig.~\ref{oo:1} 
1048: suggests that a removal of the heavy/light
1049: hole degeneracy can substantially increase $P_n$ \cite{Dyakonov:1984}, 
1050: up to the limit of complete spin polarization. 
1051: An increase in $P_n$ and $P_{\rm circ}$
1052: in GaAs strained due to a lattice mismatch with a substrate, or due
1053: to confinement in quantum well heterostructures,
1054: has indeed been demonstrated \cite{Oskotski1997:PLDS,Vasilev1993:SM},
1055: detecting $P_n$ greater than 0.9. 
1056: 
1057: While photoexcitation with  circularly polarized light creates 
1058: spin-polarized electrons, the nonequilibrium spin decays due to both 
1059: carrier recombination and spin relaxation. The steady-state degree
1060: of spin polarization depends on the balance between the spin 
1061: excitation and decay. Sometimes a distinction is made 
1062: \cite{Pierce1976:PRB,Meier:1984}
1063: between the terms optical {\it spin orientation} and optical {\it spin 
1064: pumping}. The former term is used in relation to the minority carriers
1065: (such as electrons in p-doped samples) and represents the 
1066: orientation of the excited carriers. The latter term is reserved for the
1067: majority carriers (electrons in n-doped samples), representing
1068: spin polarization of the ``ground'' state. Both spin orientation
1069: and spin pumping were demonstrated in the early investigations
1070: on p-GaSb \cite{Parsons1969:PRL} and  p- and n-Ga$_{0.7}$Al$_{0.3}$As 
1071: \cite{Ekimov1970:ZhETF,Ekimov1971:JETPL,Zakharchenya1971:JETPL}. Unless
1072: specified otherwise, we shall use the term optical orientation to 
1073: describe both
1074: spin orientation and spin pumping.
1075: 
1076: 
1077: To derive the steady-state expressions for the spin polarization due to optical
1078: orientation, consider the simple model of carrier recombination and spin 
1079: relaxation
1080: (see Sec.~\ref{sec:IVA4}) in a homogeneously doped semiconductor.
1081: The balance between direct electron-hole recombination and optical 
1082: pair creation can be written as 
1083: \begin{equation} 
1084: \label{eq:np}
1085: r(np-n_0p_0)=G,
1086: \end{equation}
1087: where $r$ measures the recombination rate, the electron and hole densities are 
1088: $n$ and
1089: $p$, with index zero denoting the equilibrium values, 
1090: and $G$ is the electron-hole
1091: photoexcitation rate. Similarly, the balance between spin relaxation and 
1092: spin generation is expressed by
1093: \begin{equation} 
1094: \label{eq:sp}
1095: rsp+s/\tau_s=P_{n}(t=0) G,
1096: \end{equation}
1097: where $s=n_+-n_-$ 
1098: is the electron spin density and $P_{n}(t=0)$ is the spin polarization
1099: at the moment of photoexcitation, given by Eq.~(\ref{eq:pn}). 
1100: Holes are assumed to lose their
1101: spin orientation very fast, so they are treated as unpolarized. The first
1102: term in Eq.~(\ref{eq:sp}) describes the disappearance of the spin density 
1103: due to carrier recombination, while the second term describes the
1104: intrinsic spin relaxation. From Eqs.~(\ref{eq:np}) and (\ref{eq:sp}) we 
1105: obtain the
1106: steady-state electron polarization as \cite{Zutic2001:PRB}
1107: \begin{equation} \label{eq:p}
1108: P_n=P_n(t=0)\frac{1-n_0p_0/np}{1+1/\tau_srp}.
1109: \end{equation}
1110: 
1111: 
1112: In a p-doped sample $p\approx p_0$, $n\gg n_0$, and Eq.~(\ref{eq:p}) gives 
1113: \begin{equation}
1114: \label{eq:orient}
1115: P_n=P_n(t=0)/(1+\tau/\tau_s),
1116: \end{equation}
1117: where $\tau=1/rp_0$ is the electron lifetime.\footnote{After the illumination
1118: is switched off, the electron spin density, or equivalently the nonequilibrium
1119: magnetization, will decrease exponentially with the inverse time constant
1120: $1/T_s=1/\tau+1/\tau_s$ \cite{Parsons1969:PRL}.}
1121: The steady-state polarization is independent of the illumination intensity,
1122: being reduced from the initial spin polarization $P_n(t=0)$.
1123: \footnote{The effect of a finite length for the light absorption on $P_n$ 
1124: is discussed 
1125: by \textcite{Pierce:1984}. The absorption length $\alpha^{-1}$ is
1126: typically a micron for GaAs.   It varies with frequency
1127: roughly as $\alpha(\hbar \omega) \propto (\hbar \omega - E_g)^{1/2}$ 
1128: \cite{Pankove:1971}.}
1129: The polarization of the photoluminescence is 
1130: $P_{\rm circ}=P_n(t=0) P_n$ \cite{Parsons1969:PRL}.
1131: Early measurements of  $P_n=0.42 \pm 0.08$ in  GaSb \cite{Parsons1969:PRL} 
1132: and 
1133: $P_n=0.46 \pm 0.06$  in  Ga$_{0.7}$Al$_{0.3}$As \cite{Ekimov1970:ZhETF}
1134: showed an effective spin orientation close to the maximum 
1135: value of $P_n(t=0)=1/2$ for a bulk unstrained  zincblende structure,
1136: indicating that $\tau/\tau_s \ll 1$. 
1137: 
1138: For spin pumping in an n-doped sample, where $n\approx n_0$ and $p\gg p_0$, 
1139: Eqs.~(\ref{eq:np}) and (\ref{eq:p}) give \cite{Dyakonov1971:JETPL}
1140: \begin{equation}
1141: \label{eq:pump}
1142: P_n=P_n(t=0)/(1+n_0/G\tau_s).
1143: \end{equation}
1144: In contrast to the previous case, the carrier (now hole) lifetime 
1145: $\tau=1/rn_0$
1146: has  no effect on $P_n$. However, $P_n$ depends on the
1147: photoexcitation intensity $G$, as expected for a pumping process. 
1148: The effective
1149: carrier lifetime is $\tau_J=n_0/G$, 
1150: where $J$ represents the intensity of the illuminating light. 
1151: If it is comparable to or shorter than $\tau_s$, 
1152: spin pumping is very effective. Spin pumping works because the photoexcited 
1153: spin-polarized electrons do not need to recombine with holes. There
1154: are plenty of unpolarized electrons in the conduction band available
1155: for recombination. The spin is thus pumped in to the electron system.
1156: 
1157: 
1158: When magnetic field {\bf B} is applied 
1159: perpendicular to the axis of spin orientation (transverse magnetic field),
1160: it will induce spin precession with the 
1161: Larmor frequency $\Omega_L=\mu_B g B/\hbar$, where $\mu_B$ is the Bohr
1162: magneton and $g$ is the electron $g$ factor.\footnote{In our convention
1163: the $g$ factor of free electrons is positive, 
1164: $g_0=2.0023$ \cite{Kittel:1996}.}
1165: The spin precession, together with the random character of carrier generation
1166: or diffusion, leads to the spin dephasing (see Sec.~\ref{sec:IIIA1}).
1167: Consider spins excited by circularly polarized light (or by any means
1168: of spin injection) at a steady rate. In a steady rate a balance between
1169: nonequilibrium spin generated and spin relaxation is maintained, resulting
1170: in a net magnetization. If a transverse magnetic field is applied,
1171: the decrease of the steady-state magnetization can have two sources:
1172: (a) spins which were excited at random time and (b) random diffusion
1173: of spins towards a detection region. Consequently, spins precess along
1174: the applied field acquiring random phases relative to those which were
1175: excited or have arrived at different times.
1176: As a result, the projection
1177: of the electron spin along the exciting beam will decrease with the 
1178: increase of transverse magnetic field, leading to depolarization of the  
1179: luminescence. 
1180: This is also known as the Hanle effect \cite{Hanle1924:ZP}, in analogy
1181: to the depolarization of the resonance fluorescence of gases. The Hanle
1182: effect was first measured in semiconductors by \textcite{Parsons1969:PRL}. 
1183: The steady-state spin polarization of the precessing electron spin 
1184: can be calculated by solving the Bloch-Torrey equations 
1185: \cite{Bloch1946:PR,Torrey1956:PR}, 
1186: Eqs.~(\ref{eq:relax:bloch1})--(\ref{eq:relax:bloch3}) describing the 
1187: spin dynamics
1188: of diffusing carriers.
1189: 
1190: In p-doped semiconductors the Hanle curve 
1191: shows a Lorentzian decrease of the polarization \cite{Parsons1969:PRL},
1192: $P_n(B)=P_n(B=0)/(1+\Omega_L T_s)^2$,
1193: where $P_n(B=0)$ is the polarization at $B=0$ from Eq.~(\ref{eq:orient})
1194: and $T_s^{-1}$ is the effective spin lifetime given by 
1195: $1/T_s=1/\tau+1/\tau_s$;
1196: see footnote 26.
1197: Measurements of the Hanle curve in GaAlAs were used 
1198: by  \textcite{Garbuzov1971:ZhETF}    
1199: to separately determine both $\tau$ and $\tau_s$ at various temperatures.
1200: The theory of the Hanle effect in n-doped semiconductors was developed by
1201: \textcite{D'yakonov1976:FTP} who showed the  
1202: non-Lorentzian decay of the luminescence for the
1203: regimes  both of low ($\tau_J / \tau_s \gg  1$) 
1204: and high ($\tau_J / \tau_s \ll 1$) intensity of the exciting light. 
1205: At high fields $P_n(B)\propto 1/B^{1/2}$, consistent
1206: with the experiments of \textcite{Vekua1976:FTP} in Ga$_{0.8}$Al$_{0.2}$As, 
1207: showing a Hanle curve different from the usual $P_n(B)\propto 1/B^2$ Lorentzian 
1208: behavior \cite{Dyakonov:1984}. Recent findings on the Hanle effect in
1209: nonuniformly doped GaAs and reanalysis of some earlier studies are
1210: given by \textcite{Dzihoev2003:FTT}.
1211: 
1212: \subsection{\label{sec:IIC} Theories of spin injection}
1213: %-------------------------------------------------------------------------------
1214: 
1215: Reviews on spin injection have covered materials ranging from semiconductors 
1216: to high temperature superconductors and have 
1217: addressed the implications for device 
1218: operation as well as for fundamental studies in solid 
1219: state systems.\footnote{See, for example, 
1220: \cite{Johnson2002:SST,Schmidt2002:SST,Jedema2002:JS,Wei2002:JS,%
1221: Goldman1999:JMMM,Goldman2001:JS,Tang:2002,Johnson2001:JS,Osofsky2000:JS,%
1222: Maekawa2001:MSE}}.  
1223: In addition to degenerate conductors, examined in these works, we also give
1224: results for nondegenerate semiconductors in which the violation of local 
1225: charge neutrality, electric fields, and carrier band bending require
1226: solving the Poisson equation. The notation introduced here emphasizes the
1227: importance of different (and inequivalent) spin polarizations arising in 
1228: spin injection.
1229: 
1230: \subsubsection{\label{sec:IIC1} F/N junction}
1231: %---------------------------------------------------------------------------
1232: 
1233: 
1234: A theory of spin injection across a ferromagnet/normal metal (F/N) interface
1235: was first offered by \textcite{Aronov1976:JETPL}. Early work also
1236: included spin injection into a semiconductor (Sm) 
1237: \cite{Aronov1976:SPS,Masterov1979:SPS}
1238: and a superconductor (S) \cite{Aronov1976:SPJETP}. Spin injection in 
1239: F/N junctions
1240: was subsequently studied in detail by 
1241: \textcite{Johnson1987:PRB,Johnson1988:PRBa},\footnote{Johnson and Silsbee base 
1242: their approach on irreversible thermodynamics and consider also  the effects of 
1243: a temperature gradient on spin-polarized transport, omitted in this section.} 
1244: \textcite{vanSon1987:PRL},
1245: \textcite{Valet1993:PRB}, \textcite{Hershfield1997:PRB}, and others.
1246: Here we follow the approach of \textcite{Rashba2000:PRB,Rashba2002:EPJ} and 
1247: consider
1248: a steady-state\footnote{Even some dc spin injection experiments are actually
1249: performed at low (audio-frequency) bias. Generalization to ac spin injection, 
1250: with a harmonic time dependence, was studied by \textcite{Rashba2002:APL}.}
1251: flow of electrons along the $x$ direction in a three-dimensional (3D) 
1252: geometry consisting of a metallic ferromagnet (region $x<0$) and a 
1253: paramagnetic metal or
1254: a degenerate semiconductor (region $x>0$).
1255: 
1256: \begin{figure}
1257: \centerline{\psfig{file=zutic_fig07.eps,width=0.9\linewidth,angle=0}}
1258: \caption{Spatial variation of the electrochemical potentials near a
1259: spin-selective resistive interface at an F/N junction. At the interface $x=0$
1260: both the spin-resolved electrochemical potentials ($\mu_\lambda$, 
1261: $\lambda=\uparrow,\downarrow$, denoted with solid lines) and the average 
1262: electrochemical potential ($\mu_{F}$, $\mu_{N}$, dashed lines) are 
1263: discontinuous. The spin diffusion length $L_{sF}$ and $L_{sN}$
1264: characterizes the decay of  $\mu_s=\mu_\uparrow - \mu_\downarrow$ 
1265: (or equivalently the decay of spin accumulation and  
1266: the nonequilibrium magnetization) away from the interface and into the bulk 
1267: F and N regions, respectively.} 
1268: \label{inj:2}
1269: \end{figure}
1270: The two regions, F and N, form a contact at $x=0$,
1271: as depicted in Fig.~\ref{inj:2}.
1272: The relative magnitudes of three characteristic resistances  
1273: per unit area\footnote{For this simple
1274: geometry various resistances have a common factor of the cross-sectional area, 
1275: which
1276: can be factored out. This is no longer possible for a more complicated geometry 
1277: \cite{Takahashi2003:PRB}.} determine the degree of current polarization 
1278: injected 
1279: into a nonmagnetic material. These are the contact resistance $r_c$ 
1280: and the two characteristic 
1281: resistances $r_N$ $r_F$, each given by the ratio of the spin diffusion 
1282: length and the 
1283: effective bulk conductivity in the corresponding region.
1284: Two limiting cases  correspond to the transparent limit, 
1285: where $r_c\rightarrow 0$, and the low-transmission limit, where 
1286: $r_c \gg r_N,r_F$.
1287: 
1288: Spin-resolved quantities are labeled by 
1289: $\lambda=1$ or $\uparrow$ for spin up, $\lambda=-1$ or $\downarrow$ 
1290: for spin down along the chosen quantization axis. 
1291: For a free electron, spin angular momentum
1292: and magnetic moment are in opposite directions, and what
1293: precisely is denoted by ``spin up'' varies in the 
1294: literature \cite{Jonker2003:P}.
1295: Conventionally, in metallic systems 
1296: \cite{Tedrow1973:PRB,Gijs1997:AP},
1297: spin up refers to carriers with majority spin. This means that the
1298: spin (angular momentum) of such carriers is
1299: antiparallel to the magnetization.
1300: Spin-resolved charge current (density)
1301: in a diffusive regime can be expressed as 
1302: \begin{equation}
1303: \label{eq:jq}
1304: j_\lambda=\sigma_\lambda \nabla \mu_\lambda,
1305: \end{equation}
1306: where $\sigma_\lambda$ is conductivity and the electrochemical potential is
1307: \begin{equation}
1308: \label{eq:elchem}
1309: \mu_\lambda=(q D_\lambda/\sigma_\lambda) \delta n_\lambda-\phi,
1310: \end{equation}
1311: with $q$ proton charge, $D_\lambda$ diffusion coefficient, 
1312: $\delta n_\lambda=n_\lambda-n_{\lambda 0}$ the change of electron density 
1313: from the equilibrium value for spin $\lambda$, 
1314: and $\phi$ the electric potential.~\footnote{More generally, 
1315: for a noncollinear magnetization, 
1316: $j_\lambda$ becomes a second-rank tensor 
1317: \cite{Johnson1988:PRBa,Margulis1994:PB,Stiles2002:PRB}.}
1318: 
1319: In the steady state the continuity equation is
1320: \begin{equation}
1321: \label{eq:jcont}
1322: \nabla j_\lambda=
1323: \lambda q \left[\frac{\delta n_\lambda}{\tau_{\lambda-\lambda}}
1324:                -\frac{\delta n_{-\lambda}}{\tau_{-\lambda \lambda}} \right],
1325: \end{equation}
1326: and $\tau_{\lambda\lambda'}$ is the average time for flipping a $\lambda$-spin
1327: to $\lambda'$-spin.
1328: For a degenerate conductor\footnote{In the nondegenerate case of
1329: Boltzmann statistics, the Einstein relation implies 
1330: that the ratio of the diffusion coefficient and the mobility is $k_BT/q$.} 
1331: the Einstein relation 
1332: is
1333: \begin{equation}
1334: \sigma_\lambda=q^2 {\cal N}_\lambda D_\lambda, 
1335: \label{eq:einstein}
1336: \end{equation}
1337: where $\sigma=\sigma_\uparrow+\sigma_\downarrow$ and 
1338: ${\cal N}={\cal N}_\uparrow+{\cal N}_\downarrow$
1339: is the density of states. Using a detailed balance 
1340: ${\cal N}_\uparrow/ \tau_{\uparrow \downarrow}=
1341: {\cal N}_\downarrow/ \tau_{\downarrow \uparrow}$  
1342: \cite{Hershfield1997:PRB,Kravchenko2002:JETP} 
1343: together with Eqs.~(\ref{eq:elchem}) and (\ref{eq:einstein}),
1344: the continuity equation can be expressed as
1345: \begin{equation}
1346: \label{eq:jq'}
1347: \nabla j_\lambda=\lambda q^2 \frac{{\cal N}_\uparrow {\cal N}_\downarrow}
1348: {{\cal N}_\uparrow+{\cal N}_\downarrow}
1349: \frac{\mu_\lambda-\mu_{-\lambda}}{\tau_s},
1350: \end{equation}
1351: where $\tau_s=\tau_{\uparrow \downarrow} \tau_{\downarrow \uparrow}/ 
1352: (\tau_{\uparrow \downarrow}+\tau_{\downarrow \uparrow})$ 
1353: is the spin relaxation time. Equation (\ref{eq:jq'})  
1354: implies the conservation of charge 
1355: current $j=j_\uparrow+j_\downarrow=const.$, while the spin counterpart,
1356: the difference of the spin-polarized currents 
1357: $j_s=j_\uparrow-j_\downarrow$ is position dependent. 
1358: Other ``spin quantities,'' $X_s$,
1359: unless explicitly defined, are analogously expressed 
1360: with the corresponding (spin) polarization given by $P_X=X_s/X$.
1361: For example, the current polarization\footnote{This is 
1362: also referred to as a
1363: spin injection coefficient \cite{Rashba2000:PRB,Rashba2002:EPJ}.} 
1364: $P_j=j_s/j$, generally different
1365: from the density polarization
1366: $P_n=(n_\uparrow-n_\downarrow)/n$, is related to the conductivity polarization 
1367: $P_\sigma$ as 
1368: \begin{equation}
1369: \label{eq:Pj}
1370: P_j=2(\sigma_\uparrow \sigma_\downarrow/\sigma) \nabla \mu_s/j+P_\sigma
1371: \end{equation}
1372: where $\mu_s=\mu_\uparrow-\mu_\downarrow$. In terms of the
1373: average electrochemical potential $\mu=(\mu_\uparrow+\mu_\downarrow)/2$, 
1374: $P_\sigma$ further satisfies
1375: \begin{equation}
1376: \label{eq:mu}
1377: \nabla \mu=-P_\sigma \nabla \mu_s/2+j/\sigma.
1378: \end{equation}
1379: From Eqs.~(\ref{eq:elchem}) and (\ref{eq:jq'}) it follows that $\mu_s$
1380: satisfies the diffusion equation
1381: \cite{vanSon1987:PRL,Valet1993:PRB,Hershfield1997:PRB,Schmidt2000:PRB} 
1382: \begin{equation}
1383: \label{eq:mus}
1384: \nabla ^2 \mu_s=\mu_s/L_s^2,
1385: \end{equation}
1386: where the spin diffusion length is $L_s=(\overline{D}\tau_s)^{1/2}$ with
1387: the spin averaged diffusion coefficient 
1388: $\overline{D}=(\sigma_\downarrow D_\uparrow 
1389: +\sigma_\uparrow D_\downarrow)/\sigma
1390: ={\cal N}({\cal N}_\downarrow/D_\uparrow+{\cal N}_\uparrow/D_\downarrow)^{-1}$. 
1391: Using Eq.~(\ref{eq:elchem}) and the local charge quasineutrality 
1392: $\delta n_\uparrow+\delta n_\downarrow=0$
1393: shows that $\mu_s$ is proportional to the nonequilibrium 
1394: spin density $\delta s=\delta n_\uparrow - \delta n_\downarrow$ 
1395: ($s=s_0+\delta s=n_\uparrow-n_\downarrow$)
1396: \begin{equation}
1397: \label{eq:spin}
1398: \mu_s=\frac{1}{2 q} \frac{{\cal N}_\uparrow+{\cal N}_\downarrow} 
1399: {{\cal N}_\uparrow {\cal N}_\downarrow}
1400: \delta s.
1401: \end{equation}                                                                 
1402: Correspondingly,  $\mu_s$ is often referred to as the (nonequilibrium) 
1403: {\it spin 
1404: accumulation}\footnote{Spin accumulation is also relevant to a number of
1405: physical phenomena outside the scope of this article, for example, to the
1406: tunneling rates in the quantum Hall regime 
1407: \cite{MacDonald1999:PRL,Chan1999:PRL}.} 
1408: and is used to explain the GMR effect in CPP structures 
1409: \cite{Johnson1991:PRL,Valet1993:PRB,Hartmann:2000,Hirota:2002,Gijs1997:AP}. 
1410: 
1411: The preceding equations are simplified for the N region 
1412: by noting that 
1413: $\sigma_\lambda=\sigma/2$, $\sigma_s=0$, and 
1414: $D_\lambda=\overline{D}$.
1415: Quantities pertaining to a particular region are denoted by the index 
1416: F or N.
1417: 
1418: Equation ~(\ref{eq:mus}) has also been used to study the diffusive 
1419: spin-polarized transport and spin accumulation in  ferromagnet/superconductor 
1420: structures \cite{Jedema1999:PRB}. 
1421: Some care is needed to establish the appropriate
1422: boundary conditions at the F/N interface.
1423: In the absence of spin-flip scattering\footnote{The effects
1424: of non-conserving interfacial scattering on spin injection were considered 
1425: in \cite{Valet1993:PRB,Fert1996:PRB,Rashba2002:EPJ}.}
1426: at the F/N interface (which can arise, for example, 
1427: due to spin-orbit coupling or magnetic impurities) 
1428: the spin current is continuous 
1429: and thus $P_{jF}(0^-)=P_{jN}(0^+)\equiv P_j$ (omitting
1430: $x=0^\pm$ for brevity, and superscripts $\pm$ in other quantities). 
1431: These boundary conditions were used by 
1432: \textcite{Aronov1976:SPS,Aronov1976:JETPL} 
1433: without relating $P_j$ to the effect of the F/N contact or 
1434: material parameters in the F region.
1435: 
1436: Unless the F/N contact is highly transparent, $\mu_\lambda$ is discontinuous
1437: across the interface 
1438: \cite{Johnson1988:PRL,Valet1993:PRB,Hershfield1997:PRB,Rashba2000:PRB} 
1439: and the boundary condition is
1440: \begin{equation}
1441: \label{eq:jbc}
1442: \ j_\lambda(0)=\Sigma_\lambda [\mu_{\lambda N}(0)-\mu_{\lambda F}(0)],
1443: \end{equation}
1444: where 
1445: \begin{equation}
1446: \label{eq:sigma}
1447: \Sigma=\Sigma_\uparrow+\Sigma_\downarrow
1448: \end{equation}
1449: is the contact conductivity. For a free-electron model 
1450: $\Sigma_\uparrow \neq \Sigma_\downarrow$ can be simply 
1451: inferred from the effect of the exchange energy, which would 
1452: yield spin-dependent Fermi wave vectors and transmission
1453: coefficients. A microscopic determination of the 
1454: corresponding contact resistance [see Eq.~(\ref{eq:rc})]
1455: is complicated by the influence of disorder,
1456: surface roughness, and different scattering mechanisms and
1457: is usually obtained from  model calculations 
1458: \cite{Schep1997:PRB,Stiles2000:PRB}. Continued work on
1459: the first-principles calculation of F/N interfaces
1460: \cite{Stiles1996:JAP,Erwin2002:PRB} is needed for 
1461: a more detailed understanding of spin injection.
1462: From Eqs.~(\ref{eq:jbc}) and (\ref{eq:sigma})  
1463: it follows that
1464: \begin{eqnarray}
1465: \label{eq:musbc}
1466: \mu_{sN}(0)-\mu_{sF}(0)=2r_c(P_j-P_\Sigma)j, \\
1467: \label{eq:mubc}
1468: \mu_N(0)-\mu_F(0)=r_c ( 1-P_\Sigma P_j ) j,
1469: \end{eqnarray}
1470: where the effective contact resistance is 
1471: \begin{equation}
1472: \label{eq:rc}
1473: r_c=\Sigma/4\Sigma_\uparrow \Sigma_\downarrow.
1474: \end{equation}
1475: The decay of $\mu_s$, away from the interface, is characterized by the
1476: corresponding spin diffusion length
1477: \begin{equation}
1478: \label{eq:decay}
1479: \mu_{sF}=\mu_{sF}(0) e^{x/L_{sF}}, \quad
1480: \mu_{sN}=\mu_{sN}(0) e^{-x/L_{sN}}.  
1481: \end{equation}
1482: A nonzero value for $\mu_{sN}(0)$ implies the existence of nonequilibrium
1483: magnetization $\delta M$ in the N region (for noninteracting electrons
1484: $q \mu_s=\mu_B \delta M/\chi$, where $\chi$ is 
1485: the magnetic susceptibility).
1486: Such  a $\delta M$, as a result of electrical spin injection, was proposed
1487: by \textcite{Aronov1976:SPS} and first measured in metals by 
1488: \textcite{Johnson1985:PRL}.
1489: 
1490: By applying Eq.~(\ref{eq:Pj}), separately, to the F and N regions, 
1491: one can obtain the amplitude of spin accumulation in terms 
1492: of the current and density of states spin polarization 
1493: and the effective resistances $r_F$ and $r_N$, 
1494: \begin{equation}
1495: \label{eq:mus0}
1496: \mu_{sF}(0)=2r_F \left [P_j-P_{\sigma F} \right ]j, \quad
1497: \mu_{sN}(0)=-2r_N P_j j, 
1498: \end{equation}
1499: where
1500: \begin{equation}
1501: \label{eq:rs}
1502: r_N=L_{sN}/\sigma_N, \quad
1503: r_F=L_{sF} \sigma_F/(4\sigma_{\uparrow F}\sigma_{\downarrow F}).
1504: \end{equation}
1505: From Eqs.~(\ref{eq:mus0}) and 
1506: (\ref{eq:musbc}) the current polarization can be obtained as 
1507: \begin{equation}
1508: \label{eq:gamma}
1509: P_j=\left [r_c P_\Sigma+ r_F P_{\sigma F} \right] /r_{FN},
1510: \end{equation}
1511: where $r_{FN}=r_F+r_c+r_N$ is the effective equilibrium
1512: resistance of the F/N junction. 
1513: It is important to emphasize that a measured highly polarized current,
1514: representing an efficient spin injection, does not itself imply
1515: a large spin accumulation or a large density polarization, typically
1516: measured by optical techniques. In contrast to the derivation of $P_j$
1517: from Eq.~(\ref{eq:gamma}), determining $P_n$
1518: requires using 
1519: Poisson's
1520: equation or a condition of the local charge 
1521: quasineutrality.\footnote{Carrier density
1522: will also be influenced by the effect of screening, which changes with
1523: the dimensionality of the spin injection geometry \cite{Korenblum2002:PRL}.} 
1524: 
1525: It is useful to note\footnote{\textcite{Rashba2002:PC}.}
1526: that Eq.~(\ref{eq:gamma}), written as Eq.~(18) in \cite{Rashba2000:PRB}
1527: can be mapped to Eq.~(A11) from \cite{Johnson1987:PRB}, where it
1528: was first derived.\footnote{The substitutions are $P_j \rightarrow \eta^*$, 
1529: $P_\sigma \rightarrow p$,
1530: $P_\Sigma \rightarrow \eta$, $r_c \rightarrow [G(\xi-\eta^2)]^{-1}$,
1531: $r_N \rightarrow \delta_n/\sigma_n\zeta_n$,
1532: $r_F \rightarrow \delta_f/\sigma_f(\zeta_f-p_f^2)$, 
1533: $L_{sN,F} \rightarrow  \delta_{n,F}$,
1534: and $n,f$ label N and F region, respectively. 
1535: $\eta$, $\zeta_n$, and $\zeta_f$ are of the order of unity.
1536: To ensure that resistances and the spin diffusion lengths in 
1537: \textcite{Johnson1987:PRB} are positive, one must additionally have 
1538: $(\xi-\eta^2)>0$ and
1539: $(\zeta_i-p_i^2)>0$, $i=n,f$ (for normal and ferromagnetic regions, 
1540: respectively). In particular, assuming $\xi=\zeta_n=\zeta_f=1$ 
1541: a detailed correspondence between  Eq.~(\ref{eq:gamma}) and  Eq.~(A11) in 
1542: \cite{Johnson1987:PRB} is recovered. For example,  
1543: $r_c \rightarrow [G(\xi-\eta^2)]^{-1}$
1544: yields Eq.~(\ref{eq:rc}), where $\Sigma \rightarrow G$.}
1545: An equivalent form for $P_j$ in Eq.~(\ref{eq:gamma}) was obtained by 
1546: \textcite{Hershfield1997:PRB} and for $r_c=0$ results from 
1547: \textcite{vanSon1987:PRL} are recovered.
1548: 
1549: In contrast to normal metals \cite{Johnson1985:PRL,Johnson1988:PRBb} 
1550: and superconductors, for which  injection has been reported in 
1551: both  conventional 
1552: \cite{Johnson1994:APL}, and high temperature superconductors 
1553: \cite{Hass1994:PC,Vasko1997:PRL,Dong1997:APL,Yeh1999:PRB},
1554: creating a substantial current polarization by direct electrical spin 
1555: injection 
1556: from a metallic ferromagnet
1557: into a semiconductor proved to be more difficult
1558: \cite{Monzon1999:JMMM,Hammar1999:PRL,Filip2000:PRB,Zhu2001:PRL}.
1559: 
1560: By examining  
1561: Eq.~(\ref{eq:gamma}) we can both infer some possible limitations and
1562: deduce several experimental strategies for effective spin injection
1563: i.e. to increase $P_j$ into semiconductors.
1564: For a perfect Ohmic contact $r_c=0$, the typical resistance mismatch 
1565: $r_F \ll r_N$
1566: (where F is a metallic ferromagnet)
1567: implies inefficient spin injection with $P_j \approx r_F/r_N \ll 1$, 
1568: referred to
1569: as the {\it conductivity mismatch} problem by \textcite{Schmidt2000:PRB}. 
1570: Even in  the absence of the resistive contacts, effective spin injection into 
1571: a semiconductor can be achieved if the resistance mismatch is reduced 
1572: by using for spin injectors either a magnetic semiconductor or a highly 
1573: spin-polarized ferromagnet.\footnote{From Eq.~(\ref{eq:rs}) a half-metallic 
1574: ferromagnet implies a large $r_F$.}  
1575: 
1576: While there was early experimental evidence \cite{Alvarado1992:PRL}
1577: that employing resistive (tunneling) contacts could lead to 
1578: an efficient spin injection\footnote{The influence of the resistive contacts on 
1579: spin injection can also be inferred by 
1580: explicitly considering resistive contacts
1581: \cite{Johnson1987:PRB,Hershfield1997:PRB}.}
1582: a systematic understanding was provided by \textcite{Rashba2000:PRB}
1583: and supported with the subsequent experimental and theoretical studies 
1584: \cite{Smith2001:PRB,Fert2001:PRB,Rashba2002:EPJ,Takahashi2003:PRB,%
1585: Johnson2003:PRB,Johnson:2003}.
1586: As can be seen from Eq.~(\ref{eq:gamma}) the spin-selective resistive 
1587: contact $r_c \gg r_F,r_N$ (such as a tunnel or Schottky 
1588: contact) would contribute to effective spin injection with 
1589: $P_j\approx P_\Sigma$ 
1590: being dominated by 
1591: the effect
1592: $r_c$ and not the ratio $r_F/r_N$.\footnote{A similar result was stated
1593: previously by \textcite{Johnson1988:PRBa}.}
1594: This limit is also instructive to illustrate the principle of 
1595: spin filtering \cite{Esaki1967:PRL,Moodera1988:PRL,Hao1990:PRB,Filip2002:APL}.
1596: In a spin-discriminating transport process
1597: the resulting degree of spin polarization is changed. Consequently
1598: the effect of spin filtering, similar to spin injection, leads  
1599: to the generation of (nonequilibrium) spin 
1600: polarization. \footnote{While most of the schemes resemble a CPP geometry 
1601: [Fig.~\ref{gmr:1}(b)], there are also proposals for generating highly polarized 
1602: currents in a CIP-like geometry [Fig.~\ref{gmr:1}(a)] 
1603: \cite{Gurzhi2001:FNT,Gurzhi2003:P}.}
1604: For example, 
1605: at low temperature EuS and EuSe, discussed in 
1606: Sec.~\ref{sec:IVC}, can act as spin-selective barriers.
1607: In the extreme case,
1608: initially spin-unpolarized carriers (say, injected from a nonmagnetic material)
1609: via spin-filtering could attain a complete polarization.
1610: For a strong spin-filtering contact $P_\Sigma > P_{\sigma F}$, the sign of the
1611: spin accumulation (nonequilibrium magnetization) is reversed in the F and
1612: N regions, near the interface [recall Eq.~(\ref{eq:musbc})], in contrast
1613: to the behavior sketched in Fig.~\ref{inj:2}, where $\mu_{s F,N} > 0$.
1614: 
1615: The spin injection process alters the potential drop across the F/N interface
1616: because differences of spin-dependent electrochemical potentials on
1617: either side of the interface generate an effective resistance $\delta R$.
1618: By integrating Eq.~(\ref{eq:mu}) for N and F regions, separately,
1619: it follows  that $Rj=\mu_N(0)-\mu_F(0)+P_{\sigma F}\mu_{sF}(0)/2$,
1620: where $R$ is the junction resistance.
1621: Using Eqs.~(\ref{eq:mubc}), (\ref{eq:rs}), and (\ref{eq:gamma}) 
1622: allows us to express 
1623: $R=R_0+ \delta R$, where $R_0=1/\Sigma$ ($R_0=r_c$ if 
1624: $\Sigma_\uparrow=\Sigma_\downarrow$) 
1625: is the equilibrium resistance, in the absence of spin injection, and 
1626: \begin{eqnarray} 
1627: \delta R= [r_N ( r_F P^2_{\sigma F} +r_c P^2_{\Sigma} ) 
1628:         +r_F r_c ( P_{\sigma F} - P_{\Sigma})^2]  / r_{FN}, 
1629: \label{eq:delR} 
1630: \end{eqnarray}
1631: where $\delta R>0$ is the nonequilibrium resistance.
1632: Petukhov has shown \cite{Jonker2003:MRS} that Eqs.~(\ref{eq:gamma})
1633: and (\ref{eq:delR}) could be obtained by considering an equivalent
1634: circuit scheme with two resistors 
1635: $\tilde{R}_\uparrow$,
1636: $\tilde{R}_\downarrow$ connected in parallel, where 
1637: $\tilde{R}_\lambda=L_{sF}/\sigma_{\lambda F}+1/\Sigma_\lambda+2L_{sN}/\sigma_N$
1638: and $\tilde{R}_\uparrow + \tilde{R}_\downarrow=4r_{FN}$.
1639: For such a resistor scheme, by noting that 
1640: $j_\uparrow \tilde{R}_\uparrow= j_\downarrow \tilde{R}_\downarrow$,
1641: Eq.~(\ref{eq:gamma}) is obtained as 
1642: $P_j=-P_{\tilde{R}}\equiv-(\tilde{R}_\uparrow-\tilde{R}_\downarrow)
1643: /(\tilde{R}_\uparrow+\tilde{R}_\downarrow)$. $\delta R$ in
1644: Eq.~({\ref{eq:delR}) is then obtained as the difference between the
1645: total resistance of the nonequilibrium spin-accumulation region of the
1646: length $L_{sF}+L_{sN}$ [given by the equivalent resistance 
1647: $\tilde{R}_\uparrow \tilde{R}_\downarrow/ 
1648: (\tilde{R}_\uparrow + \tilde{R}_\downarrow)$] and the equilibrium 
1649: resistance for the same region, $L_{sF}/\sigma_F + L_{sN}/\sigma_N$.
1650: %\textcite{Rashba2002:EPJ} has pointed out that the nontrivial 
1651: %contribution of individual conductivities in Eq.~(\ref{eq:delR})
1652: %is likely to rule out a simple equivalent scheme that would represent
1653: %the effect of spin injection.  In a resistor scheme \cite{Gijs1997:AP}  
1654: %based on the two-current model (see \ref{sec:IB1}), 
1655: %extensively used to describe GMR,
1656: %individual resistances connected in series or in parallel yield 
1657: %positive contributions, unlike the 
1658: %term $-2r_F r_c P_{\sigma F} P_{\Sigma}$, which contributes to $\delta R$ 
1659: %in Eq.~(\ref{eq:delR}).
1660: 
1661: The concept of the excess resistance $\delta R$ can also be explained as 
1662: a consequence of the Silsbee-Johnson spin-charge coupling 
1663: \textcite{Silsbee1980:BMR,Johnson1985:PRL,Johnson1987:PRB}
1664: and illustrated by considering the simplified schemes in  
1665: Figs.~\ref{inj:1} and \ref{inj:2}.
1666: Accumulated spin 
1667: near the F/N interface, 
1668: together with a finite spin relaxation  and a finite spin diffusion,
1669: impedes the 
1670: flow of spins and acts as a ``spin bottleneck'' 
1671: \cite{Johnson1991:PRL}. A rise of $\mu_{sN}$ must be accompanied by the rise of 
1672: $\mu_{sF}$ [their precise alignment at the interface is given in 
1673: Eq.~(\ref{eq:musbc})] 
1674: or there will be a backflow of the nonequilibrium spin back into the F region. 
1675: Because both spin and charge are carried by electrons 
1676: in spin-charge coupling 
1677: the backflow of spin 
1678: driven by diffusion creates an additional resistance for the charge flow 
1679: across the 
1680: F/N interface. Based on an analogy with the charge transport across a 
1681: clean N/superconductor (S) interface (see Sec.~\ref{sec:IVA3}) 
1682: \textcite{vanSon1987:PRL} explained $\delta R$ by invoking
1683: the consequences of current conversion from spin-polarized, at 
1684: far to the left of the F/N interface, to completely unpolarized, at far right 
1685: in the N region.
1686: 
1687: The increase in the total resistance with spin injection can be most
1688: dramatic if the N region is taken to be a superconductor (S);
1689: see Fig.~\ref{inj:1}(c).  
1690: Spin injection depletes the superconducting condensate and can
1691: result in the switching to a normal state of much higher resistance 
1692: \cite{Vasko1997:PRL,Dong1997:APL,Yeh1999:PRB,Wei1999:JAP,Takahashi1999:PRL}.
1693: A critical review of possible spurious effects 
1694: in reported  experiments 
1695: \textcite{Gim2001:JAP} 
1696: has also stimulated the development of a novel detection technique
1697: which uses scanning tunneling spectroscopy combined with 
1698: pulsed quasiparticle  spin injection to minimize 
1699: Joule heating \cite{Ngai2003:P} (see Sec.~\ref{sec:IVA1}).
1700: In the S region the quasiparticle energy is 
1701: $E_k=(\xi_k^2+\Delta^2)^{1/2}$, where $\xi_k$ is the single particle
1702: excitation energy 
1703: corresponding to the wave vector ${\bf k}$
1704: and $\Delta$ is the superconducting gap
1705: [see Fig.~\ref{inj:1}(c)]. Such a dispersion relation results in a
1706: smaller diffusion coefficient and a longer spin flip time than
1707: in the N region, 
1708: while their product, the spin diffusion length, remains the same
1709: \cite{Yamashita2002:PRB}. Consequently,
1710: Eq.~(\ref{eq:mus}) also applies to  the diffusive 
1711: spin-polarized transport and spin accumulation in  ferromagnet/superconductor 
1712: structures \cite{Jedema1999:PRB,Yamashita2002:PRB}. 
1713: Opening of the superconducting gap implies that a superconductor is a
1714: low carrier system for spin, which is carried by quasiparticles 
1715: \cite{Takahashi2003:PRB}.
1716: 
1717: In the preceding analysis, appropriate for bulk, homogeneous,
1718: three-dimensional N and F regions and degenerate (semi)conductors,
1719: Poisson's equation was not invoked and the local charge neutrality
1720: $\delta n_\uparrow+\delta n_\downarrow$ was used only to derive
1721: Eq.~(\ref{eq:spin}).\footnote{For spin injection in
1722: nondegenerate semiconductors (with the carriers obeying the 
1723: Boltzmann statistics)
1724: there can be large effects due to 
1725: built-in fields and deviation
1726: from local charge neutrality, as discussed in Sec.~\ref{sec:IIC3}.}
1727: Focusing on bulk samples in which both the size of the F and N regions and the
1728: corresponding spin diffusion lengths are much larger than the Debye 
1729: screening length, one can find that 
1730: the quasineutrality condition, combined with the Eqs.~(\ref{eq:elchem})
1731: and (\ref{eq:einstein}), yields
1732: \begin{equation}
1733: \label{eq:phi}
1734: \phi=-\mu-P_{\cal N} \mu_s/2,
1735: \end{equation}
1736: where the density of states spin polarization of $P_{\cal N}$ vanishes 
1737: in the N region.
1738: At the contact $x=0$ there is a potential drop, even when $r_c=0$,
1739: which can be evaluated from
1740: Eqs.~(\ref{eq:mubc}) and (\ref{eq:phi}) as
1741: \begin{equation}
1742: \label{eq:contact}
1743: \phi_N(0)-\phi_F(0)=-r_c [1-P_\Sigma P_j] j + P_{{\cal N}F}(0)\mu_{sF}(0)/2.
1744: \end{equation}
1745: The creation of nonequilibrium spin in the N region results in the 
1746: spin EMF in the F/N structure 
1747: which 
1748: can be used to detect electrical spin injection,
1749: as depicted in Fig.~\ref{inj:1}. Within a simplified semi-infinite geometry
1750: for the F and N regions, we consider an effect of spin 
1751: pumping in the N region, realized either by electrical spin injection from
1752: another F region [as shown in Fig.~\ref{inj:1}(b)] or by optical pumping 
1753: (see Sec.~\ref{sec:IIB}).  The resulting potential drop can calculated by
1754: modifying $\mu_{sN}$ in Eq.~(\ref{eq:decay}), 
1755: \begin{equation}
1756: \mu_{sN}=\mu_{sN}(\infty)
1757: +[\mu_{sN}(0)-\mu_{sN}(\infty)]e^{-x/L_{sN}},  
1758: \end{equation}
1759: where $\mu_{sN}(\infty)$ 
1760: represents the effect of homogeneous spin pumping
1761: in the N region. To calculate the open circuit voltage ($j=0$)  the continuity
1762: of spin current at $x=0$ should be combined with the fact that $P_j j=j_s$. 
1763: From Eq.~(\ref{eq:Pj}) it follows that
1764: \begin{equation}
1765: \label{eq:js0}
1766: j_s(0)=2\frac{\sigma_\uparrow \sigma_\downarrow}{\sigma_F} 
1767: \frac{\mu_{sF}(0)}{L_{sF}}=
1768: -\frac{1}{2}\sigma_N \frac{\mu_{sN}(0)-\mu_{sN}(\infty)}{L_{sN}}, 
1769: \end{equation}                                                            
1770: while the discontinuity of $\mu_s$ in Eq.~(\ref{eq:musbc}) 
1771: yields\footnote{A missprint
1772: in $\mu_{sF}(0)$ from \textcite{Rashba2002:EPJ} has been corrected.}
1773: \begin{eqnarray} \nonumber
1774: \mu_{sF}(0)&=&(r_F/r_{FN})\mu_{sN}(\infty), \quad
1775: j_s(0)=\mu_{sN}(\infty)/2 r_{FN}, \\ 
1776: \label{eq:musNF}
1777: \mu_{sN}(0)&=&[(r_c+r_F)/r_{FN}]\mu_{sN}(\infty).
1778: \end{eqnarray} \nonumber
1779: By substituting this solution into Eq.~(\ref{eq:contact}),
1780: we can evaluate
1781: the contact 
1782: potential drop
1783: can be evaluated as
1784: \begin{equation}
1785: \label{eq:contact2}
1786: \phi_N(0)-\phi_F(0)= \left [ r_F P_{{\cal N}F}+r_c P_\Sigma \right ]
1787: \mu_{sN}(\infty)/2 r_{FN}. 
1788: \end{equation}
1789: The total potential drop (recall $j=0$) at the F/N junction\footnote{A similar
1790: potential drop was also calculated across
1791: a ferromagnetic domain wall \cite{Dzero2003:PRB}.} is \cite{Rashba2002:EPJ}
1792: \begin{equation}
1793: \label{eq:drop}
1794: \Delta \phi_{FN}=\phi_N(\infty)-\phi_F(-\infty)=P_j\mu_{sN}(\infty)/2. 
1795: \end{equation}
1796: where $P_j$ is given in Eq.~(\ref{eq:gamma}).
1797: In the context of the spin-detection scheme from Fig.~\ref{inj:1} and high
1798: impedance measurements at the N/F2 junction, the spin-coupled voltage $V_s$ 
1799: \cite{Silsbee1980:BMR,Johnson1985:PRL} was also found to be proportional
1800: to current polarization and the spin accumulation 
1801: ($\mu_s \propto \delta s \propto \delta M$) \cite{Johnson1988:JAP}.
1802: 
1803: 
1804: \subsubsection{\label{sec:IIC2} F/N/F junction}
1805: %---------------------------------------------------------------------------
1806: \label{F/N/F}
1807: 
1808: The above 
1809: analysis of the F/N bilayer can be readily extended to the geometry
1810: in which the infinite F regions are separated by an N region 
1811: of thickness $d$. 
1812: The quantities pertaining to the two ferromagnets
1813: are defined as in the case of an F/N junction and labeled
1814: by the superscripts L and R (left and right regions, respectively).
1815: It follows from Eq.~(\ref{eq:Pj}), 
1816: by assuming the continuity of the spin current at L,R, 
1817: that the difference of the spin-resolved
1818: electrochemical potential, responsible for the spin accumulation, is
1819: \begin{eqnarray} 
1820: \label{eq:musL}
1821: \mu_{sF}^L&=&2 r_F^L (P_j^L-P_{\sigma F}^L) j e^{x/L_{sF}^L}, \quad x<0, \\ 
1822: \nonumber
1823: \mu_{sN}&=&2 r_N \left [ P_j^R \cosh(x/L_{sN}) 
1824: -P_j^L\cosh[(d-x)/L_{sN}] \right  ]  \\   
1825: &\times& j/\sinh(d/L_{sN}), \quad 0<x<d, \\ 
1826: \label{eq:musN}
1827: \mu_{sF}^R&=&-2 r_F^R (P_j^R-P_{\sigma F}^R) j e^{(d-x/L_{sF}^R)}, \quad x>d, 
1828: \label{eq:musR}
1829: \end{eqnarray}
1830: where the current spin polarization $P_j^{L,R}$ at the two contacts
1831: in the F/N/F geometry can be expressed \cite{Rashba2002:EPJ} in terms of 
1832: the $P_j$ calculated for F/N junction with the infinite F and N regions
1833: in Eq.~(\ref{eq:Pj}) and the appropriate effective resistances. 
1834: By $P_{j\infty}^{L,R}$ we denote the $P_j$ calculated in Eq.~(\ref{eq:gamma})
1835: for at left
1836: and right contact (with the appropriate parameters for the F/N/F junction)
1837: as if it were surrounded by the infinite F and N regions.
1838: Analogously to the F/N junction, the consequence of the spin injection
1839: is the increase of the resistance $R=R_0+\delta R$, as compared to the 
1840: equilibrium value $R_0=(\Sigma^L)^{-1}+(\Sigma^R)^{-1}$. 
1841: The nonequilibrium resistance $\delta R$ is also always
1842: positive for spin-conserving contacts \cite{Rashba2002:EPJ,Rashba2000:PRB},
1843: in agreement with experiments on all-semiconductor trilayer structures
1844: \cite{Schmidt2001:PRL}; see Sec.~\ref{sec:IID3}.
1845: 
1846: Many applications 
1847: based on magnetic multilayers rely on the spin-valve effect
1848: in which
1849: the resistance changes due to the relative orientations of the magnetization
1850: in the two F regions. The geometry considered here is relevant for CPP GMR 
1851: \cite{Parkin:2002,Bass1999:JMMM,Gijs1997:AP} and all-metallic
1852: spin injection of \textcite{Johnson1985:PRL}.
1853: In particular, the resistance change between 
1854: antiparallel and parallel magnetization orientations in the two 
1855: ferromagnets can
1856: be expressed using current polarization of an infinite F/N junction 
1857: $P_{j\infty}^{L,R}$
1858: \cite{Rashba2002:EPJ}:
1859: \begin{equation}
1860: \label{eq:DelR}
1861: \Delta R= R_{\uparrow \downarrow} - R_{\uparrow \uparrow} =
1862: 4 P_{j\infty}^L P_{j\infty}^R \frac{r_{FN}^L r_{FN}^R r_N}{{\cal D} 
1863: \sinh(d/L_{sN})}, 
1864: \end{equation}
1865: where $r_{F}^{L,R}$, $r_c^{L,R}$, and $r_N$ are defined as in 
1866: the case of an F/N junction
1867: and
1868: \begin{eqnarray}
1869: \label{eq:det}
1870: {\cal D}&=&(r_F^L+r_c^L)(r_c^R+r_F^R)  \\ \nonumber
1871: &+& r_N^2 +r_N(r_F^L+r_c^L+r_c^R+r_F^R) \coth(d/L_{s N}).  
1872: \end{eqnarray}
1873: Up to a factor of 2, Eq.~(\ref{eq:DelR}) has also been obtained by 
1874: \textcite{Hershfield1997:PRB} using Onsager relations. 
1875: In the limit of a thin N region, 
1876: $d/L_{sN} \rightarrow 0$, 
1877: $\Delta R$ remains finite. In the opposite limit, for $d \gg L_{sN}$, 
1878: \begin{equation}
1879: \Delta R \sim  P_{j\infty}^L P_{j\infty}^R \exp(-d/L_{sN}).
1880: \label{rd}
1881: \end{equation}
1882: For a symmetric F/N/F junction, 
1883: where $r_{c,F}^{L}=r_{c,F}^{R}$, it follows that
1884: \begin{equation}
1885: \label{eq:DelRS}
1886: \Delta R=\frac{4 r_N (r_c P_\Sigma + r_F P_{\sigma F})}
1887: { {\cal D} \sinh(d/L_{sN})}. 
1888: \end{equation}
1889: 
1890: Considering the spin injection from F into a ballistic 
1891: N region in the presence of diffusive interfacial scattering,
1892: where the phase coherence is lost and the Boltzmann equation can be 
1893: applied,
1894: it is 
1895: instructive to reconsider the effect of contact 
1896: resistance \cite{Kravchenko2003:PRB}. We introduce the Sharvin resistance 
1897: $R_{\rm Sharvin}$ 
1898: \cite{Sharvin1965:ZETF}, arising in ballistic transport between the 
1899: two infinite
1900: regions connected by a contact (an orifice or a narrow and short constriction)
1901: of  radius much smaller than the mean free path, $a \ll l$. 
1902: In a 3D geometry the resistance is
1903: \begin{equation}
1904: \label{eq:sharvin}
1905: R_{\rm Sharvin}=\frac{4 \rho l}{3 \pi a^2}=\left [ \frac{e^2}{h} 
1906: \frac{k^2 A}{2 \pi} \right ]^{-1}, 
1907: \end{equation}
1908: where 
1909: $h/e^2\approx 25.81$ k$\Omega$ is the quantum of resistance per spin, 
1910: $A$ is the contact
1911: area, and $k$ is the Fermi wave vector.
1912: The opposite limit,  of  diffusive transport through the contact with
1913: $a \gg l$, corresponds to the the Maxwell or Drude resistance
1914: $R_{\rm Maxwel}=\rho/2 a$. The studies of intermediate cases provide an
1915: interpolation scheme between the $R_{\rm Maxwell}$ and $R_{\rm Sharvin}$ 
1916: for  various ratios of $a/l$
1917: \cite{Wexler1966:PPSL,Jansen1980:JPCSSP,deJong1994:PRB,Nikolic1999:PRB}.
1918: Following \textcite{Kravchenko2003:PRB} the effective contact resistance 
1919: $r_c=r_{c\uparrow}+r_{c\downarrow}$ 
1920: (recall that it is defined per unit area)
1921: is obtained as
1922: \begin{equation}
1923: \label{eq:contact22}
1924: r_{c\lambda}=(4 R_{\rm Sharvin}/A) (1-t^L_\lambda-t^R_\lambda)/t^L_\lambda,
1925: \end{equation}
1926: where $t^{L,R}_\alpha$ represent the transmission coefficients 
1927: for the electrons 
1928: reaching the contact from the left and from the right and satisfy
1929: $t^L+t^R \le 1$.
1930: For $r_c$ which would exceed the resistance of the N and F bulk regions 
1931: the spin injection efficiency can attain 
1932: $P_j\sim (r_{c \uparrow}-r_{c \downarrow})/r_c$
1933: \cite{Kravchenko2003:PRB},
1934: showing, similarly to the diffusive regime, the importance of the 
1935: resistive contacts
1936: to efficient spin injection. Connection with the results in the 
1937: diffusive regime can be obtained \cite{Kravchenko2003:PRB}
1938: by identifying 
1939: $r_{c\lambda}=1/4 \Sigma_\lambda$, where the contact conductivity 
1940: $\Sigma_\lambda$ was introduced in  Eq.~\ref{eq:sigma}. 
1941: 
1942: While most of the experimental results on spin injection are feasible in the 
1943: diffusive regime, there are many theoretical studies treating the
1944: ballistic case and phase-coherent transport both in F/N and
1945: F/N/F junctions
1946: \cite{Mireles2001:PRB,Matsuyama2002:PRB,Hu2001:PRB,Hu2001:PRL}.
1947: Simple models in which the N region is a degenerate semiconductor often adopt an
1948: approach developed first for charge transport in  junctions involving 
1949: superconductors, discussed in Sec.~\ref{sec:IVA3}. Considering spin-orbit
1950: coupling and the potential scattering at the F/N interface modeled by the 
1951: $\delta$-function, \textcite{Hu2001:PRL} have examined ballistic
1952: spin injection in the F/N junction. They show that even a 
1953: spin-independent barrier 
1954: can be used to enhance the spin injection and lead to an increase in the
1955: conductance polarization. First-principles calculations were also used for 
1956: ballistic spin injection from a ferromagnetic metal into a 
1957: semiconductor \cite{Wunnicke2002:PRB,Mavropoulos2002:PRB,Zwierzycki2003:PRB}.
1958: In the limit of coherent (specular) scattering\footnote{The 
1959: wave-vector 
1960: component 
1961: along the interface is conserved during scattering.}
1962: and high interfacial quality it was shown that 
1963: different band structure in the F and the N regions would contribute to a 
1964: significant 
1965: contact resistance and  an efficient spin injection \cite{Zwierzycki2003:PRB}.
1966: 
1967: 
1968: \subsubsection{\label{sec:IIC3} Spin injection through the space-charge region}
1969: %----------------------------------------------------------------------------
1970: 
1971: 
1972: Interfaces making up a  semiconductor often
1973: develop a space-charge region---a region of local macroscopic charges.
1974: Typical examples are the Schottky contact and the depletion layer in 
1975: {\it p-n} junctions. 
1976: While phenomenological models, such as the one introduced 
1977: in Sec.~\ref{sec:IIC1}, capture a remarkable wealth of spin injection physics, 
1978: they carry little information about spin-dependent processes
1979: right at the interfaces. Microscopic studies of spin-polarized transport and
1980: spin-resolved tunneling through space-charge regions are still 
1981: limited in scope. The difficulty lies in the need
1982: to consider self-consistently simultaneous charge accumulation and
1983: electric-field generation (through Poisson's equation), both affecting 
1984: transport. 
1985: Non-self-consistent analyses of a Schottky barrier spin injection 
1986: were performed in \cite{Prins1995:JPCM,Albrecht2002:PRB,Albrecht2003:PRB}, 
1987: while
1988: \textcite{Osipov2003:P} proposed an efficient spin 
1989: injection method using a $\delta$-doped Schottky contact.
1990: 
1991: Let us now consider spin injection through the depletion
1992: layer in magnetic {\it p-n} junctions 
1993: \cite{Zutic2002:PRL,Fabian2002:PRB,Zutic2003:APL}.
1994: The physics is based on drift and diffusion\footnote{Tunneling or field 
1995: emission becomes important, for example, 
1996: in thin Schottky barriers
1997: or in {\it p-n} junctions and heterostructures at large reverse biases 
1998: \cite{Kohda2001:JJAP,Johnston-Halperin2002:PRB,vanDorpe2003:P}.} 
1999: limited
2000: by carrier recombination and spin relaxation, as described in more detail
2001: in Sec.~\ref{sec:IVA4}. The transport equations are solved
2002: self-consistently  with Poisson's equation, 
2003: taking full account of electric field due to accumulated charges.
2004: Additional examples of magnetic {\it p-n} junctions are discussed
2005: in Sec.~\ref{sec:IVD}. 
2006: 
2007: The system is depicted in Fig.~\ref{deplete}. The {\it p-n} junction 
2008: has a magnetic 
2009: $n$-region\footnote{Equilibrium magnetization can be 
2010: a consequence of doping with magnetic impurities, yielding large carrier 
2011: $g$ factors, and applying magnetic
2012: field, or of using a ferromagnetic semiconductor 
2013: \cite{Ohno1998:S,Pearton2003:JAP}.}
2014: with a net equilibrium electron spin $P_{n0}^R$, where $R$ stands
2015: for the right (here $n$) region. Holes are assumed to be unpolarized. 
2016: An important issue to be resolved is whether
2017: there will be spin accumulation in the 
2018: $p$-region 
2019: if a forward bias is applied to the junction.
2020: In other words, will spin be injected across the depletion layer? Naively 
2021: the answer is yes, since spin is carried by electrons, but the result shown 
2022: in Fig.~\ref{deplete} suggests a more complicated answer. 
2023: At small biases there is
2024: no spin injection. This is the normal limit of diode operation,
2025: in which
2026: where the
2027: injected carrier density through the depletion region is still smaller 
2028: than the equilibrium carrier density. Only with
2029: bias increasing to the high-injection limit (typically above 1 V)  
2030: is spin injected.
2031:  
2032: \begin{figure}
2033: \centerline{\psfig{file=zutic_fig08.eps,width=1\linewidth,angle=0}}
2034: \caption{Spin injection through the space-charge region of a magnetic 
2035: {\it p-n}
2036: junction. The geometry 
2037: is depicted in the inset, which shows  
2038: a junction
2039: with a spin-split conduction band in the 
2040: $n$-region
2041: with spin-polarized electrons (solid circles) and 
2042: unpolarized holes (empty circles).
2043: Under applied forward bias $V$ the charge current flows to the right. 
2044: The curves, labeled by $V$, show the electron density polarization profiles 
2045: $P_n(x)$ for the depicted geometry and GaAs materials parameters. 
2046: The equilibrium density polarization in the $n$-region 
2047: is about 0.5.
2048: At low bias (0.8 V) 
2049: there is no
2050: spin injection. Spin injection, manifested by the
2051: increase of $P_n$ in the 
2052: $p$-region, 
2053: appears only at large biases
2054: (1.2 and 1.5 V), 
2055: where it is driven by electric drift \cite{Zutic2002:PRL}.
2056: Spin polarization of the current is discussed by  
2057: \textcite{Zutic2001:APL,Fabian2002:PRB}.
2058: Adapted from \onlinecite{Zutic2002:PRL}. 
2059: }
2060: \label{deplete}
2061: \end{figure}
2062: 
2063: The explanation for the absence of spin injection at small biases 
2064: and for nondegenerate doping levels (Boltzmann statistics are applicable)
2065: is as follows.
2066: On the $n$ side, there are more spin up than spin down electrons,
2067: $n_\uparrow > n_\downarrow$.
2068: If $2q\zeta$ is the spin splitting of the
2069: conduction band, $n_\uparrow(\zeta)/n_\uparrow(\zeta=0)=\exp(q\zeta/k_BT)$.
2070: Under a forward bias, electrons 
2071: flow to the 
2072: $p$-region. 
2073: The flow is limited by thermal activation over
2074: the barrier (given by the built-in electrostatic potential  
2075: minus bias),
2076: which is, for the spin up electrons, greater by $q\zeta$. For Boltzmann
2077: statistics, the rate of transmission of spin up electrons
2078: over the barrier is 
2079: $\sim \exp(-q\zeta/k_BT)$.
2080: Since current is proportional to both the carrier density and the transmission
2081: rate, the two exponential factors cancel out. Similarly for spin down.
2082: As a result, the spin-resolved current is unaffected by $2q\zeta$ 
2083: and there is no spin current flowing through the depletion
2084: layer. There is no spin accumulation.
2085: Spin injection appears only at large biases, where it is driven
2086: by electric drift leading to nonequilibrium 
2087: spin population already in the $n$-region \cite{Fabian2002:PRB,Zutic2002:PRL}. 
2088: In addition to spin injection, spin extraction has also been
2089: predicted in magnetic {\it p-n} junctions with a magnetic
2090: $p$-region 
2091: \cite{Zutic2002:PRL}.  Under a large bias, spin is extracted 
2092: (depleted) from the nonmagnetic $n$-region. 
2093: 
2094: Electric field in the bulk regions next to the space charge is important only 
2095: at large biases. It affects not only spin density, but spin diffusion
2096: as well. That spin injection efficiency can increase in the presence
2097: of large electric fields due to an increase in the spin diffusion length 
2098: (spin drag) was first shown by \textcite{Aronov1976:SPS},
2099: and was later revisited by other authors.\footnote{See, for example,
2100: \cite{Zutic2001:PRB,Fabian2002:PRB,Margulis1994:PB,Martin2003:PRB,%
2101: Yu2002:PRBa,Flensberg2001:PRB,Vignale2003:SST,Bratkovsky2003:P}.} 
2102: To be important, 
2103: the electric field
2104: needs to be very large,\footnote{The critical magnitude is
2105: obtained by dividing a typical energy, such as the thermal or Fermi 
2106: energy, by $q$ and by the spin diffusion length. At room temperature the thermal 
2107: energy is 25 meV, while the spin diffusion length can be several
2108: microns.} more than 100 V/cm 
2109: at room temperature. While such large fields are usually present inside the
2110: space-charge regions, they exist in the adjacent bulk regions only at
2111: the high injection limit and  
2112: affect transport and spin injection.
2113: In addition to electric drift, magnetic drift,
2114: in magnetically inhomogeneous semiconductors, can also enhance spin injection
2115: \cite{Fabian2002:PRB}. 
2116: 
2117: The following formula was obtained for spin injection at small
2118: biases \cite{Fabian2002:PRB}:
2119: \begin{equation}
2120: P_{n}^L=\frac{P_{n0}^L[1-(P_{n0}^R)^2]+
2121: \delta P_{n}^R(1-P_{n0}^LP_{n0}^R)}
2122: {1-(P_{n0}^R)^2+\delta P_{n}^R(P_{n0}^L-P_{n0}^R)},
2123: \label{eq:aL}
2124: \end{equation}
2125: where $L$ (left) and $R$ (right) label the edges of the
2126: space-charge (depletion) region of a {\it p-n} junction. Correspondingly, 
2127: $\delta P_n^R$ represents  the nonequilibrium
2128: electron  polarization, evaluated at $R$, arising from a 
2129: spin source. The case
2130: discussed in Fig.~\ref{deplete} is for $P_{n0}^L=\delta P_n^R=0$. Then
2131: $P_n^L=0$, in accord with the result of no spin injection. For a homogeneous
2132: equilibrium magnetization ($P_{n0}^L=P_{n0}^R$), 
2133: $\delta P_{n}^L=\delta P_{n}^R$;
2134: the nonequilibrium spin polarization is the same across the depletion layer.
2135: Equation (\ref{eq:aL}) demonstrates that only {\it nonequilibrium} spin, 
2136: already
2137: present in the bulk region, can be transferred through the depletion
2138: layer at small biases \cite{Zutic2001:PRB,Fabian2002:PRB}.  
2139: Spin injection of nonequilibrium spin is also very effective if 
2140: it proceeds from the 
2141: $p$-region 
2142: \cite{Zutic2001:PRB},
2143: which is the case for a spin-polarized solar cell \cite{Zutic2001:APL}. 
2144: The resulting spin 
2145: accumulation in the $n$-region 
2146: extends the spin diffusion range,
2147: leading to spin amplification---increase of the spin population away from 
2148: the spin source. 
2149: These results were also confirmed in the junctions
2150: with two differently doped $n$-regions 
2151: \cite{Pershin2003:PRL,Pershin2003:Pb}.
2152: Note, however, that  
2153: the term ``spin polarization density'' used in
2154: \textcite{Pershin2003:PRL,Pershin2003:Pb} 
2155: is actually the spin density $s=n_\uparrow-n_\downarrow$,
2156: not the spin polarization $P_n$. 
2157: 
2158: Theoretical understanding of spin injection has focused largely on 
2159: spin density while neglecting spin phase, which is 
2160: important for some 
2161: proposed spintronic applications. The problem of
2162: spin evolution in various transport modes (diffusion, tunneling,
2163: thermionic emission) remains to be investigated. Particularly relevant
2164: is the question of whether spin phase is conserved during spin injection.
2165: \textcite{Malajovich2001:N} showed, by studying spin evolution 
2166: in transport through a 
2167: n-GaAs/n-ZnSe
2168: heterostructure, 
2169: that the phase can indeed be preserved.
2170: 
2171: 
2172: \subsection{\label{sec:IID} Experiments on spin injection}
2173: %===========================================================================
2174: 
2175: \subsubsection{\label{sec:IID1} Johnson-Silsbee spin injection}
2176: %---------------------------------------------------------------------------
2177: 
2178: 
2179: The first spin polarization of electrons by electrical spin injection 
2180: \cite{Johnson1985:PRL} was demonstrated in a ``bulk wire'' of 
2181: aluminum on which 
2182: an array of thin film permalloy (Py) pads (with 70 \% nickel and 30 \% iron)  
2183: was deposited spaced in multiples of 50 $\mu$m, center to center 
2184: \cite{Johnson1988:PRBb}
2185: to serve as spin injectors and detectors. In one detection scheme
2186: a single ferromagnetic pad was used as a spin injector while the distance
2187: to the spin detector was altered by selecting different Py pads to
2188: detect $V_s$ and through the spatial decay of this spin-coupled voltage 
2189: infer $L_{sN}$.\footnote{The spin relaxation time in a ferromagnet is 
2190: often assumed to be
2191: very short.  Correspondingly, in the analysis of the experimental data, 
2192: both the spin 
2193: diffusion length and $\delta M$ are taken to vanish in the F region  
2194: \cite{Silsbee1980:BMR,Johnson1985:PRL,Johnson1988:PRBa,Johnson1988:PRBb}.}
2195: This procedure is illustrated in 
2196: Fig.~\ref{injexp:1}, where 
2197: the separation between the spin injector and detector $L_x$ is variable. 
2198: 
2199: \begin{figure}
2200: \centerline{\psfig{file=zutic_fig09.eps,width=1.\linewidth,angle=0}}
2201: \caption{Schematic top view of nonlocal, quasi-one-dimensional 
2202: geometry used by \textcite{Johnson1985:PRL}: F1 and F2, 
2203: the two metallic ferromagnets having magnetizations in the $x-z$ plane;  
2204: dotted lines, equipotentials characterizing electrical current flow;
2205: grey shading, diffusing population of nonequilibrium spin-polarized
2206: electrons injected at $x=0$, with darker shades corresponding to higher density 
2207: of polarized electrons. From \onlinecite{Johnson2002:SST}.} 
2208: \label{injexp:1}
2209: \end{figure}
2210: \textcite{Johnson1985:PRL} 
2211: point out that in the depicted geometry  
2212: there
2213: is no flow of the charge current for $x >0$ and that 
2214: in the absence of 
2215: nonequilibrium spins a voltage measurement between $x=L_x$ and $x=b$ gives zero.
2216: Injected spin-polarized electrons will diffuse symmetrically (at low current
2217: density the effect of electric fields can be neglected), and the 
2218: measurement of voltage will give a spin-coupled signal $V_s$ related to the 
2219: relative orientation of magnetizations in F1 and F2.\footnote{This method for
2220: detecting the effects of spin injection is also referred to as a
2221: potentiometric method.}
2222: The results, corresponding to the polarizer-analyzer detection and the
2223: geometry of Fig.~\ref{injexp:1}, 
2224: are given in Fig.~\ref{injexp:2}. An in-plane field 
2225: (${\bf B}$ $\|$ ${\bf \hat{z}}$), 
2226: of a magnitude several times larger
2227: than a typical field for magnetization reversal, $B_0 \approx 100$ G,
2228: is applied to define the direction of magnetization in the injector 
2229: and detector.  
2230: As the field sweep is performed, from negative to positive values, 
2231: at $B_{01}$ there is 
2232: a reversal of magnetization in one of the ferromagnetic films accompanied 
2233: by a sign change in the spin-coupled signal. As $B_z$ is further
2234: increased, at approximately $B_{02}$, there is 
2235: another reversal of magnetization,
2236: resulting in parallel orientation of F1 and F2 and a $V_s$ 
2237: of magnitude similar to that for the previous parallel orientation 
2238: when $B_z < B_{01}$. 
2239: 
2240: 
2241: \begin{figure}
2242: \centerline{\psfig{file=zutic_fig10.eps,width=1.\linewidth,angle=0}}
2243: \caption{Spin injection data from bulk Al wire sample.
2244: Negative  magnetic field is applied parallel to the magnetization (-$z$ axis) 
2245: in the two ferromagnetic regions. As the field is increased, at $B_{0,1}$ 
2246: magnetization in one of the ferromagnetic regions is reversed, and at $B_{0,2}$
2247: the magnetization in the other region is also reversed 
2248: (both are along +$z$ axis).
2249: Inset: amplitude of the observed Hanle signal as a function of orientation 
2250: angle $\phi$ of magnetic field. From \onlinecite{Johnson1985:PRL}.} 
2251: \label{injexp:2}
2252: \end{figure}
2253: 
2254: A more effective detection of the spin injection 
2255: is realized through measurements of the Hanle effect, 
2256: also discussed Secs.~\ref{sec:IIB} and \ref{sec:IIIA2}, and described by 
2257: the Bloch-Torrey equations \cite{Bloch1946:PR,Torrey1956:PR}
2258: [see Eqs.(~\ref{eq:relax:bloch1})--(\ref{eq:relax:bloch3})].
2259: The inset of Fig.~\ref{injexp:2} summarizes results from a
2260: series of Hanle experiments on a single sample. For the
2261: Hanle effect ${\bf B}$ must have a component perpendicular to the orientation 
2262: axes
2263: of the injected spins. Only projection of ${\bf B}$ perpendicular to 
2264: the spin axis
2265: applies a torque and dephases spins. The magnitude of ${\bf B}$, applied at
2266: an angle $\phi$ to the $z$-axis in the $y-z$ plane, is small enough that the
2267: magnetizations in ferromagnetic thin films remain in the $x-z$ plane (see 
2268: Fig.~\ref{injexp:1}). If, at ${\bf B}=0$, injected nonequilibrium magnetization
2269: is $\delta M(0) \hat{z}$ 
2270: then at finite field $\delta {\bf M}$ precesses about
2271: ${\bf B}$ with the cone of angle $2\phi$. After averaging over several cycles,
2272: only $\delta M(0) \cos \phi$, 
2273: the component $\|$ ${\bf B}$, will survive.
2274: The voltage detector\footnote{Recall from the discussion leading to 
2275: Eq.~(\ref{eq:drop}) that the spin-coupled signal $\propto \delta M$.}
2276: senses the remaining part of the magnetization projected on the axis
2277: of the detector 
2278: $\delta M(0) \cos \phi \times \cos\phi$ \cite{Johnson1988:PRBa}.
2279: The predicted angular dependence for the  amplitude of the Hanle signal 
2280: (proportional to the  depolarization of $\delta M$ in a finite field) 
2281: $[\delta M(0) -\delta M(0) \cos^2\phi]$ 
2282: is plotted in the inset together
2283: with the measured data.\footnote{The range of the angle $\phi$, 
2284: in the inset, is corrected from the one originally given in Fig.~3 of 
2285: \textcite{Johnson1985:PRL}.} Results confirm the first application of
2286: the Hanle effect to dc spin injection. 
2287: 
2288: The Hanle effect was also studied theoretically by solving
2289: the Bloch-Torrey equations for an arbitrary orientation, characterized by the
2290: angle $\alpha$, between the magnetization
2291: in F1 and F2 \cite{Johnson1988:PRBa}. From the Hanle curve [$V_s(B_\bot)$]
2292: measured at $T=4.3$ ($36.6$) K,  
2293: the parameters $L_s=450$ ($180$) $\mu$m and 
2294: $P_\Sigma=0.06$ ($0.08$) were extracted.\footnote{The fitting parameters are 
2295: $\tau_s$, $P_\Sigma$, and $\alpha$
2296: \cite{Johnson1988:PRBb}, and since the diffusion coefficient is obtained from 
2297: Einstein's relation $L_s$ is known.}
2298: This spin injection technique using a few pV resolution of
2299: a  superconducting quantum interference device (SQUID) 
2300: and with an 
2301: estimated $P_\Sigma\approx 0.07$ provided an accuracy able to detect 
2302: $P_n \approx 5\times 10^{-12}$, causing 
2303: speculating on that
2304: a single-spin sensitivity might be possible in  smaller samples 
2305: \cite{Johnson1985:PRL,Johnson1988:PRBb}.
2306: While in a good conductor, such as Al, the observed 
2307: resistance  change 
2308: $\Delta R$ was small ($\sim n\Omega$), 
2309: the relative change at low temperatures and for 
2310: $L_x \ll L_s$ was $\Delta R/R \approx 5 \%$, where $\Delta R$  
2311: is defined as in Eq.~(\ref{eq:tmr}), determined by the relative orientation of 
2312: the magnetization in F1 and F2, and $R$ is the
2313: Ohmic resistance \cite{Johnson2002:SST}.
2314: Analysis from Sec.~\ref{sec:IIC2} shows that the measurement 
2315: of $\Delta R$ could 
2316: be used to determine the product of injected current polarizations 
2317: in the two F/N junctions.
2318: 
2319: The studies of spin injection were extended to the thin-film geometry, 
2320: also known as the ``bipolar spin switch'' or ``Johnson spin transistor''
2321: \cite{Johnson1993:S,Johnson1993:PRL} 
2322: similar to the one depicted in Fig.~\ref{inj:1}(a). 
2323: The measured spin-coupled signals
2324: \footnote{$d \sim 100$ nm 
2325: was much smaller then the separation between F1 and F2 in bulk Al wires 
2326: \cite{Johnson1985:PRL}, and the amplitude of the Hanle effect
2327: was about $10^4$ larger \cite{Johnson2002:SST}.}
2328: in Au films were larger than the values obtained in bulk Al 
2329: wires \cite{Johnson1985:PRL,Johnson1988:PRBb}.
2330: A similar trend, 
2331: $V_s\sim 1/d$, potentially important for applications,
2332: was already anticipated by \textcite{Silsbee1980:BMR}.
2333: The saturation of this increase can be 
2334: inferred from Eqs.~(\ref{eq:DelR}) and (\ref{eq:det}) for 
2335: $d \ll L_{sN}$ 
2336: and has been discussed by \textcite{Hershfield1997:PRB} and
2337: \textcite{Fert1996:PRB}. 
2338: 
2339: When polarizer-analyzer detection was used, one of the fitting parameters
2340: from the measured data $P_\Sigma$ sometimes exceeded 1---which 
2341: corresponds to complete
2342: interfacial polarization. The origin of this discrepancy remains to 
2343: be fully resolved
2344: \cite{Hershfield1997:PRB,Fert1996:PRB,Johnson1993:PRL,Johnson2002:SST,Geux2000:APPA}.
2345: Results obtained from the Hanle effect, on similar samples, gave the 
2346: expected  $P_\Sigma <1$ values \cite{Johnson2002:SST}.\footnote{Theoretical 
2347: estimates for $V_s$  from which $P_\Sigma > 1$ was inferred
2348: are modified when one considers the Coulomb interaction and 
2349: proximity effects---near the N/F interface the spin splitting 
2350: of the carrier bands in the N region will be finite even at equilibrium.
2351: Model calculations \cite{Chui1995:PRL,Chui1995:PRB}, 
2352: which treat the F/N/F junction as a whole, show that the magnetic 
2353: susceptibility 
2354: $\chi$ in N can be much smaller than the free electron value and can increase 
2355: the predicted 
2356: $V_s \propto 1/\chi$. These corrections to the free-electron picture 
2357: of an F/N/F junction
2358: are smaller for larger $d$, as in the bulk-wire geometry of 
2359: \textcite{Johnson1985:PRL},
2360: where theoretical estimates of $V_s$ did not lead to $P_\Sigma > 1$.}  
2361: 
2362: A modification of the bipolar spin switch structure was used to demonstrate the
2363: spin injection into a niobium film \cite{Johnson1994:APL},
2364: realizing the theoretical assertion of \textcite{Aronov1976:SPJETP}
2365: that nonequilibrium spin could be injected into a superconductor.
2366: Two insulating Al$_2$O$_3$ films were inserted between F1 and F2 
2367: (both made of Py) 
2368: and a Nb film [see Fig.~\ref{inj:1}(b)]. The measurements were performed 
2369: near the superconducting transition temperature $T_c$
2370: with the data qualitatively similar, above and below $T_c$, to the spin-coupled
2371: voltage, as obtained in the magnetic-field sweep from Fig.~\ref{injexp:2}.
2372: The results were interpreted as support for enhanced  depletion of the 
2373: superconducting condensate (and correspondingly the reduction of 
2374: the critical current $I_c$) by spin-polarized  quasiparticles, as compared 
2375: to the usual spin-unpolarized quasiparticle injection. Related measurements 
2376: were recently performed in a CPP geometry \cite{Gu2002:PRB}, and the
2377: penetration depth of the 
2378: quasiparticle in the Nb films was measured to be $\sim 16$ nm, 
2379: as compared to $2$ nm in 
2380: \cite {Johnson1994:APL}. The corresponding temperature dependence of CPP GMR is
2381: well explained by the theory of \textcite{Yamashita2003:PRB} and the 
2382: modification
2383: of Andreev reflection (see Sec.~\ref{sec:IVA3}) by spin polarization.
2384: 
2385: The spin injection technique of Johnson and Silsbee was also applied to 
2386: semiconductors.
2387: Initial experiments on using a metallic ferromagnet to inject spin into a 
2388: two-dimensional
2389: electron gas (2DEG) showed only a very low ($\sim 1 \%$) efficiency
2390: \cite{Hammar1999:PRL}  
2391: for which various
2392: explanations were offered \cite{Monzon2000:PRL,vanWees2000:PRL,Hammar2000:PRL}. 
2393: However, stimulated by the proposal of \textcite{Rashba2000:PRB}
2394: to employ spin-selective diffusive contacts (Sec.~\ref{sec:IIC1}), 
2395: the subsequent 
2396: measurements have showed substantially more efficient spin injection 
2397: into a 2DEG
2398: after an insulating layer was inserted \cite{Hammar2001:APL,Hammar2002:PRL}. 
2399: The geometry
2400: employed is depicted in Fig.~\ref{injexp:1}. 
2401: In interpreting 
2402: the results, the spin-orbit coupling and the energy-independent density
2403: of states 
2404: at the Fermi level were taken into account \cite{Silsbee2001:PRB}. 
2405: This topic is reviewed by \textcite{Tang:2002}. 
2406: 
2407: \begin{figure}
2408: \centerline{\psfig{file=zutic_fig11.eps,width=0.5\linewidth,angle=-90}}
2409: \caption{Schematic representation of (a) local and (b) nonlocal geometry
2410: used to measure the effects of spin injection and spin accumulation.}
2411: \label{exp:3}
2412: \end{figure} 
2413: 
2414: \subsubsection{\label{sec:IID2} Spin injection into metals}
2415: %---------------------------------------------------------------------------
2416: 
2417: An important part of the operation of  CPP GMR structures is the presence 
2418: of nonequilibrium spin polarization in nonmagnetic metallic regions.
2419: Studies of spin-injection parameters
2420: in such systems have been reviewed by \textcite{Bass1999:JMMM} and
2421: \textcite{Gijs1997:AP}.   
2422: However, until recently, except for the work of Johnson and Silsbee,
2423: there were few  other experimental studies directly concerned with 
2424: spin injection into metals. A series of 
2425: experiments, \cite{Jedema2001:N,Jedema2002:JS,Jedema2002:Na,Jedema2002:APL} 
2426: at both low (4.2 K) and room temperature,
2427: were performed using  
2428: the van der Pauw geometry depicted in Fig.~\ref{exp:3}.
2429: In various structures \cite{Jedema2002:T}
2430: the two ferromagnetic regions (made of Py, Co, or Ni) 
2431: were chosen to be of different sizes to provide different  
2432: coercive fields, allowing an independent reversal of magnetization 
2433: in F1 and F2.
2434: The cross-shaped nonmagnetic region was made of Al or Cu \cite{Jedema2002:T}. 
2435: Nonlocal measurements, similar to the approach shown in Figs.~\ref{inj:1} and
2436: \ref{injexp:1} [discussed in \cite{Johnson1988:PRBb,Johnson1993:PRL}], were
2437: shown to simplify the extraction of spurious effects (for example,
2438: anisotropic magnetoresistance  and the Hall signal)
2439: from effects intrinsic to spin injection, as compared
2440: to the local or conventional  spin-valve geometry. 
2441: 
2442: In the first type of experiment the cross-shaped region was deposited directly
2443: over the F region 
2444: (Fig.~\ref{exp:3}), 
2445: and the spin-coupled resistance $\Delta R$, defined
2446: analogously to Eq.~(\ref{eq:tmr}),
2447: was measured as a function of an in-plane magnetic filed. A theoretical 
2448: analysis
2449: \cite{Jedema2001:N, Jedema2002:JS} was performed assuming no interfacial
2450: resistance ($r_c=0)$ and the continuity of the electrochemical 
2451: potentials at the F/N interface (see Sec.~\ref{sec:IIC1}).
2452: For a spin injection from Py into Cu, the maximum current polarization 
2453: was obtained to be $P_j \approx 0.02$ at 4.2 K. The results for $\Delta R$
2454: \cite{Jedema2001:N} scaled to the size of the samples
2455: used by \textcite{Johnson1993:S,Johnson1993:PRL} were interpreted to be 
2456: 3-4 orders of magnitude smaller. As discussed in Secs.~\ref{sec:IIC1} and 
2457: \ref{sec:IIC2}, the presence of interfacial
2458: spin-selective resistance can substantially change the spin injection 
2459: efficiency 
2460: and influence the resistance mismatch between the F and N regions
2461: [see Eq.~(\ref{eq:Pj})]. 
2462: Estimates of how these considerations would  affect the results of 
2463: \textcite{Jedema2001:N} 
2464: were given by \textcite{Jedema2002:Nb}
2465: as well as by others 
2466: \cite{Takahashi2003:PRB,Johnson2003:PRB}, 
2467: who analyzed 
2468: the importance 
2469: of multidimensional geometry. 
2470: In addition to
2471: comparing characteristic
2472: values of contact resistance
2473: obtained on different samples,\footnote{For example, the measured
2474: resistance of clean F/N contacts in CPP GMR \cite{Bussmann1998:IEEETM} was 
2475: used to infer that there is also a large contact resistance in 
2476: all-metal spin injection experiments \cite{Johnson2002:N}.}
2477: for a conclusive
2478: understanding it will be crucial to have 
2479: {\it in situ} measurements.
2480: 
2481: In analyzing data for the van der Pauw cross, a two-dimensional geometry 
2482: has an important effect---while the electric current is following  
2483: the paths depicted in Fig.~\ref{exp:3}, the spin current, through 
2484: the diffusion of nonequilibrium spin,
2485: would have similar flow in all four arms \cite{Johnson2002:SST}. This
2486: is different from the usual (quasi)one-dimensional analysis in which 
2487: spin and charge currents flow along the same paths. 
2488: For a full
2489: understanding of the van der Pauw cross geometry,  two-dimensional 
2490: modeling might be necessary \cite{Johnson2003:PRB,Takahashi2003:PRB}.
2491: 
2492: In the second type of experiment, tunneling contacts were fabricated 
2493: by inserting Al$_2$O$_3$ as an insulator into the regions where F1 and F2 
2494: overlapped with the cross. By applying a transverse field B$_z$ 
2495: (see Fig.~\ref{exp:3}) the precession of the injected nonequilibrium spin 
2496: was controlled and the amplitude of the Hanle effect was measured
2497: \cite{Jedema2002:Na,Jedema2002:APL},  as  outlined in Sec.~{\ref{sec:IID1}.
2498: From Co/Al$_2$O$_3$/Al/Al$_2$O$_3$/Co structures
2499: $L_s \approx 0.5$ $\mu$m was extracted at room temperature.
2500: The analysis of the Hanle signal was performed  by averaging contributions
2501: of different lifetimes \cite{Dyakonov:1984b}. This proved to be
2502: equivalent to the 
2503: \textcite{Johnson1988:PRBb}
2504: solution to the Bloch-Torrey equations.
2505: 
2506: 
2507: \subsubsection{\label{sec:IID3}  All-semiconductor spin injection}
2508: \label{magneticSM}
2509: 
2510: 
2511: \begin{figure}
2512: \centerline{\psfig{file=zutic_fig12.eps,width=0.8\linewidth,angle=0}}
2513: \caption{Schematic device geometry and band diagram of a spin LED:
2514: (a) Recombination of
2515:  spin-polarized electrons injected from
2516: the (II,Mn)VI spin aligner and  unpolarized holes injected from the 
2517: p-doped GaAs, in the intrinsic GaAs quantum well, 
2518: producing circularly polarized  light; (b) conduction and valence bands of 
2519: a spin aligner in an external magnetic field; 
2520: (c) sketch of the corresponding
2521: band edges and band offsets in the device geometry.
2522: In the quantum well, spin down electrons and unpolarized holes
2523: are depicted by solid and empty circles, respectively. 
2524: Adapted from \onlinecite{Fiederling1999:N}.}
2525: \label{injexp:3}
2526: \end{figure} 
2527: 
2528: If a magnetic semiconductor could be used as a robust spin injector 
2529: (spin aligner) 
2530: into a  nonmagnetic semiconductor it would facilitate the integration of 
2531: spintronics 
2532: and semiconductor-based electronics. Comparable resistivities of magnetic 
2533: and nonmagnetic semiconductors could provide efficient spin injection 
2534: [see Eq.~(\ref{eq:gamma}), with $r_F\approx r_N$] even without using  
2535: resistive contacts. Ultimately, for a wide range of applications and 
2536: for compatibility with
2537: complementary metal-oxide semiconductors (CMOS) 
2538: \cite{Wong1999:PIEEE}, it would be 
2539: desirable to be able to inject spin into silicon at room temperature.
2540: 
2541: Early studies \cite{Viglin1991:FTT,Osipov1990:PZETF,Viglin1997:PLDS,Osipov1998:PL}, 
2542: which have since largely been ignored, used a Cr- and Eu-based 
2543: chalcogenide ferromagnetic semiconductor (FSm) \cite{Nagaev:1983} as the spin 
2544: injector.\footnote{These materials, while more difficult to fabricate 
2545: than the subsequent 
2546: class of III-V ferromagnetic semiconductors,  have desirable properties of 
2547: providing injection of spin-polarized electrons 
2548: (with spin lifetimes typically
2549: much longer than for holes) and large spin splitting 
2550: [$\sim 0.5$ eV at 4.2 K for  n-doped HgCr$_2$Se$_4$ \cite{Nagaev:1983}]
2551: with nearly complete spin polarization 
2552: and a Curie temperature $T_C$ of up to 130 K 
2553: (HgCr$_2$Se$_4$) \cite{Osipov1998:PL}.}
2554: The experiments were motivated
2555: by the theoretical work of \textcite{Aronov1976:SPS,Aronov1976:SPJETP} 
2556: predicting that the ESR signal, proportional to the steady-state
2557: magnetization, would be changed by spin injection. The measurements of 
2558: \textcite{Viglin1991:FTT,Osipov1990:PZETF} prompted a 
2559: related prediction \cite{Margulis1994:PB} that spin injection could be 
2560: detected through changes in electric dipole spin resonance (EDSR).
2561: EDSR is the spin-flip resonance absorption for conduction electrons
2562: at Zeeman frequency, which is excited by the electric-field 
2563: vector of
2564: an incident electromagnetic wave. The theory of EDSR, developed 
2565: by \textcite{Rashba1961:SPSS} is extensively reviewed by
2566: \textcite{Rashba:1991}.    
2567: 
2568: Ferromagnetic semiconductor 
2569: spin injectors 
2570: formed {\it p-n} and {\it n-n} heterostructures with a 
2571: nonmagnetic semiconductor
2572: InSb. The choice of InSb is  very suitable due to its large 
2573: negative ($\sim -50$) $g$ factor \cite{McCombe1971:PRB}, 
2574: for detecting the effects of spin injection through ESR. The observed absorption
2575: and emission of microwave power \cite{Osipov1998:PL} was tuned by an 
2576: applied magnetic field (from 35 GHz at $\approx$ 400 G up to 1.4 THz at 20 kG)
2577: and only seen when electrons flowed  from FSm into an Sm region. 
2578: The injection to the lower Zeeman level increased the ESR absorption,
2579: while injection to the higher Zeeman level, leading to  population 
2580: inversion,
2581: generated  microwave emission. 
2582: 
2583: 
2584: The most recent experiments using semiconductor spin injectors can be grouped 
2585: into two different classes, In one approach (II,Mn)VI paramagnetic 
2586: semiconductors 
2587: were employed as the spin aligners. 
2588: These included
2589: CdMnTe \cite{Oestreich1999:APL}, BeMnZnSe \cite{Fiederling1999:N}, and
2590: ZnMnSe \cite{Jonker2000:PRB}.  
2591: In the second approach 
2592: ferromagnetic semiconductors like 
2593: (Ga,Mn)As \cite{Ohno1999:N,Chun2002:PRB,Mattana2002:PRL} were used.
2594: Both approaches were also employed to inject spins into 
2595: CdSe/ZnSe \cite{Seufert2004:PRB} and InAs
2596: \cite{Chye2002:PRB} quantum dots, respectively.
2597: 
2598: In (II,Mn)VI materials, at low Mn-concentration and at low temperatures,
2599: there is a  giant Zeeman splitting $\Delta E = g^* \mu_B H$ 
2600: \cite{Furdyna1988:JAP,Gaj:1988}
2601: of the conduction  band,
2602: in which $g^*$ is the effective electron  $g$ factor.
2603: Such splitting arises due to 
2604: {\it sp-d} exchange between the spins of
2605: conduction electrons and the S=5/2 spins of the localized Mn$^{2+}$ ions.
2606: The $g^*$ factor for $H \ne 0$
2607: can exceed\footnote{At low temperatures ($\sim1$ K)
2608: Cd$_{0.95}$Mn$_{0.05}$Se has $|g^*|>500$
2609: \cite{Dietl:1994}, while in n-doped (In,Mn)As $|g^*|>100$
2610: at 30 K \cite{Zudov2002:PRB}. Such large $g$ factors, in the
2611: presence of a highly inhomogeneous magnetic filed could lead to the
2612: charge carrier localization \cite{Berciu2003:PRL}.} 100
2613: and is given by \cite{Furdyna1988:JAP,Brandt1984:AP}
2614: \begin{equation}
2615: g^*=g+\alpha M/(g_{Mn} \mu_B^2 H),
2616: \label{eq:geff}
2617: \end{equation}
2618: where $g$ is the $H=0$ II-VI ``band'' value $g$, generally different from the
2619: free-electron value, magnetization
2620:  M$\propto$ $\langle S_z \rangle\propto B_s[(g_{Mn} \mu_B S H)/(k_B T)]$, B$_s$ 
2621: is the Brillouin function 
2622: \cite{Ashcroft:1976}, and $\alpha$ 
2623: is the exchange integral for $s$-like $\Gamma_6$ electrons
2624: (see Table I in Sec.~\ref{sec:IIB}), given by 
2625: \cite{Furdyna1988:JAP} 
2626: \begin{equation}
2627: \alpha\equiv\langle S|J_{sp-d}|S \rangle/V_0,
2628: \label{eq:alpha}
2629: \end{equation}
2630: where $J_{sp-d}$ is the electron-ion exchange coupling,
2631: and V$_0$ is the volume of an elementary cell.
2632: From Eqs.~(\ref{eq:geff}) and (\ref{eq:alpha}) it follows that 
2633: $g^*=g^*(H)$
2634: can even change its sign. 
2635: Similar analysis applies also to $g$ factors of holes, with the 
2636: Zeeman splitting of a valence band being typically several times larger
2637: than that of a conduction band \cite{Brandt1984:AP}.
2638: 
2639: (II,Mn)VI materials can be incorporated in high quality heterostructures 
2640: with different optically active III-V nonmagnetic semiconductors which,
2641: by providing 
2642: circularly polarized luminescence, can also serve
2643: as spin detectors. In this case 
2644: carriers
2645: are excited by electrical means and we speak of 
2646: electroluminescence  rather then photoluminescence.
2647: The selection rules for the recombination light are
2648: the same as discussed in Sec.~\ref{sec:IIB}.
2649: 
2650: Figure \ref{injexp:3} depicts a scheme for
2651: realization of all-semiconductor electrical spin injection and  
2652: optical detection \cite{Fiederling1999:N,Jonker2000:PRB}. 
2653: Displayed is a spin light-emitting diode (LED) \cite{Jonker1999:PA}
2654: in a Faraday geometry where both the applied B-field and the 
2655: direction of propagation of the
2656: emitted light 
2657: lie along the growth direction.
2658: Similar to a an ordinary
2659: LED \cite{Sze:1981}, electrons and holes recombine
2660: (in a quantum well or a {\it p-n} junction) and produce electroluminescence. 
2661: However,
2662: in a spin LED,  as a consequence of radiative recombination of spin-polarized 
2663: carriers, the emitted light is circularly polarized. 
2664: In experiments of  
2665: \textcite{Fiederling1999:N,Jonker2000:PRB}, at B$\approx 1$ T, 
2666: $T\approx 4$ K, 
2667: and forward bias, electrons entering from the $n$-contact were almost 
2668: completely polarized 
2669: in the spin down state as they left the spin aligner and are injected 
2670: across the (II,Mn)VI/AlGaAs interface. The electrons further traveled
2671: (by drift and diffusion) to an intrinsic GaAs quantum well (QW) 
2672: where they recombined with the unpolarized holes, which were injected from the 
2673: p-doped GaAs.\footnote{The spatial separation and  spin relaxation
2674: between the spin injection and the point of spin detection (in QW)
2675: make a fully quantitative analysis of the injected polarization more difficult. 
2676: It would be valuable to perform realistic calculations of a spin-polarized
2677: transport and spin injection which would treat the whole spin LED as a single
2678: entity.} 
2679: 
2680: The efficiency of  electrical spin injection across 
2681: the (II,Mn)VI/AlGaAs interface was studied \cite{Fiederling1999:N}
2682: using $P_{\rm circ}$ (defined in Sec.~\ref{sec:IIB}) of electroluminescence, 
2683: as a function of 
2684: B and  the thickness of the magnetic spin aligner 
2685: (0 nm, 3 nm, and 300 nm, respectively).
2686: $P_{\rm circ}$ increased with the
2687: thickness of the magnetic layer, suggesting the finite spin relaxation time
2688: needed for initially unpolarized electrons to relax into the lower 
2689: (spin down) Zeeman level. 
2690: The results of \textcite{Jonker2000:PRB} were similar to those of 
2691: \textcite{Fiederling1999:N} 
2692: for the thickest magnetic region. The behavior of $P_{\rm circ}$(B), up to the 
2693: saturation value (B$\approx 3$ T), 
2694: could be well explained by the magnetization
2695: described with the Brillouin function \cite{Furdyna1988:JAP,Gaj:1988}, 
2696: expected
2697: for the (II,Mn)VI semiconductors. In Fig.~\ref{injexp:3} the injected spin down 
2698: electrons are majority electrons with their magnetic
2699: moments parallel to the applied magnetic field.
2700: The principles of optical orientation discussed 
2701: in Sec.~\ref{sec:IIB} and the 
2702: selection 
2703: rules for GaAs sketched in Fig.~\ref{oo:1} are used to infer $P_n$ in a QW.
2704: 
2705: For a QW of approximately the same width (150 nm) the conversion of 
2706: $P_{\rm circ}$ to $P_n$  used by
2707: \textcite{Fiederling1999:N}  differed by a factor of 2 
2708: from that used by \textcite{Jonker2000:PRB}.
2709: \textcite{Fiederling1999:N}
2710: assumed that confinement effects were negligible, leading to the selection
2711: rules for a bulk GaAs 
2712: (recall $P_{\rm circ}=-P_n/2$, from Sec.~\ref{sec:IIB}).
2713: The maximum $P_{\rm circ}\approx 43 \%$ was interpreted 
2714: as implying nearly $90$\% 
2715: polarized injected electrons. \textcite{Jonker2000:PRB} 
2716: inferred $|P_n| \approx 50$\%, 
2717: from  $P_{\rm circ}=-P_n$ \cite{Weisbuch:1991}, as a consequence of QW 
2718: confinement and 
2719: lifting of the degeneracy 
2720: between light and heavy hole states 
2721: in the valence band ($\approx 5-6$ meV), see Fig.~\ref{oo:1}.
2722: Both results clearly demonstrated a 
2723: robust low-temperature
2724: spin injection using the spin LED's. 
2725: Subsequent studies 
2726: \cite{Park2000:APL,Jonker2001:APL,Stroud2002:PRL}
2727: have supported the lifting of degeneracy 
2728: between the light and heavy hole bands. The corresponding data are 
2729: shown in Fig.~\ref{injexp:4}.
2730: Similar values
2731: of $P_{\rm circ}$ were also measured in a resonant tunneling diode based
2732: on ZnMnSe \cite{Waag2001:JS,Gruber2001:APL}. Spin injection using the
2733: spin LED's,
2734: described above, is not limited to structures grown by molecular-beam epitaxy
2735: (MBE). It is also feasible using air-exposed interfaces \cite{Park2000:APL}
2736: similar to the actual fabrication conditions employed in conventional
2737: electronics. 
2738: 
2739: \begin{figure} 
2740: \centerline{\psfig{file=zutic_fig13.eps,width=1.0\linewidth}}
2741: \caption{(a)  Electroluminescence in a spin light-emitting diode (LED):
2742: (a) Electroluminescence (EL) spectra from a surface-emitting spin LED 
2743: with a Zn$_{0.94}$Mn$_{0.06}$Se contact for selected values of applied 
2744: magnetic field, 
2745: analyzed for  $\sigma^{\pm}$ (positive and negative helicity);
2746: the magnetic field is applied along the 
2747: surface normal (Faraday geometry) and the spectra are dominated by the heavy 
2748: hole exciton;
2749: (b)  magnetic field dependence of the EL circular polarization. 
2750: Adapted from \onlinecite{Jonker2001:APL}.}         
2751: \label{injexp:4}
2752: \end{figure} 
2753: 
2754: The robustness of measured $P_{\rm circ}$ was studied by intentionally
2755: changing the density of linear defects, from stacking faults at
2756: the ZnMnSe/AlGaAs  
2757: interface \cite{Stroud2002:PRL}. An approximate linear 
2758: decrease of $P_{\rm circ}$ with the density of stacking faults
2759: was shown to be
2760: consistent with the influence of spin-orbit interaction as
2761: modeled by Elliot-Yafet scattering (see Sec.~\ref{sec:IIIB1}) at the interface.
2762: The nonspherically symmetric defect potential (entering the spin-orbit 
2763: interaction) causes  a highly anisotropic loss of 
2764: spin polarization.  At small angles to the axis of growth
2765: [see Fig.~(\ref{injexp:3})], the probability of the spin flip
2766: of an injected electron is very high,  leading to a small spin polarization.
2767: These findings illustrate the importance of
2768: interface quality and the effect of defects on the spin
2769: injection efficiency, an issue not limited to semiconductor heterostructures.
2770: Related information is currently being sought by spatial imaging of the spin 
2771: polarization 
2772: in spin LED's \cite{Thruber2002:APL,Thruber2003:JMR} 
2773: using magnetic resonance
2774: force microscopy \cite{Sidles1995:RMP}.
2775: 
2776: (III,Mn)V ferromagnetic semiconductors are also used to inject spin  
2777: in a spin-LED structures as depicted in 
2778: Fig.~\ref{injexp:3}. Spin injection can be achieved with no external
2779: field, and the reports of high $T_C$ in some compounds suggest  
2780: that all-semiconductor spin LED's could operate at room temperature.
2781: The drawback, however, 
2782: is that the most 
2783: (III,Mn)V's have spin-polarized holes 
2784: (rather than electrons) as the main carriers 
2785: which, due to spin-orbit coupling, lose their polarization
2786: very quickly after being injected into a nonmagnetic semiconductor. 
2787: Consequently, results of the spin injection show only a small degree 
2788: of hole polarization. 
2789: 
2790: In the experiment of \textcite{Ohno1999:N}, 
2791: an intrinsic GaAs spacer of thickness $d$ was 
2792: introduced
2793: between the spin aligner (Ga,Mn)As 
2794: and the (In,Ga)As quantum well. 
2795: The electroluminescence  
2796: in a QW was measured perpendicular to
2797: the growth direction [the easy magnetization axis of (Ga,Mn)As and the
2798: applied magnetic field were both perpendicular to the growth
2799: direction]. The corresponding relation between the $P_{\rm circ}$ 
2800: and
2801: hole density  polarization 
2802: $P_p$ is not
2803: straightforward; 
2804: the analysis was  performed only on the 
2805: electroluminescence 
2806: [for possible difficulties see \textcite{Fiederling2003:APL}].
2807: A small measured signal ($P_{\rm circ}\sim 1 \%$ at $5$ K), 
2808: consistent with the 
2809: expectation for holes as the injected spin-polarized carriers,
2810: was also obtained in an additional experiment \cite{Young2002:APL}.
2811: $P_{\rm circ}$ was approximately independent of the GaAs thickness 
2812: ($d=20-420$ nm), a 
2813: behavior which remains to be understood considering that the 
2814: hole spins 
2815: should relax fast \cite{Hilton2002:PRL} 
2816: as they are
2817: transfered across the nonmagnetic semiconductor.\footnote{A possible 
2818: exception is QW, in which the effects of quantum confinement and 
2819: quenching spin-orbit coupling lead to longer $\tau_s$.}
2820: In contrast, for a repeated experiment \cite{Young2002:APL}
2821: using a Faraday
2822: geometry (as in Fig.~\ref{injexp:3}), with both measured electroluminescence  
2823: and $B$ along the growth direction, the same change of thickness
2824: $P_{\rm circ}$ was reduced from $7 \%$ to $0.5\%$. A highly
2825: efficient spin injection of $P_n \approx 80\%$ in GaAs has been realized
2826: using (Ga,Mn)As
2827: as a spin injector in a Zener diode structure \cite{vanDorpe2003:P}. 
2828: The detection
2829: employed the technique of an oblique Hanle effect, discussed in 
2830: the next section.
2831: 
2832: All-electrical spin injection studies of trilayer structures
2833: (II,Mn)VI/II-VI/(II,Mn)VI have displayed up to 25\% MR at $B\approx5$ T and
2834: $T=4$ K \cite{Schmidt2001:PRL}. A strong suppression of this MR signal
2835: at applied bias of $\sim10$ mV was attributed to the nonlinear regime
2836: of spin injection, in  which the effects of band bending 
2837: and charge accumulation
2838: at the (II,Mn)VI/II-VI interface were important \cite{Schmidt2002:P}.
2839: It would be instructive to analyze these measurements by adopting 
2840: the approach discussed in the context of magnetic {\it p-n} junctions
2841: (Secs.~\ref{sec:IIC3} and \ref{sec:IVD}), which self-consistently
2842: incorporates the effects of band bending and deviation from 
2843: local charge neutrality.
2844: 
2845: 
2846: \subsubsection{\label{sec:IID4} Metallic ferromagnet/semiconductor junctions}
2847: %---------------------------------------------------------------------------
2848: 
2849: A large family of  metallic ferromagnets, some of them highly spin polarized,
2850: offer the 
2851: possibility of spin injection at room temperature, even in the 
2852: absence of
2853: applied magnetic field. Spin injection into  (110) GaAs at room temperature 
2854: has been already 
2855: demonstrated 
2856: using vacuum tunneling from a polycrystalline Ni STM tip and 
2857: optical detection via circularly polarized luminescence 
2858: \cite{Alvarado1992:PRL,Alvarado1995:PRL}. 
2859: It was shown that the minority spin electrons (spin $\downarrow$ in 
2860: the context of
2861: metals; see Sec.~\ref{sec:IIA}) 
2862: in 
2863: Ni produced the dominant contribution to the tunneling current, and the    
2864: resulting polarization was inferred to be $P_n=(-31 \pm 5.6)\%$ 
2865: \cite{Alvarado1992:PRL}. 
2866: Even thought the spin injection in  future spintronic 
2867: devices will likely be implemented by some means other than 
2868: vacuum tunneling, this result 
2869: supports the importance of the tunneling contact for efficient 
2870: spin injection, as discussed in Sec.~\ref{sec:IIC1}. 
2871: Similar studies of spatially 
2872: resolved spin injection,
2873: sensitive to the topography of the GaAs surface, have employed 
2874: a single-crystal 
2875: Ni (100) 
2876: tip \cite{LaBella2001:S}.  At 100 K nearly fully 
2877: spin-polarized injection of electrons
2878: was reported.
2879: However,
2880: further analyses 
2881: of the measurements of 
2882: $P_{\rm circ}$ have substantially reduced these estimates 
2883: to $\approx 25$\% \cite{Egelhoff2002:S,LaBella2002:S}.
2884: 
2885: Direct spin injection from a ferromagnet into 
2886: a 2DEG,\footnote{For a 
2887: comprehensive review of 2DEG,  
2888: see \textcite{Ando1982:RMP}.} 
2889: motivated by the proposal of \textcite{Datta1990:APL} proposal,
2890: initially  
2891: showed only small effects \cite{Lee1999:JAP,Hammar1999:PRL,Gardelis1999:PRB}, 
2892: with $\Delta R/R \sim 1\%$,
2893: or effects within the noise \cite{Filip2000:PRB}. Such inefficiency could be 
2894: attributed to the resistance mismatch in the F and N regions, discussed 
2895: in Secs.~\ref{sec:IIC1} and \ref{sec:IIC2}. 
2896: The possibility of spurious effects arising from the Hall and 
2897: anisotropic magnetoresistance signals in
2898: similar structures was suggested earlier  \cite{Monzon1997:APL}
2899: as well as after the initial 
2900: experiments \cite{Monzon2000:PRL,vanWees2000:PRL,Tang:2002}.
2901: Control measurements have been performed to address these issues
2902: \cite{Hammar2000:PRL,Hammar2000:PRB}. This debate about the 
2903: presence/absence of spin injection effects via Ohmic contacts stimulated
2904: further studies, 
2905: but the experimental focus has since shifted to  other 
2906: approaches.
2907: 
2908: Spin injection via Schottky contacts at room temperature was demonstrated
2909: in a Fe/GaAs junction by \textcite{Zhu2001:PRL}, 
2910: who reported 
2911: detection of 
2912: $P_{\rm circ} \approx 2\%$ using spin LED structures and optical detection as 
2913: described in \ref{sec:IID3}.
2914: These studies were extended \cite{Ramsteiner2002:PRB}
2915: by using MBE to grow 
2916: MnAs,  a ferromagnetic metal,  
2917: on top of the GaAs to provide high-quality interfaces \cite{Tanaka2002:SST}. 
2918: There was no preferential behavior for spin injection using different 
2919: azimuthal orientations of the epitaxial MnAs layer, which could have been
2920: expected from the symmetry between 
2921: conduction-band wave functions in MnAs and GaAs.
2922: The tunneling properties of a Schottky barrier
2923: were discussed by 
2924: \textcite{Meservey1982:JAP,Gibson1985:JAP,Prins1995:JPCM,Kreuzer2002:APL}.
2925: The measured I-V curves display a complicated behavior 
2926: \cite{Hirohata2001:PRB,Isakovic2001:PRB} which can be significantly affected
2927: by the interface (midgap) states at the Schottky barrier \cite{Jonker1997:PRL}.
2928: A theoretical explanation of this behavior is still lacking.  
2929:  
2930: \begin{figure} 
2931: \centerline{\psfig{file=zutic_fig14.eps,width=1.0\linewidth}}
2932: \caption{Electroluminescence (EL) and polarization due to spin injection
2933: from Fe Schottky contact: (a) EL spectra 
2934: from a surface-emitting 
2935: spin LED with an Fe/AlGaAs
2936: Schottky tunnel contact  for selected values of applied magnetic field, 
2937: analyzed for $\sigma^{\pm}$ 
2938: circular polarization. The large difference in intensity
2939: between these components indicates successful spin injection from the Fe 
2940: into the GaAs
2941: QW, and reveals an electron spin polarization in the QW of 32\%.  
2942: The magnetic field is
2943: applied along the surface normal (Faraday geometry).  
2944: The spectra are dominated by the heavy
2945: hole (HH) exciton.  Typical operating parameters are 1 mA and 2 V.
2946: (b)  Magnetic-field dependence of the EL circular polarization of the 
2947: HH exciton.  The
2948: polarization tracks the hard axis magnetization of the Fe contact, 
2949: and saturates at an
2950: applied magnetic-field value 4$\pi$M = 2.2 T, at which the Fe 
2951: magnetization is entirely along
2952: the surface normal. From \onlinecite{Hanbicki2003:P}.}  
2953: \label{injexp:5}
2954: \end{figure} 
2955: 
2956: As discussed in Sec.~\ref{sec:IIC1}, tunnel contacts formed between a metallic 
2957: ferromagnet and a semiconductor can provide effective spin injections.
2958: Optical detection in spin LED structures, as discussed 
2959: in Sec.~\ref{sec:IID3},
2960: was used to show carrier polarization of $P_n \approx 30 \%$ using 
2961: Fe as a spin injector \cite{Hanbicki2003:P}.
2962: Experimental results are given in Fig.~\ref{injexp:5}.
2963: Some care has to be taken in defining the efficiency of the spin injection,
2964: normalized to the polarization of a ferromagnet, as used in related previous
2965: experiments~\cite{Hanbicki2002:APLa,Jansen2002:APL,Hanbicki2002:APLb}. 
2966: Furthermore,
2967: there are often different conventions for defining the sign of $P_n$ 
2968: \cite{Hanbicki2003:P},  used in the context of semiconductors and ferromagnetic 
2969: metals, as discussed in Sec.~\ref{sec:IIA}. 
2970: \textcite{Jiang2004:P} have demonstrated that MgO can be a suitable choice of an
2971: insulator for highly efficient spin injection into GaAs. Spin LED with 
2972: GaAs/AlGaAs quantum well was used to detect $P_n \approx 50 \%$
2973: at 100 K injected from CoFe/MgO (100) tunnel injector. While quantum
2974: well emission efficiency limits detection at higher temperatures ($>100$ K),
2975: the same tunnel injector should also be suitable for efficient spin
2976: injection even at room temperature. 
2977: 
2978: 
2979: An oblique Hanle effect \cite{D'yakonov1975:SPJETP} (see also 
2980: Secs.~\ref{sec:IIB}
2981: and \ref{sec:IID1})
2982: was used \cite{Motsnyi2002:APL,Motsnyi2003:PRB,VanDorpe2002:P}
2983: to detect spin injection, giving up to $P_n\approx 16 \%$, 
2984: at room temperature.
2985: The geometry used is similar to that sketched in Fig.~\ref{injexp:3}(a),
2986: with an insulating layer (AlO$_x$) separating the ferromagnetic spin injector
2987: and the (Al,Ga)As/GaAs spin LED.
2988: The magnetization easy axis lies in the plane of 
2989: the ferromagnet. 
2990: An oblique
2991: magnetic field is applied to give a net out-of-plane component of injected
2992: spin which could contribute to the emission of circularly polarized light.
2993: This approach allows one to apply a magnetic field several times smaller 
2994: than would be 
2995: needed to pull the magnetization out of plane
2996: [for Fe it is $\approx 2$ T \cite{Hanbicki2002:APLa}] and to measure polarized
2997: luminescence in a Faraday geometry. Furthermore, using standard measurements
2998: of the Hanle curve, one can extract separately the spin lifetime
2999: and carrier recombination time.
3000: 
3001: Hot-electron spin injection above the Schottky barrier is another method for 
3002: using a high polarization of metallic ferromagnets  
3003: to create a nonequilibrium 
3004: spin in a semiconductor even at room temperatures. 
3005: Typically such injection is performed in three-terminal, transistor-like
3006: devices, as discussed in Sec.~\ref{sec:IVE3}.
3007: 
3008: Direct electrical spin injection has also been demonstrated in organic 
3009: semiconductors \cite{Dediu2002:SSC}. MR measurements were performed in 
3010: an F/N/F junction, where F is 
3011: La$_{0.7}$Sr$_{0.3}$MnO$_3$ 
3012: (LSMO), a colossal magnetoresistive manganite, and N is 
3013: is sexithienyl (T$_6$), a  
3014: $\pi$-conjugated rigid-rod oligomer organic semiconductor. The decrease of MR
3015: with increasing thickness of the N region was used to infer 
3016: $L_{sN} \sim 100$ nm at room temperature. The resulting spin 
3017: diffusion length is a combination of low mobility, 
3018: $\sim$ 10$^{-4}$cm$^2$V$^{-1}$s$^{-1}$
3019: (about $\sim$ 10$^7$ times smaller than for the bulk GaAs),
3020: and long spin relaxation times, $\sim$ $\mu$s,\footnote{This is
3021: also a typical value
3022: for other organic semiconductors \cite{Krinichnyi2000:SM}, 
3023: a consequence of weak spin-orbit coupling
3024: \cite{Davis2003:JAP}.} as compared to the usual III-V inorganic semiconductors.
3025: Motivated by these findings \textcite{Xiong2004:N} have replaced one of
3026: the LSMO electrodes by Co. Different coercive fields in the two ferromagnetic
3027: electrodes allowed them to measure a spin-valve effect with
3028: $\Delta R / R \sim 40 \%$ at 11 K.
3029: Related theoretical studies of the ferromagnetic metal/conjugated polymer
3030: interfaces were reported by \textcite{Xie2003:PRB}. 
3031: 
3032: \section{\label{sec:III} Spin relaxation and spin dephasing}
3033: %===========================================================================
3034: 
3035: \subsection{\label{sec:IIIA} Introduction}
3036: %------------------------------------------------------------------------
3037: 
3038: Spin relaxation and spin dephasing are processes that lead to spin 
3039: equilibration
3040: and are thus of great importance for spintronics. The fact that 
3041: nonequilibrium
3042: electronic spin in metals and semiconductors lives relatively long (typically a
3043: nanosecond), allowing for spin-encoded information to travel macroscopic 
3044: distances, is 
3045: what makes spintronics a viable option for technology. After introducing the 
3046: concepts of spin relaxation and spin 
3047: dephasing times $T_1$ and $T_2$, 
3048: respectively, 
3049: which
3050: are commonly called $\tau_s$ throughout this paper, we
3051: discuss four major physical mechanisms responsible for spin equilibration in 
3052: {\it nonmagnetic} electronic
3053: systems: Elliott-Yafet, D'yakonov-Perel', Bir-Aronov-Pikus, and 
3054: hyperfine-interaction 
3055: processes. We then survey recent works on electronic spin relaxation in 
3056: nonmagnetic metals and semiconductors, using the important examples of
3057: Al 
3058: and GaAs for illustration.
3059: 
3060: \subsubsection{\label{sec:IIIA1} $T_1$ and $T_2$}
3061: 
3062: Spin relaxation and spin dephasing of a spin ensemble are traditionally defined 
3063: within the
3064: framework of the Bloch-Torrey equations~\cite{Bloch1946:PR,Torrey1956:PR} 
3065: for magnetization dynamics. For mobile electrons, spin 
3066: relaxation time $T_1$ (often called longitudinal or spin-lattice time) and
3067: spin dephasing time $T_2$ (also transverse or decoherence time) are
3068: defined via the equations for the spin precession, decay, and diffusion of 
3069: electronic 
3070: magnetization $\bf M$ in an applied magnetic field ${\bf B}(t)=B_0\hat{\bf 
3071: z}+{\bf B}_1(t)$, 
3072: with a static longitudinal component $B_0$ (conventionally in the $\hat{z}$ 
3073: direction) and, frequently, a transverse oscillating part ${\bf B}_1$ 
3074: perpendicular to $\hat{\bf z}$
3075:  ~\cite{Torrey1956:PR,Kaplan1959:PR}:
3076: \begin{eqnarray} 
3077: \label{eq:relax:bloch1}
3078: \frac{\partial M_{x}}{\partial t}&=&\gamma ({\bf M} \times {\bf B})_x-
3079: \frac{M_x}{T_2}+D\nabla^2 M_x, \\
3080: \label{eq:relax:bloch2} 
3081: \frac{\partial M_{y}}{\partial t}&=&\gamma ({\bf M} \times {\bf B})_y-
3082: \frac{M_y}{T_2}+D\nabla^2 M_y, \\
3083: \label{eq:relax:bloch3}
3084: \frac{\partial M_{z}}{\partial t}&=&\gamma 
3085: ({\bf M} \times {\bf B})_z-\frac{M_z-M^0_z}{T_1}+D\nabla^2 M_z.
3086: \end{eqnarray}
3087: Here $\gamma=\mu_B g/\hbar$ is the electron gyromagnetic ratio 
3088: ($\mu_B$ is the Bohr 
3089: magneton 
3090: and $g$ 
3091: is the electronic $g$ factor), $D$ is the diffusion coefficient (for simplicity
3092: we assume an isotropic or a cubic solid with scalar $D$), and 
3093: $M_z^0=\chi B_0$ is the thermal equilibrium magnetization with $\chi$ 
3094: denoting the system's static susceptibility. 
3095: The Bloch equations are phenomenological, describing quantitatively 
3096: very well the dynamics of mobile electron spins (more properly, magnetization)
3097: in experiments such as conduction electron spin resonance and optical 
3098: orientation. 
3099: Although relaxation and decoherence processes in a 
3100: many-spin system are generally
3101: too complex to be fully described by only two parameters, 
3102: $T_1$ and
3103: $T_2$ are nevertheless an extremely robust and convenient measure 
3104: for quantifying 
3105: such processes in many cases of interest. 
3106: To obtain microscopic expressions for spin relaxation and dephasing times,
3107: one starts with a microscopic description of the spin system (typically
3108: using the density-matrix approach), derives the magnetization dynamics, 
3109: and compares 
3110: it 
3111: with the Bloch equations to extract $T_1$ and $T_2$.
3112: 
3113: Time $T_1$ is the time it takes for the longitudinal magnetization to reach 
3114: equilibrium. 
3115: Equivalently, it is the time of thermal equilibration of the spin population 
3116: with 
3117: the lattice.  
3118: In $T_1$ processes an energy has to 
3119: be taken from the spin system, usually by phonons, to the lattice. 
3120: Time $T_2$ is classically the time it takes for an 
3121: ensemble of transverse electron spins, initially precessing in phase  
3122: about the 
3123: longitudinal
3124: field, to lose their phase due to spatial and temporal fluctuations of the 
3125: precessing frequencies. 
3126: For an ensemble of mobile electrons the measured $T_1$ and $T_2$ come about by
3127: averaging spin over the thermal distribution of electron momenta. Electrons
3128: in different momentum states have not only different spin-flip characteristics,
3129: but also slightly different $g$ factors and thus different precession 
3130: frequency. This is analogous to precession frequencies fluctuations
3131: of localized spins due to inhomogeneities in the static field $B_0$. 
3132: However, since momentum scattering (analogous to intersite hopping or exchange 
3133: interaction 
3134: of localized spins) typically proceeds much faster than spin-flip scattering,
3135: the $g$-factor-induced broadening is inhibited by motional 
3136: narrowing\footnote{Motional (dynamical) 
3137: narrowing is an inhibition of  phase change
3138: by random fluctuations \cite{Slichter:1989}. 
3139: Consider a spin rotating with frequency
3140: $\omega_0$. The spin phase changes by 
3141: $\Delta \phi = \omega_0 t$ 
3142: over time $t$. If the spin is subject to a random force that 
3143: makes spin precession equally likely clockwise and anticlockwise,
3144: the average spin phase does not change, but the root-mean-square phase change
3145: increases with time as 
3146: $(\langle \Delta^2 \phi \rangle)^{1/2} \approx 
3147: (\omega_0 \tau_c) (t/\tau_c)^{1/2}$,
3148: where $\tau_c$ is the correlation time of the random force, or the average
3149: time of spin precession in one direction. 
3150: This is valid for rapid fluctuations, $\omega_0 \tau_c \ll 1$.
3151: The phase relaxation time  
3152: $t_\phi$ is defined as the time over which the phase 
3153: fluctuations reach unity: $1/t_\phi=\omega_0^2 \tau_c$.}
3154: and
3155: need not be generally considered as contributing to $T_2$ 
3156: [see, however,~\cite{Dupree1967:PSS}]. Indeed, motional narrowing of the 
3157: $g$-{factor
3158: fluctuations, $\delta g$, 
3159: gives a contribution to $1/T_2$ of the order of 
3160: $\Delta \omega^2\tau_p$, where the $B_0$-dependent precession frequency
3161: spread is 
3162: $\Delta \omega = (\delta g/g) \gamma B_0$ 
3163: and $\tau_p$ is the momentum scattering time. For $B_0$ fields of the order 
3164: 1 T, 
3165: scattering times of 1 ps, and $\delta g$ as large as 0.01, 
3166: the ``inhomogeneous broadening'' is a microsecond, which is much 
3167: more than 
3168: the observed values for $T_2$. Spatial inhomogeneities of $B_0$, 
3169: like those coming 
3170: from hyperfine fields, are inhibited by motional narrowing, too, due to the
3171: itinerant nature 
3172: of electrons. For localized electrons (e.g., for donor states in 
3173: semiconductors),
3174: spatial inhomogeneities play an important role 
3175: and are often observed to affect 
3176: $T_2$. 
3177: To describe such reversible phase losses, which can potentially be 
3178: eliminated by spin-echo experiments, sometimes the symbol $T_2^*$ 
3179: ~\cite{Hu2001:lanl}
3180: is used to describe spin dephasing of ensemble spins, while the symbol $T_2$ is
3181: reserved for irreversible loss of the ensemble spin phase. 
3182: In general, $T_2^* \le T_2$, 
3183: although for conduction electrons to a very good approximation $T_2^*=T_2$. 
3184: 
3185: In isotropic and cubic solids $T_1=T_2$ if $\gamma B_0 \ll 1/\tau_c$, 
3186: where $\tau_c$ is the so-called correlation or interaction time: $1/\tau_c$ is
3187: the rate of change of the effective dephasing magnetic field 
3188: (see footnote 70).  
3189: Phase losses occur during time intervals of $\tau_c$.
3190: As shown below, 
3191: in electronic systems $\tau_c$ is given either by $\tau_p$ or by the
3192: time of the interaction of electrons with phonons and holes. Those
3193: times are typically smaller than a picosecond, so $T_1=T_2$ is fulfilled for 
3194: magnetic
3195: fields up to 
3196: several tesla. The equality between the relaxation and 
3197: dephasing
3198: times was noticed first in the context of 
3199: NMR~\cite{Bloch1946:PR,Wangsness1953:PR}
3200: and later extended to electronic spin systems 
3201: ~\cite{Pines1955:PR,Andreev1958:JETP}.
3202: A qualitative reason for $T_1=T_2$ is that if the phase acquires a random 
3203: contribution 
3204: during a short time interval $\tau_c$, the energy uncertainty of the spin 
3205: levels 
3206: determining the longitudinal spin is greater than the Zeeman splitting $\hbar 
3207: \gamma B_0$ 
3208: of the levels. The splitting then does not play a role in dephasing, and 
3209: the dephasing field will act equally on the longitudinal and
3210: transverse spin. In classical terms, spin that is 
3211: oriented along the direction of the 
3212: magnetic
3213: field can precess a full period about the perpendicular fluctuating field, 
3214: feeling
3215: the same dephasing fields as transverse components. As the external field 
3216: increases,
3217: the precession angle of the longitudinal component is reduced, inhibiting 
3218: dephasing.
3219: 
3220: The equality of the two times is very convenient for comparing
3221: experiment and theory, since measurements usually yield $T_2$, 
3222: while theoretically it is often more convenient to calculate $T_1$. In many
3223: cases a single symbol $\tau_s$ is used for spin relaxation and dephasing
3224: (and called indiscriminately either of these terms), 
3225: if it does not matter what 
3226: experimental situation is involved, or if one is working at small magnetic 
3227: fields.\footnote{Sometimes one finds as spin relaxation
3228: time $2\tau_s$. 
3229: While this is correct for just one spin state, 
3230: conventionally by spin relaxation we mean magnetization relaxation, 
3231: in which each spin flip adds
3232: to the equilibration of both spin up and spin down states, doubling the
3233: rate for magnetization relaxation.} Throughout this paper we adopt this
3234: notation to avoid unnecessary confusion.
3235: 
3236: If the system is anisotropic, the equality $T_1=T_2$ no longer holds, even in
3237: the case of full motional narrowing of the spin-spin
3238: interactions and $g$-factor 
3239: broadening. Using simple qualitative analysis
3240: \textcite{Yafet:1963} showed that, while there is no general relation
3241: between the two times, the inequality $T_2 \le 2 T_1$ holds, and that
3242: $T_2$ changes with the direction by at most a factor of 2. In the motionally
3243: narrowed case this difference between $T_1$ and $T_2$ can be considered
3244: as arising from the tensorial nature of the spin relaxation rate. Specific
3245: examples of this will be discussed in studying spin relaxation in 
3246: two-dimensional semiconductor heterostructures (Sec.~\ref{sec:IIIB2}).
3247: 
3248: Finally, we discuss the connection between $\tau_s$ and the single-spin 
3249: decoherence 
3250: time\footnote{To distinguish ensemble and individual spin dephasing, we use
3251: the term decoherence in relation to single, or a few, spins.},
3252: $\tau_{sc}$, which is the single-spin 
3253: correlation time. Time $\tau_{sc}$ becomes important for applications 
3254: in spin-based 
3255: quantum computing \cite{Loss1998:PRA,Hu2000:PRA}, 
3256: where spin coherence needs to last for at least $10^4-10^6$ gate
3257: operations for the computation to be fault tolerant \cite{Preskill1998:PRSL}. 
3258: The relative magnitudes of $\tau_s$ and $\tau_{sc}$ depend on many factors. 
3259: Often
3260: $\tau_{sc} < \tau_s$, as is the case for a direct exchange interaction causing 
3261: single-spin decoherence, while contributing to ensemble dephasing only as a
3262: dynamical averaging factor (the exchange interaction preserves total spin).  
3263: An
3264: analogy with momentum scattering may be helpful. Electron-electron collisions 
3265: lead to
3266: individual momentum equilibration while conserving the total momentum 
3267: and hence 
3268: do not contribute 
3269: to charge current
3270: (unless umklapp processes are taken into account).
3271: A momentum scattering time obtained 
3272: from conductivity (analogous to $\tau_s$) 
3273: would thus
3274: be very different from a single-state momentum time (analogous to 
3275: $\tau_{sc}$).
3276: It is not clear at present how much $\tau_s$ and $\tau_{sc}$ differ for 
3277: different 
3278: materials under different conditions, although it is often, with little 
3279: justification, 
3280: assumed that they are similar. As \textcite{Dzhioev2002:PRB} 
3281: recently suggested,
3282: $\tau_{sc}$ can be smaller than $\tau_s$ by several orders of magnitude 
3283: in n-GaAs at low doping densities where electrons are localized in 
3284: donor states, see also Sec.~\ref{sec:IIID3a}. 
3285: We note that it is $\tau_s$ that is relevant for spintronics (spin transport) 
3286: applications, while $\tau_{sc}$ is relevant for solid-state quantum computing. 
3287: Much remains to be learned
3288: about $\tau_{sc}$.
3289: 
3290: 
3291: \subsubsection{\label{sec:IIIA2} Experimental probes}
3292: %---------------------------------------------------------------------------
3293: 
3294: Experiments detecting spin relaxation and decoherence of conduction electrons 
3295: can be grouped into two
3296: broad categories: (a) those measuring spectral characteristics of magnetization 
3297: depolarization
3298: and (b) those measuring time or space correlations of 
3299: magnetization.
3300: 
3301: Experiments of type (a) are exemplified by conduction-electron spin 
3302: resonance (CESR) 
3303: and optical orientation 
3304: in combination with the Hanle effect. CESR was the first technique used to 
3305: detect the spin
3306: of conduction electrons in metals ~\cite{Griswold1952:PR,Feher1955:PR}
3307: and donor states in semiconductors like Si \cite{Feher1959b:PR}. 
3308: In GaAs, spin resonance techniques are aided by other measurements, e.g., 
3309: optical detection \cite{Weisbuch1977:PRB}, photoconductivity 
3310: \cite{Stein1983:PRL},
3311: or magnetoresistance \cite{Dobers1988:PRB}.
3312: The technique measures signatures of resonant absorption of microwaves by a 
3313: Zeeman-split
3314: spin system of conduction electrons. Typically changes in surface 
3315: impedance and 
3316: in 
3317: the transmission coefficient of the sample are observed. By comparing the 
3318: absorption
3319: resonance curve with theory~\cite{Dyson1955:PR,Kaplan1959:PR} 
3320: one can obtain
3321: both  the carrier $g$ factor and $T_2$. 
3322: 
3323:  
3324: Optical spin orientation~\cite{Meier:1984} is a method
3325: of exciting spin-polarized carriers (electrons and holes) in direct-gap 
3326: semiconductors like GaAs by absorption of circularly polarized light 
3327: (see Sec.~\ref{sec:IIB}). 
3328: The injected spin polarization can be detected by observing circularly 
3329: polarized luminescence resulting from recombination of the spin-polarized 
3330: electrons and 
3331: holes.  
3332: Since in a steady state the excited spin polarization depends not only on 
3333: $\tau_s$
3334: but also on the carrier recombination time, the Hanle effect,
3335: polarization of luminescence by a transverse magnetic field 
3336: (see Sec.~\ref{sec:IID1}),
3337: is employed to deduce 
3338: $\tau_s$ unambiguously.
3339: 
3340: The Hanle effect has also been a great tool for investigating $T_2$ in metals,
3341: in connection with electrical spin injection. The advantage of optical 
3342: orientation
3343: over CESR is that, if carrier lifetime is known, zero-field 
3344: (or zero-$g$-factor 
3345: material) 
3346: data can be measured. In addition, smaller $\tau_s$ values can be detected. 
3347: 
3348: Type (b) experiments measure magnetization in a time or space domain. 
3349: The most important examples are the Johnson-Silsbee spin injection  experiment,
3350: the time-resolved (pump-probe) photoluminescence (in which the probe
3351: is used on the same principle as optical orientation),
3352: and time-resolved
3353: Faraday and (magneto-optic) Kerr rotation. The last two methods can
3354: follow coherent (in the ensemble sense) dynamics of electron spin 
3355: precession. 
3356: 
3357: Spin injection experiments \cite{Johnson1985:PRL} detect the amount of 
3358: nonequilibrium magnetization by observing charge response 
3359: (see Sec.~\ref{sec:IID1}). 
3360: In the Johnson-Silsbee scheme, electrons
3361: are first injected by electrical spin injection from a ferromagnetic
3362: electrode into a metal or semiconductor. As the spin diffuses throughout the 
3363: sample, another ferromagnetic electrode detects the amount of spin as a 
3364: position-dependent quantity. The detection is by means of 
3365: spin-charge coupling,
3366: whereby an EMF appears across the paramagnet/ferromagnet
3367: interface in proportion to the nonequilibrium magnetization 
3368: \cite{Silsbee1980:BMR}. 
3369: By fitting the
3370: spatial dependence of magnetization to the exponential decay formula, 
3371: one can extract the
3372: spin diffusion length $L_s$ and thus the spin relaxation 
3373: time $T_1=L_s^2/D$. 
3374: The Hanle effect, too, can be used in combination with
3375: Johnson-Silsbee spin injection, yielding directly $T_1=T_2$, which
3376: agrees with $T_1$ determined from the measurement of $L_s$.
3377: A precursor to the Hanle effect in spin injection was 
3378: transmission-electron
3379: spin resonance (TESR), in which
3380: nonequilibrium electron spin excited in the skin 
3381: layer of a metallic sample is transported to the other side, emitting 
3382: microwave
3383: radiation. In very clean samples and at low temperatures, electrons
3384: ballistically propagating 
3385: through
3386: the sample experience Larmor precession
3387: resulting in Larmor waves, seen as an oscillation of the transmitted
3388: radiation amplitude with changing static magnetic field~\cite{Janossy1980:PRB}.
3389: 
3390: Time-resolved photoluminescence
3391: detects, after creation of spin-polarized carriers, circular polarization of 
3392: the recombination light. This technique was used, for example, 
3393: to detect a 500 ps spin coherence time $T_2$ of free excitons in a 
3394: GaAs quantum well \cite{Heberle1994:PRL}.
3395: The Faraday and (magneto-optic) Kerr effects are the rotation of the 
3396: polarization
3397: plane of a linearly polarized light, transmitted through (Faraday) 
3398: or reflected by
3399: (Kerr)  a magnetized sample. The Kerr effect is more useful for 
3400: thicker and nontransparent samples or for thin films fabricated on thick 
3401: substrates.  The angle of rotation is proportional
3402: to the amount of magnetization in the direction of the incident light. 
3403: Pump (a circularly polarized laser pulse) and probe experiments employing 
3404: magneto-optic effects can now follow, with 100 fs resolution, the evolution
3405: of magnetization as it dephases in a transverse magnetic field. 
3406: Using lasers for spin excitation has the
3407: great advantage of not only detecting, 
3408: but also 
3409: manipulating spin dephasing, as shown using Faraday rotation on bulk 
3410: GaAs and GaAs/ZnSe 
3411: heterostructures \cite{Kikkawa2001:PE,Awschalom1999:PT}. 
3412: The Kerr effect was used, for example, to investigate the spin dynamics of bulk 
3413: CdTe \cite{Kimel2000:PRB}, 
3414: and Faraday rotation was used to study spin coherence in nanocrystals of 
3415: CdSe \cite{Gupta2002:PRB} and coherent control of spin dynamics of
3416: excitons in GaAs quantum wells \cite{Heberle1996:IEEE}.
3417: 
3418: 
3419: \subsection{\label{sec:IIIB} Mechanisms of spin relaxation}
3420: %==========================================================================
3421: 
3422: Four  mechanisms for spin relaxation of conduction electrons have been found
3423: relevant for metals and semiconductors: the Elliott-Yafet (EY),
3424: D'yakonov-Perel' (DP), Bir-Aronov-Pikus (BIP), and hyperfine-interaction 
3425: (HFI) 
3426: mechanisms.\footnote{We do not 
3427: consider magnetic scattering, that is, 
3428: scattering due to an exchange interaction between conduction electrons
3429: and magnetic impurities.} In the EY
3430: mechanism electron spins relax because the electron wave functions normally
3431: associated with a given spin have an admixture of the
3432: opposite spin states, due to spin-orbit coupling induced by ions. The DP
3433: mechanism explains spin dephasing in solids without a center of symmetry.
3434: Spin dephasing occurs because electrons feel an effective
3435: magnetic field, resulting from the lack of inversion symmetry and from the
3436: spin-orbit interaction, which 
3437: changes in random directions every time the electron
3438: scatters to a different momentum states. The BIP mechanism is 
3439: important for p-doped semiconductors, in which the electron-hole exchange 
3440: interaction
3441: gives rise to fluctuating local magnetic fields flipping electron spins. 
3442: Finally,
3443: in semiconductor heterostructures (quantum wells and quantum dots) based on
3444: semiconductors with a nuclear magnetic moment, 
3445: it is the hyperfine interaction 
3446: of the electron spins and nuclear moments which dominates spin dephasing of 
3447: localized or confined electron spins. An informal review of spin 
3448: relaxation of conduction 
3449: electrons can be found 
3450: in \textcite{Fabian1999:JVST}.
3451: 
3452: \subsubsection{\label{sec:IIIB1} Elliott-Yafet mechanism}
3453: %-------------------------------------------------------------------------
3454: 
3455: \textcite{Elliott1954:PR} was the first to realize that conduction-electron 
3456: spins can relax via
3457: ordinary momentum scattering (such as by phonons or impurities) 
3458: if the lattice ions induce spin-orbit
3459: coupling in the electron wave function. In the presence 
3460: of 
3461: the spin-orbit interaction 
3462: \begin{equation}\label{eq:relax:vso}
3463: V_{so}=\frac{\hbar}{4 m^2 c^2} \nabla V_{sc} \times \hat{\bf p} \cdot 
3464: {\bf \hat{\bf \sigma}},
3465: \end{equation}
3466: where $m$ is the free-electron mass, $V_{sc}$ is the scalar (spin-independent) 
3467: periodic lattice 
3468: potential, $\hat{\bf p}\equiv -i\hbar \nabla$
3469: is the linear momentum operator, and ${\bf \hat{\bf \sigma}}$ 
3470: are the Pauli matrices, 
3471: single-electron (Bloch) wave functions in a solid are no longer the 
3472: eigenstates of 
3473: $\hat{\sigma}_z$, but rather a mixture of the Pauli spin up 
3474: $|\uparrow \rangle$ 
3475: and spin down $|\downarrow \rangle$
3476: states. If the solid possesses a center of symmetry, the case of 
3477: elemental metals which Elliott considered, the Bloch states 
3478: of ``spin up'' and ``spin down'' electrons with the lattice momentum 
3479: $\bf k$ and band index $n$  
3480: can be written as \cite{Elliott1954:PR}
3481: \begin{eqnarray}
3482: \Psi_{{\bf k}n\uparrow}({\bf r})&=&\left [a_{{\bf k}n} ({\bf r}) 
3483: |\uparrow\rangle
3484: +b_{{\bf k}n} ({\bf r}) |\downarrow\rangle  \right ]e^{i{\bf k}\cdot {\bf r}}, 
3485: \\
3486: \Psi_{{\bf k}n\downarrow}({\bf r})&=&\left [a^*_{-{\bf k}n} ({\bf r}) 
3487: |\downarrow\rangle  
3488: -b^*_{-{\bf k}n} ({\bf r}) |\uparrow\rangle  \right ] e^{i{\bf k}\cdot {\bf r}}, 
3489: \end{eqnarray}
3490: where we write the explicit dependence of the complex 
3491: lattice-periodic 
3492: coefficients $a$ and $b$ 
3493: on the radius vector $\bf r$.
3494: The two Bloch states are degenerate: they are connected by spatial 
3495: inversion and time reversal~\cite{Elliott1954:PR}. That it makes sense to call
3496: $\Psi_{{\bf k}n\uparrow}$ and  $\Psi_{{\bf k}n\downarrow}$, respectively,
3497: spin up 
3498: and spin down 
3499: states follows from the fact that in most cases the typical values of $|a|$ are 
3500: close to unity,
3501: while $|b| \ll 1$. 
3502: 
3503: Indeed, consider a band structure in the absence of spin-orbit coupling.
3504: Turning $V_{so}$ on couples electron states of opposite 
3505: Pauli spins which are 
3506: of the same ${\bf k}$ (because $V_{so}$ has the period of the lattice), 
3507: but different $n$. Because the spin-orbit interaction is normally much 
3508: smaller than the
3509: distance between the bands, perturbation theory gives 
3510: \begin{equation} \label{eq:relax:lambda}
3511: |b|\approx \lambda_{so}/\Delta E \ll 1,
3512: \end{equation}
3513: where $\Delta E$ is the energy distance between the band state in question and 
3514: the
3515: state (of the same momentum) 
3516: in the nearest band, and $\lambda_{so}$ is the amplitude of the matrix element 
3517: of $V_{so}$ 
3518: between the two states. 
3519: The spin-orbit coupling alone does not lead to spin relaxation. However, in 
3520: combination 
3521: with
3522:  momentum scattering, the spin up [Eq.~(56)] and spin down  [Eq.~(57)] %iz lab
3523: states
3524: can couple and lead to spin relaxation.
3525: 
3526: Momentum scattering is typically caused by impurities (at low $T$) and phonons 
3527: (at high
3528: $T$). There is another important spin-flip scattering mechanism that involves 
3529: phonons. 
3530: A periodic, lattice ion-induced spin-orbit interaction is modified by phonons 
3531: and can 
3532: directly couple the (Pauli) spin up and spin down states. Such processes 
3533: were first 
3534: considered 
3535: for a jellium model by \textcite{Overhauser1953:PR} [see also 
3536: \cite{Grimaldi1997:PRB}], 
3537: and for band structure systems by \textcite{Yafet:1963}. They must be combined
3538: with the Elliott processes discussed above to form a consistent picture of 
3539: phonon-induced spin relaxation, especially at low 
3540: temperatures~\cite{Yafet:1963}, 
3541: where the
3542: two processes have similar magnitudes. At larger $T$,
3543: phonon-modified $V_{so}$ 
3544: is not 
3545: important for polyvalent metals, whose spin relaxation is dominated by spin hot 
3546: spots---states
3547: with anomalously large $|b|$---as shown by the explicit calculation 
3548: of \textcite{Fabian1999:PRL}. Spin hot spots are discussed in more detail in
3549: Sec.~\ref{sec:IIIC}.  
3550: 
3551: We now give a recipe, useful for elemental metals and semiconductors, 
3552: for calculating phonon-induced $1/\tau_s$ 
3553: from the known band and phonon structure. 
3554: The corresponding theory was systematically developed by \textcite{Yafet:1963}.
3555: The spin relaxation rate due to phonon scattering according 
3556: to the EY mechanism
3557: can be expressed through the spin-flip Eliashberg function 
3558: $\alpha_S^2F(\Omega)$
3559: as ~\cite{Fabian1999:PRL}
3560: \begin{equation} \label{eq:relax:ey}
3561: 1/\tau_s=8\pi T\int_{0}^{\infty} d\Omega \alpha_s^2F(\Omega)\partial 
3562: N(\Omega)/\partial T, 
3563: \end{equation}
3564: where $N(\Omega)=[\exp(\hbar \Omega/k_B T)-1]^{-1}$ is the phonon distribution 
3565: function.
3566: The spin-flip Eliashberg function gives the contribution of the phonons with 
3567: frequency
3568: $\Omega$ to the spin-flip electron-phonon interaction, 
3569: \begin{equation} \label{eq:relax:ef}
3570: \alpha_s^2F(\Omega)=\frac{g_S}{2M\Omega}\sum_{\nu}\langle\langle 
3571: g_{{\bf k}n\uparrow, {\bf k}'n'\downarrow}^{\nu}\delta (\omega_{{\bf q}\nu}-
3572: \Omega)
3573: \rangle_{{\bf k}n}\rangle _{{\bf k}'n'}.
3574: \end{equation}
3575: Here $g_S$ is the number of electron states per spin and atom at the Fermi 
3576: level, 
3577: $M$ is the ion mass, $\omega_{{\bf q}\nu}$ is the frequency of phonons
3578: with momentum ${\bf q}={\bf k}-{\bf k}'$ and branch index $\nu$,
3579: and the spin-flip matrix element is
3580: \begin{equation} 
3581: g_{{\bf k}n\uparrow, {\bf k}'n'\downarrow}^{\nu}\equiv |{\bf u}_{{\bf 
3582: q}\nu}\cdot
3583: \left (\Psi_{{\bf k}n\uparrow}, \nabla V \Psi_{{\bf k}'n'\downarrow} 
3584: \right )  |^2,
3585: \end{equation}
3586: where ${\bf u}_{{\bf q}\nu}$ is the polarization of the phonon with momentum 
3587: $\bf q$ and in branch $\nu$. 
3588: The brackets $\langle ... \rangle_{{\bf k}n}$ in Eq.~(\ref{eq:relax:ef}) 
3589: denote Fermi-surface averaging. The Bloch wave functions for this 
3590: calculation are chosen 
3591: to satisfy $(\Psi_{{\bf k}n\uparrow}, \hat {\sigma}_z
3592: \Psi_{{\bf k}n\uparrow})=-(\Psi_{ {\bf k}n\downarrow}, \hat {\sigma}_z 
3593: \Psi_{{\bf k}n\downarrow})$.
3594: The periodic lattice-ion interaction $V$ contains both scalar and 
3595: spin-orbit parts: $V=V_{sc}+V_{so}$.
3596: 
3597: There are two important relations giving  an order-of-magnitude estimate
3598: of $\tau_s$, as well as 
3599: its temperature dependence: the Elliott and the
3600: Yafet relations.
3601: The Elliott relation relates $\tau_s$ to the shift
3602: $\Delta g$ of the electronic $g$ factor from the free-electron value of
3603: $g_0=2.0023$~\cite{Elliott1954:PR}:
3604: \begin{eqnarray} \label{eq:relax:elliott}
3605: 1/\tau_s \approx (\Delta g)^2/\tau_p,
3606: \end{eqnarray}
3607: where $\tau_p$ is the momentum relaxation time.
3608: This relation follows from the fact that for a momentum scattering
3609: interaction $V_i$ the spin-flip scattering probability in the
3610: Born approximation is proportional to
3611: $|(\Psi_{{\bf k}n \uparrow},
3612: V_i \Psi_{{\bf k}'n' \downarrow})|^2 \approx |b|^2\times|(\Psi_{{\bf k}n
3613: \uparrow},
3614: V_i \Psi_{{\bf k}'n' \uparrow})|^2$. Realizing that the spin-conserving
3615: scattering
3616: probability gives the momentum relaxation rate, after a 
3617: Fermi-surface averaging we
3618: get the  estimate
3619: \begin{equation} \label{eq:relax:aux}
3620: 1/\tau_s \approx \langle b^2 \rangle  /\tau_p.
3621: \end{equation}
3622: 
3623: On the other hand, $\Delta g$ is determined by the expectation value of 
3624: $\hat{l}_z$, 
3625: the z-component of the orbital momentum 
3626: in a Bloch state. Without the spin-orbit 
3627: interaction this expectation value is 
3628: zero. 
3629: Considering the spin-orbit interaction to be a small parameter,
3630: we find by 
3631: perturbation theory $\Delta g \approx |b|$, 
3632: which combines with Eq.~(\ref{eq:relax:aux}) to give the Elliott relation 
3633: Eq.~(\ref{eq:relax:elliott}). An empirical test of the Elliott relation 
3634: for simple
3635: metals when spin relaxation is due to thermal phonons 
3636: ~\cite{Beuneu1978:PRB} gives the revised estimate 
3637: \begin{equation} \label{eq:revisedE}
3638: 1/\tau_s \approx 10\times (\Delta g)^2 /\tau_p. 
3639: \end{equation}
3640: 
3641: The Elliott relation is only a very rough estimate of $\tau_s$. The
3642: experimentally relevant ratio  $\tau_p/\tau_s$ depends on the scattering
3643: mechanism. The ratio is different for scattering off impurities,
3644: boundaries, or phonons, although one would expect it to be
3645: within an order of magnitude. For example, scattering by heavy impurities 
3646: induces an
3647: additional spin relaxation channel where spin flip is due to the 
3648: spin-orbit interaction induced by the impurities. 
3649: Equation (\ref{eq:relax:aux}) then
3650: does not hold.  
3651: Scattering by phonons is too complex to be simply equated 
3652: with the ratio
3653: $\tau_p/\tau_s$ 
3654: for impurity or boundary scattering. However, the ratio is comparable
3655: for scattering by light impurities and by boundaries. The ratio
3656: $\tau_p/\tau_s$ for impurity and phonon scattering in Al and Cu is
3657: compared by \textcite{Jedema2003:PRB}.
3658: 
3659: The Yafet relation is only qualitative, connecting the temperature 
3660: dependence of 
3661: $1/T_1$ with 
3662: that of the resistivity $\rho$:
3663: \begin{equation} \label{eq:relax:yafet}
3664: 1/T_1(T) \sim  \langle b^2 \rangle \rho (T),
3665: \end{equation}
3666: as follows directly from Eq.~(\ref{eq:relax:aux}) after realizing that 
3667: $1/\tau_p(T)\sim \rho (T) $. By careful symmetry considerations 
3668: \textcite{Yafet:1963} 
3669: proved that 
3670: $1/T_1 \sim T^5$ at low temperatures, similarly to the Bloch-Gr\"{u}neisen 
3671: law for resistivity,
3672: justifying Eq.~(\ref{eq:relax:yafet}) over a large temperature range.   
3673: Yafet's $T^5$ law stems from the nontrivial fact that for spin-flip 
3674: electron-phonon scattering 
3675: $g_{{\bf k}n\uparrow,{\bf k}'n\downarrow}^\nu \sim
3676: ({\bf k}-{\bf k}')^4$ as ${\bf k} \rightarrow {\bf k}'$ ~\cite{Yafet:1963}, 
3677: while only a quadratic dependence holds for the spin-conserving scattering. 
3678: This corresponds to the long-wavelength
3679: behavior $\alpha_S^2 F(\Omega)\sim \Omega^4$ of the spin-flip Eliashberg 
3680: function. The Yafet relation was tested experimentally by 
3681: \textcite{Monod1979:PRB}. This work led to a deeper understanding of spin 
3682: relaxation processes
3683: in polyvalent metals \cite{Silsbee1983:PRB,Fabian1998:PRL}.
3684: 
3685: Realistic calculations of the EY $\tau_s$ in semiconductors  
3686: can be performed analytically using approximations of the 
3687: band and phonon structures, as
3688: most important states are usually  around high symmetry points. Here we give 
3689: a formula for the spin relaxation of conduction electrons with energy 
3690: $E_{\bf k}$
3691: in the frequently studied case of III-V semiconductors 
3692: ~\cite{Chazalviel1975:PRB,Pikus:1984}:
3693: \begin{equation} \label{eq:relax:chazalviel}
3694: \frac{1}{\tau_s(E_{\bf k})}= A\left 
3695: (\frac{\Delta_{so}}{E_g+\Delta_{so}}\right )^2\left 
3696: (\frac{E_{\bf k}}{E_g}\right )^2\frac{1}
3697: {\tau_p(E_{\bf k})}, 
3698: \end{equation}
3699: where $\tau_p(E_{\bf k})$ 
3700: is the momentum scattering time at energy $E_{\bf k}$, 
3701: $E_g$ is the energy gap,
3702: and $\Delta_{so}$ is the spin-orbit splitting of the valence band 
3703: (see Fig.~\ref{oo:1}). 
3704: The Numerical factor
3705: $A$, which is of order 1, depends on the dominant scattering 
3706: mechanism (charge or neutral impurity, phonon, electron-hole).
3707: Analytic formulas
3708: for the EY mechanism due to electron-electron scattering 
3709: are given by
3710: \textcite{Boguslawski1980:SSC}.
3711: 
3712: Equation~(\ref{eq:relax:chazalviel}) shows that the EY mechanism is important 
3713: for small-gap semiconductors with large spin-orbit splitting (the prototypical 
3714: example is InSb). For degenerate electron densities
3715: the spin relaxation time is given by Eq.~(\ref{eq:relax:chazalviel}) 
3716: with $E_{\bf k}=E_F$,
3717: while for nondegenerate densities the thermal averaging leads to a substitution 
3718: of thermal energy $k_B T$ for $E_{\bf k}$ and thermal-averaged momentum 
3719: relaxation
3720: time for $\tau_p$. To estimate $\tau_s$ from 
3721: Eq.~(\ref{eq:relax:chazalviel}) one needs to know $\tau_p$. 
3722: It often suffices to know the doping or temperature dependence of $\tau_p$ 
3723: to decide on the relevance of the EY mechanism \cite{Pikus:1984}.  
3724: 
3725: The temperature dependence of $1/\tau_s$ for metals and degenerate  
3726: semiconductors follows the dependence of $1/\tau_p$. 
3727: In metals this means a constant at low $T$ 
3728: and a linear increase at large $T$. 
3729: For nondegenerate semiconductors $1/\tau_s(T)\sim T^2/\tau_p(T)$. In the 
3730: important case
3731: of scattering by charged impurities ($\tau_p\sim T^{3/2}$) $1/\tau_s\sim 
3732: T^{1/2}$. 
3733: The magnetic field dependence of the EY spin relaxation  has not been
3734: systematically investigated. At low temperatures, where cyclotron 
3735: effects become important, 
3736: one needs to average over cyclotron trajectories on the Fermi 
3737: surface need to obtain $1/\tau_s$. 
3738: We expect that such averaging leads, in general, to an 
3739: increase in $1/T_1$, especially in systems where spin hot 
3740: spots are important (see Sec.~\ref{sec:IIIC}).
3741: 
3742: 
3743: \subsubsection{\label{sec:IIIB2} D'yakonov-Perel' mechanism}
3744: %---------------------------------------------------------------------------
3745: 
3746: 
3747: An efficient mechanism of spin relaxation due to spin-orbit coupling
3748: in systems lacking inversion symmetry 
3749: was found by \textcite{Dyakonov1972:SPSS}. Without
3750: inversion symmetry the momentum states of the spin up and spin down electrons 
3751: are not degenerate: $E_{{\bf k}\uparrow}\ne E_{{\bf k}\downarrow}$.  
3752: Kramer's theorem still dictates that $E_{{\bf k}\uparrow}= E_{-{\bf 
3753: k}\downarrow}$. Most prominent examples of materials without inversion
3754: symmetry 
3755: come from groups
3756: III-V (such as GaAs) and II-VI (ZnSe) semiconductors, where 
3757: inversion symmetry is broken by the presence of two distinct atoms in the
3758: Bravais lattice. Elemental semiconductors like Si possess inversion symmetry
3759: in the bulk, so the DP mechanism does not apply to them. In heterostructures
3760: the symmetry is broken by the presence of asymmetric confining potentials.  
3761: 
3762: Spin splittings 
3763: induced by  
3764: inversion asymmetry 
3765: can be described by introducing an
3766: intrinsic ${\bf k}$-dependent magnetic field ${\bf B}_{i}({\bf k})$
3767: around which electron spins 
3768: precess with Larmor frequency ${\bf \Omega}({\bf k})=(e/m) {\bf B}_i({\bf k})$.
3769: The intrinsic field derives from the spin-orbit coupling in the band-structure.
3770: The corresponding Hamiltonian term describing the precession of
3771: electrons in the conduction band is
3772: \begin{equation} \label{eq:HDP}
3773: H({\bf{k}})=\frac{1}{2}\hbar {\bf \hat{\sigma}} \cdot{\bf \Omega}(\bf{k}),
3774: \end{equation}
3775: where ${\bf \hat{\sigma}}$ are the Pauli matrices. Momentum-dependent 
3776: spin precession 
3777: described
3778: by $H$, together with momentum scattering 
3779: characterized by momentum 
3780: relaxation time $\tau_p$,\footnote{In the qualitative reasonings below we use 
3781: $\tau_p$
3782: instead of the effective correlation time $\tilde{\tau}$ for ${\bf \Omega}$ 
3783: during momentum 
3784: scattering; $\tilde\tau$ is defined later, in Eq.~(\ref{eq:relax:taul}).}
3785: leads to spin dephasing. While the microscopic expression for 
3786: $\bf \Omega({\bf k})$
3787: needs to be obtained from the band structure, treating the effects of 
3788: inversion asymmetry by introducing intrinsic precession helps to give a 
3789: qualitative understanding
3790: of spin dephasing. It is important to 
3791: note, however, that the analogy with real
3792: Larmor precession is not complete. An applied magnetic field induces 
3793: a macroscopic spin polarization and magnetization, while $H$ of 
3794: Eq.~(\ref{eq:HDP})
3795: produces an equal number of spin up and spin down states.  
3796: 
3797: Two limiting 
3798: cases can be considered: (i) $\tau_p\Omega_{av}\agt 1$ and (ii) 
3799: $\tau_p\Omega_{av}\alt 1$, 
3800: where $\Omega_{av}$ is an average magnitude of the intrinsic 
3801: Larmor frequency $\Omega({\bf k})$ over the actual momentum
3802: distribution. Case (i) corresponds to the situation in which
3803: individual electron 
3804: spins precess a full cycle before
3805: being scattered to another momentum state. The total spin in this regime 
3806: initially dephases 
3807: reversibly due to the anisotropy in ${\bf \Omega}({\bf k})$. The spin dephasing 
3808: rate,\footnote{The reversible decay need not be exponential.}
3809: which depends on the
3810: distribution of values of ${\bf \Omega}({\bf k})$, is in general
3811: proportional to the width $\Delta \Omega$ of the distribution: 
3812: $1/\tau_s \approx \Delta \Omega$. The spin is irreversibly lost after time 
3813: $\tau_p$, when randomizing scattering takes place.
3814: 
3815: Case (ii) is what is usually meant by the D'yakonov-Perel' 
3816: mechanism. This regime can be viewed from the point of view of individual 
3817: electrons as a 
3818: spin precession about fluctuating magnetic fields, whose magnitude and 
3819: direction change randomly with the average time step of $\tau_p$. 
3820: The electron spin rotates about 
3821: the intrinsic field at an angle $\delta \phi = \Omega_{av}\tau_p$, 
3822: before experiencing 
3823: another field
3824: and starting to rotate with a different speed and in a different direction. 
3825: As a result, the spin phase follows a random walk: after time $t$, which 
3826: amounts to
3827: $t/\tau_p$ steps of the random walk, the phase
3828: progresses by $\phi(t) \approx \delta\phi\sqrt{t/\tau_p}$. Defining $\tau_s$ as 
3829: the time
3830: at which $\phi(t)=1$, the usual motional narrowing result is obtained:
3831: $1/\tau_s=\Omega_{av}^2 \tau_p$ (see footnote 70).
3832: The faster the momentum relaxation, 
3833: the slower 
3834: the
3835: spin dephasing. The difference between cases (i) and (ii) is that in 
3836: case (ii)
3837: the electron spins form an ensemble that directly samples the 
3838: distribution of 
3839: $\Omega({\bf k})$, while
3840: in case (ii) it is the distribution of the {\it sums} 
3841: of the intrinsic 
3842: Larmor frequencies (the 
3843: total 
3844: phase of a spin after many steps consists of a sum of randomly selected 
3845: frequencies
3846: multiplied by $\tau_p$), which, according to the central limit theorem, has a 
3847: significantly reduced variance. Both limits (i) and (ii) 
3848: and the transition between
3849: them have been experimentally demonstrated in n-GaAs/AlGaAs quantum wells by 
3850: observing temporal spin oscillations over a large range of temperatures 
3851: (and thus $\tau_p$)
3852:  \cite{Brand2002:PRL}.
3853: 
3854: A more rigorous expression for $\tau_s$ in regime (ii) 
3855: has been obtained by solving the kinetic rate equation for a spin-dependent 
3856: density matrix 
3857: \cite{Dyakonov1971:SPJETP,Dyakonov1972:SPSS}.
3858: If the band structure is isotropic and scattering is both elastic and 
3859: isotropic, 
3860: evolution 
3861: of the z-component of spin ${\bf s}$
3862: is  
3863: \cite{Pikus:1984}
3864: \begin{equation}\label{eq:relax:sz}
3865: \dot{s}_z =
3866: -\tilde{\tau_l}\left [s_z \overline{\left ( \Omega^2 -\Omega_z^2 \right )}
3867: -s_x \overline{\Omega_x\Omega_z}-s_y \overline{\Omega_y\Omega_z}\right ],
3868: \end{equation}
3869: where the bar denotes averaging over directions of $\bf k$. Analogous 
3870: expressions
3871: for $\dot{s}_x$ and $\dot{s}_y$ can be written by index permutation. The 
3872: effective
3873: momentum scattering time is introduced as
3874: \begin{equation}\label{eq:relax:taul}
3875: 1/\tilde{\tau_l}=\int_{-1}^1 W(\theta)\left [1-P_l\left (\cos\theta\right 
3876: )\right ] d\cos\theta,
3877: \end{equation}
3878: where $W(\theta)$ is the rate of momentum scattering through angle $\theta$ at 
3879: energy
3880: $E_{\bf k}$, and $P_l$ is the Legendre polynomial whose order $l$ 
3881: is the power 
3882: of $\bf k$ in
3883: $\Omega({\bf k})$. [It is assumed that $\Omega({\bf k})\sim k^l$ in 
3884: Eq.~(\ref{eq:relax:sz}).] 
3885:  In two dimensions $P_l(\cos\theta)$ is replaced by $\cos(l\theta)$
3886: in Eq.~(\ref{eq:relax:taul}) for the $l$th polar harmonic of 
3887: $\bf \Omega({\bf k})$
3888: ~\cite{Pikus1995:PRB}.
3889: Since it is useful to express the results in terms of the known momentum 
3890: relaxation
3891: times\footnote{In fact,  normal (not umklapp) electron-electron collisions 
3892: should also be included in the effective spin randomization time 
3893: $\tilde{\tau}$, though they 
3894: do not
3895: contribute to the momentum relaxation time which appears in the measured 
3896: mobility~\cite{Glazov2002:JETPL,Glazov2003:P}.}
3897:  $\tau_p=\tilde{\tau}_1$, one defines 
3898: \footnote{ \textcite{Pikus:1984} 
3899: initially define $\gamma_l$ as here, but 
3900: later evaluate it, inconsistently, as the inverse $\gamma_l \rightarrow 
3901: \gamma_l^{-1}$.} 
3902: $\gamma_l=\tau_p/\tilde{\tau_l}$ to measure the effectiveness of momentum 
3903: scattering
3904: in randomizing Larmor frequencies; $\tilde{\tau}_l$ accounts for the relative 
3905: angle between 
3906: ${\bf \Omega}$ before and after scattering. Generally $\gamma_l > 1$ for $l>1$, 
3907: that is, momentum scattering is more effective in randomizing spins than 
3908: in randomizing momentum. 
3909: 
3910: 
3911: Comparing with the Bloch-Torrey equations 
3912: (\ref{eq:relax:bloch1})--(\ref{eq:relax:bloch3}), for $\bf{B}=0$ and no spin 
3913: diffusion, we see that spin decay is described by the tensor 
3914: $1/\tau_{s,ij}$ 
3915: (here $i$ and $j$ are the Cartesian coordinates) whose diagonal 
3916: $1/\tau_{s,ii}$
3917: and off-diagonal $1/\tau_{s,ij}$, for $i\ne j$, terms
3918: are
3919: \begin{equation} \label{eq:relax:tauii}
3920: 1/\tau_{s,ii}=\gamma_l^{-1}\tau_p(\overline{\Omega^2}-
3921: \overline{\Omega_i^2}),\,\,\,
3922: 1/\tau_{s,ij}=-\gamma_l^{-1}\tau_p\overline{\Omega_i\Omega_j}.
3923: \end{equation}
3924: In general, spin dephasing depends on the spin direction and on the 
3925: dephasing rates of 
3926: the perpendicular
3927: spin components. Equations (\ref{eq:relax:tauii}) are valid for small magnetic 
3928: fields, 
3929: satisfying $\Omega_0 \tau_p \ll 1$, where $\Omega_0$ is the Larmor frequency of 
3930: the
3931: external field. 
3932: 
3933: The most important difference between the EY and the DP mechanism is their 
3934: opposite 
3935: dependence on $\tau_p$. While increased scattering intensity makes the EY 
3936: mechanism 
3937: more effective, it decreases the effectiveness of the
3938: the DP processes. In a sense the two 
3939: mechanisms are 
3940: similar to collision broadening and motional narrowing in 
3941: NMR \cite{Slichter:1989}. Indeed, in the EY 
3942: process
3943: the precession frequency is conserved between collisions and the loss of
3944: phase occurs only in 
3945: the short time {\it during} collision. The more 
3946: collisions there are,
3947: the greater is the loss of phase memory, in analogy with collision 
3948: broadening of 
3949: spectral
3950: lines. On the other hand, in DP spin dephasing, spin phases are randomized 
3951: {\it between} collisions, since electrons precess with different frequencies 
3952: depending
3953: on their momenta. Spin-independent collisions with impurities or phonons do not 
3954: lead
3955: to phase randomization during the collision itself, but help to establish the 
3956: random-walk-like evolution of the phase, leading to motional narrowing.
3957: While these two mechanisms coexist in systems lacking inversion symmetry,
3958: their relative strength depends on many factors. Perhaps the most robust trend
3959: is that the DP mechanism becomes more important with increasing band
3960: gap and increasing temperature. 
3961: 
3962: In the rest of the section we apply Eq.~(\ref{eq:relax:tauii})
3963: to the study of spin dephasing in bulk and two-dimensional III-V semiconductor 
3964: heterostructures.
3965: 
3966: \paragraph{\label{sec:IIIB2a} Bulk III-V semiconductors.}
3967: %--------------------------------------------------------------------------- 
3968: 
3969: In bulk III-V semiconductors the intrinsic Larmor frequency vector of
3970: Eq.~(\ref{eq:HDP}) 
3971: due
3972: to the
3973: lack of inversion symmetry is \cite{Dyakonov1971:SPJETP}
3974: \begin{equation} \label{eq:relax:dress}
3975: {\bf \Omega}({\bf{k}})=\alpha\hbar^2(2m_c^3E_g)^{-1/2}\bf{\kappa},
3976: \end{equation}
3977: where 
3978: \begin{eqnarray}
3979: {\bf\kappa} = \left [k_x(k_y^2-k_z^2),k_y(k_z^2-k_x^2),
3980: k_z(k_x^2-k_y^2)\right ].
3981: \end{eqnarray}
3982: Here $k_i$ are the lattice wave-vector components along the crystal principal 
3983: axes.
3984: The material-specific parameters are the band gap $E_g$ and the 
3985: conduction electron mass
3986: $m_c$; $\alpha$ is a dimensionless parameter specifying
3987: the strength of the spin-orbit interaction.
3988: The spin splitting described by Eq.~(\ref{eq:relax:dress})
3989: is proportional to the cube of the lattice momentum, as was first found by 
3990: \textcite{Dresselhaus1955:PR}. For GaAs $\alpha\approx 
3991: 0.07$~\cite{Marushchak1984:SPSS}. Spin splitting of conduction and
3992: heavy and light holes in GaAs quantum wells, due to bulk inversion
3993: asymmetry was calculated by \textcite{Rashba1988:PLA}.
3994: 
3995: Substituting Eq.~(\ref{eq:relax:dress}) for ${\bf\Omega}$ in 
3996: Eq.~(\ref{eq:relax:tauii}), 
3997: and using $\overline{\kappa_i\kappa_j}=(4/105)k^6\delta_{ij}$, we obtain the 
3998: expected
3999: result that the off-diagonal elements of $1/\tau_{s,ij}$ vanish for cubic 
4000: systems
4001: and the diagonal elements are all equal to~\cite{Pikus:1984} 
4002: \begin{equation}\label{eq:relax:DP}
4003: 1/\tau_s(E_{\bf k})=
4004: \frac{32}{105}\gamma_3^{-1}\tau_p(E_{\bf k}) \alpha^2\frac{E_{\bf 
4005: k}^3}{\hbar^2 E_g}.
4006: \end{equation}
4007: The above expression describes DP spin dephasing of degenerate ($E_{\bf k}=E_F$) or 
4008: hot\footnote{This is strictly true only if the spin relaxation of 
4009: the hot electrons 
4010: is faster
4011: than energy relaxation by optical phonon emission, which is rarely the case. 
4012: One has to consider either the spin relaxation at different energy levels 
4013: during
4014: the cascade process of optical phonon emission or, if the optical 
4015: phonon emission is
4016: particularly fast, spin relaxation only during the final stages of energy 
4017: relaxation
4018: by acoustic phonon emission (see \textcite{Pikus:1984}.} electrons
4019: in bulk III-V semiconductors. For impurity scattering $\gamma_3\approx 
4020: 6$, for
4021: acoustic phonons $\gamma_3\approx 1$, while for optical polar phonons 
4022: $\gamma_3\approx 41/6$.
4023: The temperature dependence of $1/\tau_s$ comes from the temperature dependence 
4024: of 
4025: $\tau_p$. For the important case of charged impurity scattering ($\tau_p\sim 
4026: T^{3/2}$),
4027: $1/\tau_s\sim T^{3/2}$. 
4028: 
4029: Compared to the EY expression, Eq.~(\ref{eq:relax:chazalviel}),
4030: the DP spin dephasing increases much faster with 
4031: increasing
4032: electron energy and is expected to be dominant at large donor doping levels and
4033: at high temperatures. The EY mechanism can be dominant in small-band-gap and 
4034: large spin-orbit-splitting materials. 
4035: The two mechanisms can also be easily distinguished by 
4036: their
4037: opposite dependence on momentum relaxation. Contrary to the EY mechanism, 
4038: greater impurity density will decrease the importance of the DP processes. 
4039: The  most frequently used ways
4040: to distinguish between various methods of spin relaxation are comparing the
4041: electron density (through the variation of the Fermi energy) 
4042: and the temperature dependences of $1/\tau_s$ with the
4043: theoretical estimates. Since the prefactors may vary with different scattering
4044: mechanisms, it is best to deduce $\tau_p(E_{\bf k})$  and $\tau_p(T)$ 
4045: from mobility measurements and use Eqs.~(\ref{eq:relax:chazalviel}) and 
4046: (\ref{eq:relax:DP})
4047: or the equations given below for $1/\tau_s(T)$.
4048: 
4049: Another interesting distinction between the two mechanisms is revealed 
4050: by the 
4051: dependence
4052: of their spin diffusion length $L_s=\sqrt{D\tau_s}$ on momentum scattering. 
4053: Since $D\sim \tau_p$, for EY $L_s\sim \tau_p$, while for DP $L_s$ does 
4054: not depend on 
4055: the momentum
4056: scattering time and for a degenerate electron system should be  a constant 
4057: independent of
4058: $T$, of the order of $v_F/\Omega_{av}$.  
4059: We do not know of an experimental verification of this distinction.
4060: 
4061: If the electrons obey nondegenerate statistics, which is the usual case of 
4062: p-doped 
4063: materials, thermal averaging over the Boltzmann distribution 
4064: gives~\cite{Pikus:1984}. 
4065: \begin{equation}\label{eq:relax:DPt}
4066: 1/\tau_s=Q \tau_m \alpha^2 \frac{(k_B T)^3}{\hbar^2 E_g},
4067: \end{equation} 
4068: where $\tau_m=\langle \tau_p(E_{\bf k}) E_{\bf k}\rangle/\langle E_{\bf k} 
4069: \rangle$.
4070: The coefficient $Q$, which is of order 1, is 
4071: \begin{equation}
4072: Q=\frac{16}{35}\gamma_3^{-1} \left ( \nu+\frac{7}{2} \right )
4073: \left (\nu+\frac{5}{2} \right ),
4074: \end{equation}
4075: where the power law $\tau_p\sim E^{\nu}_{\bf k}$ is assumed for momentum 
4076: relaxation time.
4077: For scattering by ionized impurities $Q\approx 1.5$, while scattering by polar 
4078: optical
4079: or piezoelectric phonons gives $Q\approx 0.8$, and scattering by acoustic 
4080: phonons 
4081: (deformation potential) $Q \approx 2.7$~\cite{Pikus:1984}. The temperature 
4082: behavior of
4083: DP spin dephasing in nondegenerate samples is 
4084: $1/\tau_s\sim T^3\tau_m(T)$. 
4085: For 
4086: scattering
4087: by charged impurities $1/\tau_s\sim T^{9/2}$.
4088: 
4089: Application of longitudinal (to the initial spin direction) 
4090: magnetic field suppresses the DP mechanism~\cite{Pikus:1984} for two 
4091: reasons: (i) The $B$-field 
4092: suppresses
4093: precession along the transverse intrinsic 
4094: fluctuating fields when $\Omega_L\tau_p > 1$, where $\Omega_L$ is 
4095: the Larmor precession due to $B$. (ii)  $\Omega_{\bf k}$ 
4096: is orbitally averaged,
4097: which 
4098: has a similar effect to averaging by random scattering,  when 
4099: $\Omega_c\tau_p>1$, where $\Omega_c$ is the cyclotron frequency. 
4100: Since for conduction electrons  $m_c \ll m_e$, it follows that
4101: $\Omega_c \gg \Omega_L$, the orbital motion induced by $B$ is the cause for 
4102: suppression
4103: of spin relaxation in semiconductors.    
4104: %
4105: %Theoretically spin relaxation and spin dephasing in bulk n-type GaAs was 
4106: %investigated
4107: %in \cite{Wu2000:PSS,Wu2001:JS} by solving numerically kinetic equations. Only 
4108: %the DP effect was studied. Acoustic and longitudinal phonon and impurity 
4109: %scattering 
4110: %were included. 
4111: %The main findings are: (i) Spin dephasing increases with increasing impurity 
4112: %density up to a critical density after which the dephasing time decreases. 
4113: %(ii) As a function of temperature, spin dephasing time increases with 
4114: %increasing 
4115: %temperature at small impurity content, but decreases with increasing 
4116: %temperature
4117: %at large impurity density (the latter point is in contrast to previous 
4118: %calculations).
4119: %It is also found that spin dephasing time increases with decreasing electron
4120: %density. (iii) Spin dephasing time decreases with increasing electron density,
4121: %and (iv) spin dephasing time increases with increasing magnetic field (only 
4122: %spin,
4123: %no orbital effects included).
4124: 
4125: 
4126: \paragraph{\label{sec:IIIB2b} Two-dimensional III-V semiconductor systems.}
4127: %----------------------------------------------------------------------
4128: 
4129: In two-dimensional III-V semiconductor systems (quantum wells and 
4130: heterostructures) there are 
4131: two distinct Hamiltonian terms that contribute
4132: to DP spin dephasing: the bulk inversion asymmetry term $H_{\rm BIA}$  
4133: and the structure inversion asymmetry term, $H_{\rm SIA}$, which appears 
4134: only in 
4135: asymmetric systems. Both $H_{\rm BIA}$ and
4136: $H_{\rm SIA}$ lead to spin splitting of the conduction band linear in 
4137: ${\bf k}$. The two 
4138: terms, however, predict a different dependence of $\tau_s$ on the 
4139: quantum-well 
4140: orientation relative to the principal axes. Figure \ref{fig:omega} shows 
4141: the vector field patterns of the intrinsic magnetic fields for both
4142: bulk and spin inversion asymmetry.
4143: 
4144: The bulk inversion asymmetry term comes from the bulk Dresselhaus 
4145: spin splitting, Eq.~(\ref{eq:relax:dress}). 
4146: Treating wave vectors ${\bf k}$ in Eq.~(\ref{eq:relax:dress}) as operators
4147: $\hat{\bf k}=-i\nabla$, and evaluating $\bf\Omega$ as the expectation value in 
4148: the 
4149: confined states, leads to momentum quantization along the confinement  
4150: unit vector
4151: ${\bf n}$ of the quantum well (QW). In the following, ${\bf k}$ denotes the 
4152: wave vector for a Bloch state propagating in the plane, and 
4153: ${k_n^2}\equiv\langle (\hat{\bf k}\cdot {\bf n})^2\rangle$ 
4154: denotes the expectation value 
4155: of the square 
4156: of the component of the wave-number 
4157: operator 
4158: normal to the plane 
4159: in the lowest-subband state. For a rectangular QW of width $a$, 
4160: $k_n^2=(\pi/a)^2$.
4161: For a triangular well of confining potential $eEz$, 
4162: $k_n^2 \approx 0.7794 (2m_e E/\hbar^2)^{2/3}$ [see, for example,
4163: \textcite{deSousa2003:PRBc}].
4164: Quantum averaging of $\kappa$ can be done using the formula
4165: \begin{equation}
4166: \langle \hat{k}_i \hat{k}_j \hat{k}_l \rangle 
4167: = k_n^2 \left (k_i n_j n_l + k_j n_l n_i + k_l n_i n_j  \right ) 
4168: + k_i k_j k_l. 
4169: \end{equation}
4170: This readily gives~\cite{Dyakonov1986:SPS}
4171: \begin{equation}\label{eq:relax:kappa2d}
4172: \kappa_x= k_n^2  \left [2 n_x\left (n_y k_y -n_z k_z \right ) + 
4173: k_x \left ( n_y^2 -n_z^2 \right ) \right ],
4174: \end{equation}
4175: and similarly for $\kappa_y$ and $\kappa_z$ by index permutation. 
4176: Terms cubic in $k$ were omitted from the above equation, assuming that for 
4177: narrow QW's 
4178: $k^2\ll k_n^2$. The explicit knowledge of $\kappa$ is useful in qualitative 
4179: analysis
4180: of spin dephasing for particular orientations of QW's.
4181: 
4182: 
4183: The spin dephasing tensor $1/\tau_{s,ij}$, defined in 
4184: Eq.~(\ref{eq:relax:tauii}), 
4185: is readily evaluated using 
4186: Eqs.~(\ref{eq:relax:dress}) and 
4187: (\ref{eq:relax:kappa2d})\footnote{Averaging over the directions
4188: of $\bf k$ in a plane perpendicular to $\bf n$ can be performed by using 
4189: $\overline{k_ik_j}=
4190: (k^2/2)(\delta_{ij}-n_in_j)$.}
4191: \begin{eqnarray}\label{eq:relax:nu}
4192: 1/\tau_{s,ij}=\left (\delta_{ij}{\rm Tr} \hat{\nu} 
4193: - \nu_{ij}  \right )/\tau_{s}^0(E_{\bf k}),  
4194: \end{eqnarray}
4195: where 
4196: \begin{equation}\label{eq:relax:DK}
4197: \frac{1}{\tau_{s}^0(E_{\bf k})}=\frac{\alpha^2 \hbar^2 
4198: \left (k_n^2\right) ^2}{2 m_c^2 E_g} 
4199: E_{\bf k}\tau_p(E_{\bf k}).
4200: \end{equation}
4201: The tensor $\hat{\nu}$ depends on the orientation of $\bf n$ with
4202: respect to the principal crystal 
4203: axes~\cite{Dyakonov1986:SPS}\footnote{A trivial
4204: typo in $\nu_{xx}$ is corrected.} 
4205: \begin{eqnarray}
4206: \nu_{xx}&=&4n_x^2(n_y^2+n_z^2)-(n_y^2-n_z^2)^2(9n_x^2-1), \\
4207: \nu_{xy}&=&n_xn_y\left[9(n_x^2-n_z^2)(n_y^2-n_z^2)-2(1-n_z^2)\right],
4208: \end{eqnarray}
4209: and analogously for other components. 
4210: 
4211: \begin{figure}
4212: \centerline{\psfig{file=zutic_fig15.eps,width=1\linewidth,angle=0}}
4213: \caption{Vector fields ${\bf\Omega}({\bf k})\sim \kappa({\bf k})$ 
4214: on the Fermi surface (here a circle) 
4215: for the structure 
4216: inversion asymmetry
4217: (SIA)
4218: and bulk 
4219: inversion asymmetry (BIA). 
4220: Since 
4221: ${\bf\Omega}({\bf k})$
4222: is also the spin quantization axis, the vector pattern is also the 
4223: pattern of the spin on 
4224: the Fermi surface. As 
4225: the opposite spins have different energies, the Fermi circle becomes 
4226: two concentric circles with opposite spins. This is shown here only for 
4227: the SIA case,
4228: but the analogy extends to all examples. The field for BIA [110] lies 
4229: perpendicular to
4230: the plane, with the magnitude varying along the Fermi surface. All other 
4231: cases have 
4232: constant fields lying in the plane.  
4233: }
4234: \label{fig:omega}
4235: \end{figure}
4236: 
4237: We follow \textcite{Dyakonov1986:SPS} in discussing the three important 
4238: cases of 
4239: [001], [111], and [110] quantum wells. For [001], 
4240: \begin{equation}
4241: {\bf\Omega}({\bf k})\sim\kappa=k_n^2 (-k_x,k_y,0).
4242: \end{equation}
4243: While the magnitude of ${\bf\Omega}({\bf k})$ is  
4244: constant over the Fermi surface,
4245: the directions follow a ``breathing'' pattern as shown in Fig.~\ref{fig:omega}. 
4246: The spin relaxation times follow from 
4247: Eq.~(\ref{eq:relax:DK}): 
4248: $1/\tau_{s,xx}=1/\tau_{s,yy}=1/2\tau_{zz}=1/\tau_{s}^0$.
4249: Defining $1/\tau_{s\parallel}$ and $1/\tau_{s,\perp}$ as spin dephasing 
4250: rates of 
4251: spins parallel and perpendicular to the plane, one obtains
4252: \begin{equation}
4253: 1/\tau_{s,\parallel}=1/2\tau_{s,\perp}=1/\tau_{s}^0.
4254: \end{equation}
4255: As expected for the case of the in-plane field, the lifetime of a spin 
4256: parallel to 
4257: the plane is twice that of the a spin perpendicular to the plane.
4258: 
4259: For [111] QW's,
4260: \begin{equation}
4261: {\bf\Omega}({\bf k})\sim\kappa = 2/\sqrt{3} k_n^2 ({\bf k}\times {\bf n}).
4262: \end{equation}
4263: The intrinsic magnetic field lies in the plane, having a constant magnitude 
4264: (refer to Fig.~\ref{fig:omega}). 
4265: Spin relaxation rates are now 
4266: $1/\tau_{s,ii}=16/9\tau_{s}^0$ and $1/\tau_{s,i\ne j}=4/9\tau_{s}^0$. 
4267: By diagonalizing $1/\tau_{ij}$  we obtain
4268: \begin{equation}
4269: 1/\tau_{s\parallel}=1/2\tau_{s,\perp}=4/3\tau_{s}^0.
4270: \end{equation}
4271: As for the [001] case, a perpendicular spin dephases twice as fast as a 
4272: parallel one, since 
4273: ${\bf\Omega}({\bf k})$ lies in the plane.
4274: 
4275: The most interesting case is the [110] orientation for which 
4276: $1/\tau_{s,xx}=1/\tau_{s,yy}=1/2\tau_{s,zz}=-1/\tau_{s,xy}=1/8\tau_{s}^0$. 
4277: Other
4278: off-diagonal components vanish. Diagonalizing the tensor  gives
4279: \begin{equation}
4280: 1/\tau_{s,\parallel}=1/4\tau_{s}^0,\,\,\,1/\tau_{s,\perp}=0.
4281: \end{equation}
4282: The perpendicular spin does not dephase. This is due to the fact that $\kappa$, 
4283: unlike the previous cases, is always normal to the plane 
4284: (see Fig.~\ref{fig:omega}), 
4285: and thus cannot
4286: affect the precession of the perpendicular spin. Indeed, 
4287: \begin{equation}
4288: {\bf\Omega}({\bf k})\sim\kappa = k_n^2 \left (k_x/2) (-1,-1, 0 \right ),
4289: \end{equation}
4290: where it is used that ${\bf k}\cdot {\bf n}=0$. Spin dephasing in [110] QW's can 
4291: still be 
4292: due to the cubic terms in $k$ left out of Eq.~(\ref{eq:relax:kappa2d}) or to 
4293: other spin relaxation mechanisms. Note that the magnitude of 
4294: ${\bf\Omega}({\bf k})$
4295: changes along the Fermi surface. 
4296: Electrons moving along [001] experience
4297: little spin dephasing.
4298: 
4299: The structure inversion asymmetry term arises from the Bychkov-Rashba spin 
4300: splitting ~\cite{Bychkov1984:JETPL,Bychkov1984:JPC,Rashba1960:SPSS} 
4301: occurring in asymmetric QW's 
4302: or in deformed bulk systems. The corresponding Hamiltonian is that
4303: of Eq.~(\ref{eq:HDP}), with the precession vector
4304: \begin{equation} \label{eq:relax:BR}
4305: {\bf\Omega}({\bf k})=2\alpha_{BR} ({\bf k}\times {\bf n}). 
4306: \end{equation}
4307: Here $\alpha_{BR}$ is a parameter depending on spin-orbit coupling and the 
4308: asymmetry
4309: of the confining electrostatic potentials arising from the growth 
4310: process of the
4311: heterostructure. The splitting can also arise in nominally symmetric 
4312: heterostructures
4313: with fluctuations in doping density \cite{Sherman2003:APL}.  
4314: The Bychkov-Rashba field always lies
4315: in the plane, having a constant magnitude. As for the bulk inversion 
4316: asymmetry case, the
4317: structure inversion asymmetry leads to a splitting of the Fermi surface, 
4318: according
4319: to the direction of the spin pattern---parallel or antiparallel to 
4320: $\bf{\Omega}({\bf k})$,
4321: as shown in Fig.~\ref{fig:omega}. Perhaps the most appealing fact about 
4322: structure
4323: inversion asymmetry is that $\alpha_{BR}$ can be tuned electrostatically, 
4324: potentially
4325: providing an effective spin precession control without the need for magnetic 
4326: fields 
4327: \cite{Rashba2003:PRL,Levitov2003:PRB}.
4328: This has led to one of the pioneering spintronic proposals 
4329: by \textcite{Datta1990:APL} (see Sec.~\ref{sec:IVE1}). 
4330: Note that for the [111] orientation the bulk
4331: and structure inversion asymmetry terms have the same form. 
4332: 
4333: 
4334: Using the same procedure as for bulk inversion asymmetry, 
4335: we describe the 
4336: spin relaxation rate by Eq.~(\ref{eq:relax:nu}) as 
4337: \begin{equation}
4338: 1/\tau_{s}^0=4\alpha_{BR}^2\frac{m_c}{\hbar^2} E_{\bf k} \tau_p
4339: \end{equation}
4340: and
4341: \begin{equation}\label{eq:relax:nusia}
4342: \nu_{ij}=1-n_in_j.
4343: \end{equation}
4344: Since the intrinsic precession vector $\sim {\bf \Omega}({\bf k})$ for the 
4345: structure inversion
4346: asymmetry always lies in the plane, 
4347: a perpendicular spin should dephase 
4348: twice as fast as a spin in the plane. Indeed,
4349: by diagonalizing $1/\tau_{s,ij}$ one finds that 
4350: \begin{equation}
4351: 1/\tau_{s,\parallel}=1/2\tau_{s,\perp}=1/\tau_s^0
4352: \end{equation}
4353: holds for all QW orientations ${\bf n}$. 
4354: This interesting fact qualitatively
4355: distinguishes structure from bulk inversion asymmetry and can be used in 
4356: assessing the 
4357: relative importance of
4358: the Dresselhaus and Rashba spin splittings in III-V heterostructure systems. 
4359: If bulk and structure inversion asymmetry are of similar importance, 
4360: the interference terms from the cross 
4361: product $\overline{\Omega_{\rm BIA} \Omega_{\rm SIA}}$ can lead to spin 
4362: dephasing 
4363: anisotropies within the plane, as was shown for [001] 
4364: QW's  
4365: in~ \cite{Averkiev1999:PRB,Kainz2003:PRB}. This plane anisotropy can be 
4366: easily seen by adding
4367: the corresponding vector fields in Fig.~\ref{fig:omega}. Another interesting
4368: feature of bulk and structure inversion asymmetry fields is that injection
4369: of electrons along a quasi-one-dimensional channel can lead to large relaxation
4370: times for spins oriented along $\bf{\Omega}({\bf k})$, where ${\bf k}$ is
4371: the wave vector for the states in the channel \cite{Hammar2002:PRL}.
4372: 
4373: Model spin dephasing calculations based on structure inversion asymmetry 
4374: were carried out 
4375: by \textcite{Pareek2002:PRB}. 
4376: Calculations of $\tau_s$ based on the DP mechanism, 
4377: with structural asymmetry due to doping fluctuations in the heterostructure 
4378: interface were performed by \textcite{Sherman2003:APL}.
4379: 
4380: Research on spin inversion asymmetry is largely motivated by Datta-Das
4381: spin field-effect transistor proposal (see Secs.~\ref{sec:IA} and
4382: \ref{sec:IVE1}) in which $\alpha_{BR}$ is tailored by a gate.
4383: This tailoring, however, has been controversial and the microscopic
4384: origin of the Bychkov-Rashba Hamiltonian, and thus the interpretation
4385: of experimental results on splitting in semiconductor heterostructures, 
4386: has been debated. The Bychkov-Rashba Hamiltonian is often interpreted 
4387: as arising from the electric field of the confining potential, assisted by
4388: external bias, which acts on a moving electron in a transverse direction.
4389: The relativistic transformation then gives rise to a magnetic field
4390: (spin-orbit coupling) acting on the electron spin. The parameter 
4391: $\alpha_{BR}$ is then assumed to be directly proportional to the 
4392: confining electrical field. This is in general wrong, since the average
4393: electric force acting on a confined particle of uniform effective mass
4394: is zero.
4395: 
4396: The asymmetry that gives rise to structure inversion asymmetry is the
4397: asymmetry in the band structure (including spin-orbit coupling)
4398: parameters of a heterostructure, such as the effective mass, or the
4399: asymmetry in the penetration of the electron wave function into the
4400: barriers \cite{Silva1997:PRB}. The difficulty in understanding the 
4401: influence of the external gates is caused by the lack of the
4402: understanding of the influence of the gate field on the asymmetry
4403: of the well. For a clear qualitative explanation of the involved
4404: physics see \textcite{Pfeffer1999:PRB} and \textcite{Zawadzki2001:PRB}.
4405: Band-structure {\bf $k \cdot p$} calculations of $\alpha_{BR}$ for
4406: quantum wells in GaAs/AlGaAs heterostructures can be found in 
4407: \textcite{Pfeffer1995:PRB}, \textcite{Pfeffer1997:PRB},
4408: \textcite{Wissinger1998:PRB}, and \textcite{Kainz2003:PRB};
4409: a calculation for InGaAs/InP quantum wells is reported in 
4410: \textcite{Engels1997:PRB} and \textcite{Schapers1998:JAP},
4411: in InSb/InAlSb asymmetric quantum well it can be found in 
4412: \textcite{Pfeffer2003:PRB}, and in p-InAs metal-oxide-semiconductor
4413: field-effect transistor channel in \textcite{Lamari2003:PRB}.
4414: Adding to the controversy, \textcite{Majewski:2003} have recently
4415: calculated the structure inversion asymmetry by local density functional
4416: methods and concluded that the induced spin splitting arises from
4417: microscopic electric fields in asymmetric atomic arrangements at the 
4418: interfaces, so that a large Bychkov-Rashba term can be present in 
4419: otherwise symmetric quantum wells with no common atom.
4420: 
4421: Interpretation of experimental data on structure inversion asymmetry is
4422: difficult, especially at determining the zero magnetic field spin splitting
4423: (usually seen in Shubnikov- de Haas oscillations), which is masked by
4424: Zeeman splitting at finite fields. In addition, the splittings are small,
4425: typically less than 1 meV. The Bychkov-Rashba parameter was measured in
4426: GaSb/InAs/GaSb quantum wells 
4427: ($\alpha_{BR} \approx 0.9 \times 10^{-11}$ eV m for 75 $\AA$  
4428: thick well)
4429: by \textcite{Luo1990:PRB}; in InAlAs/InGaAs/InAlAs quantum wells (20 nm),
4430: where also the gate voltage is obtained: $\alpha_{BR}$ ranged from
4431: $10^{-11}$ eV m at the depleting voltage of -1 V, to $5 \times 10^{-12}$
4432: eV m at +1.5 V. Weak antilocalization studies of InAlAs/InGaAs/InAlAs
4433: quantum wells have recently been used to study electron density dependence
4434: of $\alpha_{BR}$ by \textcite{Koga2002:PRL}. Gate dependence of $\alpha_{BR}$
4435: was also measured in modulation-doped InP/InGaAs/InP quantum wells
4436: \cite{Engels1997:PRB,Schapers1998:JAP}. The observed values are several
4437: $10^{-12}$ eV m. On the other hand, there are experimental reports that
4438: either fail to observe the expected spin splitting due to 
4439: Bychkov-Rashba field, or interpret the splitting differently [see, for example,
4440: \textcite{Brosig1999:PRB} and \textcite{Rowe2001:PRB}].
4441: Furthermore, measurements of \textcite{Heida1998:PRB} show a constant
4442: $\alpha_{BR} \approx 0.6 \times 10^{-11}$ eV m, independent of gate
4443: voltage, in asymmetric AlSb/InAS/AlSb quantum wells, demonstrating that
4444: control of $\alpha_{BR}$ may be difficult. In order to unify the
4445: different views on what exactly the Bychkov-Rashba spin splitting means
4446: and how the spin splitting can be tuned with gate voltage, more
4447: experimental efforts need to be devoted to this interesting topic.
4448: 
4449: \subsubsection{\label{sec:IIIB3} Bir-Aronov-Pikus Mechanism} 
4450: %---------------------------------------------------------------------------
4451: 
4452: Spin relaxation of conduction electrons in p-doped semiconductors can also 
4453: proceed through scattering, accompanied by spin exchange, by holes, 
4454: as was first 
4455: shown by \textcite{Bir1976:SPJETP}. 
4456: 
4457: The exchange interaction between electrons and holes is governed by the 
4458: Hamiltonian
4459: \begin{equation}
4460: H= A{\bf S}\cdot {\bf J} \delta({\bf r}),
4461: \end{equation}
4462: where $A$ is proportional to the exchange integral between the conduction and 
4463: valence states,
4464: ${\bf J}$ is the angular momentum operator for holes, ${\bf S}$ is the electron 
4465: spin operator,
4466: and ${\bf r}$ is the relative position of electrons and holes.
4467: 
4468: The spin-flip scattering probability depends on the state of the holes 
4469: (degenerate or
4470: nondegenerate, bound on acceptors or free, fast or slow). We present below the 
4471: most 
4472: frequently used formulas when assessing the relevance of the BAP mechanism. 
4473: The formulas are 
4474: valid for 
4475: the usual cases of heavy holes $m_v \gg m_c$. For electron spin relaxation 
4476: due to exchange with 
4477: nondegenerate
4478: holes, 
4479: \begin{equation}\label{eq:relax:BAP1}
4480: \frac{1}{\tau_s}= \frac{2}{\tau_0}N_a a_B^3 \frac{v_{\bf k}}{v_B} \left 
4481: [\frac{p}{N_a}
4482: |\psi(0)|^4 + \frac{5}{3}\frac{N_a-p}{N_a} \right],
4483: \end{equation}
4484: where $a_B$ is the exciton Bohr radius $a_B=\hbar^2\epsilon/e^2m_c$,
4485: $p$ is the density of free holes, $\tau_0$ is an exchange splitting parameter:
4486: $\hbar/\tau_0=(3\pi/64)\Delta_{\rm ex}^2/E_B$ (with $E_B$ denoting the Bohr 
4487: exciton
4488: energy, $E_B=\hbar^2/2m_ca_B^2$ and $\Delta_{\rm ex}$ the exchange splitting 
4489: of the
4490: excitonic ground state), and $v_B=\hbar/m_ca_B$; $|\psi(0)|^2$ is Sommerfeld's 
4491: factor, which
4492: enhances the free hole contribution. 
4493: For an unscreened Coulomb potential
4494: \begin{equation}
4495: |\psi(0)|^2 = \frac{2\pi}{\kappa} \left 
4496: [1-\exp\left(-\frac{2\pi}{\kappa}\right) 
4497: \right ]^{-1},
4498: \end{equation}
4499: where $\kappa=E_{\bf k}/E_B$. For a completely screened 
4500: potential $|\psi(0)|^2=1$.
4501: 
4502: If holes are degenerate and the electrons' velocity $v_{\bf k}$ 
4503: is greater than the
4504: Fermi velocity of  the holes', then  
4505: \begin{equation}\label{eq:relax:BAP2}
4506: \frac{1}{\tau_s}=\frac{3}{\tau_0}p a_B^3 \frac{v_{\bf k}}{v_B}\frac{k_B 
4507: T}{E_{Fh}},
4508: \end{equation}
4509: where $E_{Fh}$ is the hole Fermi energy. For degenerate holes $|\psi(0)|^2$ is 
4510: of order 1.
4511: If electrons are thermalized, $v_{\bf k}$ needs to be replaced by the thermal 
4512: velocity $v_e=(3k_B T/m_c)^{1/2}$.
4513: 
4514: The temperature dependence of $\tau_s$ is dominated by the temperature 
4515: dependence of $|\psi(0)|^2$
4516: as well as by $p$. The dependence on the 
4517: acceptor density
4518: is essentially $1/\tau_s \sim N_a$ for nondegenerate/bound holes from 
4519: Eq.~(\ref{eq:relax:BAP1}) and $1/\tau_s \sim N_a^{1/3}$ for degenerate holes 
4520: from 
4521: Eq.~(\ref{eq:relax:BAP2}). In between, $1/\tau_s$ is only weakly dependent 
4522: on $N_a$.
4523: For GaAs $a_B\approx 114$ \AA, $E_B \approx 4.9$ meV, 
4524: $v_B\approx 1.7 \times 10^7$ cm$\cdot$s$^{-1}$, 
4525: $\tau_0\approx 1\times 10^{-8}$ s, and 
4526: $\Delta_{\rm ex}\approx 4.7 \times 10^{-5}$ eV~\cite{Aronov1983:SPJETP}.
4527: 
4528: The BAP mechanism coexists with the the EY and DP mechanisms 
4529: in p-doped materials lacking
4530: inversion symmetry. The three mechanisms can be distinguished by their
4531: unique density and temperature dependences. A general trend is that the
4532: BAP dominates in heavily doped samples at small temperatures. At large
4533: temperatures even for large acceptor densities, the DP mechanism can become
4534: more important, due to its increased importance at large electron energies. 
4535: Specific examples of the domain of importance for the three
4536: mechanisms are discussed in Sec.~\ref{sec:IIID1}.  
4537: Model calculations of BAP processes for electrons in p-doped bulk and quantum 
4538: wells were
4539: performed by \textcite{Maialle1996:PRB,Maialle1997:PRB}. 
4540: 
4541: Another potentially relevant mechanism for spin relaxation of donor-bound 
4542: electrons in p-doped semiconductors is the
4543: exchange interaction with holes bound to acceptors~\cite{Dyakonov1974:SPJETP}. 
4544: The exchange interaction provides an effective magnetic field for electron 
4545: spins 
4546: to precess, leading to inhomogeneous dephasing. Both electron hopping and hole 
4547: spin-flip motionally narrow the precession.
4548: 
4549: \subsubsection{\label{sec:IIIB4} Hyperfine-interaction mechanism}
4550: %--------------------------------------------------------------------------
4551: 
4552: The hyperfine interaction, which is the magnetic interaction between 
4553: the magnetic moments of electrons and nuclei, provides an important 
4554: mechanism~\cite{Dyakonov1974:SPJETP} for ensemble spin dephasing
4555: and single-spin decoherence of localized electrons, such as those confined 
4556: in quantum dots (QD) or bound on donors. The interaction is too weak to cause 
4557: effective 
4558: spin relaxation
4559: of free electrons in metals or in bulk semiconductors~\cite{Overhauser1953:PR},
4560: as it is strongly dynamically narrowed by the itinerant nature of 
4561: electrons (see Sec.~\ref{sec:IIIA1}).
4562: In addition to spin dephasing, the hyperfine interaction
4563: is relevant for spintronics as a means to couple, in a controlled way, 
4564: electron 
4565: and nuclear spins \cite{Dyakonov:1984}.
4566: 
4567: Localized electrons are typically spread over many lattice sites 
4568: ($10^4$--$10^6$),
4569: experiencing the combined magnetic moments of many nuclei. In GaAs all the 
4570: lattice nuclei carry the
4571: magnetic moment of 3/2 spin, while in Si the most abundant isotope, $^{28}$Si,
4572: carries no spin and the hyperfine interaction is due to $^{29}$Si 
4573: (natural abundance 4.67\%) 
4574: or the frequent donor $^{31}$P, both of nuclear spin $1/2$. As a result, an 
4575: electron bound
4576: on a shallow donor in Si experiences only around 100 magnetic nuclei, and the 
4577: effects of the hyperfine interaction are considerably smaller than in GaAs.
4578: 
4579: The effective Hamiltonian for the hyperfine interaction is the 
4580: Fermi contact potential 
4581: energy~\cite{Slichter:1989} 
4582: \begin{equation}
4583: H=\frac{8\pi}{3} \frac{\mu_0}{4\pi}g_0 \mu_B \sum_i \hbar \gamma_{n,i} 
4584: {\bf S}\cdot{\bf I}_i
4585: \delta ({\bf r}- {\bf R}_i),
4586: \end{equation}
4587: where $\mu_0$ is the vacuum permeability, $g_0=2.0023$ is the free-electron 
4588: $g$ factor, $\mu_B$
4589: is the Bohr magneton, $i$ is the label for nuclei at positions 
4590: ${\bf R}_i$, ${\bf S}$,  and ${\bf I}_i$ are, respectively,  electron and 
4591: nuclear spin operators expressed in the units of $\hbar$, and $\gamma_{n,i}$ 
4592: is the nuclear 
4593: gyromagnetic ratio.
4594: We stress that it is the electron $g$ factor $g_0$ and not the effective
4595: $g$ that appears in the hyperfine interaction, Eq.~(96), %label iz
4596: as shown by \textcite{Yafet1961:JPCS} [see also \textcite{Paget1977:PRB}].
4597: It follows that the spin of an electron in an orbital state $\psi({\bf r})$ 
4598: experiences 
4599: magnetic field 
4600: \begin{equation}
4601: {\bf B}_n=\frac{2\mu_0}{3} \frac{g_0}{g} \sum_i \hbar \gamma_{n,i} {\bf I}_i 
4602: |\psi({\bf R}_i)|^2,
4603: \end{equation}
4604: where $g$ is the effective $g$ factor of the electron. The electron Zeeman 
4605: splitting 
4606: due to the average $B_n$  
4607: corresponds to a field of  $\sim 1$ 
4608: T or thermal energy of 1 K, for a complete nuclear 
4609: polarization~\cite{Paget1977:PRB}.
4610: 
4611: There are three important regimes in which the hyperfine interaction leads to 
4612: spin dephasing of 
4613: localized
4614: electrons:
4615: 
4616: (i) In the limit of small orbital and spin correlation between 
4617: separated 
4618: electron states and nuclear spin states, spatial variations in ${\bf B}_n$ 
4619: lead to 
4620: inhomogeneous dephasing of 
4621: the spin ensemble, with the rate proportional to the r.m.s. of $B_n$, 
4622: given by the 
4623: corresponding
4624: thermal or nonequilibrium distribution of the nuclear spins. 
4625: Such inhomogeneous dephasing is seen by electron spin resonance (ESR) 
4626: experiments 
4627: on donor states both in Si \cite{Feher1959b:PR} and in GaAs \cite{Seck1997:PRB}. 
4628: This
4629: effect can be removed by spin echo experiments (in Si donor states performed, 
4630: for example, by
4631:  \textcite{Gordon1958:PR}). The spread in the Larmor precession period due to 
4632: the variance in 
4633: $B_n$ in GaAs is estimated to be around 1 ns 
4634: \cite{Merkulov2002:PRB,Dzhioev2002:PRL}).
4635: 
4636: (ii) Temporal fluctuations in $B_n$, 
4637: which can occur due to nuclear dipole-dipole 
4638: interactions, lead to irreversible spin dephasing and decoherence of electron 
4639: spins.
4640: Such processes are sometimes referred to as spectral diffusion, since the 
4641: electron Zeeman levels split by $B_n$ undergo random shifts \cite{Sousa2002:P}.
4642: The typical time scale for the fluctuations in GaAs is 
4643: given by the nuclear Larmor precession period in the field of neighboring 
4644: nuclei and
4645: is of order 100 $\mu$s~\cite{Merkulov2002:PRB}. Nuclear moments 
4646: also precess (and orient) in the magnetic fields of polarized electrons, an 
4647: effect important in 
4648: optical orientation \cite{Meier:1984}, where the feedback from this precession 
4649: can 
4650: be directly observed through the modulated precession of electron spins. The 
4651: time
4652: scale for the Larmor precession of nuclear spins in hyperfine fields is 1 $\mu$s in 
4653: GaAs 
4654: \cite{Merkulov2002:PRB}, so this effect does not lead to motional narrowing of 
4655: $B_n$;
4656: electron spins precess many times before the nuclear spin flips.
4657: 
4658: (iii) In the presence of strong orbital correlations (electron hopping or 
4659: recombination with
4660: acceptor hole states) or spin (direct exchange interaction) between
4661: neighboring electron states, spin precession due to $B_n$ is motionally 
4662: narrowed.
4663: While the direct spin exchange interaction does not cause ensemble spin 
4664: relaxation 
4665: (the total 
4666: spin is preserved in spin flip-flops), it leads to individual spin decoherence, 
4667: which 
4668: can be much faster than what is inferred from $T_2$. This 
4669: effect 
4670: is much more pronounced in GaAs than in Si, since the donor states spread to 
4671: greater
4672: distances, and thus even in the low-doping limits ($\approx 10^{14}$ cm$^{-3}$ 
4673: donors)
4674: the exchange interaction can be rather large, masking the effects of temporal 
4675: fluctuations
4676: of $B_n$ (see Sec.~\ref{sec:IIID3}). 
4677: Many useful parameters for evaluating effective magnetic fields and precession 
4678: frequencies
4679: due the HFI mechanism in GaAs are given by \textcite{Paget1977:PRB}.
4680: 
4681: Ensemble spin dephasing due to the HFI mechanism in an external magnetic 
4682: field has been studied 
4683: by \textcite{Dyakonov1974:SPJETP}, who found suppression of $1/\tau_s$ 
4684: if the external field 
4685: is greater than $B_n$ for regime (i), or a smaller Larmor precession 
4686: period than the correlation time 
4687: for random changes 
4688: in $B_n$, in regime (iii).
4689: due to the external field. 
4690: 
4691: Calculations of $\tau_s$ using the HFI mechanism  were performed for shallow 
4692: donor states in Si 
4693: at low 
4694: temperatures and magnetic fields~\cite{Saikin2002:NL}, for electron spins in 
4695: QD's~\cite{Merkulov2002:PRB,Semenov2002:P}, and even for the case of 
4696: conduction 
4697: electrons in semiconductors (revisited by 
4698: \textcite{Pershin2003:P}). Spin relaxation processes due 
4699: to phonon-assisted HFI  were investigated in GaAs QD's, but were found to be 
4700: ineffective \cite{Erlingsson2001:PRB}. Unfortunately, there are still 
4701: too few experimental 
4702: data to make conclusions about the merits of specific models of the 
4703: HFI mechanism. 
4704: 
4705: Spin decoherence times for single-electron spins were recently computed 
4706: for  
4707: case 
4708: (ii) by \textcite{Khaetskii2002:PRL,Khaetskii2002:P},
4709: who studied spin coherence time $\tau_{sc}$ of a single electron spin 
4710: in the regime in which the electron Larmor period due to $B_n$ is much
4711: shorter than the correlation time of the nuclear magnetic field fluctuations.
4712: Realistic estimates 
4713: of HFI spin dephasing in GaAs QD's were given 
4714: by \textcite{Sousa2003:PRB,Sousa2002:P}, who offer
4715: reasons why mechanism (ii)
4716: should dominate spin decoherence in GaAs QD's of radius smaller than 100 nm.
4717: For instance,  in a 50 nm wide QD the estimated $\tau_{sc}$
4718: is $\approx 50$ $\mu$s, 
4719: large enough 
4720: for quantum computing applications (see Sec.~\ref{sec:IVF}).
4721: This claim is supported by the recent measurement of the spin dephasing
4722: time of about 60 ms of an isolated spin in a phosphorus donor in
4723: isotopically pure $^{28}$Si, by spin echo measurements 
4724: \cite{Tyryshkin2003:PRB}. 
4725: 
4726: 
4727: \subsection{\label{sec:IIIC} Spin relaxation in metals}
4728: %--------------------------------------------------------------------------
4729: 
4730: The spin relaxation time of conduction electrons 
4731: in metals has been measured by both CESR and spin injection techniques. 
4732: Typical values of $\tau_s$ were found to be 0.1 to 1 ns, but the range
4733: of observed values is large, from pico- to microseconds. To our knowledge the 
4734: longest $\tau_s$ reported for a metal--a microsecond--was found in high-purity
4735: sodium below 10 K~\cite{Kolbe1971:PRB}. 
4736: 
4737: The majority of simple metals are believed to follow the EY mechanism of spin 
4738: relaxation,
4739: with the possible exception of Li~\cite{Damay1976:PRB}. This is
4740: supported by several facts: 
4741: 
4742: (i) The EY processes give the right order of 
4743: magnitude
4744: for $\tau_s$ ~\cite{Elliott1954:PR,Yafet:1963},
4745: while other known possible spin relaxation mechanisms lead to 
4746: much greater $\tau_s$ than what is observed~\cite{Overhauser1953:PR}.
4747: 
4748: (ii) The temperature dependence of $\tau_s$ is consistent with
4749: the EY mechanism: $1/\tau_s$ is constant at low temperatures, indicating 
4750: impurity
4751: spin-flip scattering, while at high temperatures $1/\tau_s$ grows linearly with 
4752: increasing
4753: $T$, consistent with phonon-induced spin relaxation. 
4754: 
4755: (iii) The Elliott relationship, 
4756: Eq.~(\ref{eq:relax:elliott}), has been tested for many important metals and 
4757: found 
4758: to be valid over many orders of magnitude of $\Delta g$~\cite{Beuneu1978:PRB} 
4759: (this
4760: reference contains a useful collection of data for $\Delta g$). For the 
4761: majority of metals tested 
4762: (alkali and noble), a best fit gives the quantitative formula, 
4763: Eq.~(\ref{eq:revisedE}) 
4764: \cite{Beuneu1978:PRB}. 
4765: (iv) The Yafet relation, Eq.~(\ref{eq:relax:yafet}), is satisfied for 
4766: most metals 
4767: with the known 
4768: temperature dependence of $\tau_s$~\cite{Monod1979:PRB,Fabian1999:JVST}.
4769: The initially suggested deviation from the Yafet relation for several 
4770: polyvalent 
4771: metals 
4772: (Al, Pd, Be, and Mg) was later resolved by spin-hot-spot theory 
4773: \cite{Silsbee1983:PRB,Fabian1998:PRL,Fabian1999:JVST}, to be described
4774: below.
4775: This work showed that the magnitudes of the spin-mixing probabilities 
4776: $b^2$, taken from 
4777: atomic physics to test Eq.~(\ref{eq:relax:yafet}), should not be used in the
4778: solid-state environment. Various band-structure anomalies (spin hot spots), 
4779: such as crossing of the Brillouin-zone boundaries, 
4780: accidental degeneracy points, 
4781: or symmetry points
4782: on the Fermi surface, can increase the atomic-physics-derived $b^2$ by several 
4783: orders of magnitude, strongly enhancing spin relaxation in polyvalent metals 
4784: as compared to simple 
4785: estimates~\cite{Fabian1998:PRL,Fabian1999:JAP}.
4786: 
4787: (v) A realistic, first-principles calculation for Al \cite{Fabian1999:PRL} 
4788: (see Sec.~\ref{sec:IIIC}) using
4789: Eq.~(\ref{eq:relax:ey}) shows  
4790: excellent
4791: agreement with experiment. 
4792:  
4793: CESR measurements of $\tau_s$ in various metals are numerous.\footnote{A list
4794: of selected metal includes
4795: Li \cite{Feher1955:PR,Orchard-Webb1970:PSS,Damay1976:PRB};
4796: Na \cite{Feher1955:PR,Vescial1964:PR,Kolbe1971:PRB},
4797: K \cite{Walsh1966:PR};
4798: Rb \cite{Schultz1966:PRL,Walsh1966:PRL};
4799: Cs  \cite{Schultz1966:PRL,Walsh1966:PRL};
4800: Be \cite{Cousins1965:PL,Orchard-Webb1970:PSS}, Mg \cite{Bowring1971:PSS};
4801: Cu \cite{Schultz1965:PRL,Dunifer1976:PRB,Monod1982:JP};
4802: Au \cite{Monod1977:JLTP};
4803: Zn \cite{Stesmans1981:PRB}; 
4804: Al \cite{Lubzens1976:PRL}; 
4805: graphite \cite{Wagoner1960:PR,Matsubara1991:PRB};
4806: Rb$_{3}$C$_{60}$ \cite{Janossy1993:PRL}; MgB$_2$ \cite{simon2001:PRL}.
4807: Various data on CESR $\tau_s$ are collected in 
4808: \cite{Beuneu1978:PRB,Monod1979:PRB}.} 
4809: 
4810: The spin injection technique (see Sec.~\ref{sec:II}) was also used to measure 
4811: $\tau_s$ for various metals, 
4812: including Al 
4813: \cite{Johnson1985:PRL,Jedema2002:Na,Jedema2002:APL,Jedema2003:PRB}, 
4814: Au \cite{Elezzabi1996:PRL},
4815: Cu \cite{Jedema2001:N,Jedema2003:PRB}, and Nb \cite{Johnson1994:APL}. 
4816: In addition to CESR and spin injection, information about spin-orbit
4817: scattering 
4818: times 
4819: $\tau_{so}$ (see below)
4820: in various (but mostly noble) 
4821: metals at low temperatures has been also 
4822: obtained 
4823: from weak localization magnetoresistance measurements 
4824: on thin films~\cite{Bergmann1982:ZP} and 
4825: tunneling spectroscopy of metallic nanoparticles~\cite{Petta2001:PRL}.
4826: Surface-scattering spin relaxation times in normal metals and 
4827: superconductors 
4828: are
4829: collected by \textcite{Meservey1978:PRL}. Interesting results were
4830: obtained by injecting spin into superconductors. Using YBa$_2$Cu$_3$O$_{7-
4831: \delta}$, for
4832: example, data were interpreted \cite{Fu2002:PRB} to infer that the 
4833: in-plane spin relaxation time is unusually long, 
4834: about 100 $\mu$s at low temperatures to 1 $\mu$s close to the superconducting
4835: transition temperature. For quasiparticles moving along the c-axis, 
4836: $\tau_s$ is more likely to be the usual spin-orbit-induced spin 
4837: relaxation time, having the values of 10-100 ps.
4838: The microscopic origin of quasiparticle spin relaxation in cuprate 
4839: superconductors is not yet known. 
4840: 
4841: There is one more important time scale, the spin-orbit scattering
4842: time $\tau_{so}$, that is often invoked in mesoscopic transport
4843: as a characteristic of spin relaxation processes. We discuss it briefly
4844: in connection to $\tau_s$.
4845: The spin-orbit scattering time is the scattering time of
4846: Bloch electrons by the spin-orbit potential induced by impurities.
4847: The spin-orbit
4848: part of the Fourier transform of the impurity potential can be written 
4849: as $i c({\bf k}-{\bf k}') ({\bf k} \times {\bf k'})\cdot {\bf \sigma}$,
4850: where $c({\bf q})$ is proportional to the Fourier transform of the 
4851: impurity potential.
4852: The spin-orbit scattering time then is \cite{Werthamer:1969}
4853: \begin{equation}
4854: 1/\tau_{so}=\frac{2\pi}{\hbar} N_i \langle |c({\bf k}-{\bf k}')|^2 
4855: |{\bf k}\times {\bf k}' | \rangle^2
4856: {\cal N}(E_F),
4857: \end{equation}
4858: where $N_i$ is the impurity concentration, ${\cal N}(E_F)$   
4859: is the density of states per spin at
4860: the Fermi level, and the angle brackets denote Fermi-surface averaging. 
4861: As a parameter $\tau_{so}$ also includes the spin-orbit coupling of the host 
4862: lattice, 
4863: in the
4864: sense of the EY mechanism. Note, however, that $\tau_s \ne \tau_{so}$, even 
4865: at low
4866: temperatures where the impurity scattering dominates spin relaxation, 
4867: since the spin-orbit
4868: scattering includes both spin-flip and spin-conserving processes, which, 
4869: for
4870: isotropic scattering rates are in the ratio 2:1. In addition, 
4871: the spin relaxation rate is twice the spin-flip scattering rate, since each 
4872: spin flip
4873: equilibrates both spins equally.  For isotropic systems 
4874: $1/\tau_s\approx 4/(3\tau_{so})$.
4875: For a discussion of the effects of the DP processes on weak localization, 
4876: see \textcite{Knap1996:PRB}.}
4877: 
4878: 
4879: 
4880: 
4881: We illustrate spin relaxation in metals on the case of Al, whose $\tau_s$ was
4882: measured by CESR and spin injection, and numerically calculated
4883: from first principles. The case is instructive since it illustrates 
4884: both the general principles of the EY mechanism as well as the predicting 
4885: power 
4886: of, and the need for realistic band-structure calculations of $\tau_s$. 
4887: 
4888: Spin relaxation in Al was originally observed in CESR experiments, in which 
4889: $\tau_s$
4890: was measured at low temperatures, from 1 to 100 K \cite{Lubzens1976:PRL}.
4891: The spin relaxation rate $1/\tau_s$ was found to be independent of temperature 
4892: below 10-20 K; at higher temperatures $1/\tau_s$ increases linearly with 
4893: increasing $T$. 
4894: The same behavior was later observed in the original spin injection 
4895: experiment~\cite{Johnson1985:PRL,Johnson1988:PRBb}.
4896: Recently, $\tau_s$  was measured by spin injection at room temperature
4897: ~\cite{Jedema2002:Na,Jedema2002:APL,Jedema2003:PRB}. Unlike the CESR 
4898: and the original spin injection experiments, which were performed on 
4899: bulk samples, the room-temperature measurement
4900: used  thin Al films, observing strong spin relaxation due to surface
4901: scattering.
4902: 
4903: 
4904: 
4905: 
4906: Spin relaxation in Al, as well as in  other polyvalent metals, at first 
4907: appeared \cite{Monod1979:PRB}, in that a simple application
4908: of the Yafet relation, Eq.~(\ref{eq:relax:yafet}), yielded estimates for 
4909: $1/\tau_s$ 
4910: two orders of magnitude smaller than the observed data. Consider Na as a 
4911: reference.
4912: The atomic $\lambda_{so}/\Delta E$ [cf. Eq.~(\ref{eq:relax:lambda})] for 
4913: Na and Al are 
4914: within about 10\% of each other
4915: ~\cite{Beuneu1978:PRB,Monod1979:PRB}, yet the corresponding $\tau_s$ for Al is 
4916: about two 
4917: orders of magnitude smaller than that for 
4918: Na~\cite{Feher1955:PR,Vescial1964:PR}. 
4919: This anomaly extends to the $g$ factors as well.
4920: For Na, $\Delta g_{Na} \approx -8 \times 10^{-4}$ and 
4921: for
4922: Al it is six times greater, $\Delta g_{Al}\approx -5\times 10^{-3}$, while
4923: one would expect them to differ also by about 10\%. Note, however, that
4924: the Elliott relation, Eq.~(\ref{eq:relax:elliott}), is unaffected by this
4925: discrepancy, as it predicts that $\tau_s$(Na})/$\tau_s$(Al)$\approx 40$.  
4926: It was later suggested~\cite{Silsbee1983:PRB} that this is due to 
4927: accidental degeneracies in the two-band Fermi surface of Al. 
4928: 
4929: A full theoretical description, supported by first-principles
4930: calculations, of spin relaxation in Al and other polyvalent metals 
4931: led to the
4932: spin-hot-spots 
4933: theory~\cite{Fabian1998:PRL,Fabian1999:PRL,Fabian1999:JAP,Fabian1999:JVST}. 
4934: Spin hot spots are states on the Fermi surface that have anomalously 
4935: large spin mixing probabilities $|b|^2\approx 
4936: (\lambda_{so}/\Delta E)^2$, 
4937: arising from small energy gaps $\Delta E$. Quite generally, such
4938: states occur near Brillouin-zone boundaries and accidental degeneracy points,
4939: or at high-symmetry points. The condition for a spin hot spot is both a small 
4940: band gap
4941: $\Delta E$ and nonvanishing  
4942: $\lambda_{so}$.\footnote{At some symmetry points 
4943: $|b|$ may be very small. This occurs in the noble metals which 
4944: have Fermi states at 
4945: the Brillouin-zone boundaries, where $\Delta E$ is large,
4946: but the corresponding $\lambda_{so}$ is very small 
4947: due to symmetry.} 
4948: 
4949: If an electron hops in or out of a spin hot spot, the chance of spin flip 
4950: dramatically increases. Although the total area of spin hot spots on the 
4951: Fermi 
4952: surface
4953: is small, their contribution to $1/\tau_s$ is dominant, due to the large value
4954: of their $|b|^2$ in the Fermi surface average $\langle |b|^2 \rangle$, as
4955: was shown by analytical arguments~\cite{Fabian1998:PRL,Fabian1999:JAP}.
4956: A realistic numerical calculation~\cite{Fabian1999:PRL} for Al, 
4957: also showed that both
4958: the accidental degeneracies considered by~\cite{Silsbee1983:PRB} and states 
4959: close to 
4960: the Brillouin-zone
4961: boundaries dominate spin relaxation. 
4962: 
4963: A realistic calculation of $\tau_s$ in Al, based on pseudopotentials 
4964: and a realistic phonon description, has been performed by 
4965: \textcite{Fabian1999:PRL} and 
4966: compared to the experimental data available for $T<100$ 
4967: K~\cite{Johnson1985:PRL,Johnson1988:PRBb}. 
4968: Figure \ref{fig:T1_Al} shows both the experiment and the theory. 
4969: In the experimental data only the phonon contribution to $1/\tau_s$ is 
4970: retained
4971: \cite{Johnson1985:PRL}; the constant background impurity scattering is removed.
4972: The figure shows a rapid decrease of $\tau_s$ with increasing $T$ at low $T$, 
4973: where
4974: the agreement between experiment and theory is very good. 
4975: Above 200 K (the Debye temperature $T_D\approx 400$ K)
4976: the calculation predicts a linear dependence $\tau_s [{\rm ns}] 
4977: \approx 24\times
4978: T^{-1}[\rm K^{-1}]$.
4979: In the phonon-dominated linear regime the EY mechanism predicts that 
4980: the ratio $a^{ph}=\tau_p/\tau_s$ does not depend on $T$
4981: (see Sec.~\ref{sec:IIIB1}). 
4982: The calculated value is 
4983: $a^{ph}_{th}=1.2\times 10^{-4}$~\cite{Fabian1999:PRL}, 
4984: showing that $10^4$ phonon scatterings are needed to randomize electron spin.
4985: 
4986: 
4987: An important step towards extending spin injection capabilities was undertaken 
4988: recently
4989: by achieving spin injection into Cu and Al at room temperature
4990: ~\cite{Jedema2001:N,Jedema2002:Na,Jedema2002:APL,Jedema2003:PRB}; the 
4991: measured data 
4992: are unique
4993: in providing reliable values for spin diffusion lengths and spin relaxation 
4994: times in 
4995: these two important metals at room temperature. 
4996: The measured values for Al are 
4997: somewhat 
4998: sensitive to the experimental procedure/data analysis: $\tau_s= 85$ ps
4999: ~\cite{Jedema2002:Na} and $\tau_s=124$ ps ~\cite{Jedema2003:PRB}, as compared 
5000: to $\tau_s=90$ ps predicted by the theory at $T=293$ K. The room temperature 
5001: experimental
5002: data are included in Fig.~\ref{fig:T1_Al} for comparison. They nicely confirm 
5003: the 
5004: theoretical prediction. Less sensitive to data analysis is the ratio $a^{ph}$, 
5005: for
5006: which the experiments give $1.1\times 10^{-4}$~\cite{Jedema2002:Na} and 
5007: $1.3\times 10^{-4}$ \cite{Jedema2003:PRB}, comparing favorably with the
5008: theoretical  $a^{ph}_{th}=1.2\times 10^{-4}$.    
5009:  
5010: \begin{figure}
5011: \centerline{\psfig{file=zutic_fig16.eps,width=1\linewidth,angle=0}}
5012: \caption{Measured and calculated $\tau_s$ in Al. The low-$T$ measurements are 
5013: CESR \cite{Lubzens1976:PRL} and 
5014: spin injection \cite{Johnson1985:PRL}. 
5015: Only the phonon contribution is shown, as adapted from
5016: \textcite{Johnson1985:PRL}. 
5017: The solid line is the first-principles calculation, not a fit to the 
5018: data, \cite{Fabian1998:PRL}.
5019: The data at $T=293$ K are results from room-temperature spin injection 
5020: experiments of \textcite{Jedema2002:Na,Jedema2003:PRB}. 
5021: Adapted from \onlinecite{Fabian1999:PRL}.}
5022: \label{fig:T1_Al}
5023: \end{figure}
5024: 
5025: 
5026: Spin relaxation in Al depends rather strongly on magnetic fields at low $T$.
5027: CESR measurements ~\cite{Lubzens1976:PRL,Dunifer1976:PRB} 
5028: show that at temperatures below 100 K, $1/\tau_s$ increases linearly
5029: with increasing $B$. A specific sample~\cite{Lubzens1976:PRL}
5030: showed a decrease of $\tau_s$ from about 
5031: 20 ns to 1 ns, upon increase in $B$ from 0.05  to 1.4 T. 
5032: It was proposed that the observed behavior was due to cyclotron motion 
5033: through spin hot spots ~\cite{Silsbee1983:PRB}. The reasoning is as follows.
5034: Assume that there is considerable spread (anisotropy) 
5035: $\delta g \approx \Delta g$ 
5036: of the $g$ factors over the Fermi surface. Such a situation is common in
5037: polyvalent metals, whose spin hot spots have anomalously 
5038: large spin-orbit coupling.
5039: In a magnetic field the electron spins precess correspondingly with rates
5040: varying by $\delta \Omega_L \approx (\delta g/g) \Omega_L$, where 
5041: $\Omega_L$ is the Larmor frequency. 
5042: Motional narrowing leads to $1/\tau_s \approx (\delta\Omega_L)^2 \tau_c$,
5043: where $\tau_c$ is the correlation time for the random changes in $g$. 
5044: At small magnetic fields $\tau_c=\tau_p$ and $1/\tau_s \sim B^2 \tau_p$. 
5045: Such 
5046: a quadratic dependence of $1/\tau_s$ on $B$ is a  typical motional 
5047: narrowing case
5048: and has been observed at low temperatures in Cu \cite{Lubzens1976:PRL}.  
5049: As the field increases $\tau_c$ becomes the time of flight through 
5050: spin hot spots, in which case $\tau_c \sim 1/B$. As a result $1/\tau_s$ 
5051: acquires
5052: a component linear in $B$, in accord with experiment. 
5053: 
5054: 
5055: In an effort to directly detect phonon-induced spin flips in Al, an interesting 
5056: experiment was devised ~\cite{Lang1996:PRL,Grimaldi1996:PRL} 
5057: using the Zeeman splitting of the energy gap in Al superconducting tunnel 
5058: junctions. 
5059: Although the experiment failed, due to overwhelming spin-flip 
5060: boundary scattering,
5061: it showed the direction for future research in studying spin-flip 
5062: electron-phonon
5063: interactions.
5064: 
5065: \subsection{ \label{sec:IIID} Spin relaxation in semiconductors}
5066: %--------------------------------------------------------------------------
5067: 
5068: Although sorting out different spin-relaxation mechanisms of conduction 
5069: electrons in semiconductors
5070: is a difficult task, it has generally been 
5071: observed that the EY mechanism is relevant in small-gap and large-spin-orbit 
5072: coupling semiconductors, while the DP processes are responsible for spin 
5073: dephasing in middle-gap 
5074: materials and at high temperatures. In heavily p-doped samples the BAP  
5075: mechanism 
5076: dominates at lower temperatures, while DP at higher. In low-doped systems the
5077: DP dominates over the whole temperature range 
5078: where electron states are extended.  
5079: Spin relaxation of bound electrons proceeds 
5080: through the hyperfine interaction. Finally, spin relaxation of holes is due to 
5081: the EY processes.
5082: In bulk III-V or II-VI materials, for holes $\tau_s\approx \tau_p$, 
5083: since the valence spin and orbital states are completely
5084: mixed. However, in two-dimensional systems, where the heavy and light hole 
5085: states are split,
5086: hole spin relaxation is much less effective.
5087: 
5088: \subsubsection{\label{sec:IIID1} Bulk semiconductors}
5089: %--------------------------------------------------------------------------
5090: 
5091: There is a wealth of useful data on $\tau_s$ in 
5092: semiconductors.\footnote{References for selected semiconductors include
5093: the following:
5094: {p-GaAs}: 
5095: \cite{Fishman1977:PRB,Marushchak1984:SPSS,Seymour1981:PRB,Aronov1983:SPJETP,%
5096: Zerrouati1988:PRB,Sanada2002:APL};
5097: {n-GaAs}: See \ref{sec:IIID3};
5098: {p-Al$_{x}$Ga$_{1-x}$As}: \cite{Garbuzov1971:ZhETF,Clark1975:PRB};
5099: {p-GaSb}: \cite{Aronov1983:SPJETP,Safarov1980:JPSJ,Sakharov1981:SPSS};
5100: {n-GaSb}: \cite{Kauschke1987:PRB};
5101: {n-InSb}: \cite{Chazalviel1975:PRB};
5102: {InAs}: \cite{Boggess2000:APL};
5103: {p-InP}: \cite{Gorelenok1986:SPS};
5104: {n-InP}: \cite{Kauschke1987:PRB};
5105: {n-GaN}: \cite{Fanciulli1993:PRB,Beschoten2001:PRB}.}
5106: 
5107: 
5108: A comprehensive theoretical investigation of spin dephasing in bulk 
5109: semiconductors
5110: (both p- and n-types), applied to GaAs, GaSb, 
5111: InAs, and InSb, has been carried out by \textcite{Song2002:PRB} by using the 
5112: EY, DP, and BAP mechanisms.
5113: The calculation uses analytical formulas like 
5114: Eq.~(\ref{eq:relax:chazalviel}), 
5115: while
5116: explicitly evaluating $\tau_p$ for different momentum scattering processes 
5117: at different control parameters (temperature and density), but only 
5118: for nondegenerate 
5119: electron systems (Boltzmann statistics) and zero magnetic field. 
5120: The main results are as follows: in n-type III-V 
5121: semiconductors DP dominates at $T \agt 5$ K. At lower $T$ 
5122: the EY mechanism becomes relevant. This crossover temperature appears to be 
5123: quite 
5124: insensitive to the electron density, being between 1 and 5 K in most 
5125: investigated III-V 
5126: semiconductors for donor densities greater than $10^{14}$ cm$^{-3}$ 
5127: \cite{Song2002:PRB}. 
5128: For p-type materials the dominant mechanisms are DP and BAP, with the 
5129: crossover 
5130: temperature 
5131: sensitive to the acceptor density. For example, in p-GaAs at room temperature, 
5132: the DP mechanism dominates below $10^{18}$ cm$^{-3}$, while at 
5133: large densities 
5134: the BAP 
5135: mechanism dominates. In small-gap InSb the DP mechanism appears to be 
5136: dominant for all 
5137: acceptor densities at temperatures above 50 K. The strong disagreement with 
5138: experiment found at low T (at 5 K, to be specific) points to our still limited  
5139: theoretical understanding of spin relaxation in semiconductors. 
5140: The discrepancy likely arises 
5141: from neglect of
5142: HFI effects. 
5143: 
5144: Spin relaxation times as long as 300 ns were recently obtained in bulk, 
5145: $\approx 100$ nm wide, GaAs at low 4.2 K, placed in the proximity of QW's
5146: ~\cite{Dzhioev2001:JETPL,Dzhioev2002:PRL}. The samples were low doped 
5147: ($\approx 10^{14}$ cm$^{-3}$ uncompensated donor density), so that optical 
5148: orientation detected $\tau_s$ 
5149: of
5150: electrons bound on donors. At such low donor concentrations the HFI 
5151: mechanism is 
5152: responsible
5153: for spin relaxation. The unusually large $\tau_s$ is attributed to the 
5154: presence 
5155: of additional
5156: conduction electrons in the structure, coming from the barriers separating 
5157: the 
5158: sample and
5159: the nearest QW. The hyperfine interaction is then motionally narrowed by the 
5160: exchange 
5161: interaction
5162: between the donor-bound and conduction electrons. Upon depletion of the 
5163: conduction electrons
5164: from the sample by resonant excitations in the QW,  
5165: $\tau_s$ decreased to 5 ns~\cite{Dzhioev2002:PRL}, implying that the 
5166: effects of the
5167: static hyperfine fields on bound-electrons spin precessions are not 
5168: reduced by 
5169: motional narrowing. 
5170: 
5171: 
5172: Spin relaxation of holes in bulk III-V materials is very fast due to a
5173: complete mixing of orbital and spin degrees of freedom in the valence band. 
5174: The 
5175: EY mechanism predicts that hole $\tau_s$ is similar to hole $\tau_p$;
5176: this is a common assumption when considering hole contribution 
5177: to spin-polarized
5178: transport. Hole spin lifetime in undoped GaAs has been measured by optical 
5179: orientation and time-resolved spectroscopy \cite{Hilton2002:PRL}. 
5180: The observed value
5181: at room temperature is $\tau_s \approx 110$ fs, consistent with the 
5182: theoretical 
5183: assumption. 
5184: 
5185: Spin relaxation of conduction electrons in strained III-V crystals was studied 
5186: experimentally and theoretically by 
5187: \textcite{Dyakonov1986:SPJETP}. Spin relaxation under strain is enhanced 
5188: and becomes
5189: anisotropic due to the 
5190: strain-induced spin splitting of the conduction band, which is linear in $k$, 
5191: similarly to the bulk inversion asymmetry in two-dimensional systems 
5192: (see Sec.~\ref{sec:IIIB2}). 
5193: It was found that $1/\tau_s\sim\sigma^2$, 
5194: where $\sigma$ is the applied stress, and that $\tau_s$ is only weakly 
5195: temperature dependent. Spin relaxation of photoholes in strain crystals has 
5196: been 
5197: studied in 
5198: \cite{Dyakonov1974:SPS}, with the conclusion that the hole spin along 
5199: the strain axis can 
5200: relax (by the EY processes) on time scales much longer than in 
5201: unstrained samples, 
5202: due to 
5203: the lifting of heavy and light hole degeneracy. 
5204: 
5205: 
5206: Compared to III-V or II-VI, much less effort has been devoted to 
5207: investigation of $\tau_s$ in bulk Si. The reason is that CESR is thus far
5208: the only technique capable of effective detection of spin relaxation 
5209: in Si. Optical orientation is rather weak~\cite{Lampel1968:PRL} due to the 
5210: indirect band-gap structure, while robust 
5211: spin injection in Si is yet to be demonstrated. 
5212: Spin relaxation in Si is slow due to the presence of inversion symmetry 
5213: (the DP mechanism is not applicable) and lack of a nuclear moment for 
5214: the main Si isotope. Earlier experimental studies~\cite{Feher1959b:PR} were
5215: concerned with the hyperfine-interaction-dominated spin dephasing in donor 
5216: states.
5217: 
5218: 
5219: A comprehensive experimental study of low-doped Si (P donors were present 
5220: at the 
5221: levels
5222: $7.5\times 10^{14} \le N_d \le 8\times 10^{16}$ cm$^{-3}$), 
5223: at temperatures $20 < T < 300$ K,
5224: was performed by \textcite{Lepine1970:PRB}. Three distinct temperature regimes 
5225: were 
5226: observed:
5227: 
5228: (a) ($20 < T < 75$ K) Here $\tau_s$ decreases with increasing $T$.
5229: The HFI mechanism dominates: electrons are bound to the ground donor states, 
5230: while
5231: thermal excitations to higher states and the exchange interaction with 
5232: conduction
5233: electrons motionally narrows the hyperfine interaction. 
5234: 
5235: (b) ($75 < T < 150$ K) In this temperature range 
5236: $\tau_s$ continues to decrease with increasing $T$, the effect caused by the 
5237: spin-orbit interaction in the first excited donor state being motionally 
5238: narrowed by thermal motion. 
5239: 
5240: (c) ($T> 150$ K) Here
5241: $1/\tau_s$ increases with $T$, in accord with the EY mechanism. 
5242: The observed room-temperature CESR linewidth is
5243: about 8 G, corresponding to the electron spin
5244: lifetime of 7 ns.
5245: 
5246:  
5247: \subsubsection{\label{sec:IIID2} Low-dimensional semiconductor structures}
5248: %-------------------------------------------------------------------------
5249: 
5250: 
5251: The importance of low-dimensional semiconductor systems (quantum wells, wires, 
5252: and dots)
5253: lies in their great flexibility in manipulating charge and, now, also spin 
5254: properties
5255: of the electronic states. Studies of spin relaxation in those systems are driven
5256: not only by the need for 
5257: fundamental understanding of spin relaxation and decoherence, but also 
5258: by the goal of finding ways to reduce or otherwise control spin 
5259: relaxation and coherence in general. For a survey of spin relaxation 
5260: properties 
5261: of semiconductor 
5262: quantum wells, see \textcite{Sham1993:JPCM}.
5263: 
5264: 
5265: Spin relaxation in semiconductor heterostructures is caused by random magnetic
5266: fields originating either from the base material or from the heterostructure 
5267: itself. All four mechanisms of spin relaxation can be important, depending on 
5268: the material, doping, and geometry. The difference from the bulk is 
5269: the localization of the wave function into two, one, or zero dimensions
5270: and the appearance of structure-induced random magnetic fields. 
5271: Of all the mechanisms, the DP and HFI are believed to be most relevant. 
5272: 
5273: The most studied systems are GaAs/AlGaAs QW's. 
5274: The observed $\tau_s$ varies from nanoseconds
5275: to picoseconds, depending on the range 
5276: of control parameters such as temperature, QW width or confinement 
5277: energy $E_1$, carrier 
5278: concentration, mobility, magnetic field, or bias.\footnote{Here 
5279: is a list of selected references with useful data on $\tau_s$ in 
5280: GaAs/AlGaAs QW's:
5281: confinement energy dependence has been studied by 
5282: \textcite{Tackeuchi1996:APL,Britton1998:APL,Ohno1999:PRL,Malinowski2000:PRB,%
5283: Ohno2000:PE,Endo2000:JJAP};
5284: temperature dependence is treated by 
5285: \textcite{Wagner1993:PRB,Ohno1999:PRL,Malinowski2000:PRB,Ohno2000:PE,%
5286: Adachi2001:PE};
5287: carrier concentration dependence is studied by \textcite{Sandhu2001:PRL};
5288: dependence on mobility is examined by \textcite{Ohno1999:PRL}; 
5289: and dependence on magnetic field 
5290: in studied by \textcite{Zhitomirskii1993:JETPL}.}
5291: 
5292: Spin relaxation has also been investigated in In/GaAs 
5293: \cite{Paillard2001:PRL,Cortez2002:PE}, in an InAs/GaSb superlattice 
5294: \cite{Olesberg2001:PRB}, in InGaAs \cite{Guettler1998:PRB}, in 
5295: GaAsSb multiple QW's by \cite{Hall1999:APL}.
5296: II-VI QW's (specifically ZnCdSe) were
5297: studied by \textcite{Kikkawa1997:S}, finding $\tau_s\approx 1$ ns, 
5298: weakly dependent
5299: on both mobility and temperature (in the range $5 < T < 270$ K). Electron and 
5300: hole
5301: spin dephasing have also been investigated in dilute magnetic 
5302: semiconductor QW's doped
5303: with Mn ions \cite{Crooker1997:PRB,Camilleri2001:PRB}. 
5304: 
5305: 
5306: Reduction of spin relaxation by inhibiting the BAP electron-hole exchange 
5307: interaction through spatially
5308: separating the two carriers has been demonstrated in 
5309: $\delta$-doped p-GaAs:Be/AlGaAs 
5310: \cite{Wagner1993:PRB}. The observed $\tau_s$ was $\approx 20$ ns 
5311: at $T < 10$ K, 
5312: which is indeed unusually large.
5313: The exchange interaction was also studied at room temperature, observing
5314: an increase of $\tau_s$ with bias voltage which increases spatial separation 
5315: between electrons
5316: and holes, reducing the BAP effects \cite{Gotoh2000:JAP}. In the fractional 
5317: quantum Hall 
5318: effect regime it was demonstrated \cite{Kuzma1998:S} that nonequilibrium spin 
5319: polarization in 
5320: GaAs QW's can survive for tens of $\mu$s. Spin lifetime was 
5321: also found to be enhanced 
5322: in GaAs QW's
5323: strained by surface acoustic waves \cite{Sogawa2001:PRL}. A theoretical study 
5324: \cite{Kiselev2000:PRB} proposed that spin dephasing in 2DEG can be 
5325: significantly  suppressed by constraining the system to finite stripes, several 
5326: mean free paths wide.  
5327: 
5328: 
5329: Theoretical studies focusing 
5330: on spin dephasing in III-V and II-VI systems include those of 
5331: \cite{Wu2000:PRB,Wu2001:JS,Wu2002:SSC,Lau2001:PRB,Puller2002:P,Lau2002:JAP,%
5332: Bronold2002:PRB,Krishnamurthy2003:APL}.
5333: Spin relaxation due to the DP mechanism with bulk inversion asymmetry term 
5334: in the important case of GaAs/AlGaAs 
5335: rectangular QW's was investigated by Monte-Carlo simulations 
5336: \cite{Bournel2000:APL} at room 
5337: temperature, including interface roughness scattering. Nice agreement with 
5338: experiment was
5339: found for $\tau_s(E_1)$, where $E_1$ is the confinement energy. 
5340: Interface roughness becomes important at large
5341: values of $E_1$, where scattering increases $\tau_s$ (see also 
5342: \textcite{Sherman2003:APL}). 
5343: 
5344: Spin relaxation and spin coherence of spin-polarized photoexcited 
5345: electrons and holes in symmetric p- and n-doped 
5346: and undoped GaAs/AlGaAs quantum 
5347: wells was investigated using rate equations 
5348: \cite{Uenoyama1990:PRL,Uenoyama1990:PRB}. 
5349: It was shown that in these heterostructures hole spin relaxation proceeds 
5350: slower 
5351: than electron-hole recombination. Hole relaxation is found to occur mostly 
5352: due to 
5353: acoustic phonon emission. The ratio of the spin-conserving to spin-flip hole 
5354: relaxation
5355: times was found to be 0.46, consistent with the fact that 
5356: luminescence is polarized even in n-doped quantum wells at times greater 
5357: than the
5358: momentum relaxation time. Similar observations hold for  strained 
5359: bulk GaAs, where hole spin relaxation is also reduced. Spin relaxation of 
5360: holes in quantum wells 
5361: was calculated \cite{Ferreira1991:PRB,Bastard1992:SS} using the interaction 
5362: with ionized
5363: impurities and s-d exchange in semimagnetic semiconductors. It was shown that 
5364: size quantization 
5365: significantly reduces 
5366: spin relaxation of holes, due to the 
5367: lifting of
5368: heavy and light hole degeneracy. The observed spin lifetimes for holes at 
5369: low 
5370: temperatures
5371: reached up to 1 ns, while at $T > 50$ K in the same samples $\tau_s$ got
5372: smaller than
5373: 5 ps \cite{Baylac1995:SSS}. 
5374: 
5375: Spin dynamics and spin relaxation of excitons in GaAs 
5376: \cite{Munoz1995:PRB,Vina2001:PB} and ZnSe \cite{Kalt2000:PSS} were
5377: investigated 
5378: experimentally and theoretically \cite{Maialle1993:PRB,Sham1988:PE}.
5379: Coherent spin dynamics in magnetic semiconductors was considered by 
5380: \textcite{Linder1998:PE}.
5381: 
5382: Spin relaxation in Si heterostructures has been investigated by electron 
5383: spin resonance in modulation doped Si/SiGe QW's.  
5384: Very high mobility (about $10^5$ cm$^2\cdot$V$^{-1}\cdot$s$^{-1}$) samples with 
5385: $n\approx 3\times
5386: 10^{11}$ cm$^{-2}$ free electrons forming a 2DEG show, 
5387: at $T=4.2$ K, 
5388: $T_1$ up to 30 $\mu$s \cite{Sandersfeld2000:TSF,Graeff1999:PRB} 
5389: and $T_2$ of the order of 100 ns~\cite{Graeff1999:PRB},
5390: depending on the orientation of $B$ with respect to the QW growth direction. 
5391: Spin relaxation
5392: was attributed to Bychkov-Rashba spin splitting in these asymmetric wells, 
5393: estimating
5394: the corresponding $\hbar\alpha_{BR}$ 
5395: in Eq.~(\ref{eq:relax:BR}) to be around $1\times 
5396: 10^{-14}$ eV$\cdot$m~\cite{Wilamowski2002:PE,Wilamowski2002:PRB}. 
5397: Si/Ge heterostructures may have enhanced 
5398: rates of spin relaxation due to 
5399: the leakage of the electron wave function to Ge, 
5400: which is heavier than Si and has 
5401: greater spin-orbit interaction. 
5402: Recent studies \cite{Wilamowski2004:PRB} have confirmed the dominant role
5403: of the DP spin relaxation mechanism, leading to the microsecond spin 
5404: relaxation times. The spin dephasing is argued to be strongly suppressed
5405: by cyclotron motion in high-mobility samples (see Sec.~\ref{sec:IIIB2a}
5406: for a brief discussion of the influence of magnetic field on $\tau_s$).
5407: Spin-orbit coupling in symmetric Si/SiGe quantum wells has been studied 
5408: theoretically by \textcite{Sherman2003:PRB}.
5409: 
5410: In quantum dots the relevant spin relaxation mechanism is still being 
5411: debated,
5412: as the mechanisms 
5413: (EY and DP) effective for conduction electrons are ineffective for states
5414: localized in QD's~\cite{Khaetskii2000:PRB,Khaetskii2001:PRB,Khaetskii2001:PE}.
5415: It is believed, however, that similar to electrons bound on donors, the dominant
5416: mechanism is a HFI process
5417: \cite{Merkulov2002:PRB,Semenov2002:P,Sousa2003:PRB,Sousa2002:P,%
5418: Khaetskii2002:PRL}.
5419: Unfortunately, experiments on CdSe QD's (of diameter 22-80 \AA) show strong 
5420: inhomogeneous dephasing
5421: ($\tau_s\approx 3$ ns at $B=0$, while $\tau_s \approx 100$ ps at 4 T) 
5422: \cite{Gupta1999:PRB}, masking the intrinsic spin dephasing processes.
5423: Recently a lower bound, limited by the signal-to-noise ratio, on $T_1$
5424: of 50 $\mu$s has been measured at 20 mK in a one-electron quantum dot
5425: defined in 2DEG GaAs/AlGaAs heterostructure by 
5426: \textcite{Hanson2003:lanl}. The magnetic field of 7.5 T was oriented 
5427: parallel to the plane heterostructure. While the actual value of
5428: $T_1$ may be orders of magnitude larger, the observed bound suffices
5429: for performing elementary quantum gates (see Sec.~\ref{sec:IVF}).
5430: 
5431: \subsubsection{\label{sec:IIID3} Example: spin relaxation in GaAs}
5432: %---------------------------------------------------------------------
5433: 
5434: We review recent experimental results on spin relaxation in bulk n-GaAs% 
5435: \footnote{p-GaAs is extensively discussed by \textcite{Meier:1984}}
5436: and GaAs-based low-dimensional systems.
5437: 
5438: 
5439: 
5440: \paragraph{\label{sec:IIID3a} Bulk n-GaAs.}
5441: %---------------------------------------------------------------------------
5442: 
5443: The importance of GaAs for spintronics and quantum computing applications 
5444: has been 
5445: recently
5446: underlined by the discovery of rather long spin relaxation times (of the order 
5447: of 100 ns)
5448: in n-doped samples, as well as by the development of experimental 
5449: techniques to 
5450: manipulate spin precession in this semiconductor in a coherent 
5451: manner~\cite{Awschalom2001:PE,Oestreich2002:SST}.
5452: 
5453: 
5454: Both optical orientation and time-resolved Faraday rotation spectroscopy have 
5455: been used to measure $\tau_s$ in bulk n-GaAs. 
5456: In the earliest observations of optical spin orientation of
5457: electrons, in 
5458: n-Ga$_{0.7}$Al$_{0.3}$As with $N_d\approx 1\times 10^{16}$ cm$^{-3}$ at 4.2 K,
5459: it was found that $\tau_s\approx 1.2 $ ns~\cite{Ekimov1971:JETPL}.
5460: A much larger spin lifetime was found by optical orientation on n-GaAs 
5461: ~\cite{Dzhioev1997:PSS}, 
5462: where for $N_d\approx 1\times 10^{15}$ cm$^{-3}$ the observed
5463: $\tau_s\approx 42$ ns. 
5464: Faraday rotation studies~\cite{Kikkawa1998:PRL,Awschalom2001:PE} 
5465: found even longer  spin lifetimes. At the doping density 
5466: $N_d=1\times 10^{16}$ cm$^{-3}$ of Si donors and  $T=5$ K, the observed 
5467: $\tau_s\approx 130$ ns at zero magnetic field. 
5468: At greater and smaller doping densities, spin relaxation time is significantly 
5469: reduced: for both a nominally undoped sample and for $N_d=1\times 10^{18}$ 
5470: cm$^{-3}$, $\tau_s\approx 0.2$ ns. A comprehensive theoretical investigation
5471: of $\tau_s$ in bulk n-GaAs is reported by \textcite{Wu2000:PSS}, and
5472: \textcite{Wu2001:JS}, who solved numerically kinetic equations in the
5473: presence of magnetic filed. Only the DP mechanism was considered, 
5474: acting with longitudinal phonon and impurity scattering.
5475: 
5476: A recent comprehensive study of $\tau_s$ based on optical orientation 
5477: revealed a nice, albeit 
5478: complex, picture of spin relaxation in bulk n-GaAs over a large range of doping 
5479: levels~\cite{Dzhioev2002:PRB}. Figure \ref{fig:Dzhioev} summarizes these 
5480: findings. 
5481: The spin relaxation time rises with increasing
5482: $N_d$ at small doping levels, reaching its first maximum (180 ns) at around 
5483: $3\times 
5484: 10^{15}$ cm$^{-3}$; $\tau_s$ then decreases until $N_d=N_{dc}=2\times 10^{16}$ 
5485: m$^{-3}$,
5486: where a sudden increase brings $\tau_s$ to another maximum, reaching $\approx 
5487: 150$ ns. 
5488: At still higher doping levels $\tau_s$ decreases strongly with increasing 
5489: doping.
5490: 
5491: \begin{figure}
5492: \centerline{\psfig{file=zutic_fig17.eps,width=1\linewidth,angle=0}}
5493: \caption{Spin relaxation time in n-GaAs as a function of donor density $N_d$ 
5494: (labeled as $n_D$ here) at
5495: low temperatures:  empty symbols, the optical orientation data of 
5496: \textcite{Dzhioev2002a:PRB}; solid circles are the time-resolved 
5497: Faraday 
5498: rotation data
5499: of  \textcite{Kikkawa1998:PRL,Awschalom2001:PE};
5500: open triangles, single-spin decoherence times 
5501: $\tau_{sc}\approx \tau_c$ due to the exchange interaction 
5502: between electron spins on 
5503: neighboring donors;  solid lines, parameter-free theoretical estimates 
5504: with
5505: labels indicating the dominant spin relaxation mechanisms;
5506: dotted line a 
5507: fit to the experimental data on the exchange correlation time 
5508: (triangles) $\tau_c$, 
5509: using a simple model of the exchange coupling between donor states; 
5510: dashed vertical 
5511: line, the metal-insulator transition at 
5512: $N_{dc}=2\times 10^{16}$ cm$^{-3}$. 
5513: From \onlinecite{Dzhioev2002a:PRB}. 
5514: }
5515: \label{fig:Dzhioev}
5516: \end{figure}
5517: 
5518: The above picture is valid at $T \le 5$ K, where isolated shallow donors 
5519: are not normally ionized, and the sample is a Mott insulator at small dopings. 
5520: Conductivity
5521: is due to hopping between donor states. Beyond the critical density 
5522: $N_{dc}\approx 
5523: 2\times 10^{16}$ cm$^{-3}$ (the dashed vertical line in Fig.~\ref{fig:Dzhioev})
5524: the donor states start to overlap and form an impurity 
5525: conduction band---electronic
5526: states delocalize and the sample becomes metallic. Figure~\ref{fig:Dzhioev} 
5527: shows that
5528: it is the rather narrow window around the metal-to-insulator transition where 
5529: the largest $\tau_s$ are found.  
5530: 
5531: \begin{figure}
5532: \centerline{\psfig{file=zutic_fig18.eps,width=1\linewidth,angle=0}}
5533: \caption{Measured magnetic-field dependence of the spin dephasing time 
5534: (here denoted as $T_2^*$ to indicate the likely presence of inhomogeneous 
5535: broadening; 
5536: see Sec.~\ref{sec:IIIA1}) 
5537: for bulk n-GaAs
5538: at 5 K. Doping levels, varying from insulating 
5539: ($N_d < N_{dc}=
5540: 2\times 10^{16}$ cm$^{-3}$) to metallic ($N_d > N_{dc}$), are
5541: indicated. Adapted from \onlinecite{Kikkawa1998:PRL}.} 
5542: \label{fig:KA1}
5543: \end{figure}
5544: 
5545: 
5546: At $N_d > N_{dc}$ the DP mechanism dominates.  Equation
5547: (\ref{eq:relax:DP}) for degenerate electrons explains the observed data
5548: rather well.  Indeed, considering that $E_F\sim N_d^{2/3}$ and
5549: assuming the Brooks-Herring formula for the impurity scattering
5550: $(1/\tau_p \sim N_d/E_F^{3/2}$), one obtains $\tau_s\sim 1/N_d^2$,
5551: which is observed in Fig.~\ref{fig:Dzhioev}. The EY mechanism, 
5552: Eq.~(\ref{eq:relax:chazalviel}), would give $\tau_s\sim N_d^{-4/3}$.  The
5553: data on the insulating side are consistent with the HFI mechanism: the
5554: precession due to local random magnetic fields from the nuclear moments is
5555: motionally narrowed by the exchange interaction, which increases with
5556: increasing $N_d$ (that is, with increasing overlap between donor
5557: states). The theoretical estimates~\cite{Dzhioev2002:PRB} agree well
5558: with the data.  The behavior of $\tau_s$ in the intermediate regime, 
5559: $3\times 10^{15}$ cm$^{-3} < N_d < N_{dc}$, where $\tau_s$ decreases
5560: with increasing $N_d$, was proposed by~\cite{Kavokin2001:PRB} to be
5561: due to motional narrowing of the antisymmetric exchange interaction
5562: \footnote{The anisotropic exchange interaction of the
5563: Dzyaloshinskii-Moriya form, $\bf{S}_1\times {\bf S}_2$ 
5564: \cite{Dzyaloshinskii1958:PCS,Moriya1960:PR}
5565: appears as a result of spin-orbit coupling in semiconductors 
5566: lacking inversion symmetry.} 
5567: between bound electrons, with the correlation time $\tau_c$
5568: provided by the usual ${\bf S}_1\cdot{\bf S}_2$, direct exchange. This new
5569: mechanism of spin relaxation, which should be generally present for
5570: bound electrons in systems lacking inversion symmetry (such as III-V
5571: and II-VI), although still being
5572: investigated~\cite{Gorkov2003:PRB,Kavokin2002:P}, appears to give
5573: a satisfactory explanation for the experimental data.
5574: 
5575: In addition to the doping dependence of $\tau_s$, both the temperature and
5576: the magnetic field dependences of spin relaxation in bulk n-GaAs have been
5577: studied by \textcite{Kikkawa1998:PRL}.  Figure \ref{fig:KA1} shows
5578: $\tau_s(B)$ for samples with varying doping levels at $T=5$ K. The spin
5579: relaxation time increases with $B$ in the metallic regime, the
5580: behavior qualitatively consistent with the predictions of the DP
5581: mechanism.  In contrast, $\tau_s$ in the insulating samples decreases
5582: with increasing $B$.  Bound electrons are more susceptible to $g$-factor
5583: anisotropies (due to the distribution of electron energies over donor
5584: states) and local magnetic field variations (due to the hyperfine
5585: interactions).  These anisotropies are amplified by increasing $B$
5586: and motionally narrowed by the exchange interaction. It is thus
5587: likely that $\tau_s\sim B^{-2}\tau_c(B)$, where the exchange
5588: correlation time $\tau_c$ depends on $B$ through magnetic orbital
5589: effects on the bound electron wave functions (magnetic confinement
5590: reduces the extent of the bound orbital, thus reducing the exchange
5591: integrals between neighboring donor states). However, no satisfactory
5592: quantitative explanation for $\tau_s(B)$ in insulating samples
5593: exists.
5594: 
5595: \begin{figure}
5596: \centerline{\psfig{file=zutic_fig19.eps,width=1\linewidth,angle=0}}
5597: \caption{Measured temperature  dependence of the spin dephasing time 
5598: for bulk n-GaAs doped with $N_d=1\times 10^{16}$ cm$^{-3}$ Si donors, at 
5599: $B=0$ and $B=4$ T. 
5600: The sample is insulating at low
5601: $T$ and nondegenerate at high $T$ ($T\agt 50$ K, assuming $\approx 4$ meV for
5602: the donor binding energy), where donors are ionized.
5603: Adapted from \onlinecite{Kikkawa1998:PRL}. 
5604: }
5605: \label{fig:KA2}
5606: \end{figure}
5607: 
5608: 
5609: Figure \ref{fig:KA2} plots $\tau_s(T)$ for an insulating sample with
5610: $N_d=1\times 10^{16}$ cm$^{-3}$ at $B=0$ and $B=4$ T.   
5611: For the
5612: zero-field data the initial decrease of $\tau_s$ with $B$ is very
5613: rapid, dropping from 130 ns at 5 K to less than 1 ns at 150 K. However, 
5614: the sample held at $B=4$ T 
5615: shows at first a rapid
5616: increase with increasing $T$, and then a decrease at $T\approx 50$ K.
5617: The decrease of $\tau_s$ with increasing $T$ above 50 K has been found
5618: to be consistent with the DP mechanism~\cite{Kikkawa1998:PRL}, taking
5619: $\tau_s\sim T^{-3}$ in Eq.~(\ref{eq:relax:DPt}), while extracting
5620: the temperature dependence of $\tau_p$ from the measurement of mobility. 
5621: The DP
5622: mechanism for conduction electrons was also observed in p-GaAs in the
5623: regime of nondegenerate hole densities $N_a\approx 10^{17}$ cm$^{-3}$
5624: at temperatures above  100 K~\cite{Aronov1983:SPJETP},
5625: after the contribution from the BAP mechanism was subtracted using a
5626: theoretical prediction. From the observed mobility
5627: it was found that $\tau_p(T)\sim T^{-0.8}$, so that 
5628: according to the
5629: DP mechanism $\tau_s\sim
5630: T^{-2.2}$, which is indeed consistent with the experimental data.  The
5631: origin of $\tau_s(T)$ below 50 K in Fig.~\ref{fig:KA2} is less
5632: obvious.  At low $T$, electrons are localized, so in order to explain
5633: the experimental data the theory should include ionization of donors.
5634: The increase with increasing $T$ of $\tau_s$ at 4 T  
5635: invokes a
5636: picture of motional narrowing in which the correlation time decreases
5637: with increasing $T$ much faster than the dispersion of local Larmor
5638: frequencies. We do not know of a satisfactory quantitative explanation
5639: for these experimental results.\footnote{There is a 
5640: discrepancy in the data presented in Figs. \ref{fig:KA1}
5641: and \ref{fig:KA2}. Take the $N_d=1\times 10^{16}$ cm$^{-3}$ sample.
5642: While Fig.~\ref{fig:KA1} reports $\tau_s\approx 3$ ns at 5 K and 4 T, 
5643: $\tau_s$ is only about 1 ns in Fig.~\ref{fig:KA2}.
5644: The reason for this difference \cite{Kikkawa2003:PC}
5645: turns out to be electronically-induced nuclear polarization 
5646: \cite{Kikkawa2000:S}.
5647: At low temperatures and large magnetic fields, nuclear polarization
5648: develops via the Overhauser effect inhomogeneously throughout the
5649: electron spin excitation region. The inhomogeneous magnetic
5650: field due to polarized nuclei causes inhomogeneous broadening
5651: of the electronic $\tau_s$. The measured spin dephasing time is indeed
5652: $T_2^*$, rather than the intrinsic $T_2$. Furthermore, since nuclear
5653: polarization typically takes minutes to develop, the measured
5654: $T_2^*$ depends on the measurement ``history.'' This is the reason
5655: why two different measurements, reported in Figs. \ref{fig:KA1}
5656: and \ref{fig:KA2}, show different $T_2^*$ under otherwise equivalent
5657: conditions. The nuclear polarization effect is also part of the
5658: reason why the $T_2(T)$ at 4 T 
5659: sharply deviates from that
5660: at zero field at small $T$. The technique should give consistent
5661: results at small fields and large temperatures, as well as 
5662: in heavily doped samples where the nuclear fields are motionally
5663: narrowed by the itinerant nature of electrons.
5664: }
5665: Similar behavior of
5666: $\tau_s(T)$ 
5667: in insulating samples was found in GaN~\cite{Beschoten2001:PRB}.
5668: 
5669: 
5670: The temperature dependence of $\tau_s$ for samples with 
5671: $N_d \gg N_{dc}$ has been reported \cite{Kikkawa1998:PRL} to be very weak, 
5672: indicating, for
5673: these degenerate electron densities, that $\tau_p(T)$ is only  weakly
5674: dependent on $T$. What can be expected for $\tau_s$ at room temperature?  The
5675: answer will certainly depend on $N_d$. Recent experiments on time-
5676: resolved Kerr rotation~\cite{Kimel2001:PRB} suggest that 5 ps$ < \tau_s <
5677: 10 $ ps for undoped GaAs and 15 ps $< \tau_s < 35$ ps for a heavily doped
5678: n-GaAs with $N_d=2\times 10^{18}$ cm$^{-3}$.
5679: 
5680: For spintronic applications to make use of the large $\tau_s$
5681: observed in bulk n-GaAs one is limited to both very small temperatures
5682: and small doping levels. Although this may restrict the design of 
5683: room-temperature spintronic devices, such a regime seems acceptable for
5684: spin-based quantum computing (see Sec.~\ref{sec:IVF}), 
5685: where one is interested in the spin lifetime
5686: of single (or a few) electrons, bound to impurities or confined to
5687: quantum dots.  How close is $\tau_s$ to the 
5688: individual spin lifetime $\tau_{sc}$?
5689: There is no clear answer yet.  Ensemble spin dephasing seen for
5690: insulating GaAs samples appears to be due to motional narrowing of the
5691: hyperfine interaction. The randomizing processes are spin flips due to
5692: the direct exchange, leading to the correlation time $\tau_c$, which can
5693: be taken as a measure for the lifetime $\tau_{sc}$ of the individual spins. 
5694: Extracting
5695: this lifetime from the experiment is not easy, but the obvious trend
5696: is the smaller the $\tau_c$, the larger the $\tau_s$.  For a specific
5697: model of spin relaxation in bound electron states $\tau_c$ was
5698: extracted experimentally by \textcite{Dzhioev2002:PRB} by detecting
5699: the changes in the spin polarization due to longitudinal magnetic fields.  The
5700: result is shown in Fig.~\ref{fig:Dzhioev}.  The two times, $\tau_c$
5701: and $\tau_s$ differ by orders of magnitude. For the doping levels where
5702: $\tau_s$ is {\it greater} than 100 ns, $\tau_c$ is {\it smaller} than   
5703: 0.1 ns.
5704: Unfortunately, the useful time
5705: for spin quantum computing would be extracted in the limit of very
5706: small dopings, where the data are still sparse. For an informal recent review
5707: of $\tau_s$ in n-GaAs, see \textcite{Kavokin2002:PSS}.
5708: 
5709: Closely related to spin relaxation is spin diffusion. 
5710: \textcite{Hagele1998:APL} observed the transport of a spin 
5711: population---longitudinal spin drift---in i-GaAs over a length
5712: scale greater than 4 $\mu$m in electric
5713: fields up to 6 kV/cm and at low temperatures. 
5714: This was followed by a remarkable result of \textcite{Kikkawa1999:N},
5715: the observation of the drift of precessing electron 
5716: spins---transverse spin drift---in GaAs with $N_d=1\times 10^{16}$ cm$^{-3}$, 
5717: over 100 $\mu$m in moderate electric fields (tens of V/cm) at $T=1.6$ K,
5718: setting the length scale for the spin dephasing.
5719: By directly analyzing the spreading and drifting of the electron
5720: spin packets in time, 
5721: Kikkawa and Awschalom obtained
5722: the spin diffusion 
5723: (responsible for spreading) and electronic
5724: diffusion
5725: (drift by electric field) 
5726: coefficients.
5727: It was found that
5728: the former is about 20 times as large as the latter. These results are
5729: difficult to interpret, since the sample is just below the metal-to-insulator
5730: transition, where charge is transported via hopping, but they suggest that 
5731: spin diffusion is strongly
5732: enhanced through the exchange interaction. Investigations of this type
5733: in even smaller doping limits may prove important for understanding
5734: single-spin coherence.
5735: 
5736: \paragraph{GaAs-based quantum wells.}
5737: %-----------------------------------------------------------------
5738: 
5739: We discuss selected experimental results on spin relaxation in 
5740: GaAs/Al$_x$Ga$_{1-x}$As QW's,
5741: presenting  the temperature and confinement energy dependence of $\tau_s$.
5742: 
5743: Figure \ref{fig:QW1} plots the temperature dependence of $1/\tau_s$ in
5744: the interval of $90 < T < 300$ K for QW's of width $L$ ranging from 6
5745: to 20 nm \cite{Malinowski2000:PRB}.  The wells, with $x=0.35$ and
5746: orientation along [001], were grown on a single wafer to minimize
5747: sample-to-sample variations when comparing different wells. The
5748: reported interface roughness was less than the exciton Bohr radius of
5749: 13 nm.  In these structures the excitonic effects dominate at $T < 50$ K
5750: (with the reported $\tau_s\approx 50$ ps), while the exciton ionization is
5751: complete roughly at $T > 90$ K, so the data presented are for free
5752: electrons. Spin relaxation was studied  using pump-probe optical
5753: orientation spectroscopy with a 2 ps time resolution and the typical
5754: excitation intensity/pulse of $10^{10}$
5755: cm$^{-2}$.
5756: 
5757: \begin{figure}
5758: \centerline{\psfig{file=zutic_fig20.eps,width=1\linewidth,angle=0}}
5759: \caption{Measured temperature dependence of the conduction-electron-spin 
5760: relaxation rate 
5761: $1/\tau_s$
5762: in GaAs/AlGaAs QW's of varying widths: the dashed curve,  data for
5763: a low-doped ($N_a=4\times 10^{16}$ cm$^{-3}$) 
5764: bulk p-GaAs \cite{Meier:1984}; solid line, the 
5765: $\tau_s \sim T^2$ dependence. From \onlinecite{Malinowski2000:PRB}.} 
5766: \label{fig:QW1}
5767: \end{figure}
5768: 
5769: As Fig.~\ref{fig:QW1} shows, $\tau_s$ depends rather weakly on $T$ for
5770: the narrow wells with $L < 10$ nm.
5771: For the well with $L=15$ nm, after
5772: being approximately constant (or somewhat decreasing) as $T$ increases
5773: to about 200 K, $\tau_s$ 
5774: increases with increasing $T$ at greater temperatures.
5775: The increase is consistent with the $1/\tau_s \sim T^2$ behavior.  The
5776: thickest well increases with the same power law, $1/\tau_s\sim T^2$, over 
5777: the whole temperature range.  In order to make a reliable comparison
5778: with theoretical predictions (the expected mechanism is that of DP in
5779: two-dimensional systems), one needs to know the behavior of $\tau_p(T)$.
5780: The DP mechanism predicts, for the nondegenerate electron densities
5781: employed in the experiment, that $1/\tau_s \sim T^3 \tau_p$ [see 
5782: Eq.~(\ref{eq:relax:DPt})] in the bulk and wide QD's, the condition being that
5783: thermal energy is greater than the subband separation), and $1/\tau_s
5784: \sim T E_1^2 \tau_p$ from Eq.~(\ref{eq:relax:DK}) for the
5785: bulk inversion asymmetry
5786: after making thermal averaging ($E_{\bf k} \rightarrow k_B T$),
5787: when one realizes
5788: that confinement energy $E_1$ is $\sim \langle k_n^2 \rangle$.
5789: When one assumes that momentum relaxation in these elevated temperatures
5790: is due to scattering by phonons, $\tau_p$ should be similar in bulk
5791: and low-dimensional structures. From the observed high-temperature
5792: bulk $\tau_s(T)$ (at low temperatures $\tau_s$ is affected by BAP
5793: processes) one can estimate $\tau_p \sim 1/T$, which is consistent
5794: with the constant $\tau_s$ for the narrow wells, 
5795: and with the quadratic dependence for the wide wells.
5796: At low $T$, in addition
5797: to the BAP mechanism, $\tau_s$ will deviate from that in the bulk due
5798: to impurity scattering. The EY and BAP mechanisms were found not
5799: to be relevant to the observed data \cite{Malinowski2000:PRB}.
5800: 
5801: \begin{figure}
5802: \centerline{\psfig{file=zutic_fig21.eps,width=1\linewidth,angle=0}}
5803: \caption{Measured room-temperature dependence of $1/\tau_s$ on the
5804: confinement
5805: energy $E_1$ for GaAs/AlGaAs QW's. The solid line is a quadratic fit, 
5806: showing 
5807: behavior
5808: consistent with the DP mechanism. From \onlinecite{Malinowski2000:PRB}.} 
5809: \label{fig:QW2}
5810: \end{figure}
5811: 
5812: Figure \ref{fig:QW2} shows the dependence of $1/\tau_s$ on the
5813: experimentally determined confinement energy $E_1$ for a variety of
5814: QW's on the same wafer \cite{Malinowski2000:PRB}.  The data are 
5815: at room
5816: temperature. The spin relaxation time 
5817: varies from somewhat less than
5818: 100 ps for wide QW's, approximating the bulk data (cf.
5819: \textcite{Kimel2001:PRB} where 15 ps $< \tau_s < 35$ ps 
5820: was found for a heavily
5821: doped n-GaAs), to about 10 ps in most confined structures. The
5822: downturn for the highest-$E_1$ well (of width 3 nm) is most likely due
5823: to the increased importance of interface roughness at such small
5824: widths \cite{Malinowski2000:PRB}.  Confinement strongly enhances spin
5825: relaxation. This is consistent with the DP mechanism for
5826: two-dimensional systems, in which the spin precession about the intrinsic
5827: magnetic fields 
5828: (here induced by bulk inversion asymmetry) 
5829: increases as 
5830: $E_1^2$ with increasing
5831: confinement. The observed data in Fig.~\ref{fig:QW2} are consistent
5832: with the theoretical prediction.
5833: 
5834: Similarly to bulk GaAs, spin relaxation in GaAs QW's was found to be reduced at 
5835: carrier
5836: concentrations close to the metal-to-insulator transition
5837: ($n\approx 5\times 10^{10}$ cm$^{-2}$) \cite{Sandhu2001:PRL}.
5838: 
5839: 
5840: \section{\label{sec:IV} Spintronic devices and applications}
5841: %=========================================================================
5842: 
5843: In this section we focus primarily on the physical principles 
5844: and materials issues
5845: for various device schemes, which, while not yet commercially viable, are likely
5846: to influence future spintronic research and possible applications.
5847: 
5848: 
5849: \subsection{\label{sec:IVA} Spin-polarized transport}
5850: %--------------------------------------------------------------------------
5851: 
5852: \subsubsection{\label{sec:IVA1} F/I/S tunneling}
5853: %--------------------------------------------------------------------------
5854: 
5855: 
5856: Experiments reviewed by \textcite{Tedrow1994:PR} in
5857: ferromagnet/insulator/superconductor (F/I/S) junctions have
5858: established a sensitive technique for measuring the spin polarization $P$
5859: of magnetic thin films and, at the same time, 
5860: has demonstrated that the
5861: current will remain spin-polarized after tunneling through an
5862: insulator. 
5863: These experiments also stimulated more recent
5864: imaging techniques based on the spin-polarized STM  
5865: (see \textcite{Johnson1990:JAP,Wiesendanger1990:PRL}; and a review,
5866: \textcite{Wiesendanger:1998}) 
5867: with the ultimate goal of imaging spin
5868: configurations down to the atomic level.
5869: 
5870: The degree of spin polarization 
5871: is important for many 
5872: applications such as determining the magnitude of tunneling magnetoresistance
5873: (TMR) in magnetic tunnel junctions (MTJ) [recall Eq.~(\ref{eq:julliere})].
5874: Different probes for spin polarization generally
5875: can measure significantly different values even in experiments
5876: performed on the same homogeneous sample. 
5877: In an actual MTJ, measured polarization is {\it not} an intrinsic
5878: property of the F region and could depend on interfacial
5879: properties and the choice of insulating barrier.  Challenges in 
5880: quantifying $P$, discussed here in the context of F/I/S tunneling, even
5881: when F is a simple ferromagnetic metal, should serve as a caution for
5882: studies of novel, more exotic, spintronic materials.
5883: 
5884: F/I/S tunneling conductance is shown in 
5885: Fig.~\ref{tunnel}, where for simplicity we assume that the spin-orbit
5886: and spin-flip scattering (see Sec.~\ref{sec:IIIC})  
5887: can be neglected, a good approximation for
5888: Al$_2$O$_3$/Al \cite{Tedrow1971:PRLb,Tedrow1994:PR}, a common choice
5889: for I/S regions.  For each spin the normalized BCS density of states
5890: is ${\tilde {\cal N}}_S(E)=Re(|E|/2\sqrt{E^2-\Delta^2})$, where E is the
5891: quasiparticle excitation energy and $\Delta$ the superconducting
5892: gap.\footnote{Here we focus on a conventional $s$-wave superconductor
5893: with no angular dependence in $\Delta$.} The BCS density of states
5894: is split in a
5895: magnetic field H, applied parallel to the interface, due to a shift in
5896: quasiparticle energy $E \rightarrow E \pm \mu_B H$, for $\uparrow$
5897: ($\downarrow$) spin parallel (antiparallel) to the field, where
5898: $\mu_B$ is the Bohr magneton.  The tunneling conductance is normalized
5899: with respect to its normal state value--for an F/I/N junction,
5900: $G(V)\equiv (dI/dV)_S/(dI/dV)_N=G_\uparrow(V)+G_\downarrow(V)$, where
5901: $V$ is the applied bias. This conductance can be expressed 
5902: by generalizing analysis of
5903: \textcite{Giaever1961:PR} as
5904: \begin{eqnarray}
5905: G(V)&=&\int^{\infty}_{-\infty}\frac{1+P}{2} 
5906: \frac{{\tilde {\cal N}}_{S}(E+\mu H) \beta dE}
5907: {4 \cosh^2[\beta(E+qV)/2]}  \\ \nonumber  
5908: &+& \int^{\infty}_{-\infty} \frac{1-P}{2} 
5909: \frac{{\tilde{\cal N}}_{S}(E-\mu H) \beta dE}
5910: {4 \cosh^2[\beta(E+qV)/2]}. 
5911: \label{eq:tg}
5912: \end{eqnarray}
5913: 
5914: Here $\beta=1/k_BT$, $k_B$ is the Boltzmann constant, $T$ is the temperature, 
5915: and $q$ is the proton charge. The factors $(1\pm P)/2$ represent 
5916: the difference in
5917: tunneling probability 
5918: between for 
5919: $\uparrow$ and $\downarrow$ electrons.
5920: While a rigorous determination of $P$,
5921: in terms of materials parameters,
5922: would require a full calculation
5923: of spin-dependent tunneling, including the appropriate boundary conditions
5924: and a detailed understanding of the interface properties, it is
5925: customary to make some simplifications. Usually $P$ can be identified as 
5926: \cite{Worledge2000:PRB,Maekawa:2002}
5927: \begin{equation}
5928: P \rightarrow P_G=(G_{N\uparrow} -G_{N\downarrow})/
5929: (G_{N\uparrow}+G_{N\downarrow}),
5930: \label{eq:PG}
5931: \end{equation}
5932: the spin polarization of the normal-state conductance (proportional
5933: to the weighted average of the density of states in F and S and 
5934: the square of the  
5935: tunneling matrix element), 
5936: where $\uparrow$ is the electron spin with the magnetic moment parallel 
5937: to the applied field (majority electrons in F). 
5938: With the further simplification of spin-independent and constant 
5939: tunneling matrix element \cite{Tedrow1971:PRLa,Tedrow1994:PR},
5940: Eq.~(\ref{eq:PG}) can be expressed as
5941: \begin{equation}
5942: P \rightarrow P_{\cal N}=({\cal N}_{F\uparrow} -{\cal N}_{F\downarrow})/
5943: ({\cal N}_{F\uparrow} + {\cal N}_{F\downarrow}),
5944: \label{eq:PF}
5945: \end{equation}
5946: the spin polarization of the tunneling density of states in the F region 
5947: at the Fermi level.
5948: 
5949: \begin{figure} 
5950: \centerline{\psfig{file=zutic_fig22.eps,width=\linewidth,angle=0}}
5951: \caption{Ferromagnet/insulator/superconductor tunneling in an applied magnetic
5952: filed: (a) Zeeman splitting of the BCS density of states as a function of 
5953: applied bias; 
5954: (b) normalized spin-resolved conductance (dashed lines) and the total 
5955: conductance (solid line)
5956: at finite temperature.} 
5957: \label{tunnel}
5958: \end{figure}
5959: Spin polarization $P$ of the F electrode can be deduced \cite{Tedrow1994:PR} 
5960: from the asymmetry
5961: of the conductance amplitudes at the four peaks in Fig.~\ref{tunnel} (b) 
5962: [for $P$=0, G(V)=G(-V)].
5963: In CrO$_2$/I/S tunnel junctions, nearly complete
5964: spin polarization $P_G>0.9$ was measured \cite{Parker2002:PRL}. 
5965: Only two of the four peaks sketched
5966: in Fig.~\ref{tunnel}, have been observed, indicating no features due to the
5967: minority spin up to H=2.5 T. \textcite{Parkin2004:P} have shown that
5968: by replacing an aluminum oxide (a typical choice for an insulating 
5969: region) with magnesium oxide, one can significantly increase the 
5970: spin polarization in F/I/S junctions. Correspondingly, extraordinarily large
5971: values of TMR ($>200\%$ at room temperature) can be achieved even
5972: with conventional ferromagnetic CoFe) electrodes.
5973: 
5974: The assumption of spin-conserving tunneling can be generalized 
5975: \cite{Tedrow1994:PR,Worledge2000:PRB,Monsma2000:APLa,Monsma2000:APLb}  
5976: to extract $P$ in the presence of spin-orbit and spin-flip scattering. 
5977: Theoretical 
5978: analyses \cite{Fulde1973:AP,Maki1964:PTP,Bruno1973:PRB} using 
5979: many-body techniques
5980: show that the  spin-orbit scattering would smear the Zeeman-split 
5981: density of states, eventually merging the
5982: four peaks into two, while the magnetic impurities \cite{Abrikosov1960:ZETP}
5983: act as pair breakers and reduce the value of $\Delta$. 
5984: Neglecting the spin-orbit
5985: scattering was shown to lead to the extraction of higher $P$ 
5986: values \cite{Tedrow1994:PR,Monsma2000:APLa}.
5987: 
5988: With a few exceptions \cite{Worledge2000:PRL},
5989: F/I/S conductance measurements \cite{Tedrow1994:PR}
5990: have revealed positive $P$--the dominant contribution of 
5991: majority spin electrons for different ferromagnetic films 
5992: (for example, in Fe, Ni, Co and Gd). However, electronic
5993: structure calculations typically give  
5994: that ${\cal N}_{F\uparrow}<{\cal N}_{F\downarrow}$ and $P_{\cal N}<0$
5995: [for Ni and Co  ${\cal N}_{F\uparrow}/{\cal N}_{F\downarrow} \sim 1/10$ 
5996: \cite{Butler2001:PRB}].
5997: Early theoretical work addressed this apparent
5998: difference,\footnote{For a list of references see 
5999: \textcite{Tedrow1973:PRB,Tedrow1994:PR}.}
6000: and efforts to understand precisely what is being experimentally measured 
6001: have continued.
6002: 
6003: \textcite{Stearns1977:JMMM} suggested that only itinerant, 
6004: freelike electrons
6005: will contribute to tunneling, while nearly localized electrons, 
6006: with a large 
6007: effective mass,
6008: contribute to the total density of states but not to G(V) 
6009: [see also \textcite{Hertz1973:PRB}
6010: and, for spin-unpolarized tunneling, \textcite{Gadzuk1969:PR}].
6011: From the assumed parabolic dispersion of the spin subbands with fixed spin 
6012: splitting, Stearns related the measured polarization 
6013: to the magnetic 
6014: moment, giving positive 
6015: $P \rightarrow P_k=(k_{F\uparrow}-k_{F\downarrow})/(k_{F\uparrow}
6016: +k_{F\downarrow})$,
6017: the spin polarization of the projections of Fermi wave vectors 
6018: perpendicular
6019: to the interface.
6020: Similar arguments, for inequivalent 
6021: density-of-states contributions  
6022: to G(V), were generalized
6023: to more complex electronic structure.
6024: \textcite{Mazin1999:PRL} showed 
6025: the importance of the tunneling matrix elements which have
6026: different Fermi velocities for different bands 
6027: [see also \cite{Yusof1998:PRB},
6028: in the context of tunneling in 
6029: a high temperature superconductor (HTSC)]. 
6030: Consequently, $P_G$ could even have an opposite sign from $P_{\cal N}$--which, 
6031: for example, would be measured by spin-resolved photoemission. 
6032: 
6033: Good agreement between  
6034: tunneling data and electronic structure calculations was illustrated by 
6035: the example of Ni$_x$Fe$_{1-x}$ \cite{Nadgorny2000:PRB}, showing, however, 
6036: that $P$ is not directly related to the  magnetic moment 
6037: \cite{Meservey1976:PRL}. 
6038: The difference between bulk and the surface densities of states 
6039: of the ferromagnet 
6040: (probed in tunneling measurements) \cite{Oleinik2000:PRB}, 
6041: the choice of tunneling barrier \cite{DeTeresa1999:PRL}, and details of the
6042: interfacial properties, which can change over time \cite{Monsma2000:APLb}, have
6043: all been shown to affect the measured $P$ directly. 
6044: 
6045: Tedrow-Meservey technique is also considered as a
6046: probe to detect spin injection in Si, where optical methods, due to
6047: the indirect gap, would be ineffective. F/I/S tunneling was also studied
6048: using amorphous Si (a-Si) and Ge (a-Ge) as a barrier. While 
6049: with a-Si  some spin polarization was detected \cite{Meservey1982:JAP} 
6050: no  spin-polarized tunneling was observed using a-Ge  barrier 
6051: \cite{Gibson1985:JAP},
6052: in contrast to the first reports of TMR \cite{Julliere1975:PL}.
6053: 
6054: Spin-dependent tunneling was also studied using a 
6055: HTSC
6056: electrode as a detector
6057: of spin polarization \cite{Vasko1998:APL,Chen2001:PRB}. 
6058: While this can significantly extend the temperature
6059: range in the tunneling experiments, a lack of  understanding of 
6060: HTSC's makes such structures more a test ground for fundamental physics
6061: than a quantitative tool for 
6062: quantitatively 
6063: determining
6064: $P$.
6065: There are also several important differences between studies using
6066: HTSC's and 
6067: conventional low-temperature superconductors. The superconducting
6068: pairing symmetry no longer yields an isotropic energy gap, and even 
6069: for the BCS-like picture the density of states should be accordingly modified. 
6070: A sign change of the pair potential can result in $G(V=0) >0$
6071: for $T\rightarrow 0$ even for a strong tunneling barrier
6072: and give rise to a zero-bias conductance peak 
6073: \cite{Hu1994:PRL,Kashiwaya1995:PRL,Wei1998:PRL}.
6074: This is explained
6075: by the two-particle process of Andreev reflection (discussed further in
6076: Sec.~\ref{sec:IVA3}),
6077: which, in addition to the usual quasiparticle tunneling, 
6078: contributes to the I-V characteristics of a F/I/S junction 
6079: \cite{Zutic1999:PRBb,Zhu1999:PRB,Zutic2000:PRB,Kashiwaya1999:PRB,Hu1999:PRBb}
6080: [a simpler N/I/S case is reviewed by \textcite{Hu1998:PRB} and
6081: \textcite{Kashiwaya2000:RPP}]. 
6082: The suppression of a zero-bias conductance peak, measured by an STM,
6083: was recently used to detect spin injection into a HTSC \cite{Ngai2003:P}.
6084: 
6085: \subsubsection{\label{sec:IVA2} F/I/F tunneling}
6086: %------------------------------------------------------------------------
6087: 
6088: 
6089: In the preface to a now classic 
6090: reference on spin-unpolarized tunneling
6091: in solids, \textcite{Duke:1969} concludes that (with only a few exceptions)
6092: the study of tunneling is an art and not a science. 
6093: Perhaps this is also an apt 
6094: description
6095: for the 
6096: present state of experiment on spin-polarized tunneling between 
6097: two ferromagnetic
6098: regions. Even for MTJ's with standard ferromagnetic metals, the bias and the
6099: temperature dependence of the TMR, as identification of the relevant spin 
6100: polarization
6101: remain to be fully understood. In a brief review of current findings we
6102: intend to identify questions that could arise as new materials for
6103: MTJ's are being considered.
6104: 
6105: A resurgence in interest in the study of MTJ's, following a hiatus after the
6106: early work by \textcite{Julliere1975:PL,Maekawa1982:IEEE}, was 
6107: spurred by the 
6108: observation of large room-temperature TMR 
6109: \cite{Moodera1995:PRL,Miyazaki1995:JMMM}.
6110: This discovery has 
6111: opened the possibility of using MTJ's for fundamental studies
6112: of surface magnetism and room-temperature spin polarization in various 
6113: ferromagnetic electrodes 
6114: as well as to suggesting applications such as highly sensitive 
6115: magnetic-field sensors, magnetic read heads, 
6116: and nonvolatile magnetic memory applications. 
6117: 
6118: It is instructive to notice the similarity between the schematic geometry 
6119: and the direction of current flow in  an MTJ and that in CPP GMR 
6120: (recall Figs.~\ref{intro:1} and 
6121: \ref{gmr:1}),
6122: which only differ in the middle layer being an insulator and a metal, 
6123: respectively.
6124: By considering the limit of ballistic transport in 
6125: CPP GMR\footnote{Related
6126: applications are usually in a diffusive regime.}
6127: it is possible to give a unified picture of both TMR and CPP GMR by
6128: varying  the strength of the hopping integrals \cite{Mathon1997:PRB} in a
6129: tight-binding representation.
6130: 
6131: \textcite{Julliere1975:PL} modified Eq.~(\ref{eq:tg}) in the limit
6132: $V \rightarrow 0$, $T \rightarrow 0$,
6133: and applied it to study F/I/F tunneling. 
6134: The two F regions are treated as uncoupled with the spin-conserving 
6135: tunneling across the barrier. This effectively leads to the two-current
6136: model proposed by \textcite{Mott1936:PRCa} and also applied to 
6137: the CPP GMR geometries \cite{Valet1993:PRB,Gijs1997:AP}.
6138: The values for $P$ extracted
6139: from F/I/S measurements are in a good agreement with the observed 
6140: TMR  values (typically positive, as expected from P$_{1,2}>0$). 
6141: However, Julli{\`{e}}re's formula\footnote{For its limitations and extensions
6142: see comprehensive reviews by \textcite{Moodera1999:JMMM,Moodera1999:ARMS}.}
6143: does not provide an explicit TMR dependence on bias and temperature.
6144: 
6145: Julli{\`{e}}re's result can be obtained as a
6146: limiting case from a more general Kubo/Landauer approach \cite{Mathon1999:PRB}
6147: with the assumption that the component of the wave vector parallel to
6148: the interface ${\bf k}_\|$ is not conserved (incoherent tunneling).
6149: Such a loss of coherence is good approximation for simply capturing the
6150: effects of disorder for amorphous Al$_2$O$_3$, a common choice 
6151: for the I region
6152: with metallic ferromagnets. 
6153: Despite its simplicity, the Julli{\`{e}}re's model for the TMR 
6154: has continued to be used for interpreting
6155: the spin polarization in various MTJ's. 
6156: Recent examples include 
6157: F regions made of manganite perovskites displaying colossal 
6158: magnetoresistance (CMR) \cite{Bowen2002:P} (suggesting $P_{\cal N}$$>$0.95);
6159: magnetite (Fe$_3$O$_4$) \cite{Hu2002:PRL} (with $P$$<$0 and TMR$<$0); 
6160: III-V ferromagnetic semiconductors \cite{Chun2002:PRB}; 
6161: a nonmagnetic semiconductor used as a tunneling barrier 
6162: \cite{Kreuzer2002:APL};
6163: Co/carbon nanotube/Co MTJ \cite{Tsukagoshi1999:N};
6164: and resonant tunneling in F/I/N/F junctions \cite{Yuasa2002:S}.
6165: 
6166: For novel materials, in which the electronic structure calculations and an
6167: understanding of the interfacial properties are not available, 
6168: Julli{\`{e}}re's 
6169: formula 
6170: still provides useful insights.
6171: A quantitative understanding of MTJ's challenges similar to those 
6172: discussed for F/I/S tunneling, including determining precisely which spin 
6173: polarization is relevant and the related issue of reconciling the 
6174: (typically positive) sign 
6175: of  the observed TMR with the electronic structure 
6176: \cite{Tsymbal1997:JPCM,Oleinik2000:PRB,Bratkovsky1997:PRB,MacLaren1997:PRB,%
6177: Mathon1997:PRB,LeClair2002:PRL}.  
6178: 
6179: In an approach complementary to Julli{\`{e}}re's, 
6180: \textcite{Slonczewski1989:PRB} 
6181: considered F/I/F as a single quantum-mechanical system in a free-electron 
6182: picture. 
6183: When matching the two-component wave functions at interfaces, 
6184: coherent tunneling was  
6185: assumed, with conserved ${\bf k}_\|$, relevant to epitaxially grown MTJ's 
6186: \cite{Mathon2001:PRB} and the I region was modeled by a square 
6187: barrier.\footnote{A formally analogous problem was considered by 
6188: \textcite{Griffin1971:PRB} in an N/I/S system where the two-component
6189: wave functions represented electron-like and hole-like quasiparticles 
6190: rather then the two spin projections; see Sec.~\ref{sec:IVA3}.}
6191: The resulting TMR can be expressed as in
6192: Eq.~(\ref{eq:julliere}) but with the redefined polarization 
6193: \begin{equation}
6194: P \rightarrow P_k 
6195: (\kappa^2-k_{F\uparrow}k_{F\downarrow})/(\kappa^2
6196: +k_{F\uparrow}k_{F\downarrow}),
6197: \label{eq:slon}
6198: \end{equation}
6199: where $P_k$, as defined by \textcite{Stearns1977:JMMM}, is also
6200: $P_{\cal N}$ (in a free-electron picture) and $i \kappa$ is the usual imaginary
6201: wave vector through a square barrier. Through the dependence of
6202: $\kappa$ on $V$ the resulting polarization in Slonczewski's model can change
6203: sign. A study of a similar geometry using a Boltzmann-like approach shows
6204: \cite{Chui1997:PRB}
6205: that the spin splitting of electrochemical potentials persists in the F region
6206: all the way to the F/I interface, implying 
6207: $\kappa_\uparrow \neq \kappa_\downarrow$
6208: and an additional voltage dependence of the TMR. Variation of  
6209: the density of states [inferred from the
6210: spin-resolved photoemission data \cite{Park1998:PRL,Park1998:N}]  within the
6211: range of applied bias in MTJ's of Co/SrTiO$_3$/La$_{0.7}$Sr$_{0.3}$MnO$_3$ 
6212: (Co/STO/LSMO) 
6213: \cite{DeTeresa1999:PRL}, 
6214: together with Julli{\`{e}}re's model, was used to explain
6215: the large negative TMR (-50\% at 5 K), 
6216: which would even change sign for positive
6217: bias (raising the Co Fermi level above the corresponding one of LSMO). 
6218: The bias dependence of the TMR was also attributed to the density of states
6219: by extending the
6220: model of a trapezoidal tunneling barrier \cite{Brinkman1970:JAP} to the 
6221: spin-polarized
6222: case \cite{Xiang2002:PRB}. 
6223: 
6224: The decay of TMR with temperature can be attributed to several causes. 
6225: Early theoretical work on N/I/N tunneling 
6226: \cite{Anderson1966:PRL,Appelbaum1966:PRL}
6227: [for a detailed discussion and a review of
6228: related experimental results see \textcite{Duke:1969}]
6229: showed that the presence of magnetic impurities  in the tunneling
6230: barrier produces temperature dependent conductance--referred to as 
6231: zero-bias anomalies. 
6232: These findings, which considered both spin-dependent and 
6233: spin-flip scattering,
6234: were applied to  fit the decay of the TMR with temperature
6235: \cite{Miyazaki:2002,Jansen2000:PRB,Inoue1999:JMMM}.
6236: Hot electrons  localized at F/I interfaces were predicted, 
6237: to create magnons, or collective spin excitations,
6238: near the F/I interfaces, and suppress the TMR \cite{Zhang1997:PRL}.
6239: Magnons were observed \cite{Tsui1971:PRL} in an antiferromagnetic (AFM)
6240: NiO barrier in single crystal Ni/NiO/Pb tunnel junctions and were suggested 
6241: \cite{Moodera1995:PRL} as the cause of
6242: decreasing TMR with $T$ by spin-flip scattering.
6243: Using an $s$-$d$ exchange  (between itinerant $s$ and nearly localized $d$ 
6244: electrons)
6245: Hamiltonian, it was shown \cite{Zhang1997:PRL} that, at  $V\rightarrow 0$,
6246: $G(T)-G(0) \propto T \ln T$, for both 
6247: $\uparrow \uparrow$ and $\uparrow \downarrow$
6248: orientations.
6249: A different temperature dependence of TMR was
6250: suggested by \textcite{Moodera1998:PRL}.
6251: It was related to the decrease of the surface magnetization 
6252: \cite{Pierce1982:PRB,Pierce:1984}
6253: $M(T)/M(0) \propto T^{3/2}$. Such a temperature dependence 
6254: [known as the Bloch's law and reviewed by \textcite{Krey2003:P}],
6255: attributed to magnons, was also obtained for TMR 
6256: \cite{MacDonald1998:PRL}.
6257: An additional decrease of TMR with $T$ was expected due to the spin-independent 
6258: part of G(T)  \cite{Shang1998:PRB}, seen also in N/I/N junctions. 
6259: 
6260: Systematic studies of MTJ's containing a semiconductor (Sm) region
6261: (used as a tunneling barrier and/or  an F electrode) have begun only 
6262: recently.\footnote{The early F/Ge/F results \cite{Julliere1975:PL} were not 
6263: reproduced and other metallic structures involving Si, Ge, GaAs, and GaN 
6264: as a barrier have shown 
6265: either no \cite{Gibson1985:JAP,Loraine2000:JAP,Boeve2001:JMMM} 
6266: or only a small \cite{Meservey1982:JAP,Jia1996:IEEE,Kreuzer2002:APL} 
6267: spin-dependent signal.}
6268: To improve the performance of MTJ's it is desirable to reduce the junction
6269: resistance. A smaller $RC$ constant would allow faster switching 
6270: times in MRAM (for a detailed discussion see \textcite{DeBoeck2002:SST}). 
6271: Correspondingly, using a semiconducting barrier could prove 
6272: an alternative strategy for difficult fabrication of ultrathin 
6273: ($<$1 nm ) oxide barriers 
6274: \cite{Rippard2002:PRL}. Some F/Sm/F MTJ's have been
6275: grown epitaxially, and the amplitude of TMR can be studied as function
6276: of the crystallographic orientation of a F/Sm interface.
6277: For an epitaxially grown Fe/ZnSe/Fe MTJ electronic structure calculations
6278: have predicted  \cite{MacLaren1999:PRB} large TMR (up to $\sim$1000 \%), 
6279: increasing
6280: with ZnSe thickness. However, the observed TMR in Fe/ZnSe/Fe$_{0.85}$Co$_{0.15}$
6281: was limited  below 50 K, reaching 15 \% at 10 K for junctions of higher
6282: resistance and lower defect density\footnote{Interface defects could diminish 
6283: measured TMR. We recall (see Sec.~\ref{sec:IID3}) that at a ZnMnSe/AlGaAs interface 
6284: they limit the spin injection efficiency \cite{Stroud2002:PRL} and from 
6285: Eq.~(\ref{eq:delR}) infer a reduced spin-valve effect.}
6286: \cite{Gustavsson2001:PRB}. 
6287: Results on ZnS, 
6288: another II-VI semiconductor, demonstrated a TMR of $\sim$5 \% at room 
6289: temperature \cite{Guth2001:APL}.
6290: 
6291: \begin{figure} 
6292: \centerline{\psfig{file=zutic_fig23.eps,width=0.9\linewidth}}
6293: \caption{All-semiconductor magnetic tunnel junction:
6294: (a) magnetization of Ga$_{1-x}$Mn$_x$As (x=4.0\%,50 nm)/AlAs
6295: (3 nm)/ Ga$_{1-x}$Mn$_x$As (x=3.3\%,50 nm) trilayer measured by a SQUID at 8 K.
6296: The sample size is 3 $\times$ 3 mm$^2$. 
6297: Magnetization shown is normalized with respect to the saturation value M$_s$.
6298: (b) 
6299: TMR curves of a Ga$_{1-x}$Mn$_x$As (x=4.0\%,50 nm)/AlAs
6300: (1.6 nm)/ Ga$_{1-x}$Mn$_x$As (x=3.3\%,50 nm) tunnel junction of 200 $\mu$m
6301: in diameter. Bold solid curve, 
6302: sweep of the magnetic
6303: field from positive to negative;
6304: dashed curve, sweep from
6305: negative to positive; thin solid curve,
6306: a minor loop. From \onlinecite{Tanaka2001:PRL}.}
6307: \label{tmr0}
6308: \end{figure}
6309: 
6310: There is also a possibility of using 
6311: all-semiconductor F/Sm/F single-crystalline MTJ's
6312: where F is a ferromagnetic semiconductor. These would simplify integration
6313: with the existing conventional semiconductor-based electronics and allow
6314: flexibility of various doping profiles and fabrication of quantum structures,
6315: as compared to the conventional all-metal MTJ's. Large TMR ($>$70 \% at 8 K),
6316: shown in Fig.~\ref{tmr0},
6317: has been measured in an epitaxially grown (Ga,Mn)As/AlAs/(Ga,Mn)As junction 
6318: \cite{Tanaka2001:PRL}. The results are consistent with the $k_\|$ being 
6319: conserved in the tunneling process \cite{Mathon1997:PRB}, with the decrease of 
6320: TMR with $T$ expected from the spin-wave excitations 
6321: \cite{Shang1998:PRB,MacDonald1998:PRL},
6322: discussed above. TMR is nonmonotonic with thickness in AlAs 
6323: (with the peak at
6324: $\sim$ 1.5 nm). For a given AlAs thickness, double MTJ's were also shown
6325: to give similar TMR values and were used to determine electrically
6326: the spin injection in GaAs QW \cite{Mattana2002:PRL}.
6327: However, 
6328: a room-temperature effect 
6329: remains to be demonstrated 
6330: as 
6331: the available
6332: well-characterized ferromagnetic semiconductors do not have 
6333: as high a ferromagnetic
6334: transition temperature. 
6335: 
6336: A lower barrier in F/Sm/F MTJ's can have important implications in 
6337: determining the actual values of TMR. The standard four-probe technique for
6338: measuring I and V 
6339: has been known to give spurious values when the resistance 
6340: of the F electrodes is non-negligible to the junction resistance.
6341: The tunneling current in that regime has been shown to be highly 
6342: nonuniform\footnote{Nonuniform tunneling current has been studied
6343: in nonmagnetic junctions \cite{Pederson1967:APL}, 
6344: CPP multilayers 
6345: \cite{Lenczowski1994:JAP},    
6346: and conventional MTJ's \cite{Moodera1996:APL,Rzchowski2000:PRB}.} 
6347: and the measured apparent resistance $R_m=V/I$  (different from the
6348: actual junction resistance R$_J$)  can even attain negative values 
6349: \cite{Pederson1967:APL,Moodera1996:APL}. The important implications
6350: for MTJ are the possibility of large overestimates in the TMR amplitude
6351: \cite{Moodera1999:JMMM} and a desirable hysteresis effect--at H=0
6352: the two values of resistance can be used for various nonvolatile 
6353: applications \cite{Moodera1996:APL}. 
6354: 
6355: A detailed understanding of MTJ's will also require knowing the influence
6356: of the interface and surface roughness \cite{Itoh1999:JPSP}. Even in 
6357: the spin-unpolarized case it is known that the full quantum-mechanical 
6358: approach 
6359: \cite{Tesanovic1986:PRL} can lead to qualitatively different 
6360: results from the 
6361: usual quasiclassical picture and from averaging out the spatial information on the 
6362: length scale of the inverse Fermi wave vector.
6363: 
6364: A comprehensive review of tunneling phenomena and magnetoresistance in 
6365: granular materials, ferromagnetic single-electron transistors, 
6366: and double tunnel junctions
6367: is given by \textcite{Maekawa:2002}. 
6368: A theoretical study of F/I/F junctions,
6369: in which the I region is a quantum dot,
6370: shows the importance of Coulomb interactions, which could lead to spin 
6371: precession
6372: even in the absence of an applied magnetic field \cite{Konig2003:PRL}.
6373: 
6374: \subsubsection{\label{sec:IVA3} Andreev reflection}
6375: %------------------------------------------------------------------------
6376: 
6377: 
6378: Andreev reflection \cite{Andreev1964:SPJETP} 
6379: is a scattering process, at an interface with a superconductor, 
6380: responsible for a conversion between a dissipative quasiparticle
6381: current and a dissipationless supercurrent [see also early work
6382: by \textcite{deGennes1963:PL}]. 
6383: For a spin-singlet superconductor
6384: an incident electron (hole) of spin
6385: $\lambda$ is reflected as a
6386: hole (electron) belonging to the opposite spin subband $\overline{\lambda}$, 
6387: back
6388: to the nonsuperconducting region, while a Cooper pair is transferred to the 
6389: superconductor. 
6390: This is a phase-coherent scattering process in which the reflected particle
6391: carries the information about both the phase of 
6392: the incident particle
6393: and the macroscopic phase of the superconductor.\footnote{For instructive
6394: reviews see \textcite{Lambert1998:JPCM,Pannetier2000:JLTP}.} Andreev
6395: reflection  thus is responsible for a proximity effect where the 
6396: phase correlations are introduced to a nonsuperconducting 
6397: material \cite{Demler1997:PRB,Halterman2002:PRB,Bergeret2001:PRL,%
6398: Izyumov2002:PU,Fominov2003:T}.
6399: The probability for Andreev reflection at low bias voltage 
6400: ($qV \lesssim \Delta$), which is 
6401: related to the square of the normal-state transmission, 
6402: could
6403: be ignored for  low-transparency junctions with conventional
6404: superconductors, as discussed in Sec.~\ref{sec:IVA1}. In contrast, 
6405: for high-transparency junctions (see the discussion 
6406: of Sharvin conductance in Sec.~\ref{sec:IIC2}),
6407: single-particle tunneling vanishes [recall Eq.~(\ref{eq:tg})]
6408: at low bias and $T=0$ and Andreev
6409: reflection is the dominant process. A convenient description is
6410: provided by the 
6411: Bogoliubov-de Gennes equations \cite{deGennes:1989},
6412: \begin{eqnarray}
6413: \left[\begin{array}{cc} H_\lambda & \Delta \\
6414: \Delta^* & -H^*_{\overline{\lambda}}\end{array} \right]
6415: \left[ \begin{array}{c} u_\lambda \\
6416: v_{\overline{\lambda}} \end{array} \right]=E
6417: \left[ \begin{array}{c} u_\lambda \\
6418: v_{\overline{\lambda}} \end{array} \right],
6419: \label{eq:BdG}
6420: \end{eqnarray} 
6421: and by matching the wave functions at the boundaries (interfaces) between
6422: different regions.
6423: Here $H_\lambda$ is the single-particle Hamiltonian 
6424: for spin $\lambda=\uparrow, \downarrow$ and  $\overline{\lambda}$
6425: denotes a spin opposite to $\lambda$ \cite{deJong1995:PRL,Zutic2000:PRB}.
6426: $\Delta$ is the pair potential \cite{deGennes:1989},  $E$ the excitation energy 
6427: and $u_\lambda$, $v_{\overline{\lambda}}$ are the electronlike quasiparticle 
6428: and holelike quasiparticle amplitudes, 
6429: respectively.\footnote{Equation (\ref{eq:BdG}) 
6430: can be simply modified to include the spin flip and
6431: spin-dependent interfacial scattering \cite{Zutic1999:PRBa}.}   
6432: \textcite{Griffin1971:PRB} have solved the Bogoliubov-de Gennes 
6433: equations with 
6434: square or a $\delta$-function barriers  of  varying strength 
6435: at an N/S interface. They obtained a  
6436: result that interpolates between the clean and the tunneling limits.
6437: \textcite{Blonder1982:PRB} used a similar approach,
6438: known as the Blonder-Tinkham-Klapwijk method, in which 
6439: the two limits correspond to $Z\rightarrow0$ and $Z\rightarrow \infty$,
6440: respectively, and $Z$ is the strength of the  $\delta$-function barrier.
6441: The transparency of this
6442: approach\footnote{A  good agreement \cite{Yan2000:PRB}
6443: was obtained with the more rigorous nonequilibrium 
6444: Keldysh technique \cite{Keldysh1965:SPJETP,Rammer1986:RMP}, for an illustration 
6445: of how such a technique can be used to study spin-polarized transport in a 
6446: wide range of heterojunctions see \textcite{Melin2002:EPJB,Zeng2003:P}.}
6447: makes it suitable for the study of  
6448: ballistic spin-polarized transport and spin injection even in the
6449: absence of a  superconducting region
6450: \cite{Hu2001:PRL,Matsuyama2002:PRB,Hu2001:PRB,Schapers2001b:PRB}.
6451: 
6452: It is instructive to note a similarity between the
6453: two-component transport in N/S junctions  (for
6454: electronlike and holelike quasiparticles) and F/N junctions 
6455: (for spin $\uparrow$, $\downarrow$), which both lead to current conversion, 
6456: accompanied by the additional boundary resistance
6457: \cite{Blonder1982:PRB,vanSon1987:PRL}.  
6458: In the N/S junction Andreev reflection is responsible for the conversion 
6459: between the 
6460: normal and the 
6461: supercurrent, characterized by the superconducting coherence length,
6462: while in the F/N case a conversion between spin-polarized and unpolarized
6463: current is characterized by the spin diffusion length.
6464: 
6465: For spin-polarized carriers, with  different populations in two spin
6466: subbands, only a fraction of the incident electrons from a majority subband
6467: will have a  minority subband partner in order to be  Andreev
6468: reflected. This can be simply quantified at zero bias and $Z=0$,
6469: in terms of the total number of scattering channels (for each $k_\|$) 
6470: $N_\lambda=k_{F\lambda}^2 A/4 \pi$ at the Fermi level. Here A is the 
6471: point-contact 
6472: area, and $k_{F\lambda}$ 
6473: is the spin-resolved Fermi wave vector. 
6474: A spherical Fermi surface in the F and S regions, with no (spin-averaged)
6475: Fermi velocity mismatch, is assumed. 
6476: When S is in the normal state, the 
6477: zero-temperature Sharvin conductance is
6478: \begin{equation}
6479: G_{FN}=\frac{e^2}{h}(N_\uparrow+N_\downarrow),
6480: \label{eq:sharvin2} 
6481: \end{equation}
6482: equivalent to $R_{\rm Sharvin}^{-1}$, from Eq.~(\ref{eq:sharvin}).
6483: In the superconducting state all of the $N_\downarrow$ 
6484: and only $(N_\downarrow/N_\uparrow)N_\uparrow$ scattering channels 
6485: contribute to Andreev reflection across the F/S interface and transfer 
6486: charge $2e$, yielding \cite{deJong1995:PRL}
6487: \begin{equation}
6488: G_{FS}=\frac{e^2}{h} \left(2N_\downarrow + \frac{2N_\downarrow}{N_\uparrow} 
6489: N_\uparrow \right)=4\frac{e^2}{h}N_\downarrow.
6490: \label{eq:dejong}
6491: \end{equation}
6492: The suppression of the normalized zero-bias conductance 
6493: at $V=0$ and $Z=0$ \cite{deJong1995:PRL},
6494: \begin{equation}
6495: G_{FS}/G_{FN}=2(1-P_G)
6496: \label{eq:pc}
6497: \end{equation}
6498: with the increase in the spin polarization $P_G \rightarrow 
6499: (N_\uparrow-N_\downarrow)/(N_\uparrow+N_\downarrow)$,
6500: was used as a sensitive transport technique to detect spin polarization 
6501: in a point contact  \cite{Soulen1998:S}. 
6502: Data are 
6503: given in Fig.~\ref{point}. A
6504: similar  study, using a  thin-film nanocontact geometry 
6505: \cite{Uphaday1998:PRL}, emphasized the importance of
6506: fitting the conductance data over a wide range of applied bias, not only
6507: at $V=0$, in order to extract the spin polarization of the F region more 
6508: precisely.
6509: 
6510: \begin{figure} 
6511: \centerline{\psfig{file=zutic_fig24.eps,width=\linewidth,angle=0}}
6512: \caption{The differential conductance for several spin-polarized materials,
6513: showing the suppression of Andreev reflection with increasing $P_G$.
6514: The vertical lines denote the bulk superconducting gap for Nb: 
6515: $\Delta$($T=0$)=1.5 meV. 
6516: Note that NiMnSb, one of the Heusler alloys originally proposed 
6517: as half-metallic ferromagnets \cite{deGroot1983:PRL}, shows only
6518: partial spin polarization. From \onlinecite{Soulen1998:S}.} 
6519: \label{point}
6520: \end{figure}
6521: The advantage of such techniques is the detection of polarization
6522: in a much wider range of materials than  those which
6523: can be grown for detection in F/I/S or F/I/F tunnel junctions. 
6524: A large number of
6525: experimental results using spin-polarized Andreev reflection
6526: has since been reported 
6527: \cite{Nadgorny2001:PRB,Ji2001:PRL,Parker2002:PRL,Bourgeois2001:PRB,%
6528: Panguluri2003:Pb},  
6529: including the first direct measurements 
6530: \cite{Braden2003:P,Panguluri2003:P,Panguluri2004:P} of
6531: the spin polarization in (Ga,Mn)As and (In,Mn)Sb.\footnote{Similar 
6532: measurements were
6533: also suggested by \textcite{Zutic1999:PRBa} to yield information 
6534: about the FSm/S interface. A more complete analysis should also quantify
6535: the effects of spin-orbit coupling.}
6536: However,
6537: for a quantitative interpretation of the measured polarization, important
6538: additional factors (similar to the limitations discussed for the application
6539: of Julli{\`{e}}re's formula in Sec.~\ref{sec:IVA2}) need to be incorporated 
6540: in the picture 
6541: provided by Eq.~(\ref{eq:pc}).
6542: For example, the Fermi surface may not be spherical [see the discussion of 
6543: \textcite{Mazin1999:PRL}, specifying what type of spin polarization is
6544: experimentally measured and also that of \textcite{Xia2002:PRL}]. 
6545: The roughness or the size of the  F/S interface 
6546: may lead to a diffusive component of the transport 
6547: \cite{Jedema1999:PRB,Falko1999:PZETF,Mazin2001:JAP}.
6548: As a caution concerning the  possible difficulties in analyzing experimental data,
6549: we mention some subtleties that arise even for the simple model
6550: of a spherical Fermi surface used to describe both F and S regions.
6551: Unlike charge transport in N/S junctions \cite{Blonder1983:PRB} 
6552: in a Griffin-Demers-Blonder-Tinkham-Klapwijk approach, Fermi velocity mismatch
6553: between the F and the S regions, 
6554: does not simply increase the value of effective $Z$.
6555: Specifically, at $Z=V=0$ and normal incidence
6556: it is possible to have perfect transparency even when all the
6557: Fermi velocities differ, satisfying 
6558: $(v_{F\uparrow} v_{F\downarrow})^{1/2}=v_S$,
6559: where $v_S$ is the Fermi velocity in a superconductor 
6560: \cite{Zutic1999:PRBa,Zutic1999:PRBb,Zutic2000:PRB}. In other words, 
6561: unlike in Eq.~(\ref{eq:pc}), the spin polarization 
6562: (nonvanishing exchange energy) 
6563: can {\it increase} the subband conductance, 
6564: for fixed Fermi velocity mismatch. 
6565: Conversely, at a fixed exchange energy, an increase in Fermi velocity mismatch  
6566: could increase the subgap conductance.\footnote{Similar results were 
6567: also obtained when 
6568: F and S region were separated with a quantum dot 
6569: \cite{Zhu2002:PRB,Feng2003:PRB,Zeng2003:P} 
6570: and even in a 1D tight-binding model with  no spin polarization 
6571: \cite{Affleck2000:PRB}.}
6572: In a typical interpretation of a measured conductance, complications can
6573: then arise in trying to disentangle the influence of parameters $Z$, $P_G$,
6574: and Fermi velocity mismatch from the nature of the point contacts 
6575: \cite{Kikuchi2002:PRB} 
6576: and the role of inelastic scattering \cite{Auth2003:PRB}.
6577: Detection of $P$ in HTSC's is even possible with a large barrier or a
6578: vacuum between the F and S regions, 
6579: as proposed by \textcite{Wang2002b:P} 
6580: using resonant Andreev reflection and 
6581: a $d$-wave superconductor.\footnote{Interference effects between the
6582: quasi-electron and quasi-hole scattering trajectories that feel
6583: pair potentials of different sign lead to a large 
6584: conductance near zero bias, even at large interfacial barrier
6585: (referred to as a zero-bias conductance peak in Sec.~\ref{sec:IVA1}).
6586: } 
6587: 
6588: Large magnetoresistive effects are predicted for crossed Andreev reflection
6589: \cite{Deutscher2000:APL}, when the two F regions, separated within the 
6590: distance of
6591: the superconducting coherence length,\footnote{Recent theoretical
6592: findings suggest that the separation should not exceed the 
6593: Fermi wavelength \cite{Yamashita2003:P}.} are on the same side of the
6594: S region. Such structures have also been theoretically studied to
6595: understand the implications of nonlocal correlations 
6596: \cite{Apinyan2002:EPJB,Melin2002:EPJB}. 
6597: 
6598: \subsubsection{\label{sec:IVA4} Spin-polarized drift and diffusion}
6599: %------------------------------------------------------------------------------
6600: 
6601: Traditional semiconductor devices such as field-effect transistors, 
6602: bipolar diodes and transistors, or semiconductor solar cells rely 
6603: in great part on carriers (electrons and holes) whose motion
6604: can be described as drift and diffusion, limited by carrier recombination. In 
6605: inhomogeneous devices where charge buildup is rule, the 
6606: recombination-limited 
6607: drift-diffusion is supplied by Maxwell's equations, to be solved in a 
6608: self-consistent manner. 
6609: Many proposed spintronic devices as well as experimental 
6610: settings for spin injection (see Sec.~\ref{sec:II}) can be described by 
6611: both carrier and spin 
6612: drift and diffusion, limited by carrier recombination
6613: and spin relaxation \cite{Fabian2002:PRB,Zutic2002:PRL}. 
6614: In addition, if spin precession is important for device operation, spin 
6615: dynamics need to be explicitly incorporated into the transport equations 
6616: \cite{Qi2003:PRB}. Drift of the spin-polarized carriers can be due not only 
6617: to the electric field, but also 
6618: to magnetic fields. We illustrate spin-polarized drift and diffusion 
6619: on the transport model of spin-polarized bipolar transport, 
6620: where bipolar refers to the presence of electrons and holes, 
6621: not spin up and down. 
6622: A spin-polarized unipolar transport can be obtained as a limiting case by
6623: setting the electron-hole recombination rate to zero and considering only
6624: one type of carrier (either electrons or holes).
6625: 
6626: Consider electrons and holes whose density is commonly denoted here as 
6627: $c$ (for carriers), moving in the electrostatic potential $\phi$ which 
6628: comprises both the external bias $V$ and the internal built-in fields 
6629: due to 
6630: charge inhomogeneities. Let the equilibrium spin splitting of the 
6631: carrier band be 
6632: $2q\zeta_{c}$. The spin $\lambda$ resolved carrier charge-current density is
6633: \cite{Zutic2002:PRL}
6634: \begin{equation}
6635: \label{eq:jcl}
6636: {\bf j}_{c\lambda}=
6637: -q\mu_{c\lambda}c_{\lambda}\nabla \phi {\pm}qD_{c\lambda}\nabla c_\lambda
6638: -q\lambda\mu_{c\lambda}c_{\lambda}\nabla \zeta_c,
6639: \end{equation}
6640: where $\mu$ and $D$ stand for mobility and 
6641: diffusion
6642: coefficients, 
6643: the upper sign is for electrons  and the lower sign is for holes. The first term on the
6644: right hand side describes drift caused by the total electric field, 
6645: the second term represents diffusion, while the last term stands for 
6646: magnetic drift---carrier drift in inhomogeneously split 
6647: bands.\footnote{Equation~(\ref{eq:jcl}) can be viewed as the generalization
6648: of the Silsbee-Johnson spin-charge coupling 
6649: \cite{Johnson1987:PRB,Heide2001:PRL,Wegrowe2000:PRB} to bipolar transport 
6650: and to systems with spatially inhomogeneous charge density.}
6651: More transparent are the equations for the total charge, 
6652: $j=j_{\uparrow}+
6653: j_{\downarrow}$, and spin, $j_s=j_{\uparrow}-j_{\downarrow}$, 
6654: current densities:
6655: \begin{eqnarray} \label{eq:jc}
6656: &{\bf j}_c&=-\sigma_c \nabla\phi - 
6657: \sigma_{sc}\nabla \zeta_c \pm qD_c \nabla c \pm qD_{sc}\nabla s_c,   \\
6658: \label{eq:js}
6659: &{\bf j}_{sc}&=-\sigma_{sc} \nabla\phi - 
6660: \sigma_{c}\nabla \zeta_c \pm qD_{sc} \nabla c \pm qD_{c}\nabla s_c,
6661: \end{eqnarray}
6662: where the carrier density $c=c_\uparrow+c_\downarrow$ 
6663: and spin $s_c=c_\uparrow-c_\downarrow$, 
6664: and we introduced the carrier charge and spin conductivities 
6665: $\sigma_c=q(\mu_c c+ \mu_{sc} s_c)$
6666: and $\sigma_{sc}=q(\mu_{sc}c+\mu_cs_c)$, 
6667: where $\mu_c=(\mu_{c\uparrow}+\mu_{c\downarrow})/2$
6668: and $\mu_{cs}=(\mu_{c\uparrow}-\mu_{c\downarrow})/2$ 
6669: are charge and spin mobilities, and similarly
6670: for the diffusion coefficients. Equation (\ref{eq:jc}) 
6671: describes the spin-charge coupling 
6672: in bipolar transport in inhomogeneous magnetic semiconductors. 
6673: Spatial variations in 
6674: spin density can cause charge currents. 
6675: Similarly, it follows from Eq.~(\ref{eq:js})
6676: that spatial variations in carrier densities can lead to spin currents.  
6677: 
6678: Steady-state carrier recombination and spin relaxation processes 
6679: are described 
6680: by the continuity equations for the spin-resolved carrier densities:
6681: \begin{equation} 
6682: \label{eq:cont}
6683: \nabla\cdot \frac{{\bf j}_{c\lambda}}{q}= 
6684: \pm w_{c\lambda} (c_{\lambda}\bar{c}-
6685: c_{\lambda 0}\bar{c}_0)\pm 
6686: \frac{c_{\lambda}-c_{-\lambda}-\lambda \tilde{s}_c}{2\tau_{sc}}.
6687: \end{equation}
6688: Here $w$ is the spin-dependent recombination rate, the bar denotes
6689: a complementary carrier ($\bar{n}=p$, for example), $\tau_{sc}$ 
6690: is the spin relaxation
6691: time of the carrier $c$ (not to be confused with the single spin
6692: decoherence time discussed in Sec.~\ref{sec:IIIA1}),
6693: and $\tilde{s}_c=P_{c0}c$ is the nonequilibrium spin
6694: density, which appears after realizing that spin relaxation equilibrates 
6695: spin while
6696: preserving carrier density. Finally, the set of equations is 
6697: completed with
6698: Poisson's equation, 
6699: \begin{equation}\label{eq:poisson}
6700: \varepsilon\Delta \phi=-\rho,
6701: \end{equation} 
6702: connecting the electric field and charge $\rho=q(p-n+N_d-N_a)$,
6703: where $N_d$ and $N_a$ are the donor and acceptor densities, respectively, and
6704: $\varepsilon$ is
6705: the dielectric constant. 
6706: 
6707: In many important cases Eqs.~(\ref{eq:jcl}), (\ref{eq:cont}), 
6708: and (\ref{eq:poisson}) 
6709: need to be solved self-consistently, which usually requires
6710: numerical techniques \cite{Zutic2002:PRL}. 
6711: In some cases it is possible to extract the relevant physics in limiting 
6712: cases analytically,
6713: usually neglecting electric field or magnetic drift. 
6714: In unipolar spin-polarized transport one does not need to consider carrier 
6715: recombination. It also often suffices to study pure spin diffusion, 
6716: if the built-in electric fields are small. Unipolar spin-polarized transport 
6717: in inhomogeneous systems in the presence of electric fields was analyzed by 
6718: \textcite{Fabian2002:PRB,Yu2002:PRBa,Yu2002:PRBb,Martin2003:PRB,Pershin2003:PRL}.
6719: Spin-polarized drift and diffusion in 
6720: model GaAs quantum wires was studied by \textcite{Sogawa2000:PRB}, while 
6721: ramifications of magnetic drift for unipolar transport were 
6722: studied by \textcite{Martin2003:PRB,Fabian2002:PRB}. Bipolar transport 
6723: in the presence of electrical drift and/or 
6724: diffusion has been studied by 
6725: \textcite{Zutic2002:PRL,Fabian2002:PRB,Beck2002:PE,Flatte2000:PRL}.
6726: Transient dynamics of spin drift and diffusion was considered by 
6727: \textcite{Fabian2002b:PRB}.
6728: Recently an interesting study \cite{Saikin2003:JAP} was reported on a 
6729: Monte-Carlo simulation of quantum-mechanical 
6730: spin dynamics limited by spin relaxation, in which quasiclassical 
6731: orbital transport 
6732: was carried out for the in-plane
6733: transport in III-V heterostructures where spin precession is due to  bulk
6734: and structure inversion asymmetry (see Sec.~\ref{sec:IIIB2}).
6735: 
6736: 
6737: \subsection{\label{sec:IVB} Materials considerations}
6738: 
6739: Nominally highly spin-polarized materials, as discussed in the previous
6740: sections, could provide both effective spin injection into nonmagnetic
6741: materials and large MR effects, important for nonvolatile applications.
6742: Examples include  half-metallic oxides such as CrO$_2$, Fe$_3$O$_4$, 
6743: CMR materials, and double perovskites \cite{Kobayashi1998:N} 
6744: [for reviews of half-metallic materials see 
6745: \cite{Pickett2001:PT,Fang2002:JAP}].
6746: Ferromagnetic semiconductors \cite{Nagaev:1983}, known since  CrBr$_3$ 
6747: \cite{Tsubokawa1960:JPSJ}, have been demonstrated to be highly spin polarized.
6748: However, more recent interest in ferromagnetic semiconductors was spurred by the 
6749: fabrication of (III,Mn)V compounds.\footnote{Ferromagnetic order 
6750: with Mn-doping was
6751: obtained previously, for example, in (Sn,Mn)Te \cite{Escorne:1974}, 
6752: (Ge,Mn)Te \cite{Cochrane1974:PRB} and (Pb,Sn,Mn)Te \cite{Story1986:PRL}.}
6753: After the initial discovery of (In,Mn)As 
6754: \cite{Munekata1989:PRL,Ohno1992:PRL,Munekata1991:JCG}, 
6755: most of the research 
6756: has focused on (Ga,Mn)As \cite{Ohno1996:APL,VanEsch1997:PRB,Hayashi1997:JCG}.  
6757: In contrast to (In,Mn)As and
6758: (Ga,Mn)As with high carrier density ($\sim$ 10$^{20}$ cm$^{-3}$),
6759: a much lower carrier density in (Zn,Cr)Te \cite{Saito2002:JAP}, 
6760: a II-VI ferromagnetic 
6761: semiconductor with Curie temperature $T_C$ near room temperature 
6762: \cite{Saito2003:PRL},
6763: suggests that transport properties can be effectively controlled by
6764: carrier doping. Most of the currently studied ferromagnetic semiconductors
6765: are p-doped with holes as spin-polarized carriers, which typically
6766: leads to lower mobilities and shorter spin relaxation times than in
6767: n-doped materials. It is possible to use selective doping to substantially
6768: increase $T_C$, as compared to the uniformly doped bulk ferromagnetic
6769: semiconductors \cite{Nazmul2003:PRB}.
6770: 
6771: Early work on (Ga,Mn)As \cite{DeBoeck1996:APL} showed the low solubility 
6772: of Mn and the formation of  
6773: magnetic nanoclusters 
6774: characteristic of many 
6775: subsequent compounds and different magnetic impurities. 
6776: The presence of such 
6777: nanoclusters often complicates accurate
6778: determination of  $T_C$ as well as of whether 
6779: the compound is actually in a single phase. Consequently, the reported
6780: room-temperature ferromagnetism  in an increasing number of compounds
6781: reviewed by \textcite{Pearton2003:JAP} is not universally accepted.
6782: Conclusive evidence for intrinsic ferromagnetism in
6783: semiconductors is highly nontrivial. For example, early work reporting 
6784: ferromagnetism even at nearly 900 K in  La-doped CaBa$_6$ 
6785: \cite{Young1999:N,Ott2000:PB,Tromp2001:PRL}, was later revisited
6786: suggesting extrinsic effect \cite{Bennett2003:P}. 
6787: It remains to be understood what the limitations are for using 
6788: extrinsic ferromagnets and, for example,  whether they can be 
6789: effective spin injectors. 
6790: 
6791: \begin{figure}
6792: \centerline{\psfig{file=zutic_fig25.eps,width=1.0\linewidth,angle=0}}
6793: \caption{Photoinduced ferromagnetism in a (In,Mn)As/GaSb heterostructure:
6794: (a) light-irradiated 
6795: sample displaying photoinduced
6796: ferromagnetism--direction of light irradiation is
6797: shown by an arrow; (b) band-edge profile of (In,Mn)As/GaSb heterostructure.
6798: $E_c$, conduction band, $E_v$, valence band; $E_F$,  
6799: Fermi level, respectively. From \onlinecite{Koshihara1997:PRL}.} 
6800: \label{light:1}
6801: \end{figure}
6802: 
6803: A high $T_C$ 
6804: and almost complete spin polarization in 
6805: bulk samples are alone not sufficient for successful applications. Spintronic 
6806: devices typically rely on inhomogeneous doping, structures of reduced 
6807: dimensionality, 
6808: and/or structures containing different materials. 
6809: Interfacial properties, as discussed in the
6810: previous sections, can significantly influence the magnitude of 
6811: magnetoresistive effects\footnote{In magnetic multilayers GMR is typically
6812: dominated by interfacial scattering \cite{Parkin1993:PRL}, while in MTJ's 
6813: it is the surface rather than the bulk electronic structure which 
6814: influences the relevant spin polarization.}
6815: and the efficiency of spin injection.
6816: Doping properties and possibility of fabricating a wide range of structures
6817: allow spintronic applications beyond MR effects, for example,
6818: spin transistors, spin lasers, and spin-based quantum computers 
6819: (Sec.~\ref{sec:IVF}).
6820: Materials properties of hybrid F/Sm heterostructures, relevant to
6821: device applications, were reviewed by \textcite{Samarth2003:SSC}.
6822: 
6823: Experiments in which the ferromagnetism is induced 
6824: optically 
6825: \cite{Koshihara1997:PRL,Oiwa2002:PRL,Wang2003:P} and electrically 
6826: \cite{Ohno2000:N,Park2002:S} provide a method for distinguishing the 
6827: carrier-induced
6828: ferromagnetism, based on the exchange interaction between the 
6829: carrier and the magnetic impurity spins,
6830: %(based on the exchange interaction $J {\bf S} \cdot {\bf s}_i$ between
6831: %the carrier spins ${\bf S}$ and the localized magnetic spins ${\bf s}_i$)
6832: from ferromagnetism that originates from magnetic nanoclusters. 
6833: Such experiments also suggest a possible nonvolatile multifunctional
6834: devices with tunable, optical, electrical, and magnetic properties. 
6835: Comprehensive surveys of magneto-optical materials and applications, not
6836: limited to semiconductors, are given by \textcite{Zvezdin:2003,Sugamo:2000}.
6837: 
6838: Photoinduced ferromagnetism was demonstrated by \textcite{Koshihara1997:PRL} 
6839: in p-(In,Mn)As/GaSb 
6840: heterostructure, shown in Figs.~\ref{light:1}, and \ref{light:2}. 
6841: Unpolarized light penetrates through a thin (In,Mn)As layer and is absorbed 
6842: in the GaSb layer. A large band bending across the heterostructures separates,
6843: by a built-in field, electrons and holes. The excess holes generated in a 
6844: GaSb layer are effectively transferred to the p-doped (In,Mn)As layer where 
6845: they enhance the ferromagnetic spin exchange among Mn ions, resulting
6846: in a paramagnetic-ferromagnetic transition.
6847: 
6848: 
6849: \begin{figure}
6850: \centerline{\psfig{file=zutic_fig26.eps,width=0.85\linewidth,angle=0}}
6851: \caption{Magnetization curves for (In,Mn)As/GaSb at 5 K observed before 
6852: (open circles)
6853: and after (solid circles) light irradiation. Solid line show a theoretical 
6854: curve. (b) Hall resistivity at 5 K before (dashed lines) 
6855: and after (solid lines) 
6856: light irradiation. From \onlinecite{Koshihara1997:PRL}.} 
6857: \label{light:2}
6858: \end{figure}
6859: The increase in magnetization, measured by a SQUID, 
6860: is shown in Fig.~\ref{light:2}(a) and in
6861: Fig.~\ref{light:2}(b) the corresponding Hall resistivity 
6862: \begin{equation}
6863: \rho_{\rm Hall}=R_0 B+ R_S M,
6864: \label{eq:hall}
6865: \end{equation}
6866: is shown, where the $R_0$ is the ordinary and $R_S$ 
6867: the anomalous Hall coefficient, 
6868: respectively.
6869: Typical for (III,Mn)V compounds, $\rho_{\rm Hall}$ is dominated by the 
6870: anomalous
6871: contribution, $\rho_{\rm Hall} \propto M$. 
6872: 
6873: 
6874: \begin{figure}
6875: \centerline{\psfig{file=zutic_fig27.eps,width=1.0\linewidth}}
6876: \caption{Electric-field control of ferromagnetism.
6877: R$_{\rm Hall}$ vs field curves under three different gate biases.
6878: Application of V$_G$=0, +125, and -125 V results in a qualitatively different 
6879: field
6880: dependence of R$_{\rm Hall}$ measured at 22.5 K (sample B):
6881: almost horizontal dash-dotted line, paramagnetic response 
6882: are
6883: partially depleted from the channel (V$_G$=+125 V);
6884: dashed lines, clear hysteresis at low fields
6885: ($<$0.7 mT) as holes are accumulated in the channel (V$_G$=-125 V);
6886: solid line, R$_{\rm Hall}$ curve measured at V$_G$=0 V before 
6887: application of $\pm$ 125 V, dotted line, 
6888: R$_{\rm Hall}$ after 
6889: application of $\pm$ 125 V.
6890: Inset, the same curves shown at higher magnetic fields.
6891: From \onlinecite{Ohno2000:N}.} 
6892: \label{ohno}
6893: \end{figure}
6894: 
6895: 
6896: A different type of photoinduced magnetization was measured in ferromagnetic 
6897: (Ga,Mn)As.\footnote{Previous studies in paramagnetic (II,Mn)VI materials 
6898: have shown that
6899: nonequilibrium spin-polarized carriers can change the orientation of
6900: magnetic spins in (Hg,Mn)Te \cite{Krenn1985:PRL,Krenn1989:PRB} and in 
6901: (Cd,Mn)Te
6902: \cite{Awschalom1987:PRL}.}
6903: In a Faraday geometry (recall \ref{sec:IID3}), by changing the polarization
6904: of a circularly polarized light, 
6905: one can modulate
6906: the Hall resistance and thus the induced 
6907: magnetization
6908: by up to 15\% of the saturation value \cite{Oiwa2002:PRL}.
6909: Additional experiments on photoinduced magnetization rotation 
6910: \cite{Munekata2003:JS,Oiwa2003:JS} suggest that the 
6911: main contribution of carrier spin to such rotation is realized by generating
6912: an effective magnetic field through the $p$-$d$ exchange interaction, 
6913: rather than by spin-transfer torque as discussed in 
6914: Secs.~\ref{sec:IB1} and \ref{sec:V}
6915: \cite{Moriya2003:P}.
6916: In GaAs-Fe composite films an observation of room temperature photoenhanced 
6917: magnetization was used to demonstrate that a magnetic force can be changed
6918: by light illumination \cite{Shishi2003:APL}.
6919: 
6920: Electrically induced ferromagnetism was realized  by applying gate
6921: voltage V$_G$ to change the hole concentration in $d$=5nm thick (In,Mn)As
6922: used as a magnetic channel in a metal-insulator semiconductor FET structure.
6923: Below a metal gate and an insulator the (In,Mn)As channel was grown on top
6924: of a InAs/(Al,Ga)Sb/AlSb and GaAs substrate. 
6925: In Fig.~\ref{ohno},
6926: the corresponding data for $R_{\rm Hall}=\rho_{\rm Hall}/d \propto M$ 
6927: [recall Eq.~(\ref{eq:hall}), show that the ferromagnetism can be switched on
6928: and off, as an electric analog of the manipulation of M 
6929: from Fig.~\ref{light:2}.
6930: Subsequent work by \textcite{Park2002:S} showed that in MnGe ferromagnetism 
6931: can
6932: be manipulated at higher temperature and at significantly lower gate voltage
6933: (at $\sim$ 50 K and $\sim$ 1 V). The combination of light and electric-field
6934: control of ferromagnetism was used in modulation-doped p-type
6935: (Cd,Mn)Te QW \cite{Boukari2002:PRL}. It was demonstrated that 
6936: illumination by light in $p-i-n$ diodes would  enhance the spontaneous
6937: magnetization, while illumination in $p-i-p$ structures would 
6938: destroy ferromagnetism.
6939: 
6940: 
6941: \begin{figure}
6942: \centerline{\psfig{file=zutic_fig28.eps,width=1.0\linewidth}}
6943: \caption{Voltage-controlled spin precession: (a) time-resolved Kerr rotation
6944: measurements of electron spin precession in a quantum well at different
6945: gate voltages $V_G$ with Al concentration of 7\% at 5 K and B=6 T;
6946: displacement of the electron wave function
6947: towards the back gate into regions with more Al concentration
6948: as a positive
6949: voltage $V_G$ is applied between back and front gate;
6950: leading to an increase of $g$. 
6951: At $V_G$=2 V, no precession is
6952: observed, corresponding to $g$=0. From \onlinecite{Salis2001:N}.} 
6953: \label{mater:1}
6954: \end{figure}
6955: 
6956: 
6957: In semiconductors $g$ factors, which determine the 
6958: spin splitting of carrier bands (and consequently influence the spin dynamics 
6959: and spin resonance), can be very different from the free-electron value. 
6960: With strong spin-orbit coupling in narrow-band III-V's they 
6961: are $\approx$-50 
6962: for InSb and $\approx$-15 for InAs, while, 
6963: as discussed in Sec.~\ref{sec:IID3} the doping 
6964: with magnetic impurities can give even $|g^*|\sim 500$. Manipulation
6965: of the $g$ factor in a GaAs/AlGaAs quantum well (QW)  
6966: in Fig.~\ref{mater:1}, relies on the results for a bulk Al$_x$Ga$_{1-x}$As, 
6967: the variation of Al concentration changes the $g$ factor 
6968: \cite{Chadi1976:PRB,Weisbuch1977:PRB} 
6969: $g$=-0.44 for x=0 and $g$=0.40 for x=0.3. Related experiments on 
6970: modulation-doped 
6971: GaAs/Al$_{0.3}$Ga$_{0.7}$As have shown that by applying V$_G$ 
6972: on can shift the electron
6973: wave function in the QW and produce $\sim$1\% change
6974: in the $g$ factor \cite{Jiang2001:PRB}. Subsequently, 
6975: in an optimized Al$_x$Ga$_{1-x}$As quantum well, where
6976: $x$ varied gradually across the structure, much larger changes were
6977: measured--when $V_G$ is changed, the electron wave function  
6978: efficiently samples 
6979: different regions with different $g$ factors \cite{Salis2001:N}. 
6980: Figure~\ref{mater:1}(a) gives
6981: the time-resolved Kerr rotation data (the technique is discussed 
6982: in Sec.~\ref{sec:III}) 
6983: data which can be described as 
6984: $\propto \exp(-\Delta t/T_2^*)\cos(\Omega \Delta t)$,
6985: where $\Delta t$ is the delay time between the circularly polarized pump
6986: and linearly polarized probe pulses, T$_2^*$ is the
6987: transverse electron spin lifetime with inhomogeneous broadening, and 
6988: the angular precession frequency $\Omega=\mu_B g  B/\hbar$ can be used to 
6989: determine the $g$ factor. 
6990: It is also possible to manipulate $g$ factors dynamically using 
6991: time-dependent $V_G$ \cite{Kato2003:S}. The anisotropy of $g$ factor
6992: ($g$ tensor) allows voltage control of both the magnitude and the
6993: direction of the spin precession vector ${\bf \Omega}$.
6994: 
6995: \subsection{\label{sec:IVC} Spin filters}
6996: %-----------------------------------------------------------------------
6997: 
6998: Solid state spin filtering (recall the similarity with
6999: spin injection from Sec~\ref{sec:IIC1}) was first realized in N/F/N
7000: tunneling. 
7001: It was shown by \textcite{Esaki1967:PRL} that the magnetic tunneling through
7002: (ferro)magnetic semiconductor Eu chalcogenides 
7003: \cite{Nagaev:1983,Kasuya1968:RMP,vonMolnar1967:JAP}, 
7004: such as EuSe\footnote{At zero magnetic field
7005: EuSe is an antiferromagnet, 
7006: and at moderate fields it becomes a ferromagnet with T$_C$$\approx$ 5K.} 
7007: and EuS,\footnote{At zero magnetic field, exchange
7008: splitting of a conduction band in bulk EuS is $\approx$ 0.36 eV
7009:  \cite{Hao1990:PRB}.}
7010: could be modified by an applied magnetic field. The change in $I-V$ curves 
7011: in the
7012: N/F/N structure, where N is a normal metal and F is a ferromagnet, was 
7013: explained by the influence of the magnetic field on the height of the barrier 
7014: formed at the N/F interface (for EuSe, the barrier height was lowered by
7015: 25\% at 2 T). 
7016: The large spin splitting of the Eu chalcogenides was subsequently 
7017: employed in the absence of
7018: applied field with EuS \cite{Moodera1988:PRL} and nearly 100\%
7019: spin polarization was reached  
7020: at $B$=1.2 T with EuSe \cite{Moodera1993:PRL}. 
7021: These spin-filtering properties of the 
7022: Eu chalcogenides, used together with one-electron quantum dots, were proposed 
7023: as the basis for a method to convert 
7024: single spin into single charge 
7025: measurements\footnote{This method could already be 
7026: realized using single-electron transistors or 
7027: quantum point contacts.} and provide an important ingredient in realizing a 
7028: quantum computer \cite{DiVincenzo1999:JAP}, see Sec.~\ref{sec:IVF}.
7029: 
7030: Zeeman splitting in semiconductor heterostructures and superlattices 
7031: (enhanced by large $g$ factors) \cite{Egues1998:PRL,Guo2001:PRB},
7032: in quantum dots \cite{Recher2000:PRL,Deshmukh2002:PRL,Borda2003:PRL}, 
7033: and nanocrystals  
7034: \cite{Efros2001:PRL} provide effective spin filtering and spin-polarized
7035: currents. Predicted quantum size effects and resonance tunneling 
7036: \cite{Duke:1969}   %iz p.79 Duke
7037: also have their spin-dependent counterparts. The structures studied
7038: are typically double-barrier resonant tunneling diodes 
7039: (for an early
7040: spin-unpolarized study see \textcite{Tsu1973:APL}), with either 
7041: Zeeman splitting or using ferromagnetic materials, in which
7042: spin filtering can be tuned by an applied bias.\footnote{See, for example, 
7043: \cite{Brehmer1995:APL,Mendez1986:PRB,Giazotto2003:APL,Ohno1998:S,% 
7044: Petukhov2002:PRL,Ting2002:APL,Slobodskyy2003:PRL,petukhov1998:ASS,%
7045: Aleiner1991:JETPL, Vurgaftman2003:PRB}.} 
7046: 
7047: Several other realizations of spin filtering have been investigated,
7048: relying on  spin-orbit coupling.\footnote{These include the work of 
7049: \cite{Kiselev2001:APL,Governale2002:PRB,Voskoboynikov1998:PRB,%
7050: Koga2002b:PRL,Silva1999:PRB,Voskoboynikov1999:PRB,Perel2003:PRB}.} 
7051: or hot-electron transport across ferromagnetic regions,\footnote{See
7052: \cite{vanDijken2002:PRB,Monsma1995:PRL,Rippard2000:PRL,Upadhyay1999:APL,%
7053: Oberli1998:PRL,Cacho2002:PRL,Filipe1998:PRL}.} discussed in more detail 
7054: in Sec.~\ref{sec:IVE3}.
7055: A choice of particular atomically ordered F/Sm interfaces was suggested
7056: to give a strong spin-filtering effect \cite{Kircenzow2001:PRB},
7057: limited by the spin-orbit coupling and interfacial spin-flip scattering.
7058: 
7059: \begin{figure} 
7060: \centerline{\psfig{file=zutic_fig29.eps,width=1.0\linewidth}}
7061: \caption{Mesoscopic spin filter: 
7062: (a) micrograph and circuit showing the
7063: polarizer-analyzer configuration used in the experiment
7064: of \textcite{Folk2003:S}. The emitter (E)
7065: can be formed into either a quantum dot or a quantum point contact (QPC). 
7066: The collector
7067: (C), is a single point contact. Electrons are focused from E to C through
7068: the base region (B), using a small perpendicular magnetic field. Gates
7069: marked with ``x'' are left undepleted 
7070: when E is operated as a QPC;
7071: (b) base-collector voltage (V$_C$) showing two focusing peaks;
7072: (c) focusing peak height at B$_\|$=6 T with spin-selective collector
7073: QPC conductance ($g_C$=$0.5e^2/h$), comparing E as QPC at $2e^2/h$ (dashed
7074: curve) and E as a quantum dot with both leads at $2e^2/h$ (solid curve). 
7075: Adapted from \onlinecite{Folk2003:S}.} 
7076: \label{filter:1}
7077: \end{figure}
7078: 
7079: Mesoscopic spin filters have also been suggested 
7080: \cite{Frustaglia2001:PRL,Joshi2001:PRB,Ionicioiu2003:PRB,Avdonin2003:P}
7081: and we discuss here a particular realization.
7082: In an applied magnetic field 
7083: two quantum point contacts (QPC), an emitter (E)
7084: and a collector (C), fabricated on top of a high-mobility 
7085: 2DEG in GaAs/AlGaAs,
7086: can act as spin polarizer and analyzer 
7087: \cite{Potok2002:PRL}.\footnote{QPC was also used 
7088: to locally create and probe nonequilibrium nuclear spin
7089: in GaAs/AlGaAs heterostructures 
7090: in the quantum Hall regime \cite{Wald1994:PRL}.} 
7091: In a ballistic regime and at $T=70$ mK 
7092: (mean free path $\approx$ 45 $\mu$m $>>$ sample size $\approx$ 1.5 $\mu$m)
7093: magnetic focusing\footnote{Suggested by \textcite{Sharvin1965:ZETF}
7094: as a technique to study Fermi surfaces; see also \textcite{vanHouten1989:PRB}.} 
7095: with $B_\bot$ results in base-collector voltage peaks, when the separation
7096: of the two QPCs is an even multiple of the cyclotron radius $m^*v_F/eB_\bot$,
7097: where $m^*$ is the effective mass, and v$_F$ the Fermi velocity.
7098: These results are illustrated in Figs.~\ref{filter:1}(a) and (b) 
7099: on a slightly modified structure \cite{Folk2003:S}
7100: where, by applying a gate voltage, one can cause the emitter to form either 
7101: a quantum dot or QPC.  
7102: Effective spin filtering, due to the large in-plane field $B_\|$, can 
7103: be tuned by the gate voltage which changes the conductance of QPC.
7104: The resulting effect of spin filtering modifies the collector voltage
7105: $V_C$  \cite{Potok2002:PRL}, 
7106: \begin{equation}
7107: V_C=\alpha(h/2e^2)I_E(1+P_{I_E}P_{T_C}),
7108: \label{eq:filter}
7109: \end{equation}
7110: where $0< \alpha <1$ parameterizes the imperfections in focusing,
7111: $P_{I_E}$ and $P_{T_C}$ are the spin polarization [recall Eq.~(\ref{eq:polar})] 
7112: of the emitter current $I_E$, and the collector transmission coefficient $T_C$
7113: is related to the collector conductance by $g_C=(2e^2/h)T_C$. 
7114: In Eq.~(\ref{eq:filter}) we note a recurring form for a spin-valve
7115: effect. The measured signal involves the product of two different spin 
7116: polarizations, for example, similar to TMR in Eq.~(\ref{eq:julliere}) 
7117: or to spin-charge coupling due to nonequilibrium spin 
7118: [recall Eqs.~(\ref{eq:DelR}) and (\ref{eq:md})].
7119: Another mesoscopic spin filter with few-electron quantum dot 
7120: (GaAs/AlGaAs-based) was used to demonstrate a nearly complete spin
7121: polarization which could be reversed by adjusting gate voltages 
7122: \cite{Hanson2003:P}.
7123: 
7124: \subsection{\label{sec:IVD} Spin diodes}
7125: %--------------------------------------------------------------------------
7126: 
7127: Spin diodes are inhomogeneous two-terminal devices whose electronic or optical 
7128: properties depend on the spin polarization of the carriers. Such devices
7129: were envisaged long before the emergence of 
7130: spintronics. \textcite{Solomon1976:SSC},
7131: for example, proposed and demonstrated a silicon {\it p-n} junction whose 
7132: current was modified by changing the spin polarization of the recombination
7133: centers. In a magnetic field both the mobile carriers and the recombination
7134: centers have an equilibrium spin polarization due to the Zeeman splitting.
7135: The current in a {\it p-n} junction depends on the recombination rate,
7136: which, in turn, depends on the relative orientation of the spin 
7137: of the carriers and the centers \cite{Lepine1972:PRB}. The trick to 
7138: modifying the current is to decrease (even saturate) the spin polarization
7139: of either the electrons or the centers by electron spin
7140: resonance. Indeed, \textcite{Solomon1976:SSC} found a variation of
7141: $\approx 0.01\%$ of the saturation current at small biases where recombination
7142: in the space-charge region dominates. Similar experiments could be used to 
7143: detect nonequilibrium spin due to (potential) spin injection in Si,
7144: where optical methods are ineffective, but also in other semiconductors
7145: where electrical detection would be desirable.\footnote{Spin diodes can 
7146: also probe fundamental properties of electronic
7147: systems. The diode demonstrated by \textcite{Kane1992:PRB} is 
7148: based on a junction
7149: between two coplanar AlGaAs/GaAs 2DEG's, one with $\nu<1$ and 
7150: the other with $\nu>1$, 
7151: where $\nu$ is the Landau-level filling; that is, the two regions have
7152: opposite spins at the Fermi level. The current crossing such a junction,
7153: which has a diode property due to the existence of a built-in field in
7154: the contact, is accompanied by a spin flip. Interestingly, the
7155: current is also time dependent, due to the current-induced dynamic nuclear
7156: polarization.} 
7157: 
7158: Several spin diodes have recently been proposed or demonstrated 
7159: with the goal of either maximizing the sensitivity of the $I-V$ characteristics
7160: to spin and magnetic field, or to facilitating spin injection and its
7161: detection through semiconductor interfaces comprising a magnetic semiconductor
7162: as the injector. 
7163: Magnetic tunneling diodes have been used for spin injection 
7164: from a ferromagnetic to a nonmagnetic semiconductor, in p-GaMnAs/n-GaAs
7165: {\it p-n} 
7166: junctions~\cite{Kohda2001:JJAP,Johnston-Halperin2002:PRB,vanDorpe2003:P}.
7167: As discussed in Sec.~\ref{sec:IID3},
7168: {\it p-n} 
7169: heterostructures have combined Cr- or Eu-based ferromagnetic
7170: semiconductors and InSb  \cite{Osipov1998:PL,Viglin1997:PLDS}.
7171: Spin light-emitting diodes (recall Figs.~\ref{injexp:3}
7172: and \ref{injexp:4}) 
7173: were employed for injecting and detecting spins 
7174: in semiconductors,
7175: while resonant tunneling diodes have been demonstrated as effective spin 
7176: injectors (Sec.~\ref{sec:IID3}) and spin filters (Sec.~\ref{sec:IVC}).
7177: A magnetic unipolar diode has been proposed by
7178: \textcite{Flatte2001:APL}
7179: to simulate the working of ordinary diodes, but with 
7180: homogeneous monopolar
7181: doping (either donors or acceptors, not both). The role of 
7182: inhomogeneous doping in the {\it p-n} junction is played by the inhomogeneous 
7183: spin splitting
7184: of the carrier band, with the spin up and spin down carriers 
7185: playing roles similar
7186: to those of the electrons and holes in bipolar diodes. 
7187: Si-based {\it p-i-n} diode sandwiched between two ferromagnetic
7188: metals was suggested to allow controlling the device performance
7189: by an externally applied magnetic field \cite{Dugaev2003:PE}.
7190: Finally, 
7191: \textcite{Zutic2002:PRL} 
7192: have proposed the magnetic bipolar diode described below.
7193: 
7194: 
7195: The magnetic bipolar diode\footnote{Not to be confused with the usual 
7196: magnetic diodes
7197: which are ordinary diodes in a magnetic field. The $I-V$ characteristics of
7198: such diodes, depend on the magnetic field through small orbital effects 
7199: on diffusion coefficient, 
7200: not through the spin effects described here.} (MBD) is a {\it p-n} 
7201: junction diode with one or both regions magnetic 
7202: \cite{Zutic2002:PRL,Fabian2002:PRB}.
7203: The MBD is the prototypical device of bipolar
7204: spintronics, a subfield of spintronics in which both electrons and holes 
7205: take part 
7206: in carrier transport, while either electrons or holes (or both) are spin 
7207: polarized (see Sec.~\ref{sec:IVA4}). 
7208: Examples of nonmagnetic bipolar spintronic devices are the 
7209: spin-polarized {\it p-n} 
7210: junction \cite{Zutic2001:PRB} and the spin solar cell \cite{Zutic2001:APL}. 
7211: These devices offer 
7212: opportunities for effective spin injection, 
7213: spin amplification (see Sec.~\ref{sec:IIC3}), 
7214: or spin capacity---the effect of 
7215: changing, by voltage, nonequilibrium spin density \cite{Zutic2001:PRB}. 
7216: The advantages of 
7217: magnetic bipolar spintronic devices 
7218: \cite{Zutic2002:PRL,Fabian2002:PRB,Zutic2003:APL,Fabian2002:P}
7219: lie in the combination of equilibrium 
7220: magnetism and nonequilibrium spin and effective methods to manipulate
7221: a minority carrier population.
7222: The most useful effects of the 
7223: spin-charge coupling in MBD's are the 
7224: spin-voltaic
7225: and the giant-magnetoresistive
7226: effects, which are enhanced over those of metallic systems by the 
7227: exponential dependence of the current on bias voltage.
7228: 
7229: \begin{figure}
7230: \centerline{\psfig{file=zutic_fig30.eps,width=1\linewidth,angle=0}}
7231: \caption{Scheme of a magnetic bipolar diode. 
7232: The $p$-region (left) is magnetic, indicated by the spin
7233: splitting $2q\zeta$ of the
7234: conduction band. The $n$-region (right) is nonmagnetic, but spin 
7235: polarized by a spin source: Filled circles, spin-polarized electrons;
7236: empty circles, unpolarized holes.
7237: If the nonequilibrium spin in the 
7238: $n$-region is oriented parallel (top figure) 
7239: to the equilibrium spin in the 
7240: $p$-region, large forward current flows. 
7241: If the relative orientation is antiparallel (bottom), the current drops
7242: significantly. Adapted from \onlinecite{Zutic2002:PRL}. 
7243: }
7244: \label{fig:md}
7245: \end{figure}
7246: 
7247: A scheme of an MBD is shown in Fig.~\ref{fig:md} 
7248: (also see Fig.~\ref{deplete}). 
7249: The $p$-region is magnetic, by 
7250: which we mean that it has a spin-split conduction band with the spin
7251: splitting (Zeeman or exchange) 
7252: $2q\zeta$ $\sim k_BT$. 
7253: Zeeman splitting can be significantly enhanced by 
7254: the large $g^*$ factors of magnetically doped (Sec.~\ref{sec:IIC3})
7255: or narrow-band-gap semiconductors (Sec.~\ref{sec:IVB}).
7256: Using an MBD with a ferromagnetic semiconductor slightly above its $T_C$ 
7257: is also expected to give large $g^*$ factors.
7258: The $n$-region 
7259: is nonmagnetic, but 
7260: electrons can be 
7261: spin-polarized 
7262: by a spin source (circularly polarized light or 
7263: magnetic electrode). The interplay between the equilibrium spin of polarization 
7264: $P_{n0}=\tanh(q\zeta/k_B T)$ 
7265: in the $p$-region, and the nonequilibrium 
7266: spin source of polarization $\delta P_n$ 
7267: in the $n$-region, at the edge of the depletion layer, 
7268: determines the $I-V$  characteristics of the diodes. 
7269: It is straightforward to generalize these considerations to include 
7270: the spin-polarized holes \cite{Fabian2002:PRB}.
7271: 
7272: The dependence of the electric current $j$ on $q\zeta$ and $\delta P_n$ was 
7273: obtained by both numerical and analytical methods. 
7274: Numerical calculations \cite{Zutic2002:PRL} 
7275: were performed by self-consistently solving for the drift-diffusion, 
7276: continuity, as well as carrier recombination and 
7277: spin-relaxation equations, 
7278: discussed in  
7279: Sec.~\ref{sec:IVA4}. While the numerical calculations are indispensable in the
7280: high-injection limit,\footnote{The small bias or low-injection limit 
7281: is the regime of applied bias in which the density of the carriers injected
7282: through the depletion layer (the minority carriers) is much smaller than 
7283: the equilibrium density of the majority carriers. 
7284: Here and in Sec.~\ref{sec:IVE2} the terms majority and
7285: minority refer to the relative carrier (electron or hole) population and
7286: not to spin. The large bias or high-injection limit is the regime where 
7287: the injected carrier density
7288: becomes comparable to the equilibrium density. This occurs
7289: at forward biases comparable to the built-in potential, typically  1 V.}
7290: valuable insight and  
7291: analytical formulas can be obtained in the
7292: low-injection limit, where the Shockley theory \cite{Shockley:1950} for 
7293: ordinary {\it p-n}
7294: junctions was generalized by \textcite{Fabian2002:PRB} for the magnetic
7295: case.  
7296: 
7297: To illustrate the $I-V$ characteristics of MBD's, consider the low-injection 
7298: limit in the configuration of Fig.~\ref{fig:md}. 
7299: The electron contribution to the total electric current is
7300: \cite{Fabian2002:PRB}
7301: \begin{equation}\label{eq:md}
7302: j_n \sim n_{0}(\zeta) 
7303: \left [e^{qV/k_B T} \left (1+\delta P_n P_{n0}\right ) -1 \right ],
7304: \end{equation}
7305: where $V$ is the applied bias (positive for forward bias) 
7306: and 
7307: $n_{0}(\zeta)=(n_i^2/N_a)\cosh(q\zeta/k_B T)$ 
7308: is the equilibrium number of electrons 
7309: in the $p$-region, 
7310: dependent on the splitting, the intrinsic carrier 
7311: density $n_i$, and the acceptor doping $N_a$. Equation (\ref{eq:md}) 
7312: generalizes the 
7313: Silsbee-Johnson spin-charge coupling \cite{Silsbee1980:BMR,Johnson1985:PRL}, 
7314: originally proposed for ferromagnet/paramagnet metal interfaces, to 
7315: the case of magnetic {\it p-n} junctions.  The advantage
7316: of the spin-charge coupling in {\it p-n} junctions, as opposed to metals or 
7317: degenerate systems, is the nonlinear voltage dependence 
7318: of the nonequilibrium carrier and spin densities \cite{Fabian2002:PRB}, 
7319: allowing for 
7320: the exponential enhancement of the effect with increasing $V$. 
7321: Equation (\ref{eq:md}) can be understood qualitatively
7322: from Fig.~\ref{fig:md} \cite{Fabian2002:PRB}. 
7323: In equilibrium, $\delta P_n=0$ 
7324: and $V=0$, no current flows through the
7325: depletion layer, as the electron currents from both sides of the junction 
7326: balance out. The balance is disturbed either by applying bias or 
7327: by selectively populating different spin states,
7328: making the flow of one spin species greater than that of the other. 
7329: In the latter case, 
7330: the effective barriers for crossing
7331: of electrons from the $n$ to the $p$ side is different for spin
7332: up and down electrons (see Fig.~\ref{fig:md}).
7333: Current can flow even at $V=0$ when $\delta P_n\ne 0$. 
7334: This is an example of the spin-voltaic effect (a spin analog
7335: of the photo-voltaic effect), in which  nonequilibrium
7336: spin causes an EMF \cite{Zutic2002:PRL,Zutic2003:P}.
7337: In addition, the direction of the zero-bias current 
7338: is controlled by the relative sign of 
7339: $P_{n0}$ and $\delta P_n$. 
7340: 
7341: \begin{figure}
7342: \centerline{\psfig{file=zutic_fig31.eps,width=1\linewidth,angle=0}}
7343: \caption{Giant magnetoresistance (GMR) effect in magnetic diodes. 
7344: Current/spin-splitting characteristics ($I$-$\zeta$) 
7345: are calculated self-consistently 
7346: at $V$=0.8 V
7347: for the diode from  Fig.~\ref{fig:md}. 
7348: Spin splitting $2q\zeta$ on the p-side is normalized to $k_B T$. 
7349: The solid curve corresponds to a switched-off spin source. The current is
7350: symmetric in $\zeta$. With spin source on (the extreme case of 100\% spin 
7351: polarization injected into the 
7352: $n$-region 
7353: is shown), the current is a strongly asymmetric 
7354: function of $\zeta$, displaying large GMR, shown by the dashed curve. 
7355: Materials parameters of GaAs were applied.
7356: Adapted from \onlinecite{Zutic2002:PRL}.}
7357: \label{fig:gmrd}
7358: \end{figure}
7359: 
7360: MBD's can  display an interesting  GMR-like effect, which 
7361: follows from Eq.~(\ref{eq:md}) \cite{Zutic2002:PRL}. The 
7362: current depends strongly on the relative orientation of the 
7363: nonequilibrium spin and the equilibrium magnetization. 
7364: Figure \ref{fig:gmrd} plots $j$,
7365: which also includes the contribution from holes, as a function of 
7366: $2q\zeta/k_B T$ for 
7367: both the unpolarized, $\delta P_n=0$, and fully polarized, $\delta P_n=1$, 
7368: $n$-region. 
7369: In the first case $j$ is
7370: a symmetric function of $\zeta$, increasing exponentially with increasing 
7371: $\zeta$ due to the increase in the 
7372: equilibrium minority carrier density $n_0(\zeta)$. 
7373: In unipolar systems, where transport is due to the majority carriers, 
7374: such a modulation of the current is not likely, as the majority carrier
7375: density is fixed by the density of dopants. 
7376: 
7377: If $\delta P_n \ne 0$, the current will depend on the sign of 
7378: $P_{n0}\cdot \delta 
7379: P_n$.  For parallel nonequilibrium (in the $n$-region)
7380: and equilibrium spins (in the $p$-region), most electrons cross 
7381: the depletion layer through the 
7382: lower barrier (see Fig.~\ref{fig:md}), increasing the current. 
7383: In the opposite case of antiparallel relative orientation, 
7384: electrons experience a larger barrier and the current is inhibited. 
7385: This is demonstrated in Fig.~\ref{fig:gmrd}
7386: by the strong asymmetry in $j$. The corresponding GMR ratio, the 
7387: difference between  $j$ for parallel and antiparallel orientations,
7388: can also be calculated analytically from Eq.~(\ref{eq:md}) as
7389: $2|\delta P_n P_{n0}|/(1-|\delta P_n P_{n0}|)$ \cite{Fabian2002:PRB}. 
7390: If, for example, $|P_{n0}|=|\delta 
7391: P_n|=0.5$, the relative change is 66\%. 
7392: The GMR effect should be useful for measuring
7393: the spin relaxation rate of bulk semiconductors \cite{Zutic2003:APL},
7394: as well as for detecting nonequilibrium spin in the nonmagnetic
7395: region of the {\it p-n} junction.\footnote{This could be a
7396: way to detect spin injection into Si, where optical detection
7397: is ineffective.}
7398: 
7399: Although practical MBD's are still to be fabricated and the predicted effects
7400: tested, magnetic {\it p-n} junctions have already been demonstrated. 
7401: Indeed, \textcite{Wen1968:IEEETM}\footnote{We thank M. Field for bringing this
7402: reference to our attention.} were perhaps
7403: the first to show that a ferromagnetic {\it p-n} junction, based on the
7404: ferromagnetic semiconductor CdCr$_2$Se$_4$ doped with Ag acceptors and In 
7405: donors, could act as a diode. Heavily doped p-GaMnAs/n-GaAs junctions
7406: were fabricated 
7407: \cite{Ohno2000:ASS,Kohda2001:JJAP,Johnston-Halperin2002:PRB,Arata2001:PE,%
7408: vanDorpe2003:P} 
7409: to demonstrate tunneling interband spin injection. 
7410: Incorporation of (Ga,Mn)As layer in the intrinsic region of {\it p-i-n}
7411: GaAs diode was shown to lead to an efficient photodiode, in which the
7412: Mn ions function as recombination centers \cite{Teran2003:APL}.
7413: It would be interesting to see such devices combined with a spin
7414: injector in the bulk regions.
7415: Recently, 
7416: \textcite{Tsui2003:APL}
7417: have shown that the current in p-CoMnGe/n-Ge magnetic heterojunction diodes 
7418: can indeed be controlled by magnetic field.
7419: To have functioning MBD's at room temperature, and to
7420: observe the above predicted phenomena, several important challenges
7421: have to be met:
7422: 
7423: (i) Zeeman or exchange splitting needs to sufficiently large to 
7424: provide equilibrium spin polarization,
7425: $\gtrsim 1-10\%$. This may be difficult at room temperature, unless
7426: the effective $g$ factor is $\sim 100$ at $B \sim 1$ T 
7427: (Sec.~\ref{sec:IID3}). The use of ferromagnetic semiconductors
7428: is limited by their $T_C$ (Sec.~\ref{sec:IVB}).
7429: 
7430: (ii) For a strong spin-charge coupling [recall the discussion of 
7431: Eq.~(\ref{eq:md})] a nondegenerate carrier density
7432: is desirable, which, while 
7433: likely in (Zn,Cr)Te, is not easily realized in many other ferromagnetic
7434: semiconductors that are typically heavily doped (Sec.~\ref{sec:IVB}).
7435: 
7436: (iii) An effective integration of magnetic and nonmagnetic structures 
7437: into single devices \cite{Samarth2003:SSC} is needed.
7438: 
7439: (iv) The samples need to be smaller than the spin diffusion lengths,
7440: requiring high carrier mobility and long spin relaxation (easier to 
7441: realize for spin-polarized electrons).
7442: 
7443: (v) The effects of actual device structures, such as two- and/or 
7444: three-dimensional spin flow, interface contacts, spin-dependent band offsets 
7445: and band bendings, strong spin relaxation in the depletion layers, etc., 
7446: will need to be understood.
7447: 
7448: By combining two magnetic {\it p-n} junctions in series one can
7449: obtain a magnetic bipolar transistor (Sec.~\ref{sec:IVE2}), a 
7450: three terminal device which offers spin-dependent amplification.
7451: 
7452: \subsection{\label{sec:IVE} Spin transistors}
7453: %---------------------------------------------------------------------------
7454: 
7455: We review several proposals for spin transistors 
7456: that have at least one semiconductor region and 
7457: that aim at integrating spin 
7458: and charge transport within traditional device schemes of the field-effect and 
7459: junction transistors. 
7460: Three important cases are discussed in detail: the Datta-Das  
7461: spin field-effect transistor, the magnetic bipolar transistor, and
7462: the hot-electron spin transistor. 
7463: 
7464: Various spin transistors that contain metallic (and insulating)
7465: regions have been proposed 
7466: \cite{Johnson1993:S,Bauer2003:APL,Zvezdin2003:TP,You2000:JAP}. 
7467: There is also a large category of
7468: the spin single-electron transistors, first realized by 
7469: \textcite{Ono1996:JPSJ}, and later investigated in 
7470: \cite{Barnas1998:PRL,Korotkov1999:PRB,Ciorga2002:APL,Martinek2002:PRB}. 
7471: Spin single-electron
7472: transistors can be viewed as an extension of magnetic tunneling 
7473: (see Sec.~\ref{sec:IVA2})
7474: to double tunnel junctions, where the Coulomb blockade becomes important 
7475: \cite{Takahashi1998:PRL}. For a review of spin single-electron transistors see 
7476: \cite{Maekawa:2002}. 
7477: 
7478: \subsubsection{\label{sec:IVE1} Spin field-effect transistors}
7479: %--------------------------------------------------------------------------------------------
7480: 
7481: \textcite{Datta1990:APL} proposed what became the prototypical spintronic 
7482: device scheme, 
7483: the Datta-Das spin field-effect transistor (SFET) (see Fig.~\ref{fig:DD}). 
7484: The device is based on spin injection and spin detection by a ferromagnetic 
7485: source and 
7486: drain, and on spin precession about the built-in structure inversion asymmetry 
7487: (Bychkov-Rashba) field ${\bf\Omega}$, Eq.~(\ref{eq:relax:BR}), 
7488: in the asymmetric, quasi-one-dimensional channel 
7489: of an ordinary field-effect transistor. The attractive feature of the
7490: Datta-Das SFET is that spin-dependent device operation is controlled
7491: not by external magnetic fields, but by gate bias, which controls the 
7492: spin precession rate. 
7493: 
7494: 
7495: The structure of the Datta-Das SFET is shown in Fig.~\ref{fig:DD}. 
7496: Consider a 2DEG  confined along the plane of the unit vector $\bf n$. 
7497: The precession axis of $\bf\Omega$
7498: lies always in the channel plane (see Sec.~\ref{sec:IIIB2}), so 
7499: the results 
7500: (unlike those for bulk inversion asymmetry) 
7501: are insensitive 
7502: to the relative orientation of ${\bf n}$ and the principal crystal axes. 
7503: Equation (\ref{eq:relax:BR}) determines the evolution of the 
7504: expectation value for a spin perpendicular to the plane, 
7505: $s_{n}={\bf s}\cdot {\bf n}$, and a spin parallel to the in-plane 
7506: $\bf k$, $s_{\parallel}={\bf s}\cdot{\bf k}/k$:  
7507: \begin{equation}
7508: ds_n/dt=2\alpha_{BR} k s_{\parallel}, \,\,\,\,
7509: ds_\parallel/dt=-2\alpha_{BR} k s_{n},
7510: \end{equation}
7511: where $\alpha_{BR}$ is the structure inversion asymmetry coefficient 
7512: appearing in 
7513: Eq.~(\ref{eq:relax:BR}).
7514: The average spin component along 
7515: $\bf\Omega$, $s_{\perp}={\bf s}\cdot ({\bf k} \times {\bf n})/k$, 
7516: is constant. 
7517: As a result, $s_\parallel=s_{0\parallel}\cos(\omega t)$, where 
7518: $\omega=2\alpha_{BR} k$ 
7519: and the injected spin at the source is labeled with zero. 
7520: If $\varphi$ is the angle between $\bf k$ and the source-drain axis,
7521: the electron will reach the drain at time $t'=L m_c/(\hbar k\cos\varphi)$, 
7522: with the spin 
7523: $s_\parallel$
7524: precessing at
7525: the angle $\phi=2\alpha_{BR} m_c L/\hbar$. The average spin at the drain 
7526: in the direction
7527: of magnetization is $s_{\parallel}(t')\cos\varphi +
7528: s_{0\perp}{\bf m}\cdot ({\bf k}\times {\bf n})$, so the current is 
7529: modulated by 
7530: $1 - \cos^2\varphi\sin^2(\phi/2)$, the probability of finding the spin in 
7531: the direction of magnetization $\bf m$.   
7532: 
7533: Note that $\phi$ does not depend on the momentum (or energy) of the 
7534: carriers. As 
7535: the spread 
7536: $\varphi$ in the momenta increases, the modulation effect decreases.  
7537: The largest effect is seen for $\varphi=0$, where the current modulation 
7538: factor is $\cos^2(\phi/2)$. It was therefore proposed~\cite{Datta1990:APL} 
7539: that $\varphi$
7540: be limited by further confining the electron motion along $\varphi=0$ using 
7541: a one-dimensional channel as a waveguide. 
7542: Spin modulation of the current becomes ineffective if transport is 
7543: diffusive. 
7544: Taking typical values~\cite{Nitta1997:PRL,Koga2002:PRL}
7545: for 
7546: $\hbar\alpha_{BR}\approx 1\times 10^{-11}$ eV$\cdot$m, and 
7547: $m_c=0.1 m_e$, current modulation should be observable at source-drain 
7548: separations of $L\agt 1$ $\mu$m, setting the scale for ballistic transport.
7549: The device will work best with 
7550: narrow-gap materials ~\cite{Lommer1988:PRL}
7551: like InAs, in which the structure inversion asymmetry dominates the spin 
7552: precession \cite{Luo1988:PRB,Luo1990:PRB,Das1989:PRB}. 
7553: Another option is using Si heterostructures, in which bulk inversion 
7554: asymmetry is
7555: absent. However, the small magnitude of the spin-orbit interaction makes 
7556: $\alpha_{BR}$ in Si probably rather weak.
7557: 
7558: The Datta-Das SFET is yet to be realized. There are at least four important
7559: difficulties in observing the proposed effects. 
7560: 
7561: (i) The effective spin 
7562: injection
7563: of spin-polarized carriers from the ferromagnetic source into a 2DEG  
7564: is nontrivial (see Sec.~\ref{sec:IID4}).
7565: 
7566: (ii) Ballistic spin-polarized transport should be realized through the channel
7567: with $\it uniform$ $\alpha_{BR}$ by eliminating undesirable electric fields 
7568: due to interface inhomogeneities. 
7569: 
7570: (iii) The parameter $\alpha_{BR}$ should 
7571: be effectively controllable by the gate. 
7572: 
7573: (iv) 
7574: The structure inversion asymmetry should dominate over the bulk inversion
7575: asymmetry, and the spin precession rate must be 
7576: large enough ($\hbar\alpha_{BR} \agt 10^{-11}$ eV$\cdot$m) to allow
7577: at least a half precession during the ballistic transport. 
7578: 
7579: These four factors present a great challenge to fabricating a 
7580: Datta-Das SFET at room temperature, limiting the design to 
7581: special materials and very clean interfaces.
7582: However, the modulation of $\alpha_{BR}$ 
7583: by biasing voltage (iii) 
7584: has been already convincingly demonstrated in 
7585: In$_{0.53}$Ga$_{0.47}$As/In$_{0.52}$Al$_{0.48}$As 
7586: QW's~\cite{Nitta1997:PRL,Hu1999:PRB,Grundler2000:PRL} 
7587: [for GaAs/AlGaAs 2DEG see
7588: also \cite{Miller2003:PRL}].  Initial experimental 
7589: investigations of magnetoresistance in the Datta-Das SFET systems 
7590: were performed by \textcite{Gardelis1999:PRB}. Recently spin precession
7591: in the the Datta-Das SFET, including the bulk inversion asymmetry,
7592: was investigated by \cite{Winkler2003:lanl} using $k\cdot p$
7593: model calculations. It is not surprising that the conductance through the
7594: transistor, in the present orientation-dependent bulk inversion asymmetry,
7595: depends rather strongly on the crystallographic orientation of the
7596: two-dimensional channel \cite{Lusakowski2003:PRB}. For more discussion
7597: of the Dresselhaus bulk inversion asymmetry and the Bychkov-Rashba structure
7598: asymmetry see Sec.~\ref{sec:IIIB2b}.
7599: 
7600: The Datta-Das SFET has generated great interest in mesoscopic spin-polarized 
7601: transport
7602: in the presence of structure inversion asymmetry. 
7603: Model calculations using the tight-binding formulation of $H_{\rm SIA}$
7604: (recall Sec.~II.B.2)
7605: were reported by \textcite{Pareek2002:PRB}.
7606: Further theoretical investigations on the theme of the Datta-Das spin 
7607: transistor
7608: can be found in \textcite{Matsuyama2002:PRB,Nikolic2001:P} 
7609: and in an extensive review by \textcite{Bournel2000:APF}. 
7610: Distinct SFET's have also been suggested, even in the absence of 
7611: ferromagnetic
7612: regions which are replaced by a rotating external magnetic field of 
7613: uniform strength
7614: \cite{Wang2002:P}. \textcite{Ciuti2002:APL} proposed 
7615: a ferromagnetic-oxide-semiconductor 
7616: transistor, with a nonmagnetic source and drain, 
7617: but with two ferromagnetic gates in series above 
7618: the base channel. The relative orientation of the gates' magnetization 
7619: leads to 
7620: magnetoresistance effects. An SFET that can operate in the diffusive regime, 
7621: in the presence of both bulk and structure inversion asymmetry, 
7622: has been considered by \textcite{Schliemann2002:P}.
7623: 
7624: 
7625: \subsubsection{\label{sec:IVE2} Magnetic bipolar transistor}
7626: %--------------------------------------------------------------------------
7627: 
7628: The magnetic bipolar transistor (MBT) is a bipolar transistor
7629: with spin-split carrier bands and, in general, an injected
7630: spin 
7631: \cite{Fabian2002:P,Fabian2003:Pa,Fabian2003:Pb}. 
7632: A related device structure
7633: was already proposed by \textcite{Gregg1997:JMMM} in a push for
7634: silicon-based spintronics. In this proposal (also called
7635: SPICE for spin-polarized injection current emitter)
7636: the semiconductors have no equilibrium spin, while
7637: the 
7638: spin source 
7639: is provided by a ferromagnetic spin
7640: injector attached to the emitter, and another ferromagnetic
7641: metal, a spin detector,
7642: is attached to the base/collector junction to modulate
7643: the current flow. In both configurations the aim is to
7644: control current amplification 
7645: by spin and magnetic field.
7646: 
7647: A scheme of a particular  MBT is shown in Fig.~\ref{fig:mbt}. 
7648: Such a three-terminal device can be thought of as consisting of 
7649: two magnetic {\it p-n} junctions connected in series. Materials 
7650: considerations discussed in Sec.~\ref{sec:IVD} also apply  
7651: to an MBT in order to provide a sufficient equilibrium polarization
7652: in a magnetic base $P_{B0}$. 
7653: While nonmagnetic, the emitter has a nonequilibrium polarization $\delta P_E$ 
7654: from a spin source, similar to the magnetic diode case in Fig.~\ref{fig:md}.
7655: Only the spin polarization of electrons is assumed. Applying the generalized
7656: Shockley theory to include spin 
7657: effects \cite{Fabian2002:PRB}, a theory of MBT was developed 
7658: by \textcite{Fabian2002:P,Fabian2003:Pb}. 
7659: Later, simplified schemes of MBT [not including the effect of nonequilibrium
7660: spin ($\delta P_E=0$)] were also considered by \textcite{Flatte2003:APL}
7661: and \textcite{Lebedeva2003:JAP}.
7662: 
7663: \begin{figure}
7664: \centerline{\psfig{file=zutic_fig32.eps,width=1\linewidth,angle=0}}
7665: \caption{
7666: Scheme of an {\it n-p-n} 
7667: magnetic bipolar transistor with magnetic base (B),
7668: nonmagnetic emitter (E), and collector (C). 
7669: Conduction and valence bands are separated by the energy gap $E_g$.
7670: The conduction band has a spin splitting $2q\zeta$, leading
7671: to equilibrium spin polarization $P_{B0}=\tanh(q\zeta/k_B T)$. 
7672: Carriers and depletion regions are represented as in Fig.~\ref{fig:md}.
7673: In the so called forward active regime, where the transistor
7674: can amplify currents, the E-B junction is forward biased (here
7675: with voltage $V_{BE}>0$ lowering the built-in potential $V_{bi}$), while the
7676: B-E junction is reverse biased ($V_{BC}<0$). The directions
7677: of the current flows are indicated. Electrons flow from E to B, 
7678: where they either recombine with holes (dashed 
7679: lines) or continue to be swept by the electric field in the B-E 
7680: depletion layer towards C. Holes form mostly the base current, $I_B$,
7681: flowing to the emitter. The current amplification
7682: $\beta=I_C/I_B$ 
7683: can be controlled by $P_{B0}$ 
7684: as well as by the nonequilibrium spin in E. 
7685: Adapted from \onlinecite{Fabian2003:Pa}.
7686: } 
7687: \label{fig:mbt}
7688: \end{figure}
7689: The current amplification (gain) $\beta=I_C/I_B$ (see Fig.~\ref{fig:mbt})
7690: is typically $\sim 100$ in practical transistors.
7691: This ratio depends on many factors, such as the doping 
7692: densities, carrier lifetimes,  
7693: diffusion coefficients,
7694: and structure geometry. 
7695: In an MBT $\beta$ also depends on the spin splitting 
7696: $2q \zeta$ (see Fig.~\ref{fig:mbt})
7697: and the nonequilibrium polarization $\delta P_E$. 
7698: This additional dependence of $\beta$
7699: in an MBT is called magnetoamplification \cite{Fabian2003:Pb}.
7700: An important prediction is that the nonequilibrium spin can be 
7701: injected at low bias 
7702: all the way from the emitter, through the base, to the collector 
7703: \cite{Fabian2002:P,Fabian2003:Pb}
7704: in order to make possible an
7705: effective control of $\beta$ by $\delta P_E$. 
7706: 
7707: The calculated dependence of the gain on the spin splitting for 
7708: $\delta P_E=0.9$ is 
7709: shown in Fig.~\ref{fig:gain}, for GaAs and Si materials parameters. 
7710: The gain is
7711: very sensitive to the equilibrium magnetization in Si, while the 
7712: rapid carrier 
7713: recombination in GaAs prevents more effective control of the transport 
7714: across the base.
7715: In Si it is the spin injection at the emitter-collector depletion 
7716: layer which 
7717: controls the current. As the spin-charge coupling is most effective 
7718: across 
7719: the depletion layer (see Sec.~\ref{sec:IVD}), this 
7720: coupling is essential for the current in Si. 
7721: In the limit of slow carrier recombination \cite{Fabian2002:P},
7722: \begin{equation} \label{eq:beta}
7723: \beta \sim \cosh(q\zeta/k_B T) (1 + \delta P_E P_{B0}).  
7724: \end{equation}
7725: Both magnetic field (through $\zeta$) and nonequilibrium spin affect the 
7726: gain, an implication of the spin-voltaic effect
7727: \cite{Zutic2002:PRL,Zutic2003:P}.
7728: The sensitivity of the current to spin can be used to measure  
7729: the injected spin polarization. If no 
7730: spin source 
7731: is present ($\delta P_E=0$),
7732: there is no spin-charge coupling in the space-charge regions, unless
7733: at least two regions are magnetic. The only remaining effects 
7734: on the $I-V$ characteristics come from the sensitivity of the carrier 
7735: densities 
7736: in the equilibrium  
7737: magnetic regions to $\zeta$ [see Eq.~(\ref{eq:beta})
7738: for the case of $\delta P_E=0$]. 
7739: 
7740: \begin{figure}
7741: \centerline{\psfig{file=zutic_fig33.eps,width=1\linewidth,angle=0}}
7742: \caption{Calculated gain dependence of an MBT as a function of
7743: base spin splitting $2q\zeta$, given in units of thermal energy 
7744: $k_B T$.  The nonequilibrium
7745: spin polarization in the emitter is $\delta P_E=0.9$. 
7746: Si (solid) and GaAs (dashed) 
7747: materials parameters were applied.
7748: Adapted from \onlinecite{Fabian2002:P}. 
7749: } 
7750: \label{fig:gain}
7751: \end{figure}
7752: 
7753: The MBT is, in effect, a magnetic heterostructure transistor, since its 
7754: functionality depends
7755: on tunability of the band structure of the emitter, base, or collector. The 
7756: advantage 
7757: of MBT, however, is that the band structure is not built-in, but can be tuned 
7758: {\it during} 
7759: the device operation by magnetic field or spin injection signals. 
7760: The challenges to demonstrate the predicted phenomena in MBT are 
7761: similar to those of magnetic bipolar diodes, see Sec.~\ref{sec:IVD}.
7762: 
7763: \subsubsection{\label{sec:IVE3} Hot-electron spin transistors}
7764: %-----------------------------------------------------------------------
7765: 
7766: \label{hotspin}
7767: Spin transistors that
7768: rely on transport of hot (nonthermalized) carriers
7769: have the potential to serve of several different purposes. 
7770: On the one hand, they could be used as a diagnostic tool to characterize
7771: spin- and energy-dependent interfacial properties, scattering processes, 
7772: and electronic structure, relevant to spintronic 
7773: devices.\footnote{These efforts 
7774: are motivated in part by the success of (spin-insensitive) 
7775: ballistic-electron-emission microscopy in providing high spatial and
7776: energy resolution of 
7777: properties of metal/semiconductor interfaces 
7778: \cite{Kaiser1988:PRL,Bonnell:2001,Smith2000:PRB}. A subsequent
7779: variation--a ballistic-electron-magnetic microscopy, which also uses an STM 
7780: tip to inject hot carriers, is capable of resolving magnetic features at 
7781: a $\sim10$ nm length scale \cite{Rippard1999:APL,Rippard2000:PRL}.}
7782: On the other hand, hot-electron transistors are also of interest for
7783: their ability to sense magnetic fields, their possible
7784: memory applications, 
7785: and a their potential as a source
7786: of ballistic hot-electron spin injection.
7787: Below we discuss two representative examples,
7788: a spin-valve transistor and a magnetic tunneling transistor.
7789: 
7790: The spin-valve or Monsma transistor provided an early demonstration of a
7791: hot-electron 
7792: spin transistor and realization of a hybrid spintronic
7793: device that integrates metallic ferromagnets and semiconductors 
7794: \cite{Monsma1995:PRL,Monsma1998:S}.
7795: A three terminal structure 
7796: \footnote{Similar to other hot-electron 
7797: spin devices, the term
7798: transistor characterizes their three-terminal structure rather
7799: than the usual functionality of a conventional semiconductor transistor.
7800: In particular, a semiconductor bipolar transistor,
7801: which also has an emitter/base/collector structure, typically has
7802: a sizable current gain--a small change in the base current 
7803: leads to a large change in the collector current (see Sec.~\ref{sec:IVE2}). 
7804: However, only a small
7805: current gain $\sim2$ (due to large current in a metal base)
7806: was predicted in magnetic tunnel-junction-based devices \cite{Hehn2002:PRB}.}
7807: consisted of a metallic base (B) made
7808: of a ferromagnetic multilayers in a CPP geometry [as depicted in 
7809: Fig.~\ref{gmr:1}(a)] surrounded by a silicon emitter (E) and 
7810: collector (C) with two Schottky contacts, formed at E/B and B/C 
7811: interfaces.\footnote{Another realization of a spin-valve transistor
7812: combines a GaAs emitter with a Si collector \cite{Dessein2000:JAP}.} 
7813: Forward bias $V_{EB}$ controls the emitter current $I_E$
7814: of spin-unpolarized electrons, which are injected
7815: into a base region as hot carriers. 
7816: The scattering processes in the base,
7817: together with the reverse bias $V_{BC}$,
7818: influence how many of the injected electrons can overcome the B/C Schottky
7819: barrier and contribute to the collector current $I_C$. 
7820: Similar to the physics of GMR structures~\cite{Levy:2002,Gijs1997:AP}
7821: scattering in the base region strongly depends on the relative orientation 
7822: of the magnetizations in the ferromagnetic layers.
7823: %iz check majority direction
7824: Electrons with spin which has magnetic moment opposite (antiparallel)
7825: to the magnetization of a ferromagnetic layer typically are scattered
7826: more than electrons with parallel magnetic moments,
7827: resulting
7828: in a spin-filtering effect which can be described in terms of spin-dependent
7829: mean free path % $l_\lambda$, $\lambda=\uparrow,\downarrow$ 
7830: \cite{Rendell1980:PRL,Pappas1991:PRL,Hong2000:PRB}. 
7831: Generally, both elastic and 
7832: inelastic
7833: scattering processes determine the effective spin-dependent mean free path,
7834: sometimes also referred to as the attenuation length.\footnote{For electrons 
7835: with sufficiently high excess energy, a scattering process
7836: (influencing the mean free path) does not necessarily remove the electron
7837: from the collector current. The attenuation length, which can be determined
7838: by measuring the base layer thickness dependence of the collector current
7839: [see 
7840: \cite{Rippard1999:APL,Rippard2000:PRL,Vlutters2001:PRB,vanDijken2002:PRB}]
7841: can therefore differ from the effective mean free path.}
7842: The magnetorestive response is usually expressed using magneto current (MC),
7843: defined as the change in collector current, normalized to the 
7844: minimum value
7845: \begin{equation}
7846: MC=(I_{C\uparrow \uparrow} -I_{C\uparrow \downarrow})/
7847: I_{C\uparrow \downarrow},
7848: \label{eq:mc}
7849: \end{equation}
7850: analogous to the expression for GMR or TMR structures
7851: [recall Eq.~(\ref{eq:tmr})],
7852: where $\uparrow \uparrow$ (parallel) and $\uparrow \downarrow$ 
7853: (antiparallel) denote the relative orientation of the magnetizations.
7854: The large values of MC\footnote{These values substantially 
7855: exceed the CPP GMR value 
7856: for the same magnetic multilayer used in the base.} ($>$200\%) 
7857: and the sensitivity of $\sim$130\% per G measured  at room temperature 
7858: \cite{AnilKumar2000:JMMM} demonstrate a capability for magnetic-field sensors. 
7859: Several important challenges, raised by the operation of
7860: the spin-valve transistor, need to be addressed to better realize 
7861: the potential of hot-electron 
7862: transistors. These challenges 
7863: include increasing the small collector current and determining 
7864: whether the spin injection of hot carriers into semiconductors is 
7865: feasible. Furthermore, it would be desirable to fabricate
7866: structures in which semiconductor regions played an active
7867: role, not limited to energy selection (via Schottky barriers)
7868: of the carriers injected into the base and collector regions.
7869: 
7870: \begin{figure} 
7871: \centerline{\psfig{file=zutic_fig34.eps,width=0.85\linewidth,angle=0}}
7872: \caption{Schematic energy diagram of a magnetic tunneling transistor.
7873: Region 1 is the emitter, region 2 
7874: the Al$_2$O$_3$ tunnel barrier of height $\phi$, and region 3 the base.
7875: Together they form a magnetic tunnel junction.
7876: Region 4 is a semiconductor collector that has a Schottky barrier
7877: at the interface with the base. 
7878: From \onlinecite{vanDijken2003:PRL}. 
7879: }
7880: \label{mtt}
7881: \end{figure}
7882: 
7883: An alternative class of hot-electron transistors, often referred to as 
7884: magnetic 
7885: tunneling transistors, has a tunneling junction instead of a Schottky 
7886: barrier 
7887: emitter
7888: \cite{Mizushima1997:IEEETM,Yamauchi1998:PRB,Sato2001:APL,vanDijken2003:PRL,%
7889: vanDijken2002:PRB,Jiang2003:PRL,vanDijken2002:APL}. The
7890: addition of a tunnel junction, combined with a variable $V_{EB}$, allows
7891: exploration of   
7892: hot-electron transport over an energy range
7893: of several eV. At large $V_{EB}$ bias, the ratio $I_C/I_E$, important 
7894: for the device performance, can be substantially increased 
7895: over that of
7896: the 
7897: spin-valve transistor 
7898: \cite{vanDijken2003:APLa,vanDijken2003:APLb,Sato2001:APL}.
7899: 
7900: A particular realization is depicted in Fig.~\ref{mtt}. Different
7901: coercive fields in the regions 1 and 3 ensure independent switching of
7902: of the corresponding magnetizations in the applied
7903: magnetic field. The magnetocurrent MC, defined in Eq.~(\ref{eq:mc}),
7904: shows a nonmonotonic behavior with $V_{EB}$ \cite{vanDijken2003:PRL}
7905: influenced by the conduction-band structure of a collector. 
7906: In GaAs, in addition to the direct conduction band minimum
7907: at the $\Gamma$ point [recall Fig.~\ref{oo:1} (a)], there
7908: are indirect minima at $L$ points at higher energy \cite{Blakemore1982:JAP}.
7909: After an initial decrease of MC with electron energy, 
7910: at $V_{EB} \approx 0.3$ V 
7911: larger than the base/collector Schottky barrier
7912: there is an onset of hot-electron transport 
7913: into $L$ valleys 
7914: accompanied by an increase in MC \cite{vanDijken2003:PRL}. 
7915: 
7916: A large magnetocurrent alone, measured in various hot-electron 
7917: spin transistors 
7918: \cite{Monsma1995:PRL,Monsma1998:S,Sato2001:APL,vanDijken2003:APLb},
7919: is not sufficient to demonstrate spin injection in a semiconductor 
7920: collector. 
7921: For conclusive evidence spin detection in a collector region is needed. 
7922: This was first achieved \cite{Jiang2003:PRL} 
7923: using optical detection with a spin LED 
7924: structure\footnote{Analogous to the spin LED from Fig.~\ref{injexp:3},
7925: in which GaAs collector served as an $n$-type spin aligner and InGaAs/GaAs
7926: was used for a quantum well.}
7927: added to the collector in Fig.~\ref{mtt}.
7928: Measurements at $T=1.4$ K and $B=2.5$ T, after a background subtraction, 
7929: showed majority spin injection  with $P_{\rm circ} \approx 10$ \%.
7930: 
7931: In another realization of a magnetic tunnel transistor, more similar
7932: to the original spin-valve transistor, the emitter was nonmagnetic (Cu) 
7933: while the
7934: base was a magnetic multilayer (F1/N/F2) \cite{vanDijken2003:APLb}.
7935: The resulting strong spin-filtering effect can be inferred  from the 
7936: transmitted hot carriers with a
7937: spin-dependent exponential decay  within the F$_i$, $i=1,2$ layer.
7938: Unpolarized electrons, injected from the emitter, 
7939: after passing an F1 layer of thickness $t$ acquire an effective
7940: transmitted polarization 
7941: \begin{equation}
7942: P_{N1}=\frac{N_\uparrow-N_\downarrow}
7943: {N_\uparrow+N_\downarrow}=
7944: \frac{e^{-t/l_\uparrow}-e^{-t/l_\downarrow}}
7945: {e^{-t/l_\uparrow}+e^{-t/l_\downarrow}},
7946: \end{equation}
7947: where 
7948: $N_\uparrow$ and $N_\downarrow$ 
7949: represent the number of transmitted electrons
7950: with majority or minority spin and $l_\uparrow$ and $l_\downarrow$
7951: are the corresponding attenuation length (the polarization $P_{N2}$ has
7952: an analogous form). 
7953: The resulting magnetocurrent can be
7954: expressed as \cite{vanDijken2003:APLb}
7955: \begin{equation}
7956: MC=2 P_{N1}P_{N2}/(1-P_{N1} P_{N2}),
7957: \label{eq:mc2}
7958: \end{equation}
7959: which is analogous to Eq.~(\ref{eq:julliere}) 
7960: for TMR using Julli{\`{e}}re's model,
7961: but with the redefined definition of spin polarization. 
7962: At $V_{EB}=0.8$ V and at $T=77$ K, the measured
7963: MC exceeds $3400$ \%, while  with Eq.~(\ref{eq:mc2}) the
7964: polarization of the transmitted electrons can be estimated to exceed 90 \%,
7965: even with a ferromagnet only a $\sim 3$ nm thick \cite{vanDijken2003:APLb}.
7966: A theoretical analysis of spin injection and spin filtering in
7967: magnetic tunneling transistors was given by \textcite{Rashba2003:P}
7968: who extended the approach for ballistic spin injection 
7969: \cite{Kravchenko2003:PRB}
7970: (Sec.~\ref{sec:IIC2}) to include the effects of hot-electron transport
7971: and inelastic scattering.
7972: 
7973: Future studies of hot-electron 
7974: spin transistors are expected to
7975: result in increased spin injection even at room temperatures and
7976: to utilize other semiconductor collectors. It would be particularly 
7977: desirable to demonstrate hot-electron 
7978: spin injection in Si
7979: and facilitate an integration with the CMOS technology.
7980: 
7981: 
7982: \subsection{\label{sec:IVF} Spin qubits in semiconductor nanostructures}
7983: %-------------------------------------------------------------------------
7984: 
7985: A potentially revolutionary idea in spintronics is the possibility
7986: of using the two-level nature of 
7987: electron spin to create a solid-state quantum computer 
7988: \cite{DiVincenzo1995:S,DasSarma2001:SSC,Nielsen:2000}.  
7989: The basic unit in a quantum computer is the quantum bit (or qubit), 
7990: the quantum
7991: analog of the binary bit in a classical digital computer.
7992: A qubit is essentially
7993: a controllable quantum two-level system 
7994: \cite{DasSarma2001:AS,Nielsen:2000}.  
7995: While the dimensionality (2$^n$) of the Hilbert space of $n$ electron spins
7996: is the same as the number of  configurations of a corresponding classical
7997: system, a quantum system can be
7998: in a superposition of all the basis states, effectively performing 
7999: (via a unitary evolution) many classical computations in parallel.  
8000: Several spin-based quantum computer
8001: schemes have been proposed and extensively studied\footnote{See, for example,  
8002: \cite{Loss1998:PRA,Privman1998:PL,Kane1998:N,Hu2000:PRA,Koiller2002:PRL,%
8003: Koiller2003:PRL,Friesen2003:PRB,Piermarocchi2002:PRL,DiVincenzo2000:FP,%
8004: Kane2000:FP,Hu2001:PRA,Burkard1999:PRB,Troiani2003:PRL,%
8005: Skinner2003:PRL,Levy2002:PRL,Meier2003:PRL,Vrijen2000:PRA}.}  
8006: A common theme in these proposals is the idea of
8007: manipulating the dynamics of a single (or a few) electron spin(s) in
8008: semiconductor nanostructures (e.g., quantum dots), with the reasonable
8009: hope that the predicted behavior will extend to many-spin systems,
8010: requisite for practical quantum computation. 
8011: 
8012: \begin{figure}
8013: \centerline{\psfig{file=zutic_fig35.eps,width=1\linewidth,angle=0}}
8014: \caption{The \textcite{Loss1998:PRA} proposal for spin-based solid-state 
8015: quantum computing. Electrons are localized in
8016: electrostatically defined quantum dots, with coupling between 
8017: electron spins---via the exchange interaction---allowed by tunneling 
8018: between the dots. The tunneling is controlled by gate
8019: voltage. The figure shows two electrons localized in the regions defined
8020: by the gates (shaded). Single-qubit
8021: operations are realized by single-spin precessions (circles), 
8022: performed by applying 
8023: local magnetic fields (here perpendicular to the page) to each dot. 
8024: Two-qubit operations are done 
8025: through the exchange interaction indicated by the dashed curves. 
8026: The scheme works according to the time-dependent Hamiltonian 
8027: $H(t)=\sum_{i,j}' J_{i,j}(t) {\bf S}_i\cdot {\bf S}_j 
8028: +\mu_B g \sum {\bf B}_i(t) {\bf S}_i$, 
8029: where the first summation, over all neighboring spin pairs, 
8030: describes the local
8031: exchange interaction ($J$ is the exchange coupling), 
8032: while the second describes
8033: the single spin operations by local magnetic fields. Variations to 
8034: this scheme are described by \textcite{Burkard2000:FP}.  
8035: }
8036: \label{fig:QC}
8037: \end{figure}
8038: 
8039: The control of spin dynamics and entanglement [many-spin quantum 
8040: correlations \cite{Nielsen:2000}]
8041: at the single-spin level in a semiconductor quantum dot structure is a 
8042: formidable task,
8043: which has not been achieved even at mK temperatures, 
8044: although impressive experimental advances have recently been made
8045: \cite{Fujisawa2002:N,Elzerman2003:PRB,Hanson2003:lanl}.
8046: The current architectures for spin-based quantum computing 
8047: employ GaAs quantum dots \cite{Loss1998:PRA} or
8048: Si (or Si-Ge) systems \cite{Kane1998:N}, with different 
8049: variations. 
8050: The basic idea (see Fig.~\ref{fig:QC}) is to manipulate
8051: the spin states of a single electron using external magnetic fields (or
8052: microwaves) for single-qubit operations and to utilize the quantum exchange
8053: coupling between two neighboring electrons to carry out two-qubit
8054: operations.  
8055: 
8056: State-of-the-art 
8057: techniques, to measure a single spin in a solid,
8058: such as magnetic resonance force microscopy 
8059: \cite{Sidles1995:RMP,Mamin2003:P,Barbic2002:JAP} 
8060: or the spin-selective single-electron-transistor spectroscopy
8061: \cite{Ono2002:S}, are 
8062: still not sensitive enough 
8063: for quantum computing operations.  
8064: However, recently a single shot readout of the spin of an individual electron
8065: has been demonstrated using an
8066: electrical pump-probe measurement \cite{Kouwenhoven2004:PC}.
8067: A single electron with an unknown spin was trapped in a quantum dot
8068: for a few miliseconds. At the end of the trapping time the spin was
8069: measured by quickly shifting the Zeeman resolved spin states towards
8070: the Fermi energy. A spin to charge conversion allowed for an
8071: electrical readout of the spin.
8072: 
8073: The real motivation for using the spin
8074: as a qubit is its long coherence time, microseconds or longer
8075: in operational experimental conditions \cite{Sousa2003:PRB}, to be
8076: contrasted with typical picosecond coherence times for charge or orbital
8077: states in solids. 
8078: Interest in spin-based quantum computing
8079: will only increase as we understand
8080: more about spin coherence and relaxation from other spintronic studies. 
8081: The broad
8082: subject of spin-based quantum computing, 
8083: which is related to the areas of quantum measurement and quantum 
8084: decoherence \cite{Zurek2003:RMP}
8085: is beyond the scope of 
8086: this review.
8087: 
8088: 
8089: 
8090: \section{\label{sec:V} Outlook}
8091: 
8092: 
8093: We have reviewed selected topics on spintronics, emphasizing both the
8094: fundamental aspects of spin dynamics, transport,
8095: and relaxation, and the potential applications.
8096: While the current push for spintronics
8097: is driven by the prospect of technological applications, the fundamental
8098: spin physics, which has a longstanding tradition in the solid-state
8099: community, is by itself exciting and worth pursuing. Furthermore,
8100: even though many proposed spintronic device schemes may turn out
8101: to be impractical in the end, their importance lies in stimulating
8102: interesting experimental and theoretical research.
8103: 
8104: There are many challenges and open questions to be tackled by future
8105: research, in particular a robust
8106: spin injection into silicon.\footnote{Small signals attributed to spin 
8107: injection have already been reported \cite{Jia1996:IEEE}.}
8108: While GaAs is of great technological
8109: importance, the control of spin in silicon would raise hopes for
8110: seamless integration of spintronics with the current
8111: information technology. In addition, the small magnitude of the
8112: spin-orbit interaction and the absence of inversion symmetry
8113: lead to relatively long room-temperature spin lifetimes (of about
8114: 10 ns; see Sec.~\ref{sec:IIID1}), relaxing some constraints on the
8115: operational length and time scales. Important materials advances
8116: have been realized in improving the compatibility of Si/III-V
8117: structures \cite{Sieg1998:APL}
8118: suggesting a possibility that the existing control
8119: of spin in GaAs or in III-V ferromagnetic semiconductors might be
8120: extended to Si.
8121: 
8122: Future progress in spin-polarized transport will be largely driven
8123: by the materials advances. In the context of semiconductors, considering
8124: all-semiconductor structures rather than the hybrid structures with
8125: metallic ferromagnets will depend on the improvements in ferromagnetic
8126: semiconductors, for example, whether they 
8127: can achieve higher mobility, 
8128: higher Curie temperature,\footnote{There still remain many challenges in
8129: accurately predicting Curie temperature. First principles results 
8130: suggest that dominant models of ferromagnetism in semiconductors
8131: cannot be used to explain a variation of Curie temperature across
8132: different materials \cite{Erwin2004:NM}. For reviews of ferromagnetic
8133: semiconductor theories outside the scope of this article, see,
8134: for example, \textcite{Nagaev:1983}; \textcite{Bhat2002:JS};
8135: \textcite{Dietl2002:SST}; \textcite{Sanvito2002:JS}, 
8136: \textcite{DasSarma2003:PRB}; \textcite{Koning:2003}; 
8137: \textcite{Timm2003:JPCM}.}
8138: and a simple fabrication of high quality
8139: interfaces with nonmagnetic materials. What is missing, even in the
8140: currently available materials, is a systematic understanding of the
8141: effects of magnetic interfaces and materials inhomogeneities on 
8142: spin-polarized transport. A comprehensive transport calculation
8143: in the actual devices
8144: with realistic electronic structure of
8145: the studied materials would provide valuable insights into 
8146: both the  spin polarization being measured and
8147: how it is reduced from the moment it was generated.
8148: 
8149: Spin relaxation and spin dephasing of conduction electrons
8150: is a rather field, with the basic principles well understood.
8151: What is needed are accurate band-structure-derived calculations
8152: of spin relaxation times, in both metals and semiconductors.
8153: The same can be said of
8154: $g$ factor, calculation of which from first
8155: principles is a nontrivial task that has not been accomplished even
8156: for the elemental metals. An important and still debated issue
8157: is spin relaxation and decoherence of localized or
8158: confined electrons, where the hyperfine-interaction mechanism
8159: dominates. Furthermore, single-spin relaxation and
8160: decoherence, and their relation to the ensemble spin dephasing,
8161: need to be 
8162: pursued further 
8163: in the context of quantum computing. 
8164: A first step towards understanding
8165: single-spin relaxation  is the recent experiment of
8166: \textcite{Hanson2003:lanl} in a one-electron quantum dot.
8167: 
8168: While dynamic nuclear polarization induced by electron spin can often
8169: be a nuisance for detecting intrinsic spin dynamics 
8170: (see Sec.~\ref{sec:IIID3}),
8171: the interaction between electron and nuclear spins 
8172: \cite{Paget:1984,Fleisher:1984,Vagner2003:P,Smet2002:N} 
8173: is of fundamental
8174: importance for spintronics. An NMR of the nuclear spin polarized
8175: by spin-polarized photoexcited electrons has already been used to detect
8176: the nonequilibrium electron spin in Si \cite{Lampel1968:PRL}. On the
8177: other hand, an NMR signal can be detected optically through measuring
8178: changes in the circular polarization of photoluminescence
8179: due to resonant variations of the nuclear field 
8180: \cite{Dyakonov:1984}, as shown first in
8181: p-doped Ga$_{0.7}$Al$_{0.3}$As \cite{Ekimov1972:JETPL}. 
8182: The early work of \textcite{Lampel1968:PRL}, 
8183: and \textcite{Ekimov1972:JETPL} established the basic
8184: principles for a series of experiments that demonstrated various 
8185: realizations of an all-optical NMR. The role of the resonant 
8186: radio waves is played by periodically optically excited electron spins
8187: \cite{Salis2001:PRL,Kalevich1986:FTT,Kalevich1981:FTT,Kalevich1980:FTT,%
8188: Kikkawa2000:S,Eickhoff2002:PRB,Fleisher:1984b}.
8189: Electron-nuclear spintronics is likely to become relevant for
8190: quantum computation and for few-spin manipulations, which can benefit
8191: from long nuclear spin coherence times (even lasting minutes).
8192: 
8193: The range of potential spintronic applications 
8194: goes beyond the use of
8195: large magnetoresistive
8196: effects.
8197: \textcite{Rudolph2003:APL}, for example, have demonstrated the operation of
8198: a spin laser. The laser is a vertical-cavity surface-emitting laser 
8199: (VCSEL), optically pumped in the gain medium, here two InGaAs quantum
8200: wells, with 50\% spin-polarized electrons. The electrons recombine with 
8201: heavy holes, which are effectively unpolarized, emitting circularly
8202: polarized light (see Sec.~\ref{sec:IIB}). The threshold electrical current, 
8203: extracted from the pump power for the lasing operation, was found to be 
8204: 0.5 A$\cdot$cm$^{-2}$,
8205: which is 23\% below the threshold current of the spin-unpolarized VCSEL.
8206: Furthermore, for a fixed pump power, the emission power of the laser
8207: changed by 400\% upon changing the degree of circular polarization of
8208: the pump laser. The reason for the decrease in threshold is the selective
8209: coupling of spin-polarized electrons to photons with one helicity.
8210: While the experiment was conducted at 6 K, a room-temperature
8211: operation and an electrically pumped laser should be viable as 
8212: well.\footnote{The requirement is that the spin relaxation time be longer 
8213: than
8214: the carrier recombination time, and that the spin injection spot and the
8215: gain medium be within the spin transport length.}
8216: 
8217: \begin{figure}
8218: \centerline{\psfig{file=zutic_fig36.eps,width=1.0\linewidth}}
8219: \caption{(a) Nanopilar device: schematic diagram of
8220: a nanopilar device
8221: operating at room temperature. 
8222: The direction of magnetization
8223: is fixed (pinned) in the thick Co film and free in the thin Co film;
8224: (b) differential resistance $dV/dI$ of a nanopilar device as
8225: a function of applied field; (c) $dV/dI$ of the device as a
8226: function of applied current. 
8227: The arrows in panels (b) and (c) represent the direction of
8228: magnetic field and current sweeps, respectively. 
8229: For positive current, 
8230: electrons flow from the thin to the thick Co film.
8231: Adapted from \onlinecite{Albert2002:PRL}.} 
8232: \label{ralph}
8233: \end{figure}
8234: 
8235: The demonstration that the flow of spin-polarized carriers, rather
8236: than applied magnetic field, can also be used to manipulate
8237: magnetization of ferromagnetic materials brings the exciting
8238: prospect of a novel class of spintronic devices. In addition
8239: to reversal of magnetization, which is a key element in realizing
8240: various
8241: magnetoresitive applications, the driving of a spin-polarized current
8242: can lead to coherent 
8243: microwave oscillations in nanomagnets \cite{Kiselev2003:P}. 
8244: Spin-transfer
8245: torque (Sec.~\ref{sec:IB1}) 
8246: has already been realized in several 
8247: experimental geometries. These include nanowires 
8248: \cite{Wegrowe1999:EL,Kelly2003:PRB}, 
8249: point contacts 
8250: \cite{Tsoi1998:PRLa,Yi2003:PRL,Tsoi2000:N,Tsoi2002:PRL}, 
8251: nanoconstrictions \cite{Myers1999:S,Rippard2003:P}, 
8252: and nanopilars \cite{Katine2000:PRL,Albert2002:PRL,Urazhdin2003:PRL} 
8253: (see Fig.~\ref{ralph}), all involving metallic ferromagnets.
8254: The common feature of all these geometries is a need for very 
8255: large current densities ($\sim10^7$ Acm$^{-2}$).  
8256: Ongoing
8257: experiments \cite{Munekata2003:PC,Chiba2004:P,Yamanouchi2004:N}
8258: to demonstrate 
8259: spin-transfer torque 
8260: (together with other cooperative phenomena) 
8261: in ferromagnetic semiconductors, which have
8262: much smaller magnetization than their metallic counterparts, are expected
8263: to also require much smaller switching currents. Based on the findings
8264: in electric-field controlled ferromagnetism (see Fig.~\ref{ohno}),
8265: it has been demonstrated that the reversal of magnetization in (In,Mn)As can
8266: be manipulated by modifying the carrier density, using a gate voltage
8267: in a FET structure \cite{Chiba2003:S}.
8268: 
8269: \acknowledgments{We thank R. H. Buhrman, K. Bussmann, R. de Sousa, 
8270: M. I. D'yakonov,  
8271: S. C. Erwin, M. E. Fisher, A. M. Goldman, K. Halterman, X. Hu, S. V. Iordanskii,
8272: M. Johnson, B. T. Jonker, K. Kavokin, J. M. Kikkawa,  
8273: S. Maekawa, C. M. Marcus, I. I. Mazin, B. D. McCombe, J. S. Moodera, 
8274: H. Munekata, 
8275: B. E. Nadgorny, H. Ohno, 
8276: S. S. P. Parkin,  
8277: D. C. Ralph, E. I. Rashba, W. H. Rippard, V. I. Safarov, G. Schmidt, 
8278: E. Sherman, 
8279: R. H. Silsbee, D. D. Smith, 
8280: M. D. Stiles, O. T. Valls, S. van Dijken, 
8281: H. M. van Driel, T. Venkatesan, S. von Moln\'{a}r, and J. Y. T. Wei 
8282: for useful discussions.  
8283: We thank D. D. Awschalom, N. C. Emley, R. T. Harley, M. Johnson, B. T. Jonker, 
8284: K. Kavokin, J. M. Kikkawa, C. M. Marcus, H. Munekata, Y. Ohno, A. Oiwa, 
8285: S. S. P. Parkin, R. J. Soulen, Jr.,  and M. Tanaka for providing us with 
8286: illustrative
8287: figures from their published works, and A. Kaminski and D. J. Priour
8288: for help with preparation of the manuscript. This work was supported by 
8289: DARPA, the US ONR, and the NSF-ECS.} 
8290: 
8291: \begin{table}
8292: \begin{tabular}{ll}
8293: \hline
8294: \hline
8295: 2DEG  & two dimensional electron gas \\
8296: BAP   & Bir-Aronov-Pikus \\
8297: BCS   & Bardeen Cooper Schrieffer \\
8298: BIA   & bulk inversion asymmetry \\
8299: BTK   & Blonder-Tinkham-Klapwijk \\
8300: CESR  & conduction electron spin resonance \\
8301: CIP   & current in plane \\
8302: CMOS  & complementary metal oxide semiconductor \\
8303: CMR   & colossal magnetoresistance \\
8304: CPP   & current perpendicular to plane \\
8305: DNA   & deoxyribonucleic acid \\
8306: DP    & D'yakonov-Perel' \\
8307: EDSR  & electron dipole spin resonance \\
8308: EMF   & electromotive force \\
8309: ESR   & electron spin resonance \\
8310: EY    & Elliot-Yafet \\
8311: F     & ferromagnet \\
8312: FSm   & ferromagnetic semiconductor \\
8313: GMR   & giant magnetoresistance \\
8314: FET   & field effect transistor  \\
8315: HFI   & hyperfine interaction \\
8316: I     & insulator \\
8317: HTSC  & high temperature superconductor \\
8318: LED   & light emitting diode \\
8319: MBD   & magnetic bipolar diode \\
8320: MBE   & molecular beam epitaxy \\
8321: MBT   & magnetic bipolar transistor \\
8322: MC    & magnetocurrent \\
8323: MR    & magnetoresistance \\
8324: MRAM  & magnetic random access memory \\
8325: MTJ   & magnetic tunnel junction \\
8326: N     & normal (paramagnetic) metal \\
8327: NMR   & nuclear magnetic resonance \\
8328: QD    & quantum dot \\
8329: QPC   & quantum point contact \\
8330: QW    & quantum well \\
8331: S     & superconductor \\
8332: SET   & single electron transistor  \\
8333: SFET  & spin field effect transistor  \\
8334: SIA   & structure inversion asymmetry \\
8335: Sm    & semiconductor \\
8336: SQUID & superconducting interference quantum device \\
8337: STM   & scanning tunneling microscope \\
8338: TESR  & transmission electron spin resonance \\
8339: \hline
8340: \hline
8341: \end{tabular}
8342: \caption{List of acronyms used in the text.}
8343: \label{tab:2}
8344: \end{table}
8345: 
8346: \bibliographystyle{apsrmp}
8347: \bibliography{references}
8348: 
8349: \end{document}
8350: