1: \documentclass[prl,aps,showpacs,twocolumn,unsortedaddress]{revtex4}
2: \usepackage{graphics,bm}
3: \usepackage{amssymb}
4: \usepackage{epsfig}
5: \usepackage{epsf}
6:
7: \begin{document}
8:
9: \title{Thermally-Assisted Current-Driven Domain Wall Motion}
10:
11: \author{R.A. Duine}
12: \email{duine@physics.utexas.edu}
13: \homepage{http://www.ph.utexas.edu/~duine}
14:
15: \author{A. S. N\'u\~nez}
16: \email{alnunez@physics.utexas.edu}
17: \homepage{http://www.ph.utexas.edu/~alnunez}
18:
19: \author{A.H. MacDonald}
20: \email{macd@physics.utexas.edu}
21: \homepage{http://www.ph.utexas.edu/~macdgrp}
22:
23: \affiliation{The University of Texas at Austin, Department of
24: Physics, 1 University Station C1600, Austin, TX 78712-0264}
25: \date{\today}
26:
27: \begin{abstract}
28: Starting from the stochastic Landau-Lifschitz-Gilbert equation, we
29: derive Langevin equations that describe the nonzero-temperature
30: dynamics of a rigid domain wall. We derive an expression for the
31: average drift velocity of the domain wall $\langle \dot{r}_{\rm
32: dw} \rangle$ as a function of the applied current, and find
33: qualitative agreement with recent magnetic semiconductor
34: experiments. Our model implies that at any nonzero temperature
35: $\langle \dot{r}_{\rm dw} \rangle$ initially varies linearly with
36: current, even in the absence of non-adiabatic spin torques.
37: \end{abstract}
38:
39: \pacs{72.25.Pn, 72.15.Gd}
40:
41: \maketitle
42:
43: % definitions
44: \def\bx{{\bf x}}
45: \def\bk{{\bf k}}
46: \def\half{\frac{1}{2}}
47: \def\args{(\bx,t)}
48:
49: \noindent {\it Introduction} --- Possibilities opened up by modern
50: nanofabrication capabilities have motivated renewed interest in
51: current-induced domain wall motion, a phenomenon first predicted
52: and subsequently observed in seminal work by Berger
53: \cite{berger1984,freitas1985}. The theoretical
54: \cite{tatara2004,zhang2004,thiaville2004,barnes2005,tserkovnyak2005,kohno2006,
55: nguyen2006} and experimental
56: \cite{grollier2003,tsoi2003,yamaguchi2004,klaui2005,yamanouchi2004,yamanouchi2006}
57: study of domain wall motion induced by spin transfer torques
58: \cite{slonczewski1996,berger1996,tsoi1998,myers1999} is currently
59: one of the most active subfields of spintronics \cite{wolf2001}.
60: Recent research has highlighted a number of fundamentally
61: interesting issues that are currently under spirited debate. One
62: controversy concerns intrinsic
63: pinning\cite{tatara2004,barnes2005}, {\it i.e.}, domain walls that
64: are stationary up to a critical current in the absence of spatial
65: inhomogeneity. An intellectually distinct but phenomenologically
66: related debate surrounds theories of nonadiabatic spin torques,
67: which differ widely in their predictions
68: \cite{zhang2004,thiaville2004,barnes2005,tserkovnyak2005,kohno2006}.
69: It turns out that in the presence of these torques, a domain wall
70: is never intrinsically pinned
71: \cite{zhang2004,thiaville2004,barnes2005}.
72:
73: This Letter is motivated primarily by the recent experiments of
74: Yamanouchi {\it et al.} \cite{yamanouchi2006}, in which
75: current-induced domain wall motion was studied over five orders of
76: magnitude of average wall velocity. An important conclusion of
77: these authors is that at low temperatures the domain wall
78: undergoes creep motion, {\it i.e.}, that the domain wall does not
79: move rigidly. Yamanouchi {\it et al.} arrive at this conclusion
80: because the effects of nonzero temperature, when treated
81: \cite{tatara2005} in a rigid domain wall approximation, seem to
82: lead to results that are irreconcilable with experiment. In this
83: Letter we demonstrate that a systematic theory of the influence of
84: a thermal bath on the current-driven motion of a rigid domain wall
85: leads to results that are in qualitative agreement with
86: experiment. In the following sections we first explain our theory
87: of nonzero-temperature domain wall motion and then discuss its
88: implications for recent experiments.
89:
90:
91: \noindent {\it Drift velocity of a rigid domain wall} --- Our
92: starting point is the stochastic Landau-Lifschitz-Gilbert (LLG) equation
93: \cite{brown1963,kubo1970,ettelaie1984,garciapalacios1998,heinonen2004,safonov2005,rossi2005}
94: for the direction of magnetization $\hat \Omega$
95: \begin{equation}
96: \label{eq:stochasticLLG}
97: \frac{\partial \hat \Omega}{\partial t} =\hat \Omega \bm{\times} \left (
98: {\bf H} + \bm {\eta} \right) - \alpha \, \hat \Omega
99: \bm{\times} \frac{\partial \hat \Omega}{\partial t}~,
100: \end{equation}
101: where ${\bf H}$ is the effective field defined by ${\bf H}(\bx) =
102: - \delta E_{\rm MM}[\hat{\Omega}]/(\hbar \delta
103: \hat{\Omega}(\bx))$ and $E_{\rm MM}[\hat{\Omega}]$ is the
104: micromagnetic energy functional. In Eq.~(\ref{eq:stochasticLLG})
105: ${\bm \eta}$ is a gaussian stochastic magnetic field with zero
106: mean and correlations
107: \begin{equation}
108: \label{eq:noisecorrsfull}
109: \langle \eta_\sigma (\bx,t) \eta_{\sigma'} (\bx',t') \rangle
110: = \sigma \delta (t-t') a^3 \delta
111: (\bx-\bx') \delta_{\sigma\sigma'}~,
112: \end{equation}
113: where $a^3$ is the (local) volume of the finite element grid. The
114: strength of the noise is given by $\sigma = 2 \alpha k_{\rm B}
115: T/\hbar$, proportional to the Gilbert damping parameter $\alpha$
116: and to the thermal energy $k_{\rm B} T$. ($\hbar$ is Planck's
117: constant which relates energy and frequency.) In using this
118: expression for the strength of the fluctuations we neglected the
119: influence of current on thermal magnetization fluctuations
120: \cite{foros2005} which is higher order \cite{duine2006} than the
121: spin-torque effects studied here.
122:
123: The effective field can be separated into magnetic energy and spin transfer
124: torque contributions ${\bf H}=\left.{\bf H}\right|_{0}+\left.{\bf
125: H}\right|_j$. For a ferromagnet with an easy $z$ axis and a hard $y$ axis we have
126: \begin{equation}
127: \label{eq:efffieldstatic}
128: \left. {\bf H} \right|_0 = J \nabla^2 \hat
129: \Omega + 2 \omega_{\rm i}\Omega_z - 2 \omega_{\rm o} \Omega_y~ + H_{\rm ext} \hat{z},
130: \end{equation}
131: where $J$ is the spin stiffness, $H_{\rm ext}$ is an external
132: field in the easy-axis direction and $\omega_{\rm i}$ and
133: $\omega_{\rm o}$ are respectively the easy axis and hard-plane
134: anisotropy constants. Assuming only locality implied by smooth
135: magnetization textures, the spin-transfer torque can be separated
136: quite generally into contributions parallel and perpendicular to
137: the spatial derivative of the magnetization:
138: \begin{equation}
139: \label{eq:efffieldcurrent}
140: \left. {\bf H} \right|_j \times \hat{\Omega} = v_{\rm s} \frac{\partial \hat \Omega}{\partial r}
141: + \beta v_{\rm s} \hat \Omega \bm{\times} \frac{\partial \hat \Omega}{\partial
142: r}~,
143: \end{equation}
144: where the gradient is taken in the direction of current flow. In
145: the absence of spin-orbit coupling, it follows from total spin
146: conservation that $\beta = 0$ and that the {\em spin velocity}
147: $v_{\rm s} \equiv (j_{\uparrow} - j_{\downarrow})/(-e
148: (n_{\uparrow} - n_{\downarrow}))$, where $j_{\sigma}$ and
149: $n_{\sigma}$ are majority and minority spin contributions to the
150: currents and spin-densities in the collinear limit. For realistic
151: ferromagnets spin and orbital degrees of freedom are coupled, and
152: the microscopic theory of $v_s$ and $\beta$ is more challenging
153: and still controversial
154: \cite{zhang2004,thiaville2004,barnes2005,tserkovnyak2005,kohno2006}.
155: The term proportional to $\beta$, the nonadiabatic spin transfer
156: torque, plays a central role in the theory of current-driven
157: domain wall motion.
158:
159: In the absence of current and noise, Eq.~(\ref{eq:stochasticLLG})
160: admits time-independent solutions corresponding to domain walls.
161: In terms of the angles $\theta$ and $\phi$ defined by $\hat \Omega
162: = (\sin \theta \cos \phi, \sin \theta \sin \phi, \cos \theta)$,
163: the solution corresponding to an isolated domain wall centered at
164: $r_{\rm dw}$ is $\phi=0$ and $\cos \theta (x-r_{\rm dw}) =\tanh
165: [(x-r_{\rm dw})/\lambda]$, where the domain wall width
166: $\lambda=\sqrt{J/(2\omega_{\rm i}})$\,\cite{tatara2004}. Rigid
167: domain wall motion is described by elevating $r_{\rm dw} \to
168: r_{\rm dw}(t)$ and $\phi(x,t) \to \phi_0(t)$ to the role of
169: collective dynamical variables\cite{tatara2004}. The Langevin
170: equations which describe their stochastic dynamics are
171: \begin{eqnarray}
172: \label{eq:langevinvarpars}
173: \dot \phi_0 + \alpha \frac{\dot r_{\rm dw}}{\lambda} &=& \frac{\beta v_{\rm
174: s}}{\lambda} -H_{\rm ext} + \eta_{\phi} ~; \\
175: \label{eq:langevinvarpars2} \frac{\dot r_{\rm dw}}{\lambda} -
176: \alpha \dot \phi_0 &=&\omega_{\rm o} \sin (2\phi_0) + \frac{v_{\rm
177: s}}{\lambda} + \eta_r ~.
178: \end{eqnarray}
179: These equations can be derived heuristically by enforcing
180: consistency with the stochastic LLG equations at the center of the
181: domain wall. Alternately these equations can be derived by noting
182: that the probability distribution $P[\hat \Omega,t]$, generated by
183: Eqs.~(\ref{eq:stochasticLLG})~and~(\ref{eq:noisecorrsfull}) can be
184: written as a path integral $P[\hat \Omega,t] = \int d [\hat
185: \Omega] \delta [\hat{\Omega} \cdot \hat{\Omega} - 1]
186: e^{-S[\hat\Omega]}$, with effective action
187: \cite{duine2006,zinnjustinbook,duine2002}
188: \begin{equation}
189: \label{eq:effactionfull}
190: S[\hat \Omega] = \int^{t}dt'\int\!\frac{d\bx}{a^3} \frac{1}{2 \sigma}
191: \left(
192: \frac{\partial \hat \Omega}{\partial t'} \times \hat \Omega - {\bf H} + \alpha
193: \frac{\partial \hat \Omega}{\partial t'}
194: \right)^2~.
195: \end{equation}
196: Inserting the domain-wall solution with time-dependent $r_{\rm
197: dw}$ and $\phi_0$ into this action gives rigid domain wall
198: probabilities specified by the effective action
199: \begin{eqnarray}
200: \label{eq:effactionsrdwandphi}
201: && S[r_{\rm dw},\phi_0] = \int^t dt' \frac{N}{2\sigma}\left[
202: \left( \dot \phi_0 + \alpha \frac{\dot r_{\rm dw}}{\lambda} - \frac{\beta v_{\rm s}}{\lambda}+H_{\rm ext}\right)^2
203: \right. \nonumber \\
204: && +
205: \left( \frac{\dot r_{\rm dw}}{\lambda} - \alpha \dot \phi_0
206: -\omega_{\rm o } \sin (2\phi_0) - \frac{v_{\rm s}}{\lambda}\right)^2
207: \nonumber \\
208: && \left.
209: -\frac{4}{3} \omega_{\rm o} ^2 \sin^4 \phi_0
210: \right]~,
211: \end{eqnarray}
212: where $N=2\lambda A/a^3$ is the number of spins in a domain wall
213: with cross-sectional area $A$. If we ignore the last term in this
214: effective action, which vanishes in any case in the gaussian
215: fluctuation limit,
216: Eqs.~(\ref{eq:langevinvarpars})~and~(\ref{eq:langevinvarpars2})
217: are recovered. This minor inconsistency in the theoretical
218: treatment can be traced to the constant azimuthal angle across the
219: domain wall in the variational {\em ansatz}. The final term must
220: also be dropped if the rigid domain wall approximation action is
221: to reproduce the correct equilibrium probability distribution
222: function. (Note that even when $\beta=0$ the Boltzmann equilibrium
223: distribution is approached in the steady state.) The advantage of
224: this approach is that we can read off the strengths of the
225: gaussian noise terms $\eta_\phi$ and $\eta_r$ which are given by
226: \begin{equation}
227: \label{eq:noiseonvarpars}
228: \langle \eta_\phi (t) \eta_\phi (t') \rangle
229: = \langle \eta_r (t) \eta_r (t') \rangle
230: = \frac{\sigma}{N} \delta (t-t')~.
231: \end{equation}
232: Eqs.~(\ref{eq:langevinvarpars}),~(\ref{eq:langevinvarpars2}),~and~(\ref{eq:noiseonvarpars})
233: generalize the variational approach to current-induced domain-wall
234: motion of Ref.~\cite{tatara2004} to finite temperature. Tatara and
235: Kohno \cite{tatara2004} derive their equations from an energy
236: functional, an approach that has been criticized recently by
237: Barnes and Maekawa \cite{barnes2005}. Our derivation does not
238: appeal to an energy functional. Therefore, in addition to deriving
239: the correct form of the noise terms required for a consistent
240: treatment of thermal fluctuations, it also provides an alternative
241: justification of the zero-temperature approach of Tatara and
242: Kohno.
243:
244: These equations can be explored numerically without difficulty.
245: However, it turns out to be possible to make some analytic
246: progress. Solving for $\dot{\phi_0}$ alone specializing to purely
247: current-driven motion ($H_{\rm ext}=0$) we find that
248: \begin{equation}
249: \label{eq:langevinphionly}
250: (1 + \alpha^2) \dot \phi_0 = - \alpha \omega_{\rm o} \sin (2\phi_0)
251: + \frac{(\beta - \alpha) v_{\rm s}}{\lambda}+ \eta_{\phi} - \alpha \eta_{r}~.
252: \end{equation}
253: This is the equation of motion for an overdamped Brownian particle
254: which has been studied extensively with a large number of
255: different physical motivations. Our theory for the equation of
256: motion of the magnetization tilt should be contrasted with the
257: treatment in Ref.~\cite{tatara2005} which ultimately arrives at an
258: underdamped Brownian particle equation to describe nonzero
259: temperature domain wall motion.
260:
261: The average $\langle \dot \phi_0 \rangle$ can be calculated
262: exactly for overdamped Brownian motion in a tilted periodic
263: potential \cite{riskenbook}. Inserting this well known result in
264: Eq.~(\ref{eq:langevinvarpars}) we find that
265: \begin{widetext}
266: \begin{equation}
267: \label{eq:fulleqndwvelocity}
268: \langle \dot r_{\rm dw} \rangle
269: \!=\!\frac{\beta v_{\rm s}}{\alpha}-
270: \frac{ 2\pi \lambda k_{\rm B} T \left[1\!-\!e^{
271: 2 \pi \hbar N (\alpha-\beta) v_{\rm s}
272: /(\alpha \lambda k_{\rm B} T) }\right]}
273: {\hbar N \left\{\int_0^{2\pi}e^{ V(\phi)/(k_{\rm B} T)}
274: d\phi \int_0^{2\pi}e^{- V(\phi')/(k_{\rm B} T)} d\phi'
275: \!-\!\left[1\!-\!e^{
276: 2 \pi \hbar N (\alpha-\beta) v_{\rm s}
277: /(\alpha \lambda k_{\rm B} T) }\right] \int_0^{2\pi}e^{- V(\phi)/(k_{\rm B} T)}
278: d\phi \int_0^{\phi}e^{V(\phi')/(k_{\rm B} T)}
279: d\phi'\right\}}~,
280: \end{equation}
281: \end{widetext}
282: where
283: \begin{equation}
284: \label{eq:potentialphi} V(\phi_0)=- \hbar N \omega_{\rm o} \cos
285: (2\phi_0)/2+\hbar N(\alpha-\beta)v_{\rm s}\phi_0/(\alpha
286: \lambda)~,
287: \end{equation}
288: is the tilted-washboard potential experienced by the Brownian
289: ``$\phi_0$-particle'' \cite{tatara2004}. In the above equations we
290: have assumed that $\alpha^2 \ll 1$, for notational convenience.
291:
292: It is well-known \cite{riskenbook} that at zero temperature the
293: equation for an overdamped particle in the tilted periodic
294: potential of Eq.~(\ref{eq:langevinphionly}) has solutions with
295: $\langle \dot \phi \rangle \neq 0$ only if $|(\alpha - \beta )
296: v_{\rm s} |> \alpha \lambda \omega_{\rm o}$. For $\beta=0$ the
297: zero temperature result for the average velocity of the domain
298: wall is $\langle \dot r_{\rm dw} \rangle \propto \sqrt{(v_{\rm
299: s}/v_{\rm sc})^2-1}$ \cite{tatara2004,riskenbook}, where the
300: critical current for depinning the domain wall is $v_{\rm sc} =
301: \lambda \omega_{\rm o}$ \cite{tatara2004}. For $\beta \ne 0$, the
302: average domain wall velocity is nonzero at any finite value of
303: $v_s$ even at zero temperature.
304:
305:
306: \begin{figure}
307: \vspace{-0.5cm} \centerline{\epsfig{figure=v_j.eps}}
308: \caption{Domain wall velocity as a function of current,
309: for various temperatures and values of $\beta$. We take $\alpha=0.02$.
310: The horizontal axis is normalized to $v_{\rm sc}=\lambda \omega_{\rm o}$.}
311: \label{fig:v_j}
312: \end{figure}
313:
314: In Fig.~\ref{fig:v_j} we plot the average domain-wall velocity
315: calculated from the full expression
316: [Eq.~(\ref{eq:fulleqndwvelocity})] for various temperatures and
317: values of $\beta$. These curves were calculated for Gilbert
318: damping parameter $\alpha=0.02$. We consider only $\beta<\alpha$
319: since for $\beta>\alpha$ the domain wall velocity would decrease
320: with increasing temperature in clear disagreement with experiment
321: \cite{yamanouchi2006}. One of the main conclusions of
322: Ref.~\cite{tatara2005} is that for $\beta = 0$ and small currents
323: the average domain-wall velocity $\langle \dot r_{\rm dw} \rangle
324: \propto \exp(C v_{\rm s})$ where $C$ is a constant. In the
325: low-temperature limit we find that for $v_{\rm s}/v_{\rm sc} < 1$
326: the average domain wall velocity is given by
327: \begin{widetext}
328: \begin{equation}
329: \langle
330: \dot r_{\rm dw} \rangle
331: = \frac{\beta v_{\rm s}}{\alpha}
332: + 2 \sqrt{\left(\lambda \omega_{\rm o}\right)^2 - \left( v_{\rm s}\right)^2}
333: \exp\left\{ \left(\frac{\beta}{\alpha} -1 \right)
334: \frac{N \hbar \omega_{\rm o}}{k_{\rm B} T}\left[
335: \sqrt{1-\left( \frac{v_{\rm s}}{v_{\rm sc}}\right)^2} + \left(\frac{v_{\rm s}}{v_{\rm
336: sc}}\right)
337: \sin^{-1} \left(\frac{v_{\rm s}}{v_{\rm sc}} \right)\right]\right\}
338: \sinh \left( \frac{\pi \hbar N \left( \alpha - \beta \right) v_{\rm s}}{\alpha \lambda k_{\rm B} T}
339: \right)~.
340: \label{lowT}
341: \end{equation}
342: \end{widetext}
343: This difference in estimated domain wall velocities is closely
344: analogous to the difference between the Langer-Ambegaokar
345: \cite{LA} and Halperin-McCumber \cite{HMcC} estimates of phase
346: slip resistance in thin superconducting wires. It follows from
347: Eq.~(\ref{lowT}) that the observation of linear dependence of
348: domain wall velocity on current does not necessarily imply that
349: $\beta \neq 0$. A careful analysis of the temperature dependence
350: of $\langle \dot r_{\rm dw} \rangle$ will be necessary to
351: determine the value of $\beta$ from low current experiments,
352: especially so if $N \hbar \omega_{\rm o}$ is not very large
353: compared to $k_{\rm B} T$. Because of thermal noise, even when
354: $\beta=0$ the domain is not intrinsically pinned at any nonzero
355: temperature.
356:
357: \noindent {\it Comparison with experiment} --- The results
358: presented in Fig.~\ref{fig:v_j} look qualitatively similar to the
359: experimental results of Yamanouchi {\it et al.}
360: \cite{yamanouchi2006}, in particular to the inset in Fig.~3 of
361: Ref.~\cite{yamanouchi2006}. A direct comparison with these
362: experimental results is complicated by the fact that they are
363: performed close to the critical temperature, with the result that
364: the magnetic anisotropy energy-density and the polarization of the
365: current depend on temperature. Moreover, the Gilbert-damping
366: parameter $\alpha$ may also depend on temperature. The fact that
367: curves at different temperatures in Fig.~\ref{fig:v_j} are not as
368: strongly offset vertically from each other at $v_{\rm s}/v_{\rm
369: sc}>1$ as the experimental results is most likely due to these
370: additional temperature-dependent effects.
371:
372: \begin{figure}
373: \vspace{-0.5cm} \centerline{\epsfig{figure=v_sqrtj.eps}}
374: \caption{Test of the scaling $\ln \langle \dot r_{\rm dw} \rangle \propto \sqrt{v_{\rm s}}$,
375: for various temperatures and values of $\beta$. We take
376: $\alpha=0.02$.}
377: \label{fig:v_sqrtj}
378: \end{figure}
379:
380: One of the main results of Yamanouchi {\it et al.} is the
381: empirical finding that for small currents $\ln \langle \dot r_{\rm
382: dw} \rangle \propto \sqrt{ v_{\rm s}}$. From this the authors
383: conclude that the domain wall undergoes current-induced creep
384: motion at small currents. In Fig.~\ref{fig:v_sqrtj} we plot the
385: average domain wall velocity as a function of $\sqrt{v_{\rm s}}$
386: for various temperatures and values of $\beta$. From this we
387: observe that $\ln \langle \dot r_{\rm dw} \rangle$ is proportional
388: to the square root of the current over several decades of domain
389: wall velocity. It turns out that this approximate relationship is
390: valid for a larger range of parameter space with decreasing
391: temperature. In the experiments of Yamanouchi {\it et al.} a
392: typical velocity is $\langle \dot r_{\rm dw} \rangle \sim 1$ m
393: s$^{-1}$. Putting this equal to $\lambda \omega_{\rm o}$ we have
394: for $\lambda = 17$ nm and $T \sim 100$ K that $k_{\rm B} T/(N
395: \hbar \omega_{\rm o}) \sim 0.01$ for $N=3.4\times10^6$ Mn moments
396: (we take the density of moments $N_{\rm Mn}\sim 1$ nm$^{-3}$) in a
397: wall of cross-sectional area $A=2$ nm $\times$ $5$ $\mu$m. For
398: this value of $k_{\rm B} T/(N \hbar \omega_{\rm o})$ we conclude
399: from Fig.~\ref{fig:v_sqrtj} that $\ln \langle \dot r_{\rm dw}
400: \rangle \propto \sqrt{v_{\rm s}}$ over a large regime, and that
401: our theory is therefore in agreement with the experimental results
402: of Yamanouchi {\it et al.} \cite{yamanouchi2006}. Moreover, from
403: Fig.~\ref{fig:v_sqrtj} we also conclude that for $\beta=0.01$ this
404: scaling does not hold. Hence, an additional conclusion from the
405: experimental data is that $\beta$ is much smaller than $\alpha$
406: in these ferromagnetic semiconductors.
407:
408: A similar estimate for typical parameters of the metallic
409: nanowires used in studies of current-driven domain wall motion
410: \cite{grollier2003,tsoi2003,yamaguchi2004,klaui2005} leads to the
411: conclusion that $k_{\rm B} T/(N\hbar\omega_{\rm o})\sim0.0002$ and
412: therefore that temperature effects on rigid domain wall motion are
413: likely much less important in these systems. This difference
414: arises mainly because the density of moments is roughly $40$ times
415: higher in metallic systems.
416:
417: In conclusion, we have presented a theory of the influence of
418: nonzero temperatures on current-driven motion of a rigid domain
419: wall, and found qualitative agreement with ferromagnetic
420: semiconductor experiments. An estimate of the bending energy of
421: the domain wall shows that these degrees of freedom are also
422: thermally accessible. It is, however, difficult to assess how they
423: influence translation of the domain wall. The qualitative
424: agreement of our results with the experiments of Yamanouchi {\it
425: et al.} \cite{yamanouchi2006} indicates that thermal activation of
426: rigid domain wall motion could play an important role in these
427: experiments. We expect that accounting for thermal fluctuations
428: will also be important in assessing the impact of intended and
429: unintended extrinsic domain wall pinning. This work was
430: supported by the National Science Foundation under grants
431: DMR-0115947 and DMR-0210383, by a grant from Seagate Corporation,
432: and by the Welch Foundation.
433:
434:
435:
436:
437: \begin{thebibliography}{99}
438: %%% First by Berger:
439: \bibitem{berger1984} L. Berger, J. Appl. Phys. {\bf 55}, 1954
440: (1984).
441: \bibitem{freitas1985} P. P. Freitas and L. Berger, J. Appl. Phys. {\bf 57}, 1266
442: (1985).
443:
444:
445: %%% CI-DW theory
446: \bibitem{tatara2004} G. Tatara and H. Kohno, Phys. Rev. Lett. {\bf 92}, 086601
447: (2004); Phys. Rev. Lett. {\bf 96}, 189702 (2006).
448: \bibitem{zhang2004} S. Zhang and Z. Li, Phys. Rev. Lett. {\bf 93}, 127204 (2004).
449: \bibitem{thiaville2004} A. Thiaville, Y. Nakatani, J. Miltat, and Y
450: Suzuki, Europhys. Lett. {\bf 69}, 990 (2005).
451: \bibitem{barnes2005} S. E. Barnes and S. Maekawa, Phys. Rev. Lett. {\bf 95},
452: 107204 (2005); S. E. Barnes, Phys. Rev. Lett. {\bf 96}, 189701
453: (2006).
454: \bibitem{tserkovnyak2005} Y. Tserkovnyak, A. Brataas, and G. E. W.
455: Bauer, cond-mat/0512715.
456: \bibitem{kohno2006} H. Kohno, G. Tatara, and J. Shibata,
457: cond-mat/0605186.
458: \bibitem{nguyen2006} A. K. Nguyen, H. J. Skadsem, and A. Brataas,
459: cond-mat/0606498.
460:
461:
462: %%% CI-DW expt metals
463: \bibitem{grollier2003} J. Grollier, P. Boulenc, V. Cros, A. Hamzi, A. Vaurès,
464: A. Fert, and G. Faini, Appl. Phys. Lett. {\bf 83}, 509 (2003).
465: \bibitem{tsoi2003} M. Tsoi, R.E. Fontana, and S.S.P. Parkin, Appl. Phys. Lett. {\bf 83}, 2617
466: (2003).
467: \bibitem{yamaguchi2004} A. Yamaguchi, T. Ono, S. Nasu, K. Miyake, K. Mibu, and T.
468: Shinjo, Phys. Rev. Lett. {\bf 92}, 077205 (2004).
469: \bibitem{klaui2005} M. Kl\"aui, C. A. F. Vaz, J. A. C. Bland, W. Wernsdorfer,
470: G. Faini, E. Cambril, L. J. Heyderman, F. Nolting, and U.
471: R\"udiger, Phys. Rev. Lett. {\bf 94}, 106601 (2005).
472:
473: %%% CD-DW expt FMSCs
474: \bibitem{yamanouchi2004} M. Yamanouchi, D. Chiba, F. Matsukura, and H.
475: Ohno, Nature {\bf 428}, 539 (2004).
476: \bibitem{yamanouchi2006} M. Yamanouchi, D. Chiba, F. Matsukura, T. Dietl, and H.
477: Ohno, Phys. Rev. Lett. {\bf 96}, 096601 (2006).
478:
479: %%% STT Theory
480: \bibitem{slonczewski1996} J.C. Slonczewski, J. Mag. Mat. Mag. {\bf 159}, L1
481: (1996).
482: \bibitem{berger1996} L. Berger, Phys. Rev. B {\bf 54}, 9353 (1996).
483:
484: %%% STT Experiment
485: \bibitem{tsoi1998} M. Tsoi, A. G. M. Jansen, J. Bass, W.-C. Chiang, M. Seck, V. Tsoi, and P.
486: Wyder, \prl {\bf 80}, 4281 (1998).
487: \bibitem{myers1999} E. B. Myers, D. C. Ralph, J. A. Katine, R. N. Louie, R.
488: A. Buhrman, Science {\bf 285}, 867 (1999).
489:
490: %%% Spintronics general
491: \bibitem{wolf2001} S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S.
492: von Moln\'ar, M. L. Roukes, A. Y. Chtchelkanova, D. M. Treger,
493: Science {\bf 294}, 1488 (2001).
494:
495: %%% Tatara finite T
496: \bibitem{tatara2005} G. Tatara, N. Vernier, and J. Ferr\'e, Appl. Phys. Lett. {\bf 86},
497: 252509 (2005).
498:
499: %%% General References Stochastic LLG
500: \bibitem{brown1963} W.F. Brown, Jr., Phys. Rev. {\bf 130}, 1677
501: (1963).
502: \bibitem{kubo1970} R. Kubo and N. Hashitsume, Prog. Theor. Phys.
503: Suppl. {\bf 46}, 210 (1970).
504: \bibitem{ettelaie1984} R. Ettelaie and M.A. Moore, J. Phys. A {\bf
505: 17}, 3505 (1984).
506: \bibitem{garciapalacios1998} J. L. Garc\'ia-Palacios and
507: F. J. L\'azaro, \prb {\bf 58}, 14937 (1998).
508: \bibitem{heinonen2004} O.G. Heinonen and H.S. Cho,
509: IEEE Transactions on Magnetics {\bf 40}, 2227 (2004).
510: \bibitem{safonov2005} V. L. Safonov and H. N. Bertram, \prb {\bf
511: 71}, 224402 (2005).
512: \bibitem{rossi2005} E. Rossi, O. G. Heinonen, and A.H. MacDonald,
513: \prb {\bf 72}, 174412 (2005).
514:
515: %%% effective T refs
516: \bibitem{foros2005} J\o rn Foros, A. Brataas, Y. Tserkovnyak, and G. E. W.
517: Bauer, \prl {\bf 95}, 016601 (2005).
518: \bibitem{duine2006} R.A. Duine, A.S. N\'u\~nez, and A.H. MacDonald, to
519: be submitted.
520:
521: %%% ZJ book
522: \bibitem{zinnjustinbook} J. Zinn-Justin, {\it Quantum Field Theory
523: and Critical Phenomena} (Oxford, New York, 1989).
524:
525: %%% stochastic business with PI's
526: \bibitem{duine2002} R.A. Duine and H.T.C. Stoof, Phys. Rev. A {\bf 65}, 013603
527: (2002).
528:
529: %%% Risken book
530: \bibitem{riskenbook} H. Risken, {\it The Fokker-Planck Equation}
531: (Springer-Verlag, Berlin, 1984).
532:
533:
534: %%% relation to super wires
535: \bibitem{LA} J.S. Langer and V. Ambegaokar, Phys. Rev. {\bf 164}, 498 (1967).
536:
537: \bibitem{HMcC} D.E. McCumber and B.I. Halperin, Phys. Rev. B {\bf 1}, 1054 (1970).
538:
539:
540:
541:
542:
543:
544:
545:
546: \end{thebibliography}
547:
548:
549:
550:
551: \end{document}
552: