cond-mat0601185/KBNN.tex
1: \documentclass[prl,twocolumn,showpacs,floatfix,amsfonts,superscriptaddress]{revtex4}
2: 
3: 
4: \usepackage{graphicx,graphics,color,epsfig}% Include figure files
5: \usepackage{bm}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: %\usepackage[showrefs]{refcheck}
9: 
10: 
11: \newcommand{\bS}{{\bf S}}
12: \newcommand{\bd}{{\bf d}}
13: \newcommand{\bB}{{\bf B}}
14: \newcommand{\bQ}{{\bf Q}}
15: \newcommand{\bk}{{\bf k}}
16: \newcommand{\bv}{{\bf v}}
17: \newcommand{\bl}{{\bf l}}
18: \newcommand{\br}{{\bf r}}
19: \newcommand{\bp}{{\bf p}}
20: \newcommand{\s}{{\sigma}}
21: \newcommand{\bA}{{\bf A}}
22: \newcommand{\ba}{{\bf a}}
23: \newcommand{\bJ}{{\bf J}}
24: \newcommand{\bE}{{\bf E}}
25: \newcommand{\bb}{{\bf b}}
26: \newcommand{\bsig}{\mbox{\boldmath{$\sigma$}}}
27: \newcommand{\Det}{\,{\rm{Det}\,}}
28: \newcommand{\Tr}{\,{\rm{Tr}\,}}
29: \newcommand{\tr}{\,{\rm{tr}\,}}
30: \newcommand{\bn}{{\bf \nabla}}
31: 
32: 
33: \newcommand{\sign}{\,{\rm{sign}\,}}
34: \newcommand{\bomega}{\bar{\Omega}}
35: \newcommand{\wL}{\omega_{L}}
36: \newcommand{\COMMENT}{\large\bf}
37: \newcommand{\FigWidth}{\columnwidth}
38: \newcommand{\beqa}{\begin{eqnarray}}
39: \newcommand{\eeqa}{\end{eqnarray}}
40: \newcommand{\bI}{\bf{I}_s}
41: 
42: 
43: \renewcommand{\Re}{{\rm Re}}
44: \renewcommand{\Im}{{\rm Im}}
45: 
46: 
47: \usepackage{mathptm}    %clears up some problems for generating PDF file
48: \usepackage{dcolumn}                    % Align table columns on decimal point
49: \usepackage{bm}                         % bold math
50: \usepackage{graphicx}
51: 
52: 
53: 
54: \usepackage{times}
55: 
56: 
57: \newcommand{\PRG}{PuRhGa$_5$}
58: \newcommand{\PCG}{PuCoGa$_5$}
59: 
60: 
61: \begin{document}
62: 
63: 
64: \title{Voltage dependence of Landau-Lifshitz-Gilbert damping of a spin in
65: a current driven tunnel junction}
66: 
67: 
68: 
69: 
70: 
71: 
72: \author{Hosho Katsura}
73: \email[Electronic address: ]{katsura@appi.t.u-tokyo.ac.jp}
74: \affiliation{Department of Applied Physics, University of Tokyo,
75: Hongo, Bunkyo-ku, Tokyo 113-8656, Japan}
76: 
77: 
78: \author{Alexander V. Balatsky}
79: \email{avb@lanl.gov}  \affiliation{Theoretical Division, Los Alamos
80: National Laboratory, Los Alamos, New Mexico 87545, USA}
81: 
82: 
83: \author{Zohar Nussinov}
84: \email{zohar@wuphys.wustl.edu}
85: \affiliation{Department of Physics, Washington University, St.
86: Louis, MO 63160, USA}
87: 
88: 
89: \author{Naoto Nagaosa}
90: \affiliation{Department of Applied Physics, University of Tokyo,
91: Hongo, Bunkyo-ku, Tokyo 113-8656, Japan} \affiliation{CERC, AIST
92: Tsukuba Central 4, Tsukuba 305-8562, Japan} \affiliation{CREST,
93: Japan Science and Technology Agency (JST)}
94: 
95: 
96: 
97: \begin{abstract}
98: We present a theory of Landau-Lifshitz-Gilbert damping $\alpha$ for
99: a localized  spin ${\vec S}$ in the junction coupled to the
100: conduction electrons in both leads under an applied volatege $V$. We
101: find the voltage dependence of the damping term reflecting the
102: energy dependence of the density of states. We find the  effect is
103: linear in the voltage and cotrolled by particle-hole asymmetry of
104: the leads.
105: \end{abstract}
106: 
107: 
108: \pacs{75.80.+q, 71.70.Ej, 77.80.-e}
109: 
110: 
111: \date{\today}
112: \maketitle
113: 
114: 
115: 
116: \section{Introduction}
117: 
118: 
119: Spintronics is an emerging subfield that holds the potential to
120: replace conventional electronic devices with spintronic analogues
121: where the manipulation, control, and readout of spins will
122: enable novel functionality with no or little electronic charge
123: dynamics \cite{spintronics}. In order to realize this
124: promise, the spin dynamics of the small scale devices needs to be well
125: controlled. One of the most pressing questions concerns
126: a set up which would preserve coherence and allow a manipulation of spins.
127: In most systems, the relevant spin degrees of freedom are coupled to
128: some bath, such as a fermionic bath of electrons. The detailed
129: dynamics of single spins when in contact with such a bath
130: plays a pivotal role in addressing decoherence in potential
131: spintronic systems.
132: 
133: 
134: The conventional way to treat this problem is via a Caldeira-Leggett
135: approach where the external bath is modeled by collective
136: excitations which are capable of destroying coherent spin dynamics.
137: Often, spin dynamics is described by a Landau-Lifshitz-Gilbert equation
138: \cite{ll1,ll2}:
139: \begin{equation} {\frac {d{\vec S(t)}}{dt}} = - {\vec
140: S(t)}\times{\vec h} -\alpha {\vec S(t)}\times{\frac{d{\vec
141: S(t)}}{dt}}, \label{LLG2}
142: \end{equation}
143: where ${\vec h}$ is, up to constant prefactors, the external
144: magnetic field and the coefficient $\alpha$ captures the damping due
145: to the external bath. A caricature of the solution of this equation
146: \cite{Chika} is provided in Fig.\ref{FIG:FIG2}. There are standard
147: methods to calculate $\alpha$ in an equilibrium situation when, say,
148: one considers a spin in a Fermi liquid \cite{Heinrich,Halperin}.
149: 
150: In the current publication, we address a related novel question
151: concerning the effect of an applied voltage bias on the Gilbert
152: coefficient $\alpha$. Our work complements the recent results of
153: \cite{Onoda} wherein the effects of the ``retarded'' electronic
154: contributions in the equations of motion for a system of spins were
155: studied. Both such retarded correlations \cite{Zhu03} as well as
156: additional ``Keldysh'' correlations generally manifest themselves in
157: the single spin equations of motion, see e.g. \cite{nsabz} for
158: general spin equations of motion entailing the effects of both
159: correlations. In the current work, we examine the voltage dependence
160: of Gilbert damping. For the sake of clarity, we depart from the
161: Keldysh contour formalism of \cite{Onoda,Zhu03,nsabz}, and use
162: a Caldeira-Leggett approach.
163: 
164: 
165: In what follows, we consider the case of a junction between two
166: electrodes that contains one spin $\vec{S}$, see Fig.\ref{FIG:FIG1}.
167: This spin $\vec{S}$ may be the spin of a single magnetic impurity or
168: it may portray the spin of a cluster at low temperature when the
169: spins in the cluster are locked. Upon applying a finite bias between
170: the electrodes of Fig.\ref{FIG:FIG2}, a current flow is generated.
171: Thereafter, at long times, the system is at a steady but
172: non-equilibrium state so long as the voltage bias $V$ is applied. We
173: will focus on the voltage dependence of the damping term $\alpha(V)$
174: in Eq.(1). We find that the change in the density of states
175: associated with the chemical potential gradient across the junction
176: triggers a modification to the damping $\alpha$ that is {\em linear}
177: in voltage and is proportional to the particle-hole assymmetry of
178: the density of states. The scale of the correction is set by the
179: Fermi energy of the metal in the leads $E_F$ and by particle-hole
180: asymmetry in the density of states:
181:  \beqa \alpha(V) = \alpha_0+\alpha_1(V)
182: = \alpha_0 ( 1+ O(1)eV/E_F). \label{voltage+} \eeqa
183:  This result
184: vividly illustrates the presence of voltage induced damping in such
185: junctions. Spin unpolarized electrons tunneling across the junction
186: interact via exchange interaction with the spin $\vec{S}$ and
187: produce random magnetic fields that disorder the local spin. This
188: noise augments that already present equilibrium magnetic noise in a
189: Fermi liquid bath. Such a behavior of $\alpha(V)$ with the external
190: voltage is in line with the works of \cite{hooley}. An analysis of
191: a related single spin problem in a Josephson junction (instead of
192: the normal junction studied here) was advanced in
193: \cite{Zhu03,nsabz,bulaevskii}.
194: 
195: 
196: We will shortly derive the effective single spin action from which
197: the principle equation of motion of Eq.(\ref{LLG2}) follows. Several
198: technical details of our derivation are given in the appendices.
199: 
200: %Below we will present our results. We will start with the
201: %Hamiltonian and obtain effective action get equations of motion
202: %etc.
203: 
204: 
205: 
206: 
207: \begin{figure}
208: \centerline{\includegraphics[width=0.9\columnwidth]{fig2.eps}}
209: \caption{Sketch of the dissipative spin dynamics.
210: Panel (a) depicts a cartoon of the Larmor
211: precession of the spin about the direction of an applied magnetic
212: field (B). In panel (b), a caricature of the spin dynamics in the
213: presence of Landau-Lifshitz-Gilbert damping is shown.}
214: \label{FIG:FIG2}
215: \end{figure}
216: 
217: 
218: \begin{figure}
219: \centerline{\includegraphics[width=0.9\columnwidth]{fig1.eps}}
220: \caption{Magnetic impurity coupled to two electrodes. $\mu_{\rm L}$
221: and $\mu_{\rm R}$ denotes the chemical potentials of the left and
222: right leads respectively.  The voltage drop across the two
223: electrodes is $eV= \mu_{L} - \mu_{R}$.} \label{FIG:FIG1}
224: \end{figure}
225: 
226: 
227: 
228: 
229: 
230: 
231: 
232: \section{THE SYSTEM AND OUR PRINCIPLE RESULTS}
233: 
234: 
235: The physical system under consideration in this publication is
236: illustrated in Fig.\ref{FIG:FIG1}. It consists of two (left(L) and
237: right(R)) electrodes across which a voltage bias is applied; a
238: magnetic impurity ($\vec{S}$) is situated in between (or lies on one
239: of) the electrodes. An external magnetic field ${\vec{B}}$ is
240: present. In the absence of effects stemming from conduction
241: electrons in the tunneling barrier, the single spin would precess at
242: the Larmor precession frequency about the applied field direction
243: (see, e.g. panel (a) of Fig.\ref{FIG:FIG2}). With the external
244: circuit elements present, the spin motion becomes dissipative (as
245: schematically shown in panel (b) of Fig.\ref{FIG:FIG2}).
246: 
247: 
248: With the spin embedded in the tunneling barrier, the work function
249: is modified and the conduction electron tunneling matrix element is
250: supplanted by a Kondo like exchange term $ J(\vec{S} \cdot
251: \vec{\sigma}_{c})$, with $\vec{\sigma}_{c}$ denoting the conduction
252: electron spin. In what follows, we will dispense with the $c$
253: subscript. The Hamiltonian governing this system is given by
254: \begin{eqnarray}
255: {\cal H} &=& {\cal H}_e + {\cal H}_s + {\cal H}_T,
256: \\
257: {\cal H}_e &=& \sum_{\alpha,k,\sigma} \xi_{\alpha k} c_{\alpha k
258: \sigma}^{\dagger}c_{\alpha k \sigma}, \nonumber
259: \\
260: {\cal H}_s &=& - {\vec h} \cdot {\vec S(t)}, \nonumber
261: \\
262: {\cal H}_T &=& {\frac{1}{\Omega}}{\sum_{\alpha,k,\sigma}}{\sum_{\beta,p,\sigma'}}
263: c_{\alpha k \sigma}^{\dagger} (T_{\alpha\beta})_{\sigma \sigma'}
264: c_{\beta p \sigma'}, \label{Ham}
265: \end{eqnarray}
266: where $c_{\alpha k \sigma}^{\dagger}~(c_{\alpha k \sigma})$ creates
267: (annihilates) an electron with momentum $k$ and spin $\sigma \in \{
268: \uparrow,\downarrow \}$ in the lead $\alpha \in \{{\rm L}, {\rm
269: R}\}$. The abbreviation $\xi_{\alpha k} = \epsilon_{\alpha k}
270: -\mu_{\alpha}$, where $\epsilon_{\alpha k}$ is the energy of
271: electron with momentum $k$ in (the lead) $\alpha$ and $\mu_{\alpha}$
272: is the chemical potential in (the lead) $\alpha$. The second term in
273: Eq.(\ref{Ham}), ${\cal H}_s$, is Zeeman energy of the spin in an
274: external magnetic field $\vec B$. Here, $\vec h \equiv g \mu_B {\vec
275: B}$ with $g$ gyromagnetic ratio and $\mu_B$ the Bohr magneton. The
276: last term in Eq.(\ref{Ham}), ${\cal H}_T$, represents both Kondo
277: coupling and direct tunneling process, where the amplitudes
278: $\{T_{\alpha \beta}\}=\{ T_{{\rm L}{\rm L}},T_{{\rm L}{\rm
279: R}},T_{{\rm R}{\rm L}},T_{{\rm R}{\rm R}} \}$ are tunneling matrix
280: elements and their explicit forms are
281: \begin{eqnarray}
282: T_{LL} &=& J_{LL}({\vec \sigma} \cdot {\vec S(t)}), \nonumber
283: \\
284: T_{RR} &=& J_{RR}({\vec \sigma} \cdot {\vec S(t)}), \nonumber
285: \\
286: T_{LR} &=& T_{RL} = (T_0 \delta_{\sigma \sigma'} + J_{LR}({\vec
287: \sigma} \cdot {\vec S(t)})). \label{tunnel_amp}
288: \end{eqnarray}
289: Here, $T_0$ is the direct tunneling matrix element and $J_{\alpha
290: \beta}$ is the Kondo coupling, 'while $\Omega$ denotes the Volume of
291: each lead (assumed, for simplicity, to be the same)'. Typically, from
292: the expansion of the work function for tunneling,
293: ${J_{LR}}/{T_0} \sim J/U$, where $U$ is the height of a
294: spin-independent tunneling barrier and $J$ the magnitude of the spin
295: exchange interaction ~\cite{Zhu_Balatsky}. Typical values of the
296: ratio the spin dependent to spin independent tunneling amplitudes in
297: Eq.(\ref{tunnel_amp}), $J_{\alpha \beta}/T_{0}$, are
298: ${\cal{O}}(10^{-1})$, with a typical Fermi energy $E_{F}^{R}$ of the
299: order of several electron-volts. From the Hermiticity of the
300: Hamiltonian, we can  find that the matrix element $(T_{\alpha
301: \beta})_{\sigma \sigma'}$ satisfies $((T_{\alpha \beta})_{\sigma
302: \sigma'})* = (T_{\beta \alpha})_{\sigma' \sigma}.$
303: 
304: 
305: In the up and coming, we derive the effective action for the single
306: impurity spin via an imaginary time path integral formalism. The
307: full action is given by
308: \begin{eqnarray}
309: S = \int_0^{\beta}d{\tau} \sum_{\alpha k \sigma} c_{\alpha k
310: \sigma}^{\dagger} \partial_{\tau} c_{\alpha k \sigma} +
311: iS\omega(\vec S(\tau)) +\int_0^{\beta}d{\tau}{\cal H}(\tau),
312: \label{action}
313: \end{eqnarray}
314: where the second, Wess-Zumino-Novikov-Witten(WZNW), term in Eq.(\ref
315: {action}) depicts the Berry phase accumulated by the spin
316: \cite{Fradkin}. In our action, we have the following quadratic form
317: of fermions,
318: \begin{eqnarray}
319: && \int_0^{\beta} d{\tau} {\frac{1}{\Omega}} \sum_{\alpha k \sigma}
320: \sum_{\beta p \sigma'}  c_{\alpha k \sigma}^{\dagger}
321: (\delta_{\alpha \beta} \delta_{\sigma \sigma'}(\partial_{\tau} +
322: \xi_{\alpha k}) + (T_{\alpha \beta})_{\sigma \sigma'}) c_{\beta p
323: \sigma'} \nonumber
324: \\
325: && \equiv \int_0^{\beta} d{\tau} {\frac{1}{\Omega}} \sum_{\alpha k
326: \sigma} \sum_{\beta p \sigma'} c_{\alpha k \sigma}^{\dagger}
327: ((M_{\alpha \beta}(\tau))_{\sigma \sigma'})_{kp} ~ c_{\beta p \sigma'}.
328: \label{quad}
329: \end{eqnarray}
330: We may integrate over the lead electrons to obtain the effective
331: action for the spin
332: \begin{equation}
333: S_{\rm {eff}}(\vec S(\tau)) \sim i S \omega({\vec S(\tau)}) +
334: \int_0^{\beta} d{\tau} {\vec h} \cdot {\vec S(\tau)} -{\rm{ln}}{\rm {det}} M , \label{eff}
335: \end{equation}
336: where ${\rm {det}} M$ means \textit{functional determinant} of $M.$
337: From the third term in Eq.(\ref {eff}), we obtain a quadratic
338: non-local in time interaction of the spin with itself, $\vec
339: S(\tau)$ as
340: \begin{equation}
341: \Delta S(\vec S(\tau)) = -2 \int d{\tau} \int d{\tau'} {\vec S(\tau)}
342: \cdot {\vec S(\tau')} K(\tau-\tau'), \label{K+}
343: \end{equation}
344: where
345: \begin{eqnarray}
346: K(\tau) &=& \sum_{\alpha, \beta \in \{ \rm{L,R} \}} K_{\alpha
347: \beta}(\tau), \nonumber
348: \\
349: K_{\alpha \beta} &=& \sum_{k \in \alpha} \sum_{p \in \beta}
350: {\frac{J_{\alpha \beta} J_{\beta \alpha}}{2}} {\frac{1}{\beta}}
351: \sum_{\omega_m} {\frac {f(\xi_k)-f(\xi_p)}{i\omega_m +\xi_k -\xi_p}}
352: e^{-i \omega_m \tau}, \label{Kt}
353: \end{eqnarray}
354: with $f(\xi)$ denotes the Fermi distribution function (see APPENDIX
355: A). The effective action $\Delta S$ of Eq.(\ref{eff}) can be
356: decomposed into two (trivial and non-trivial) components as $\Delta
357: S = \Delta S_{\rm loc} + \Delta S_{\rm{dis}}$, with
358: \begin{eqnarray}
359: \Delta S_{\rm loc} &=& -2K(\omega=0) \int d{\tau} ({\vec S(\tau)})^2,
360: \nonumber
361: \\
362: \Delta S_{\rm{dis}} &=& \int d{\tau} \int d{\tau}' ({\vec
363: S(\tau)}-{\vec S({\tau})})^2 K(\tau-{\tau}'). \label{locdis}
364: \end{eqnarray}
365: Here, $K(\omega=0)$ is the zero-frequency Fourier component of
366: $K(\tau)$. The first term ($\Delta S_{\rm loc}$) is a trivial
367: constant as $S(\tau)^2 = S^2$. The nonlocal part ($\Delta
368: S_{\rm{dis}}$) represents the dissipative effect due to the coupling
369: between $S(\tau)$ and electrons bath. The integral kernel $K(\tau)$
370: is calculated in the same way as the Caldeira-Leggett theory
371: \cite{Leggett,Leggett2} leading to
372: \begin{equation}
373: K(\tau) = \int_0^{\infty} d{\omega} J(\omega) \frac{{\rm
374: cosh}({{\beta}/2}-|\tau|){\omega}}{{\rm sinh}({\beta \omega}/2)},
375: \label{K}
376: \end{equation}
377: where $J(\omega)$ is the \textit{spectral density} and its explicit
378: form is
379: \begin{equation}
380: J(\omega) = \sum_{\alpha \beta} \sum_{k \in \alpha} \sum_{p \in
381: \beta} {\frac{J_{\alpha \beta} J_{\beta \alpha}}{2}} [f(\xi_k)
382: -f(\xi_p)] \delta (\omega+ \xi_k -\xi_p). \label{spec}
383: \end{equation}
384: The details of the derivation of Eq.(\ref{spec}) are provided in
385: APPENDIX B.
386: The spectral density of Eq.(\ref{spec}), $J(\omega)$, is estimated
387: as
388: \begin{eqnarray}
389: J(\omega) &=& \sum_{\alpha \beta} {\frac{J_{\alpha \beta} J_{\beta
390: \alpha}}{2}} \int_{E_F^{\alpha}}^{\infty} d{\xi_{\alpha}}
391: N(\xi_{\alpha})\int_{E_F^{\beta}}^{\infty} d{\xi_{\beta}} N(\xi_{\beta})
392: \nonumber
393: \\
394: && \times [f(\xi_{\alpha})-f(\xi_{\beta})]
395: \delta(\omega-\xi_{\alpha}-\xi_{\beta}) \nonumber
396: \\
397: &\sim& \sum_{\alpha \beta}{\frac{J_{\alpha \beta} J_{\beta
398: \alpha}}{2}} N(\xi_{\alpha}=0) N(\xi_{\beta}=0) \omega,
399: \label{spec2}
400: \end{eqnarray}
401: where $E_F^{\alpha}$ denotes the Fermi Energy of the lead $\alpha$.
402: It is obvious that $J(\omega)$ in Eq.(\ref{spec2}) is proportional
403: to $\omega$, i.e., $J(\omega)$ is Ohmic. If spectral density is
404: expressed as $J(\omega)={\eta \omega}/{2 \pi}$, then
405: the Gilbert coefficient $\alpha$ in Eq.(\ref{LLG2}) is exactly equal
406: to $\eta$. By varying the total action with respect to the spin
407: ${\delta S}/{\delta {\vec S(\tau)}}=0$, we immediately obtain the
408: Landau-Lifshitz-Gilbert equation with $\alpha={\frac{2
409: \pi}{\omega}}J(\omega)$(see APPENDIX C). In other words, the voltage
410: dependence of $\alpha$ in Eq.(\ref{LLG2}) is identically the same as
411: that of $J(\omega)$. We next examine the voltage dependence of
412: $J(\omega)$.
413: 
414: If we apply a voltage leading to a chemical drop of
415: $\Delta\mu_L-\Delta\mu_R=eV$. Assuming, for example,that
416:  the
417: net charge on both right and left leads is unchanged, we also have
418:  $D^L(E_F)\Delta\mu_L+D^R(E_F)\Delta\mu_R=0$.
419: With these constraints we get
420: \begin{eqnarray}
421: \Delta\mu_L&=&\frac{D^R(E_F)}{D^L(E_F)+D^R(E_F)}eV,
422: \nonumber \\
423: \Delta\mu_R&=&-\frac{D^L(E_F)}{D^L(E_F)+D^R(E_F)}eV,
424: \end{eqnarray}
425: the Gilbert coefficient $\alpha$ may be approximated as\\
426: \begin{eqnarray}
427: \alpha(V) &\sim& 2\pi\Bigl(\frac{J_{LL}^2}{2} D^L(E_F+\Delta\mu_L)^2 + \frac{J_{RR}^2}{2}
428: D^R(E_F+\Delta\mu_R)^2 \nonumber\\
429: &&+J_{LR}^2 D^L(E_F+\Delta\mu_L) D^R(E_F+\Delta\mu_R)\Bigr)
430: \nonumber
431: \end{eqnarray}
432: \begin{eqnarray}
433: &\sim& 2\pi \Bigl(
434: \frac{[J_{LL}D^L(E_F)]^2}{2} + \frac{[J_{RR} D^R(E_F)]^{2}}{2} + J^2_{LR}D^L(E_F)D^R(E_F)\nonumber \\
435: &&+(J^2_{LL}D^L(E_F)\frac{\partial{D^L(E_F)}}{\partial{E_F}}
436: +   J^2_{LR}D^R(E_F)\frac{\partial{D^L(E_F)}}{\partial{E_F})})\Delta\mu_L
437: \nonumber \\
438: &&+(J^2_{RR}D^R(E_F)\frac{\partial{D^R(E_F)}}{\partial{E_F}}
439: +   J^2_{LR}D^L(E_F)\frac{\partial{D^R(E_F)}}{\partial{E_F})})\Delta\mu_R)
440: \Bigr) \nonumber
441: \end{eqnarray}
442: \begin{eqnarray}
443: &=&2\pi \Bigl(
444: \frac{[J_{LL}D^L(E_F)]^2}{2} + \frac{[J_{RR}D^R(E_F)]^{2}}{2}
445: + J^2_{LR}D^L(E_F)D^R(E_F)\nonumber \\
446: &+& \frac{eV}{D^L(E_F)+D^R(E_F)}
447: (J^2_{LL}D^L(E_F)D^R(E_F)\frac{\partial{D^L(E_F)}}{\partial{E_F}}
448: \nonumber \\
449: &-& J^2_{RR}D^L(E_F)D^R(E_F)\frac{\partial{D^R(E_F)}}{\partial{E_F}}
450: \nonumber \\
451: &+& J^2_{LR}D^R(E_F)D^R(E_F)\frac{\partial{D^L(E_F)}}{\partial{E_F}}
452: \nonumber \\
453: &-& J^2_{LR}D^L(E_F)D^L(E_F)\frac{\partial{D^R(E_F)}}{\partial{E_F}})\Bigr).
454: \label{Gil}
455: \end{eqnarray}
456: The change in the density of states associated with the chemical
457: potential gradient across the junction triggers a modification of
458: the damping $\alpha$ that is linear in voltage. For typical Fermi
459: energy $E_{F}^{L/R}$ of the order of several electron-volts, the
460: voltage dependence of $\alpha$ may become very notable. This voltage
461: driven effect may be expressed in terms of $\alpha_0$ and
462: $\alpha_1(V)$ with $ \alpha(V)=\alpha_0 + \alpha_1$
463: (Eq.(\ref{voltage+})). Here \begin{eqnarray}
464: \alpha_0= \pi \bigl(
465: J_{LL}^2[D(E_F^L)]^2 + 2J_{LR}^2 D(E_F^L)
466: D(E_F^R)+J_{RR}^2[D(E_F^R)]^2 \bigl), \nonumber
467: \\
468: \alpha_1= \frac{1}{4 e|T_{0}|^{2}[D(E_F^L) + D(E_F^R)]} \Bigl(
469: \frac{I_{o}^{L}}{[D(E_F^L)]^2} + \frac{I_{o}^{R}}{[D(E_F^R)]^2}
470: \Bigl) \nonumber
471: \\ \times \bigl[ D(E_F^L)
472: D(E_F^R) \bigl( J_{LL}^{2} \frac{\partial D^{L}(E_F)}{\partial
473: E_{F}} - J_{RR}^{2} \frac{\partial D^{R}(E_F)}{\partial E_{F}}
474: \bigl) \nonumber
475: \\+ J_{LR}^{2} \bigl( [D(E_F^R)]^2 \frac{\partial D^{L}(E_F)}{\partial E_{F}}
476: - [D(E_F^L)]^2 \frac{\partial D^{R}(E_F)}{\partial E_{F}} \bigl)
477: \bigl].
478: %\alpha_0
479: %{\frac{2(J_{LL}^2+J_{LR}^2)}{J_{LL}^2+2J_{LR}^2+J_{RR}^2}} =I_o
480: %{\frac{1}{eE_F^R}} {\frac{J_{LL}^2+J_{LR}^2}{2|T_0|^2}}.
481: \label{central++}
482: \end{eqnarray}
483: 
484: 
485: 
486: %It points out that the
487: %steady non-equilibrium current produces a correction to the
488: %Landau-Lifshiz-Gilbert damping  of the spin.
489: 
490: 
491: \section{Conclusions}
492: In conclusion, we present a theoretical study of
493: Landau-Lifshitz-Gilbert damping (Eq.(\ref{LLG2})) for a localized
494: spin ${\vec S}$ in a junction. The exchange interactions between the
495: localized spin and tunneling electrons leads to additional
496: dissipation of the spin motion, see Fig.(\ref{FIG:FIG2}). In the
497: presence of an applied voltage bias $V$, the damping coefficient,
498: i.e., Gilbert damping, is modified in linear order in $V$ for the
499: leads with particle-hole asymmetry in the Density of States. 
500: 
501: 
502: \section{Acknowledgements}
503: 
504: Work at LANL was supported by the US DOE under LDRD X1WX.
505: 
506: 
507: 
508: \section{Appendix A: DERIVATION OF THE EFFECTIVE ACTION}
509: Here we will give a detailed derivation of the effective action for
510: a spin. From Eq.(\ref {eff}), we can extract a quadratic form of
511: spins with the aid of the well known identity ${\rm {ln}}~{\rm
512: {det}}M ={\rm Tr}~{\rm {ln}}M.$  In order to tabulate the expansion
513: of ${\rm Tr}~{\rm {ln}}M$ perturbatively, we define matrices $M_0$
514: and $M_1$,
515: 
516: \begin{eqnarray}
517: M_{KP}&&=(M_0)_{KP}+(M_1)_{KP}
518: \nonumber
519: \\
520: (M_0)_{KP} &&\equiv ((-i\omega + \xi_{\alpha k})\delta_{\alpha \beta} \delta_{\sigma \sigma'})\delta_{KP}
521: \nonumber
522: \\
523: (M_1)_{KP} &&\equiv {\frac{1}{\sqrt{\beta}}}(T_{\alpha \beta}(\omega-\omega'))_{\sigma \sigma'},
524: \nonumber
525: \end{eqnarray}
526: where $K \equiv (k, \omega)$ and $P \equiv (p, \omega')$ with fermionic Matsubara frequencies $\omega$ and $\omega'$.
527: Employing the expansion ${\rm ln}(1+x) = - \sum_{n=1}^{\infty}
528: {\frac {(-x)^n}{n}}$, we can  write the effective action as
529: \begin{equation}
530: S_{\rm {eff}}(\vec S(\tau)) \sim S_0 + {\rm {Tr}}~{\rm {ln}}M_0 +
531: {\rm {Tr}}(M_0^{-1}M_1) -{\frac{1}{2}}{\rm {Tr}} (M_0^{-1}M_1)^2 +
532: \cdots, \label{Seff}
533: \end{equation}
534: where $S_0$ is the sum of the first and the second term in
535: Eq.(\ref{eff}). The third term in Eq.(\ref{eff}) (and consequent
536: last term shown in Eq.(\ref{Seff})) is the first non-trivial
537: contribution to the spin equation of motion.
538: Its evaluation is straightforward,
539: \begin{eqnarray}
540: &&{\rm Tr}(M_0^{-1}M_1)^2
541: \nonumber
542: \\
543: &&= \sum (M_0^{-1})_{K_1 K_1}(M_1)_{K_1 K_2}(M_0^{-1})_{K_2 K_2}(M_1)_{K_2 K_1}
544: \nonumber
545: \\
546: &&={\frac{1}{2 \beta}}\sum (-i\omega + \xi_{\alpha k})^{-1} (T_{\alpha \beta}(\omega-\omega'))_{\sigma \sigma'}
547: \nonumber
548: \\
549: &&(-i\omega' + \xi_{\beta p})^{-1} (T_{\beta \alpha}(\omega'-\omega))_{\sigma' \sigma},
550: \nonumber
551: \end{eqnarray}
552: where, repeated indices are implicitly summed over.
553: Then, we find
554: \begin{eqnarray}
555: &&\Delta S = {\frac{1}{2}} {\rm Tr}(M_0^{-1}M_1)^2
556: \nonumber
557: \\
558: &&={\frac{1}{2 \beta}} \sum_{\omega\omega'} \sum_{\alpha \beta} \sum_{\sigma \sigma'}
559: \sum_{k \in \alpha} \sum_{p \in \beta}(-i\omega + \xi_{\alpha k})^{-1} (T_{\alpha \beta}(\omega-\omega'))_{\sigma \sigma'}
560: \nonumber \\
561: & &(-i\omega' + \xi_{\beta p})^{-1} (T_{\beta \alpha}(\omega'-\omega))_{\sigma' \sigma}
562: \nonumber
563: \\
564: &&=-\sum_{\alpha \beta} \sum_{k \in \alpha} \sum_{p \in \beta} \sum_{\omega_m}
565: J_{\alpha \beta} J_{\beta \alpha} {\frac{f(\xi_k)-f(\xi_p)}{i\omega_m +\xi_k -\xi_p}}
566: {\vec S}(-\omega_m) \cdot {\vec S}(\omega_m)
567: \nonumber
568: \\
569: &&\equiv -\int d\tau \int d\tau' {\vec S}(\tau) \cdot {\vec S}(\tau') 2K(\tau-\tau').
570: \nonumber \end{eqnarray}
571: Here, $K(\tau-\tau')$ denotes the integral kernel defined in
572: Eq.(\ref{Kt}). Upon invoking the identity $2{\vec S(\tau)}\cdot{\vec
573: S({\tau}')} = (({\vec S(\tau)})^2+({\vec S({\tau}')})^2) -({\vec
574: S(\tau)}-{\vec S({\tau}')})^2$, the effective action becomes that of
575: Eq.(\ref{locdis}).
576: 
577: \section{APPENDIX B: THE DERIVATION OF THE SPECTRAL DENSITY}
578: We return to Eq.(\ref{Kt}) derived in Appendix A, and express the
579: sum as a contour integral following standard procedures,
580: e.g.\cite{Mahan}, to obtain
581: \begin{widetext}
582: \begin{eqnarray}
583: K_{\alpha \beta} &=& {\frac{J_{\alpha \beta} J_{\beta \alpha}}{2}}
584: \sum_{k \in \alpha}  \sum_{p \in \beta} \oint {\frac{dz}{2 \pi i}}
585: \Bigl( {\frac {e^{-z \tau}}{e^{\beta z}-1}} \theta(-\tau) + {\frac
586: {e^{-z \tau}}{1-e^{-\beta z}}} \theta(\tau) \Bigr) {\frac
587: {f(\xi_k)-f(\xi_p)}{z +\xi_k -\xi_p}} \nonumber
588: \\
589: &=& P \int_{- \infty}^{\infty} d{\omega} \Bigl( {\frac {J_{\alpha
590: \beta} J_{\beta \alpha}}{2}} \Bigr)
591:               \sum_{k \in \alpha}  \sum_{p \in \beta} [f(\xi_k)-f(\xi_p)]
592: \delta(\omega + \xi_k - \xi_p)
593:               \Bigl({\frac {e^{-\omega \tau}}{e^{\beta \omega}-1}}
594: \theta(-\tau) + {\frac {e^{-\omega \tau}}{1-e^{-\beta \omega}}}
595: \theta(\tau)  \Bigr) \nonumber
596: \\
597: &=& P \int_0^{\infty} d{\omega} \Bigl( -{\frac {J_{\alpha \beta}
598: J_{\beta \alpha}}{2}} \sum_{k \in \alpha}  \sum_{p \in
599: \beta}[f(\xi_k)-f(\xi_p)] \delta(\omega + \xi_k - \xi_p)  \Bigr)
600: \frac{{\rm cosh}({{\beta}/2}-|\tau|){\omega}}{{\rm sinh}({\beta
601: \omega}/2)} \nonumber
602: \\
603: &\equiv&  \int_0^{\infty} d{\omega} J(\omega) \frac{{\rm
604: cosh}({{\beta}/2}-|\tau|){\omega}}{{\rm sinh}({\beta \omega}/2)}
605: \label{cont}
606: \end{eqnarray}
607: \end{widetext}
608: where $J(\omega)$ denotes the \textit{spectral density} in
609: Caldeira-Leggett theory. The standard contour employed here is shown
610: in Fig.(\ref{FIG:FIG3}). The symbol $P$ in Eq.(\ref{cont}) denotes
611: the principal part of the integral.
612: \begin{figure}
613: \centerline{\includegraphics[width=0.9\columnwidth]{fig3.eps}}
614: \caption{The standard contour employed in Eq.(\ref{cont}) in order
615: to evaluate the Matsubara sum of Eq.(\ref{Kt}). The crosses along
616: the imaginary axis denote the Matsubara frequencies.}
617: \label{FIG:FIG3}
618: \end{figure}
619: 
620: 
621: 
622: 
623: %\section{Appendix C: A DERIVATION OF THE SPECTRAL DENSITY}
624: The spectral density $J(\omega)$ in Eq.(\ref{spec2})
625: \begin{widetext}
626: \begin{eqnarray}
627: J(\omega)=\frac{J_{LL}^2}{2} \int_{-E_F^L}^{\infty} d{\xi_L}
628: N(\xi_L) \int_{-E_F^L}^{\infty} d{\xi_L'} N(\xi_L')
629: [f(\xi_L)-f(\xi_L')]\delta(\omega+{\xi_L}-{\xi_L'}) \nonumber
630: \\
631: +\frac{J_{RR}^2}{2} \int_{-E_F^R}^{\infty} d{\xi_R} N(\xi_R)
632: \int_{-E_F^R}^{\infty} d{\xi_R'} N(\xi_R')
633: [f(\xi_R)-f(\xi_R')]\delta(\omega+{\xi_R}-{\xi_R'}) \nonumber
634: \\
635: +J_{LR}^2 \int_{-E_F^L}^{\infty} d{\xi_L} N(\xi_L)
636: \int_{-E_F^R}^{\infty} d{\xi_R} N(\xi_R)
637: [f(\xi_L)-f(\xi_R)]\delta(\omega+{\xi_L}-{\xi_R}) \nonumber
638: \\
639: \sim  \Bigl( \frac{J_{LL}^2}{2} N(\xi_L =0) N(\xi_L' =0) +
640: \frac{J_{RR}^2}{2} N(\xi_R =0) N(\xi_R' =0)
641: + J_{LR}^2 N(\xi_L =0) N(\xi_R =0) \Bigr) \omega
642: \nonumber
643: \\
644: =\Bigl( \frac{J_{LL}^2}{2} D^L(E_F^L) D^L(E_F^L) + \frac{J_{RR}^2}{2}
645: D^R(E_F^R) D^R(E_F^R)
646:             + J_{LR}^2 D^L(E_F^L) D^R(E_F^R) \Bigr) \omega,
647: \label{aspec}
648: \end{eqnarray}
649: \end{widetext}
650: 
651: where $D^{L/R}(E^{L/R}_F)$ denotes the \textit{density of states} at the Fermi
652: energy level of the left/right lead.
653: \\
654: If we apply a voltage leading to a chemical potential drop of
655: $E^L_F-E^R_F=(E_F +\Delta\mu_L)-(E_F+\Delta\mu_R)=eV$, then
656: Eq.(\ref{spec}) follows. This, in turn, leads to Eq.(\ref{Gil}).
657: then Eq.(\ref {spec}) follows.
658: 
659: 
660: 
661: 
662: \section{APPENDIX C: THE SPIN EQUATION OF MOTION}
663: 
664: 
665: If $J(\omega) = \frac{\alpha \omega}{2 \pi}$ then, from
666: Eq.(\ref{K}), the non-local in time kernel of the action
667: (Eq.(\ref{K+})) is $K(\tau) \sim {\frac{\alpha}{2 \pi}}
668: {\frac{1}{{\tau}^2}}$. We  thus obtain from Eq.(\ref{locdis}),
669: \begin{equation}
670: \Delta S_{dis} =  {\frac{\alpha}{2 \pi}} \int d{\tau} \int d{\tau}'
671: {\frac {({\vec S(\tau)}-{\vec S({\tau}')})^2} {(\tau-{\tau}')^2}}.
672: \end{equation}
673: The functional derivative of $\Delta S_{\rm {dis}}$ with respect to
674: $\vec S(\tau)$ is
675: \begin{eqnarray}
676: {\frac{\delta \Delta S_{dis}}{\delta S(\tau)}} =
677: {\frac{\alpha}{\pi}} \int d{\tau}'{\frac {({\vec S(\tau)}-{\vec
678: S({\tau}')})} {(\tau-{\tau}')^2}} \nonumber
679: \\ = {\frac{\alpha}{\pi}} \int d{\tau}' {\frac{1}{(\tau-{\tau}')}}
680: {\frac{d}{d{\tau}'}}({\vec S(\tau)}-{\vec S({\tau}')}) = i \alpha
681: {\frac {d}{d{\tau}}}{\vec S(\tau)} \label{corr}
682: \end{eqnarray}
683: From the free portion of the action (the first two terms of
684: Eq.(\ref{eff})), we have
685: \begin{equation}
686: {\frac{\delta S_0}{\delta {\vec S(\tau)}}} = i {\frac{1}{S^2}}
687: {\frac{d{\vec S(\tau)}}{d{\tau}}} \times {\vec S(\tau)} + \vec h.
688: \label{free}
689: \end{equation}
690: Adding Eqs.(\ref{corr},\ref{free}), equating the sum to zero, cross
691: multiplying with $\vec{S}(\tau)$, and changing $\tau \to it$, we
692: obtain Eq.(\ref{LLG2}).
693: 
694: 
695: 
696: \begin{thebibliography}{20}
697: 
698: 
699: \bibitem{spintronics} D. D. Awschalom, M. E. Flatte, and N.Samarth,
700: {\em Spintronics}, Scientific American, pp. 67-73, June 2002.
701: 
702: 
703: \bibitem{ll1} T. L. Gilbert, Phys. Rev. {\bf 100}, 1243 (1955)
704: \bibitem{ll2} L. Landau and E. M. Lifshitz, Physik Zeits. Soviets. {\bf 8},
705: 
706: 
707: 153 (1935)
708: 
709: 
710: \bibitem{Chika} S. Chikazumi, {\em Physics of Ferromagnetism}, (Oxford
711: university press, Oxford, 1997)
712: 
713: 
714: 
715: 
716: \bibitem{Heinrich} B. Heinrich, D. Fraitov\'a, and V. Kambersk\'y, Phys.
717: Status. Solidi, {\bf 23}, 501 (1967).
718: 
719: 
720: \bibitem{Halperin}  Y. Tserkovnyak, G. A. Fiete and B. I. Halperin, Appl.
721: Phys. Lett. {\bf 84}, 5234 (2004).
722: 
723: 
724: \bibitem{Onoda}  M. Onoda and N. Nagaosa,
725: cond-mat/0509058
726: 
727: 
728: \bibitem{Zhu03} Jian-Xin Zhu, Z. Nussinov, A. Shnirman and A. V.
729: Balatsky, Phys. Rev. Lett., {\bf 92}, 107001 (2004).
730: 
731: 
732: \bibitem{nsabz} Z. Nussinov, A. Shnirman, D. P. Arovas,
733: A. V. Balatsky, and J-X. Zhu, Phys. Rev. B {\bf 71}, 214520 (2005)
734: 
735: 
736: 
737: \bibitem{hooley} O. Parcollet and C. Hooley, Phys.
738: Rev. B {\bf 66}, 085315 (2002); L. N. Bulaevskii, M. Hruska, and G.
739: Ortiz, Phys. Rev. B {\bf 68}, 125415 (2003)
740: 
741: 
742: %\bibitem{SHE} S. Murakami, N. Nagaosa, and S.-C. Zhang, Science {\bf 301},
743: % 1248 (2003); J. Sinova et al., Phys. Rev. Lett. {\bf 92}, 126603 (2004)
744: 
745: 
746: 
747: \bibitem{bulaevskii}
748: L. Bulaevskii, M. Hruska, A. Shnirman, D. Smith, and Yu. Makhlin, Phys.
749: Rev. Lett. {\bf 92}, 177001 (2004)
750: 
751: 
752: \bibitem{Zhu_Balatsky}
753: J-X. Zhu, and A. V. Balatsky, Phys. Rev. B {\bf 67}, 174505 (2003)
754: 
755: 
756: \bibitem{Fradkin} E. Fradkin, {\em Field Theories of Condensed Matter
757: Systems} (Addison-Wesley, Redwood City, 1991).
758: 
759: 
760: \bibitem{Leggett} A. O. Caldeira and A. J. Leggett, Ann. Phys, {\bf 149},
761: 374(1983).
762: 
763: 
764: \bibitem{Leggett2} A. O. Caldeira and A. J. Leggett, Phys. Rev. Lett, {\bf
765: 46}, 211(1981).
766: 
767: 
768: 
769: 
770: %\bibitem{Ohm} The correction term in Eq.(\ref{Gil})  formally can be expressed as
771: 
772: 
773: 
774: %A single Ohmic current form for $\alpha_{1}$ is readily achieved
775: %($\frac{I_{o}^{L}}{[D(E_F^L)]^2} = \frac{I_{o}^{R}}{[D(E_F^R)]^2}$).
776: %a variety of other equivalent forms for $\alpha^{1}$ are
777: % a single current expression).
778: 
779: \bibitem{Mahan} G. D. Mahan, {\em Many-Particle Physics} (Kluwer, New York,
780: 2000).
781: 
782: 
783: \end{thebibliography}
784: 
785: 
786: \end{document}
787: