1: \documentclass[amsmath,amssymb,widetext,10pt,a4paper]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{bm}
4: \textheight 23truecm
5: \textwidth 16truecm
6: \evensidemargin 0.5truecm
7: \oddsidemargin 0.5truecm
8: \baselineskip 6pt
9: \topmargin 0truecm
10:
11: \def\beq{\begin{equation}}
12: \def\eeq{\end{equation}}
13:
14: \def\rmd{{\rm d}}
15: \def\rmD{{\rm D}}
16: \def\fl{}
17:
18: \begin{document}
19:
20: \title{Current-driven switching of magnetisation- theory and experiment}
21:
22:
23: \author{D. M. Edwards}
24: \affiliation{Department of Mathematics, Imperial College, London SW7
25: 2BZ, U.K.}
26: \author{ F. Federici}
27: \affiliation{NEST-INFM and Classe di Scienze, Scuola Normale
28: Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy}
29:
30: \maketitle
31:
32:
33:
34: \section{Introduction}\label{intro}
35:
36: Recently there has been a lot of interest in magnetic nanopillars of
37: 10-100 nm in diameter. The pillar is a metallic layered structure
38: with two ferromagnetic layers, usually of cobalt, separated by a
39: non-magnetic spacer layer, normally of copper. Non-magnetic leads
40: are attached to the magnetic layers so that an electric current may
41: be passed through the structure. In the simplest case the pillar may
42: exist in two states, with the magnetisation of the two magnetic
43: layers parallel or anti-parallel. The state of a pillar can be read
44: by measuring its resistance, this being smaller in the parallel
45: state than in the anti-parallel one. This dependence of the
46: resistance on magnetic configuration is the giant magnetoresistance
47: (GMR) effect \cite{1}. A dense array of these nanopillars could form
48: a magnetic memory for a computer. Normally one of the magnetic
49: layers in a pillar is relatively thick and its magnetisation
50: direction is fixed. In order to write into the memory the
51: magnetisation direction of the second thinner layer must be
52: switched. This might be achieved by a local magnetic field of
53: suitable strength and methods have been proposed \cite{2} for
54: providing such a local field by currents in a criss-cross array of
55: conducting strips. However an alternative, and potentially more
56: efficient method, proposed by Slonczewski \cite{3} makes use of a
57: current passing up the pillar itself. Slonczewski's effect relies on
58: ``spin transfer'' and not on the magnetic field produced by the
59: current which in the nanopillar geometry is ineffective. The idea of
60: spin-transfer is as follows. In a ferromagnet there are more
61: electrons of one spin orientation than of the other so that current
62: passing through the thick magnetic layer (the polarising magnet)
63: becomes spin-polarised. In general its state of spin-polarisation
64: changes as it passes through the second (switching) magnet so that
65: spin angular momentum is transferred to the switching magnet. This
66: transfer of spin angular momentum is called spin-transfer torque
67: and, if the current exceeds a critical value, it may be sufficient
68: to switch the direction of magnetisation of the switching magnet.
69: This is called current-induced switching.
70:
71: In the next section we show how to calculate the spin-transfer torque
72: for a simple model.
73:
74: \section{Spin-transfer torque in a simple model}\label{due}
75:
76: For simplicity we consider a structure of the type shown in Fig.
77: \ref{fig1}, where {\bf p} and {\bf m} are unit vectors in the
78: direction of the magnetisations. This models the layered structure
79: of the pillars used in experiments but the atomic planes shown are
80: considered to be unbounded instead of having the finite
81: cross-section of the pillar. This means that there is translational
82: symmetry in the $x$ and $z$ directions. The structure consists of a
83: thick (semi-infinite) left magnetic layer (polarising magnet), a
84: non-magnetic metallic spacer layer, a thin second magnet (switching
85: magnet) and a semi-infinite non-magnetic lead. In the simplest model
86: we assume the atoms form a simple cubic lattice, with lattice
87: constant $a$, and we adopt a one-band tight-binding model with
88: hopping Hamiltonian
89:
90: \begin{equation}
91: H_0=t\sum_{{\bf k}_{\parallel}\sigma}
92: \sum_{n}c^{\dagger}_{k_{\parallel} n\sigma} c_{k_{\parallel}
93: n-1\sigma}\,\, +\,\, {\rm h. c.}. \label{hamiltonian}
94: \end{equation}
95:
96: Here $c^{\dagger}_{k_{\parallel} n\sigma}$ creates an electron on
97: plane $n$ with two-dimensional wave-vector ${\bf k}_{\parallel}$ and
98: spin $\sigma$, and $t$ is the nearest-neighbour hopping integral.
99:
100: In the tight-binding description the operator for spin angular
101: momentum current between planes $n-1$ and $n$, which we require to
102: calculate spin-transfer torque, is given by
103: \begin{equation}
104: {\bf j}_{n-1}=-\frac{{\rm i}t}2\sum_{{\bf
105: k}_{\parallel}}\left(c^{\dagger}_{k_{\parallel},n,\sigma\uparrow},
106: c^{\dagger}_{k_{\parallel},n,\downarrow}\right){\bm\sigma}
107: \left(c_{k_{\parallel} ,n-1,\uparrow}, c_{k_{\parallel}
108: ,n-1,\downarrow} \right)^{\dagger} +\,\, {\rm h. c.}. \label{gamma}
109: \end{equation}
110: Here ${\bm\sigma}=\left(\sigma_x,\sigma_y,\sigma_z\right)$ where
111: the components are Pauli matrices. Eq. (\ref{gamma}) yields the
112: charge current $j_{n-1}^{\rm c}$ if $\frac12{\bm\sigma}$ is
113: replaced by a unit matrix multiplied by the number $e/\hbar$, where
114: $e$ is the electronic charge (negative). All currents flow in the
115: $y$ direction, perpendicular to the layers, and the components of
116: the vector ${\bf j}$
117: correspond to transport of $x$, $y$ and $z$ components of spin. The
118: justification of Eq. (\ref{gamma}) for ${\bf j}_{n-1}$ relies
119: on an equation of continuity, as pointed out in Section \ref{quattro}.
120:
121: To define the present model completely we must supplement the
122: hopping Hamiltonian $H_0$ by specifying the on-site potentials in
123: the various layers. For simplicity we assume the on-site potential
124: for both spins in non-magnetic layers, and for majority spin in
125: ferromagnetic layers, is zero. We assume an infinite exchange
126: splitting in the ferromagnets so that the minority spin potential in
127: these layers is infinite. Thus minority spin electrons are
128: completely excluded from the ferromagnets. Clearly the definition of
129: majority and minority spin relate to spin quantisation in the
130: direction of the local magnetisation. We take $\alpha=0$, so that
131: the magnetisation of the switching magnet is in the z direction and
132: take $\theta=\psi$, where $\psi$ is the angle
133: between the magnetisations.
134:
135: To describe spin transport in the structure we adopt the generalised
136: Landauer approach of Waintal {\it et. al.} \cite{4}. Thus the
137: structure is placed between two reservoirs, one on the left and one
138: on the right, with electron distributions characterised by Fermi
139: functions $f(\omega-\mu_{\rm L})$, $f(\omega-\mu_{\rm R})$
140: respectively. The system is then subject to a bias voltage $V_{\rm
141: b}$ given by $eV_{\rm b}=\mu_{\rm L}-\mu_{\rm R}$, the difference
142: between the chemical potentials. We discuss the ballistic limit
143: where scattering occurs only at interfaces, the effect of impurities
144: being negligible. We label the atomic planes so that $n=0$
145: corresponds to the last atomic plane of the polarising magnet. the
146: planes of the spacer layer correspond to $n=1,2\, \cdot\cdot\,N $
147: and $n=N+1$ is the first plane of the switching magnet.
148:
149: Consider first an electron incident from the left with wave-function
150: $|k,maj\rangle$, where $k>0$, which corresponds to a Bloch wave
151: $|k\rangle = \sum_{n}{\rm e}^{{\rm i}kna}|{\bf
152: k}_{\parallel}n\rangle$ with majority spin in the polarising magnet.
153: In this notation the label ${\bf k}_{\parallel}$ is suppressed. The
154: particle is partially reflected by the structure and finally emerges
155: as a partially transmitted wave in the lead, with spin $\uparrow$
156: corresponding to majority spin in the switching magnet. Thus the
157: wave-function is of the form
158: \begin{equation}
159: |P_k\rangle =|k,maj\rangle + B|-k,maj\rangle
160: \label{p}
161: \end{equation}
162: in the polariser and
163: \begin{equation}
164: |L_k\rangle =F|k,\uparrow\rangle
165: \label{p2}
166: \end{equation}
167: in the lead. A majority spin in either ferromagnet enters or leaves
168: the spacer without scattering, since in our simple model there is no
169: potential step. Also the minority spin wave-function entering a
170: ferromagnet is zero. The spacer wave-function may therefore be written
171: in two ways:
172: \begin{equation}
173: |S_k\rangle =F|k,\uparrow\rangle+E \left({\rm e}^{-{\rm
174: i}k(N+1)a}|k,\downarrow\rangle- {\rm e}^{{\rm
175: i}k(N+1)a}|-k,\downarrow\rangle \right)
176: \label{pp}
177: \end{equation}
178: or
179: \begin{eqnarray}\label{ppp}
180: |S_k\rangle &=& |k,maj\rangle+B|-k,maj\rangle+D\left(|k,min\rangle-
181: |-k,min\rangle\right)\nonumber\\
182: &=& \cos\left(\psi/2\right)|k,\uparrow\rangle+\sin\left(\psi/2\right)
183: |k,\downarrow\rangle
184: +B\left[\cos\left(\psi/2\right)|-k,\uparrow\rangle+\sin\left(\psi/2\right)|-k,\downarrow\rangle\right]\\
185: &&+D\left[-\sin\left(\psi/2\right)|k,\uparrow\rangle+\cos\left(\psi/2\right)
186: |k,\downarrow\rangle+\sin\left(\psi/2\right)
187: |-k,\uparrow\rangle-\cos\left(\psi/2\right)|-k,\downarrow\rangle\right].\nonumber
188: \end{eqnarray}
189: On equating coefficients of $|k,\uparrow\rangle$,
190: $|k,\downarrow\rangle$, $|-k,\uparrow\rangle$,
191: $|-k,\downarrow\rangle$ in expressions (\ref{pp}) and (\ref{ppp})
192: we have four equations which may be solved for $B$, $D$, $E$, $F$.
193: In particular the transmission coefficient $T$ is given by
194: \begin{equation}
195: \label{T}
196: T=\left|F\right|^2=\frac{4\cos^2(\psi/2)\sin^2
197: k(N+1)a}{\sin^4(\psi/2)+
198: 4\cos^2(\psi/2)\sin^2k(N+1)a}.
199: \end{equation}
200: Similarly an electron incident from the right with wave-function
201: $|-k\uparrow\rangle$ in the lead is partially reflected and finally
202: emerges as a partially transmitted wave $F^{\prime}|-k,maj\rangle$ in
203: the polarising magnet. It is found that $F^{\prime}=F$ so that the
204: transmission coefficient is the same for particles from left or right.
205:
206: The spin angular momentum current in a particular layer, which we
207: shall denote by S although it need not be the spacer layer, is the
208: sum of currents carried by left and right moving electrons. Thus we
209: have a Landauer-type formula \cite{5}
210: \begin{equation}
211: \label{landau}
212: {\bf j}_{\rm s}=\frac{a}{2\pi}\sum_{{\bf k }_{\parallel}}
213: \left\{\int_{k>0}{\rm d}k\left[\langle S_k|{\bf
214: j}_{n-1}|S_k\rangle f(\omega-\mu_L)+\langle S_{-k}|{\bf
215: j}_{n-1}|S_{-k}\rangle f(\omega-\mu_{R})\right]\right\}
216: \end{equation}
217: where $|S_k\rangle$, $|S_{-k}\rangle$ are wave-functions in the layer
218: considered corresponding to electrons incident from left and right,
219: respectively. Here $\omega$, the energy of the Bloch wave $k$, is
220: given by the tight-binding formula
221: \begin{equation}
222: \label{omega}
223: \omega=u_{{\bf k}_{\parallel}}+2t\cos ka
224: \end{equation}
225: where $u_{{\bf k}_{\parallel}}=2t(\cos k_xa+\cos k_za)$. We take
226: $t<0$ so that positive $k$ corresponds to positive velocity
227: $\hbar^{-1}\partial\omega/\partial k$ as we have assumed. The
228: current ${\bf j}_{\rm s}$ in layer $S$ calculated by Eq.
229: (\ref{landau}) does not depend on the particular planes $n-1$, $n$
230: between which it is calculated. On changing the integration variable
231: in Eq. (\ref{landau}) we find
232: \begin{equation}
233: \label{landau2} {\bf j}_{\rm s}=\frac{1}{2\pi} \sum_{{\bf
234: k}_{\parallel}}\int{\rm d}\omega\left[{\bf J}_{+}f(\omega-\mu_{\rm
235: L})+{\bf J}_{-}f(\omega-\mu_{\rm R})
236: \right]
237: \end{equation}
238: where
239: \begin{equation}
240: \label{landau3}
241: {\bf J}_{\pm}=\frac{\langle S_{\pm k}|{\bf j}_{n-1}|S_{\pm
242: k}\rangle}{-2t\sin ka}.
243: \end{equation}
244: Here $k=k(\omega,{\bf k}_{\parallel})$ is the positive root of Eq.
245: (\ref{omega}). Eq. (\ref{landau2}) may be written as
246: \begin{equation}
247: \label{landau33}
248: {\bf j}_{\rm s}=\frac{1}{4\pi}\sum_{{\bf k }_{\parallel}}
249: \int{\rm d}\omega\left\{
250: \left({\bf J}_{+}+{\bf J}_{-}\right)
251: \left[f(\omega-\mu_{\rm L})+f(\omega-\mu_{\rm R})\right]+
252: \left({\bf J}_{+}-{\bf J}_{-}\right)
253: \left[f(\omega-\mu_{\rm L})-f(\omega-\mu_{\rm R})\right]
254: \right\}.
255: \end{equation}
256:
257: Before discussing this spin current we briefly consider the charge
258: current $j^{\rm c}$, and we denote the analogues of ${\bf J}_{\pm}$ by
259: $J_{\pm}^{\rm c}$. Since the charge current is conserved
260: throughout the structure $J_{+}^{\rm c}$ and $J_{-}^{\rm c}$ can be
261: calculated in different ways, {\it e.g.} in the lead for $J_{+}^{\rm c}$ and in the polariser for $J_{-}^{\rm c}$. Since
262: $T=\left|F\right|^2=\left|F^{\prime}\right|^2$ we find $J_{+}^{\rm
263: c}+J_{-}^{\rm c}=0$ and for small bias $eV_{b}=\mu_{\rm L}-\mu_{\rm R}$
264: the charge current is given by
265: \begin{equation}
266: \label{current}
267: j^{\rm c}=\frac{2e^2V_{\rm b}}{h}\sum_{{\bf
268: k}_{\parallel}}T
269: \end{equation}
270: where the transmission coefficient $T$ is given by Eq. (\ref{T})
271: with $k=k(\mu,{\bf k}_{\parallel})$, $\mu$ being the common chemical
272: potential as $V_{\rm B}\rightarrow 0$. This is the well-known
273: Landauer formula \cite{5}.
274:
275: The spin transfer torque on the switching magnet is given by
276: \begin{equation}
277: \label{torque}
278: {\bf T}^{\rm s-t}=\langle{\bf j}_{\rm spacer}\rangle-
279: \langle{\bf j}_{\rm lead}\rangle,
280: \end{equation}
281: where $\langle{\bf j}_{\rm spacer}\rangle$ and $\langle{\bf j}_{\rm
282: lead}\rangle$ are spin currents in the spacer and lead respectively.
283: For zero bias ($\mu_{\rm L}=\mu_{\rm
284: R}$) there is clearly no charge current in the structure and
285: straight-forward calculation shows that all components of spin
286: current in the spacer and the lead vanish, except for a non-zero
287: $y$-spin current in the spacer. There is therefore a non-zero $y$
288: component of spin-transfer torque acting on the switching magnet for
289: zero bias, and its dependence on the angle $\psi$ between the
290: magnetisations is found to be approximately $\sin\psi$. This torque
291: is due to exchange coupling, analogous to an RKKY coupling, between
292: the two magnetic layers. This coupling oscillates as a function of
293: spacer thickness and tends to zero as the thickness tends to
294: infinity. For finite bias $V_{\rm B}$ the second term in the
295: integrand of Eq. (\ref{landau33}) comes into play. In general this
296: leads to finite $x$ and $y$ components of ${\bf T}^{\rm s-t}$
297: proportional to $V_{\rm b}$ (for small $V_{\rm b}$) whereas
298: $T_{z}^{\rm s-t}=0$. However for the special model considered here
299: with infinite exchange splitting in both ferromagnets it turns out
300: that $T_y^{\rm s-t}=0$. For this model the only non-zero component
301: of ${\bf T}^{\rm s-t}$ proportional to
302: $V_{\rm b}$ is found to be
303: \begin{equation}
304: \label{torquex}
305: T_x^{\rm s-t}=\frac{\hbar j^{\rm c}}{2|e|}\tan\frac{\psi}{2}
306: \end{equation}
307: where $j^{\rm c}$ is the charge current given by Eq. (\ref{current}).
308:
309: Slonczewski \cite{3} originally obtained this result for the
310: analogous parabolic band model. From Eqs. (\ref{torquex}),
311: (\ref{current}) and (\ref{T}) it follows that $T_{x}^{\rm s-t}$
312: contains an important factor $\sin\psi$ although this does not
313: represent the whole angle dependence. Clearly, from Eq.
314: (\ref{torquex}), the torque proportional to bias remains finite for
315: arbitrarily large spacer thickness, in the ballistic limit. For this
316: model, with infinite exchange splitting, the torque is independent
317: of the thickness of the switching magnet.
318:
319: From the results of this simple model we can infer a general form of
320: the spin-transfer torque ${\bf T}^{\rm s-t}$ which is independent of
321: the choice of coordinate axes. Thus we write
322: \begin{equation}
323: \label{torqueperp}
324: {\bf T}^{\rm s-t}={\bf T}_{\perp}+{\bf T}_{\parallel}
325: \end{equation}
326: where
327: \begin{eqnarray}
328: \label{torque2} {\bf T}_{\perp}&=&\left(g^{\rm ex}+g_{\perp} e
329: V_{\rm b}\right)({\bf
330: m}\times{\bf p}) \nonumber\\
331: {\bf T}_{\parallel}&=&g_{\parallel} e V_{\rm b}{\bf m}\times({\bf
332: p}\times{\bf m}).
333: \end{eqnarray}
334: With the choice of axes in Fig. \ref{fig1} ${\bf T}_{\parallel}$
335: corresponds to the $x$ component of torque, that is the component
336: parallel to the plane containing the magnetisation directions ${\bf
337: m}$ and ${\bf p}$. Similarly ${\bf T}_{\perp}$ corresponds to the
338: $y$ component of torque, this being perpendicular to the plane of
339: ${\bf m}$ and ${\bf p}$. The modulus of both the vectors ${\bf
340: m}\times{\bf p}$ and ${\bf m}\times({\bf p}\times{\bf m})$ is
341: $\sin\psi$, so that the factors $g^{\rm ex}$, $g_{\perp}$ and
342: $g_{\parallel}$ are functions of $\psi$ which contain deviations
343: from the simple $\sin\psi$ behaviour. The bias-independent term
344: $g^{\rm ex}$ corresponds to the interlayer exchange coupling, as
345: discussed above, and henceforth we assume that the spacer is thick
346: enough for this term to be negligible. Sometimes the $\sin\psi$
347: factor accounts for most of the angular dependence of $T_{\perp}$
348: and $T_{\parallel}$ so that $g_{\perp}$ and $g_{\parallel}$ may be
349: regarded as constant parameters for the given structure. In the next
350: section we use Eqs. (\ref{torque2}) for the spin-transfer torque in
351: a phenomenological theory of current-induced switching of
352: magnetisation. This phenomenological treatment enables us to
353: understand most of the available experimental data. It is more usual
354: in experimental works to relate spin-transfer torque to current
355: rather than bias. However in theoretical work, based on the Landauer
356: or Keldysh approach, bias is more natural. In practice the
357: resistance of the system considered is rather constant (the GMR
358: ratio is only a few percent) so that bias and current are in a
359: constant ratio.
360:
361:
362:
363: \section{Phenomenological treatment of current-induced switching of
364: magnetisation}\label{tre}
365: In this section we explore the consequences of the spin-transfer
366: torque acting on a switching magnet using a phenomenological Landau
367: Lifshitz equation with Gilbert damping (LLG equation). This is
368: essentially a generalisation of the approach used originally by
369: Slonczewski \cite{3} and Sun \cite{sun}. We assume that there is a
370: polarising magnet whose magnetisation is pinned in the $xz$-plane in
371: the direction of a unit vector ${\bf p}$, which is at general fixed
372: angle $\theta$ to the $z$-axis as shown in Fig. \ref{fig1}.
373: \begin{figure}
374: \includegraphics[width=11cm]{Fig1.eps}
375: \caption{Schematic picture of a magnetic layer structure for
376: current-induced switching (magnetic layers are darker, non-magnetic
377: layers lighter).}
378: \label{fig1}
379: \end{figure}
380: The pinning of the magnetisation of the polarising magnet can be due
381: to its large coercivity (thick magnet) or a strong uniaxial
382: anisotropy. The role of the polarising magnet is to produce a stream
383: of spin-polarised electrons, {\it i.e.} spin current, that is going
384: to exert a torque on the magnetisation of the switching magnet whose
385: magnetisation lies in the general direction of a unit vector ${\bf
386: m}$. The orientation of the vector ${\bf m}$ is defined by the polar
387: angles $\alpha$, $\phi$ shown in Fig. \ref{fig1}.
388: There is a non-magnetic metallic layer inserted
389: between the two magnets
390: whose role is merely to separate magnetically the two magnetic layers and
391: allow a strong charge current to pass. The total thickness of the
392: whole trilayer sandwiched between two non-magnetic leads must be
393: smaller than the spin diffusion length $l_{\rm sf}$ so that there
394: are no spin flips due to impurities or spin-orbit coupling. A
395: typical junction in which current-induced switching is studied
396: experimentally \cite{albert} is shown schematically in Fig.
397: \ref{fig2}. The thickness of the polarising magnet is 40nm, that of
398: the switching magnet 2.5nm and the non-magnetic spacer is 6nm thick.
399: The materials for the two magnets and the spacer are cobalt and
400: copper, respectively, which are those most commonly used. The
401: junction cross section is oval-shaped with dimensions
402: 60nm$\times$130nm. A small diameter is necessary so that the torque
403: due to the Oersted field generated by a charge current of
404: $10^{7}$-$10^8$ A/cm$^2$, required for current-induced switching, is
405: much smaller than the spin-transfer torque we are interested in.
406: \begin{figure}
407: \includegraphics[width=8cm,angle=-90]{Fig2.eps}
408: \caption{Schematic picture of a junction in which current-induced
409: switching is studied experimentally.}
410: \label{fig2}
411: \end{figure}
412:
413: The aim of most experiments is to determine the orientation of the
414: switching magnet moment as a function of the current (applied bias)
415: in the junction. Sudden jumps of the magnetisation direction, {\it
416: i.e.} current-induced switching, are of particular interest. The
417: orientation of the switching magnet moment ${\bf m}$ relative to
418: that of the polarising magnet ${\bf p}$, which is fixed, is
419: determined by measuring the resistance of the junction. Because of
420: the GMR effect, the resistance of the junction is higher when the
421: magnetisations of the two magnets are anti-parallel than when they
422: are parallel, In other words, what is observed are hysteresis loops
423: of resistance versus current. A typical experimental hysteresis loop
424: of this type \cite{16} is
425: reproduced in Fig. \ref{fig3}.
426: \begin{figure}
427: \includegraphics[width=8cm]{Fig3.eps}
428: \caption{Resistance vs current hysteresis loop (after Grollier {\it
429: et al.} \cite{16}). } \label{fig3}
430: \end{figure}
431: It can be seen from Fig. \ref{fig3} that, for any given current, the
432: switching magnet moment is stationary (the junction resistance has a
433: well defined value), {\it i.e.} the system is in a steady state.
434: This holds everywhere on the hysteresis loop except for the two
435: discontinuities where current-induced switching occurs. As indicated
436: by the arrows jumps from the parallel (P) to anti-parallel (AP)
437: configurations of the magnetisation, and from AP to P
438: configurations, occur at different currents. It follows that in
439: order to interpret experiments which exhibit such hysteresis
440: behaviour, the first task of the theory is to determine from the LLG
441: equation all the possible states and then investigate their
442: dynamical stability. At the point of instability the system seeks
443: out a new steady state, {\it i.e.} a discontinuous transition to a
444: new steady state with the switched magnetisation occurs. We have tacitly
445: assumed that there is always a steady state available for the system
446: to jump to. There is now experimental evidence that this is not
447: always the case. In the absence of any stable steady state the
448: switching magnet moment remains permanently in the time-dependent
449: state. This interesting case is implicit in the phenomenological LLG
450: treatment and we shall discuss it in detail later.
451:
452: In describing the switching magnet by a unique unit vector ${\bf m}$, we
453: assume that it remains uniformly magnetised during the switching
454: process. This is only strictly true when the exchange stiffness of the
455: switching magnet is infinitely large. It is generally a good
456: approximation as long as the switching magnet is small enough to
457: remain single domain, so that the switching occurs purely by rotation
458: of the magnetisation as in the Stoner-Wohlfarth theory \cite{17} of
459: field switching. This seems to be the case in many experiments
460: \cite{albert,16,18,19}
461:
462: Before we can apply the LLG equation to study the time evolution of
463: the unit vector ${\bf m}$ in the direction of the magnetisation of
464: the switching magnet, we need to determine all the contributions to
465: the torque acting on the switching magnet. Firstly, there is the
466: spin-transfer torque ${\bf T}^{\rm s-t}$ which we discussed in
467: Section \ref{due}. Secondly, there is a torque due to the uniaxial
468: in-plane and easy plane (shape) anisotropies. The easy-plane shape
469: anisotropy torque arises because the switching magnet is a thin
470: layer typically only a few nanometers thick. The in-plane uniaxial
471: anisotropy is usually also a shape anisotropy arising from an
472: elongated cross section of the switching magnet \cite{albert}. We
473: take the uniaxial anisotropy axis of the switching magnet to be
474: parallel to the $z$-axis of the coordinate system shown in Fig.
475: \ref{fig1}. Since the switching magnet lies in the $xz$-plane, we
476: can write the total anisotropy field as
477: \begin{equation}
478: \label{hh}
479: {\bf H}_{\rm A}={\bf H}_{\rm u}+{\bf H}_{\rm p}
480: \end{equation}
481: where ${\bf H}_{\rm u}$ and ${\bf H}_{\rm p}$ are given by
482: \begin{equation}
483: \label{hu}
484: {\bf H}_{\rm u}= H_{\rm u0}({\bf m}\cdot{\bf e}_{z}){\bf e}_{z},
485: \end{equation}
486:
487: \begin{equation}
488: \label{hp}
489: {\bf H}_{\rm p}=-H_{\rm p0}({\bf m}\cdot{\bf e}_{y}){\bf e}_{y}.
490: \end{equation}
491: Here ${\bf e}_x$, ${\bf e}_y$ and ${\bf e}_z$ are unit vectors in
492: the directions of the axes shown in Fig. 1. If we write the energy
493: of the switching magnet in the anisotropy field as $-{\bf H}_{\rm
494: A}\cdot\langle{\bf S}_{\rm tot}\rangle$, where $\langle{\bf S}_{\rm
495: tot}\rangle$ is the total spin angular momentum of the switching
496: magnet, them $H_{\rm u0}$,$H_{\rm p0}$ which measure the strengths
497: of the uniaxial and easy-plane anisotropies have dimensions of
498: frequency. These quantities may be converted to a field in tesla by
499: multiplying by $\hbar/2\mu_{\rm
500: B}=5.69\times 10^{-12}$.
501:
502: We are now ready to study the time evolution of the unit vector ${\bf
503: m}$ in the direction of the switching magnet moment. The LLG
504: equation takes the usual form
505: \begin{equation}
506: \label{llg}
507: \frac{{\rm d}{\bf m}}{{\rm d}t}+\gamma{\bf m}\times\frac{{\rm d}{\bf m}}{{\rm d}t}={\bm\Gamma}
508: \end{equation}
509: where the reduced total torque ${\bm\Gamma}$ acting on the switching magnet
510: is given by
511: \begin{equation}
512: \label{Gamma}
513: {\bm\Gamma}=\left[-\left({\bf H}_{\rm A}+{\bf H}_{\rm
514: ext}\right)\times\langle{\bf S}_{\rm tot}\rangle+{\bf
515: T}_{\perp}+{\bf T}_{\parallel}\right]/
516: \left|\langle {\bf S}_{\rm tot}\rangle\right|.
517: \end{equation}
518: Here ${\bf H}_{\rm ext}$ is an external field, in the same frequency
519: units as ${\bf H}_{\rm A}$, and $\gamma$ is the Gilbert damping
520: parameter. Following Sun \cite{sun}, Eq. (\ref{llg}) may be written
521: more conveniently as
522: \begin{equation}
523: \label{Gamma2}
524: \left(1+\gamma^2\right)\frac{{\rm d}{\bf m}}{{\rm
525: d}t}={\bm\Gamma}-\gamma{\bf m}\times{\bm\Gamma}.
526: \end{equation}
527: It is also useful to measure the strengths of all the torques in units
528: of the strength of the uniaxial anisotropy \cite{sun}. We shall,
529: therefore, write the total reduced torque ${\bm\Gamma}$ in the form
530: \begin{equation}
531: \label{Gamma3}
532: {\bm\Gamma}=H_{uo}\left\{({\bf m}\cdot{\bf e}_{z}){\bf m}\times{\bf e}_z-
533: h_{\rm p}({\bf m}\cdot{\bf e}_{y}){\bf m}\times
534: {\bf e}_y+v_{\parallel}(\psi)
535: {\bf m}\times({\bf p}\times{\bf m})+\left[v_{\perp}(\psi)+h_{\rm ext}\right]
536: {\bf m}\times{\bf p}\right\}
537: \end{equation}
538: where the relative strength of the easy plane anisotropy $h_{\rm
539: p}=H_{\rm p0}/H_{\rm u0}$ and
540: $v_{\parallel}(\psi)=vg_{\parallel}(\psi)$,
541: $v_{\perp}(\psi)=vg_{\perp}(\psi)$ measure the strengths of the
542: torques ${\bf T}_{\parallel}$ and ${\bf T}_{\perp}$. The reduced
543: bias is defined by $v=eV_{\rm b}/(|\langle {\bf S}_{\rm
544: tot}\rangle|H_{\rm u0})$ and has the opposite sign from the bias
545: voltage since $e$ is negative. Thus positive $v$ implies a flow of
546: electrons from the polarising to the switching magnet.
547: The last contribution to the torque in Eq. (\ref{Gamma3}) is due to the
548: external field $H_{\rm ext}$ with $h_{\rm ext}=H_{\rm ext}/H_{\rm
549: u0}$. The external field is taken in the direction of the
550: magnetisation of the polarising magnet, as is the case in most
551: experimental situations.
552:
553: It follows from Eq. (\ref{llg}) that in a steady state ${\bm\Gamma}=0$. We
554: shall first consider some cases of experimental importance where the
555: steady state solutions are trivial and the important physics is
556: concerned entirely with their stability. To discuss stability, we
557: linearise Eq. (\ref{Gamma2}), using Eq. (\ref{Gamma3}), about a steady
558: state solution ${\bf m}={\bf m}_{0}$. Thus
559: \begin{equation}
560: \label{Gamma4}
561: {\bf m}={\bf m}_0+\xi{\bf e}_{\alpha}+\eta{\bf e}_{\phi},
562: \end{equation}
563: where ${\bf e}_{\alpha}$, ${\bf e}_{\phi}$ are unit vectors in the
564: direction ${\bf m}$ moves when $\alpha$ and $\phi$ are increased
565: independently. The linearised equation may be written in the form
566: \begin{equation}
567: \label{Gamma5}
568: \frac{{\rm d}\xi}{{\rm d}\tau}=A\xi+B\eta;\quad
569: \frac{{\rm d}\eta}{{\rm d}\tau}=C\xi+D\eta.
570: \end{equation}
571: Following Sun \cite{sun}, we have introduced the natural
572: dimensionless time variable $\tau=tH_{\rm u0}/(1+\gamma^2)$. The
573: conditions for the steady state to be stable are
574: \begin{equation}
575: \label{Gamma6}
576: F=A+D\leqslant 0;\quad G=AD-BC\geqslant 0
577: \end{equation}
578: excluding $F=G=0$ \cite{20}. For simplicity we give these conditions
579: explicitly only for the case where either
580: $v_{\parallel}^{\prime}(\psi_0)=v_{\perp}^{\prime}(\psi_0)=0$, with
581: $\psi_0=\cos^{-1}({\bf p}\cdot{\bf m}_0)$, or ${\bf m}_0=\pm{\bf p}$.
582: The case ${\bf m}_0=\pm{\bf p}$ is very common experimentally as is
583: discussed below. The stability condition $G\geqslant 0$ may be written
584: \begin{eqnarray}
585: \label{stab}
586: &&Q^2v_{\parallel}^2+(Qh+\cos 2\alpha_0)(Qh+\cos^2\alpha_0)+h_{\rm
587: p}\left\{Qh(1-3\sin^2\phi_0\sin^2\alpha_0)+\cos 2\alpha_0
588: (1-2\sin^2\alpha_0\sin^2\phi_0)\right\}-\nonumber\\
589: &&h_{\rm
590: p}^2\sin^2\alpha_0\sin^2\phi_0(1-2\sin^2\phi_0\sin^2\alpha_0)
591: \geqslant 0,
592: \end{eqnarray}
593: where $v_{\parallel}=v_{\parallel}(\psi_0)$,
594: $h=v_{\perp}(\psi_0)+h_{\rm ext}$ and $Q=\cos\psi_0$. The condition
595: $F\leqslant 0$ takes the form
596: \begin{equation}
597: \label{F0}
598: -2(v_{\parallel}+\gamma h)Q-\gamma(\cos
599: 2\alpha_0+\cos^2\alpha_0)-\gamma h_{\rm
600: p}(1-3\sin^2\phi_0\sin^2\alpha_0)
601: \leqslant 0.
602: \end{equation}
603:
604: We now discuss several interesting examples, the first of these
605: relating to experiments of Grollier {\it et al.} \cite{18} and
606: others. In these experiments the magnetisation of the polarising
607: magnet, the uniaxial anisotropy axis and the external field are all
608: collinear (along the in-plane $z$-axis in our convention). In this
609: case the equation ${\bm\Gamma}=0$, with ${\bm\Gamma}$ given by
610: Eq. (\ref{Gamma3}), shows immediately that possible steady states are
611: given by ${\bf m}_0=\pm{\bf p}(\alpha_0=0,\pi)$, corresponding to the
612: switching magnet moment along the $z$-axis. These are the only
613: solutions when $h_{\rm p}=0$. For $h_{\rm p}\neq 0$ other steady-state
614: solutions may exist but in the parameter regime which has been
615: investigated they are always unstable \cite{13}. We shall assume this
616: is always the case and concentrate on the solutions ${\bf m}_0=\pm{\bf
617: p}$. In the state of parallel magnetisation (P) ${\bf m}_0={\bf p}$
618: we have $v_{\parallel}=vg_{\parallel}(0)$, $h=vg_{\perp}(0)+h_{\rm
619: ext}$, $\alpha_0=0$ and $Q=1$. The stability conditions (\ref{stab})
620: and (\ref{F0}) become
621: \begin{equation}
622: \label{stab2}
623: \left[g_{\parallel}(0)\right]^2v^2+(vg_{\perp}(0)+h_{\rm
624: ext}+1)^2+h_{\rm p}\left[vg_{\perp}(0)+h_{\rm
625: ext}+1\right]\geqslant 0
626: \end{equation}
627:
628: \begin{equation}
629: \label{F0-2}
630: g_{\parallel}(0)v+\gamma\left[vg_{\perp}(0)+h_{\rm
631: ext}+1+\frac12 h_{\rm p}\right]\geqslant 0.
632: \end{equation}
633: In the state of anti-parallel magnetisation (AP) ${\bf m}_0=-{\bf p}$ we
634: have $v_{\parallel}=vg_{\parallel}(\pi)$, $h=vg_{\perp}(\pi)+h_{\rm
635: ext}$, $\alpha_0=\pi$ and $Q=-1$. The stability conditions for the
636: AP state are thus
637: \begin{equation}
638: \label{stab2AP}
639: \left[g_{\parallel}(\pi)\right]^2v^2+(-vg_{\perp}(\pi)-h_{\rm
640: ext}+1)^2+h_{\rm p}\left[-vg_{\perp}(\pi)-h_{\rm
641: ext}+1\right]\geqslant 0
642: \end{equation}
643:
644: \begin{equation}
645: \label{F0-2AP}
646: g_{\parallel}(\pi)v+\gamma\left[vg_{\perp}(\pi)+h_{\rm
647: ext}-1-\frac12 h_{\rm p}\right]\leqslant 0.
648: \end{equation}
649:
650:
651: In the regime of low external field ($h_{\rm ext}\approx 1$, {\it
652: i.e.} $H_{\rm ext}\approx H_{u0}$) we have $H_{\rm p}>>H_{\rm ext}$
653: ($h_{\rm p}\approx 100$). Eqs. (\ref{stab2}) and (\ref{stab2AP}) may be
654: then approximated by
655: \begin{equation}
656: \label{stab2AP-approx}
657: vg_{\perp}(0)+h_{\rm ext}+1>0
658: \end{equation}
659:
660: \begin{equation}
661: \label{F0-2AP-approx}
662: vg_{\perp}(\pi)+h_{\rm ext}-1<0.
663: \end{equation}
664: Equation (\ref{stab2AP-approx}) corresponds to P stability and
665: (\ref{F0-2AP-approx}) to AP stability. It is convenient to define
666: scalar quantities $T_{\perp}$, $T_{\parallel}$ by
667: $T_{\perp}=g_{\perp}(\psi)\sin\psi$,
668: $T_{\parallel}=g_{\parallel}(\psi)\sin\psi$, these being scalar
669: components of spin-transfer torque in units of $eV_{\rm b}$
670: (cf. Eq. (\ref{torque2})). Then $g_i(0)=[{\rm d}T_i/{\rm
671: d}\psi]_{\psi=0}$ and $g_i(\pi)=-[{\rm d}T_i/{\rm
672: d}\psi]_{\psi=\pi}$ with $i=\perp,\parallel$. Model calculations
673: \cite{13} show that both $g_{\perp}$ and $g_{\parallel}$ can be of
674: either sign, although positive values are more common. Also there is no
675: general rule about the relative magnitude of $g_i(0)$ and $g_i(\pi)$.
676:
677: We now illustrate the consequences of the above stability conditions
678: by considering two limiting cases. We first consider the case
679: $g_{\perp}(\psi)=0$, $g_{\parallel}>0$, as assumed by Grollier {\it
680: et. al} \cite{18} in the analysis of their data. In Fig. 4 we
681: plot the regions of P and AP stability deduced from
682: Eqs. (\ref{F0-2}),(\ref{F0-2AP})-(\ref{F0-2AP-approx}), in the
683: ($v$,$h_{\rm ext}$)-plane. Grollier {\it et al.} plot current instead
684: of bias but this should not change the form of the figure. Theirs is
685: rather more complicated, owing to a less transparent stability
686: analysis with unnecessary approximation. The only approximations made
687: above, to obtain Eqs. (\ref{stab2AP-approx}) and(\ref{F0-2AP-approx}), can
688: easily be removed, which results in the critical field lines $h_{\rm
689: ext}=\pm1$ acquiring a very slight curvature given by $h_{\rm
690: ext}\approx 1+[vg_{\parallel}(\pi)]^2/h_{\rm p}$ and $h_{\rm
691: ext}\approx -1-[vg_{\parallel}(0)]^2/h_{\rm p}$. The critical biases
692: in the figure are give by
693: \begin{eqnarray}
694: \label{crbias}
695: v_{{\rm AP}\rightarrow{\rm P}}&=&\gamma \left[1+\frac12 h_{\rm p}-h_{\rm
696: ext}\right]/g_{\parallel}(\pi)\nonumber\\
697: v_{{\rm P}\rightarrow{\rm AP}}&=&-\gamma \left[1+\frac12 h_{\rm p}+h_{\rm
698: ext}\right]/g_{\parallel}(0).
699: \end{eqnarray}
700: A downward slope from left to right of the corresponding lines in
701: Fig. \ref{fig4} is not shown there. Since the damping parameter
702: $\gamma$ is small ($\gamma\approx 0.01$) this downward slope of the
703: critical bias lines is also small. From Fig. \ref{fig4} we can deduce
704: the behaviour of resistance versus bias in the external field
705: regimes $|h_{\rm ext}|<1$ and $|h_{\rm ext}|>1$.
706: \begin{figure}
707: \includegraphics[width=13cm]{Fig4.eps}
708: \caption{Bias-field stability diagram for $g_{\perp}(\psi)=0$,
709: $g_{\parallel}(\psi)>0$. A small downward slope of the lines $V_{{\rm
710: AP}\rightarrow{\rm P}}$,$V_{{\rm P}\rightarrow{\rm AP}}$
711: (see Eq. (\ref{crbias})) is not shown.}
712: \label{fig4}
713: \end{figure}
714:
715: Consider first the case $|h_{\rm ext}|<1$. Suppose we start in the AP
716: state with a bias $v=0$ which is gradually increased to $v_{{\rm
717: AP}\rightarrow P}$. At this point the AP state becomes unstable
718: and the system switches to the P state as $v$ increases further. On
719: reducing $v$ the hysteresis loop is completed via a switch back to
720: the AP state at the negative bias $v_{{\rm P}\rightarrow AP}$. The
721: hysteresis loop is shown in Fig. \ref{fig5}(a). The increase in resistance
722: R between the P and AP states is the same as would be
723: produced by varying the applied field in a GMR experiment.
724: \begin{figure}
725: \includegraphics[width=13cm]{Fig5.eps}
726: \caption{(a) Hysteresis loop of resistance vs bias for $|h_{\rm
727: ext}|<1$; (b) Reversible behaviour (no hysteresis) for $|h_{\rm
728: ext}|<-1$ (upper curve) and $h_{\rm ext}>1$ (lower curve). The
729: dashed lines represent hypothetical behaviour of average
730: resistance in regions of Fig. \ref{fig4} marked ``both unstable''
731: where no steady states exist.}
732: \label{fig5}
733: \end{figure}
734: Now consider the case $h_{\rm ext}<-1$. Starting again in the AP
735: state at $v=0$ we see from Fig. \ref{fig4} that, on increasing $v$
736: to $v_{{\rm AP}\rightarrow{\rm P}}$, the AP state becomes unstable but there is
737: no stable P state to switch to. This point is marked by an asterisk
738: in Fig. \ref{fig5}(b). For $v>v_{{\rm AP}\rightarrow{\rm P}}$,
739: the moment of the
740: switching magnet is in a persistently time-dependent state. However,
741: if $v$ is now decreased below $v_{{\rm P}\rightarrow{\rm AP}}$
742: the system homes in on the stable AP state and the overall behaviour
743: is reversible,
744: {\it i.e.} no switching and no hysteresis occur. When $h_{\rm
745: ext}>1$ similar behaviour, now involving the P state, occurs at
746: negative bias, as shown in Fig. \ref{fig5}(b). The dashed curves in
747: Fig. \ref{fig5}(b) show a hypothetical time-averaged resistance in the
748: regions of time-dependent magnetisation. As discussed later
749: time-resolved measurements of resistance suggest that several
750: different types of dynamics can occur in these regions.
751:
752: It is clear from Fig. \ref{fig5}(a) that the jump AP$\rightarrow$P
753: always occurs for positive bias $v$, which corresponds to flow of
754: electrons from the polarising to the switching magnet. This result
755: depends on the assumption that $g_{\parallel}>0$; if
756: $g_{\parallel}<0$ it is easy
757: to see that the sense of the hysteresis loop is reversed and the
758: jump P$\rightarrow$AP occurs for positive $v$. To our knowledge this
759: reverse jump has never been observed, although $g_{\parallel}<0$ can
760: occur in principle and is predicted theoretically \cite{13} for the
761: Co/Cu/Co(111) system with a switching magnet consisting of a single
762: atomic plane of Co. It follows from Eq. (\ref{crbias}) that
763: $|v_{{\rm P}\rightarrow{\rm AP}}/v_{{\rm AP}\rightarrow{\rm
764: P}}|=|g_{\parallel}(\pi)/g_{\parallel}(0)|$ in zero external
765: field. Experimentally this ratio, essentially the same as the ratio of
766: critical currents, may be considerably less than 1 ({\it e.g.} $<0.5$
767: \cite{albert}), greater than 1 ({\it e.g.} $\approx 2$ \cite{19}) or
768: close to 1 \cite{16}. Usually the field dependence of the critical
769: current is found to be stronger than that predicted by Eq. (\ref{crbias})
770: \cite{albert,16}.
771:
772: We now discuss the reversible behaviour shown in Fig. \ref{fig5}(b)
773: which occurs for $|h_{\rm ext}|>1$. The transition from hysteretic
774: to reversible behaviour at a critical external field seems to have
775: been first seen in pillar structures by Katine {\it et al.}
776: \cite{21}. Curves similar to the lower one in Fig. \ref{fig5}(b) are
777: reported with $|v_{{\rm P}\rightarrow{\rm AP}}|$ increasing with
778: increasing $h_{\rm ext}$, as expected from Eq. (\ref{crbias}). Plots
779: of the differential resistance ${\rm d}V/{\rm d}I$ show a peak near
780: the point of maximum gradient of the dashed curve. Similar
781: behaviour has been reported by several groups \cite{22,23,24}. It
782: is particularly clear in the work of Kiselev {\it at al.} \cite{22}
783: that the transition from hysteretic behaviour (as in Fig.
784: \ref{fig5}(a)) to reversible behaviour with peaks in ${\rm d}V/{\rm
785: d}I$ occurs at the coercive field 600 Oe of the switching layer
786: ($h_{\rm ext}=1$). The important point about the peaks in ${\rm
787: d}V/{\rm d}I$ is that for a given sign of $h_{\rm ext}$ they only
788: occur for one sign of the bias. This clearly shows that this effect
789: is due to spin-transfer and not to Oersted fields. Myers {\it et
790: al.} \cite{25} show a current-field stability diagram similar to the
791: bias-field one of Fig. \ref{fig4} with a critical field of 1500 Oe.
792: They examine the time dependence of the resistance at room
793: temperature with the field and current adjusted so that the system
794: is in the ``both unstable'' region in the fourth quadrant of Fig.
795: \ref{fig4} but very close to its top left-hand corner. They observe
796: telegraph-noise-type switching between approximately P and AP
797: states with slow switching times in the range 0.1-10 s. Similar
798: telegraph noise with faster switching times was observed by Urazhdin
799: {\it et al.} \cite{23} at current and field close to a peak in ${\rm
800: d}V/{\rm d}I$. In the region of P and AP instability Kiselev {\it et
801: al.} \cite{22} and Pufall {\it et al.} \cite{24} report various
802: types of dynamics of precessional type and random telegraph
803: switching type in the microwave Ghz regime. Kiselev {\it et al.}
804: \cite{22} propose that systems of the sort considered here might
805: serve as nanoscale microwave sources or oscillators, tunable by
806: current and field over a wide frequency range.
807:
808: We now return to the stability conditions
809: (\ref{F0-2}),(\ref{F0-2AP})-(\ref{F0-2AP-approx}) and consider the case
810: of $g_{\perp}(\psi)\neq 0$ but $h_{\rm ext}=0$. These conditions of
811: stability of the P state may be written approximately, remembering
812: that $\gamma << 1$, $h_{\rm p}>> 1$, as
813: \begin{equation}
814: \label{vgperp}
815: vg_{\perp}(0)>-1, \quad vg_{\parallel}(0)>-\frac12 \gamma h_{\rm p}.
816: \end{equation}
817: The conditions for stability of the AP state are
818: \begin{equation}
819: \label{vgperpAP}
820: vg_{\perp}(\pi)<1, \quad vg_{\parallel}(\pi)<\frac12 \gamma h_{\rm p}.
821: \end{equation}
822: In Fig. \ref{fig6} we plot the regions of P and AP stability,
823: assuming $g_{\perp}(0)=g_{\perp}(\pi)=g_{\perp}$ and
824: $g_{\parallel}(0)=g_{\parallel}(\pi)=g_{\parallel}$ for simplicity.
825: We also put $r=g_{\perp}/g_{\parallel}$.
826: \begin{figure}
827: \includegraphics[width=10cm]{Fig6.eps}
828: \caption{Stability diagram for $h_{\rm ext}=0$.} \label{fig6}
829: \end{figure}
830: For $r>0$ we find the normal hysteresis loop as in Fig.
831: \ref{fig5}(a) if we plot $R$ against $vg_{\parallel}$ (valid for
832: either sign of $g_{\parallel}$). In Fig. \ref{fig7} we plot the
833: hysteresis loops for the cases $r_{\rm c}<r<0$ and $r<r_{\rm c}$,
834: where $r_{\rm c}=-2/(\gamma h_{\rm p})$ is the value of $r$ at the
835: point $X$ in Fig. \ref{fig6}.
836: \begin{figure}
837: \includegraphics[width=13cm]{Fig7.eps}
838: \caption{Hysteresis loop for (a) $r_{\rm c}<r<0$; (b) $r<r_{\rm
839: c}$.} \label{fig7}
840: \end{figure}
841: The points labelled by asterisks have the same significance as in
842: Fig. \ref{fig5}(b). If in Fig. \ref{fig7}(a) we increase
843: $vg_{\parallel}$ beyond its value indicated by the right-hand
844: asterisk we move into the ``both-unstable'' region where the
845: magnetisation direction of the switching magnet is perpetually in a
846: time-dependent state. Thus negative $r$ introduces behaviour in zero
847: applied field which is similar to that found when the applied field
848: exceeds the coercive field of the switching magnet for $r=0$. This
849: behaviour was predicted by Edwards {\it et al.} \cite{13}, in
850: particular for a Co/Cu/Co(111) system with the switching magnet
851: consisting of a Co monolayer. Zimmler {\it et al.} \cite{26} use
852: methods similar to the ones described here to analyse their data on
853: a Co/Cu/Co nanopillar and deduce that $g_{\parallel}>0$,
854: $r=g_{\perp}/g_{\parallel}\approx -0.2$. It would be interesting to
855: carry out time-resolved resistance measurements on this system at
856: large current density (corresponding to $vg_{\perp}<-1$) and zero
857: external field.
858:
859: So far we have considered the low-field regime
860: ($H_{\rm ext}\approx$ coercive field of switching magnet) with both
861: magnetisations and the external field in-plane. There is another
862: class of experiments in which a high field, greater than the
863: demagnetising field ($>2T$), is applied perpendicular to the plane
864: of the layers. The magnetisation of the polarising magnet is then
865: also perpendicular to the plane. This is the situation in the early
866: experiments where a point contact was employed to inject high
867: current densities into magnetic multilayers \cite{27,28,29}. In this
868: high-field regime a peak in the differential resistance ${\rm
869: d}V/{\rm d}I$ at a critical current was interpreted as the onset of
870: current-induced excitation of spin waves in which the spin-transfer
871: torque leads to uniform precession of the magnetisation
872: \cite{6,27,28}. No hysteretic magnetisation reversal was observed
873: and it seemed that the effect of spin-polarised current on the
874: magnetisation is quite different in the low- and high-field regimes.
875: Recently, however, \"{O}zyilmaz {\it et al.} \cite{30} have studied
876: Co/Cu/Co nanopillars ($\approx 100$nm in diameter) at $T=4.2$K for
877: large applied fields perpendicular to the layers. They observe
878: hysteretic magnetisation reversal and interpret their results using
879: the Landau-Lifshitz equation. We now give a similar discussion
880: within the framework of this section.
881:
882: Following \"Ozyilmaz {\it et al.}, we neglect the uniaxial
883: anisotropy term in Eq. (\ref{Gamma3}) for the reduced torque ${\bm\Gamma}$
884: while retaining $H_{u0}$ as a scalar factor. Hence
885: \begin{equation}
886: \label{gamma25}
887: {\bm\Gamma}=H_{u0}\left\{\left[h_{\rm
888: ext}+v_{\perp}(\psi)-h_{\rm
889: p}\cos\psi\right]{\bf m}\times{\bf p}+v_{\parallel}(\psi){\bf m}\times({\bf p}\times{\bf m})\right\}
890: \end{equation}
891: where ${\bf p}$ is the unit vector perpendicular to the plane. When
892: $v_{\parallel}(\psi)\neq 0$ the only possible steady-state solutions
893: of ${\bm\Gamma} =0$ are ${\bf m}_0=\pm{\bf p}$. On linearizing Eq. \ref{Gamma2}
894: about ${\bf m}_0$ as before we find that the condition $G\geqslant
895: 0$ is always satisfied. The second stability condition $F<0$ becomes
896: \begin{equation}
897: \label{stabcond}
898: \left[v_{\parallel}(\psi_0)+\gamma(v_{\perp}(\psi_0)+h_{\rm
899: ext}-h_{\rm p})\right]\cos\psi_0>0
900: \end{equation}
901: where $\psi_0=\cos^{-1}({\bf m}_0\cdot{\bf p})$. Applying this to the P
902: state ($\psi_0=0$) and the AP state ($\psi_0=\pi$) we obtain the
903: conditions
904: \begin{equation}
905: \label{conditiona} v>\gamma(h_{\rm p}-h_{\rm ext})/g(0)
906: \end{equation}
907: \begin{equation}\label{conditionb}
908: v<-\gamma(h_{\rm p}+h_{\rm ext})/g(\pi),
909: \end{equation}
910: where the first condition applies to the P stability and the second
911: to the AP stability. Here $g(\psi)=g_{\parallel}(\psi)+\gamma
912: g_{\perp}(\psi)$. The corresponding stability diagram is shown in
913: Fig. \ref{fig8}, where we have assumed $g(\pi)>g(0)>0$ for
914: definiteness.
915: \begin{figure}
916: \includegraphics[width=11cm]{Fig8.eps}
917: \caption{Bias-field stability diagram for large external field
918: ($h_{\rm ext}>h_{\rm p}$) perpendicular to the layers. }
919: \label{fig8}
920: \end{figure}
921: The boundary lines cross at $h_{\rm ext}=h_{\rm c}$, where $h_{\rm
922: c}=h_{\rm p}[g(\pi)+g(0)]/[g(\pi)-g(0)]$. This analysis is only
923: valid for fields larger than the demagnetising field ($h_{\rm
924: ext}>h_{\rm p}$) and we see from the figure that for $h_{\rm
925: ext}>h_{\rm c}$ hysteretic switching occurs. This takes place for
926: only one sign of the bias (current) and the critical biases
927: (currents) increase linearly with $h_{\rm ext}$ as does the width of
928: the hysteresis loop $|v_{{\rm P}\rightarrow{\rm AP}}-v_{{\rm
929: AP}\rightarrow{\rm P}}|$. This accords with the observations of
930: \"Ozyilmaz {\it et al.} The critical currents are not larger than
931: those in the low-field or zero-field regimes (cf. Eqs.
932: (\ref{conditiona}), (\ref{conditionb}) with Eq. (\ref{crbias})) and yet the
933: magnetisation of the switching magnet can be switched against a very
934: large external field. However, in this case the AP state is only
935: stabilised by maintaining the current.
936:
937: The experiments on spin transfer discussed above have mainly been
938: carried out at constant temperature, typically $4.2$K or room
939: temperature. The effect on current-driven switching of varying the
940: temperature has recently been studied by several groups
941: \cite{23,25,31}. The standard N\'eel-Brown theory of thermal
942: switching \cite{32} does not apply because the Slonczewski in-plane
943: torque is not derivable from an energy function. Li and Zhang
944: \cite{33} have generalised the standard stochastic Landau-Lifschitz
945: equation, which includes white noise in the effective applied field,
946: to include spin transfer torque. In this way they have successfully
947: interpreted some of the experimental data. A full discussion of this
948: work is outside the scope of the present review. However it should
949: be pointed out that in addition to the classical effect of white
950: noise there is an intrinsic temperature dependence of quantum
951: origin. This arises from the Fermi distribution functions which
952: appear in expressions for the spin-transfer torque (see Eqs.
953: (\ref{torque}) and (\ref{landau33})).
954:
955: So far we have discussed steady-state solutions of the LLG equation
956: (\ref{Gamma2}). It is important to study the magnetisation dynamics
957: of the switching layer in the situation during the jumps
958: AP$\rightarrow$P and P$\rightarrow$AP of the hysteresis curve in
959: zero external field, and secondly under conditions where only
960: time-dependent solutions are possible, for example in the regions of
961: sufficiently strong current and external field marked ''both
962: unstable" in Fig. \ref{fig4}. The first situation has been studied
963: by Sun \cite{sun}, assuming single-domain behaviour of the switching
964: magnet, and by Miltat {\it et al.} \cite{34} with more general
965: micromagnetic configurations. Both situations have been considered
966: by Li and Zhang \cite{35}. In the second case they find precessional
967: states, and the possibility of ''telegraph noise" at room
968: temperature, as seen experimentally in Refs. \cite{22,24}. Switching
969: times (AP$\rightarrow$P and P$\rightarrow$AP) are estimated to be of
970: the order 1ns. Micromagnetic simulations \cite{34} indicate that the
971: Oersted field cannot be completely ignored for typical pillars with
972: diameter of the order of 100nm.
973:
974: Finally, in this section, we briefly discuss some practical
975: considerations which may ultimately decide whether current-induced
976: switching is useful in spintronics. Sharp switching, with nearly
977: rectangular hysteresis loops, is obviously desirable and this
978: demands single-domain behaviour. In experiments on nanopillars of
979: circular cross section \cite{21} multidomain behaviour was observed
980: with the switching transition spread over a range of current.
981: Subsequently the same group \cite{albert} found sharp switching in
982: pillars whose cross-section was an elongated hexagon, which
983: introduces strong uniaxial in-plane shape anisotropy. It was known
984: from earlier magnetisation studies of nanomagnet arrays \cite{36}
985: that such a shape anisotropy can result in single domain behaviour.
986: A complex switching transition need not necessarily indicate
987: multidomain behaviour. It could also arise from a marked departure
988: of $T_{\perp}(\psi)$ and/or $T_{\parallel}(\psi)$ from sinusoidal
989: behaviour, such as occurs near $\psi=\pi$ in calculations for
990: Co/Cu/Co(111) with two atomic planes of Co in the switching magnet
991: (see Fig. \ref{fig13}(b)). In the calculations of the corresponding
992: hysteresis loops (Fig. \ref{fig16}) the torques were approximated by
993: sine curves but an accurate treatment would certainly complicate the
994: AP$\rightarrow$P transition which occurs at negative bias in Fig.
995: \ref{fig16}(b). Studies of this effect are planned.
996:
997: The critical current density for switching is clearly an important
998: parameter. From Eq. (\ref{crbias}) the critical reduced bias for the
999: P$\rightarrow$AP transition is to a good approximation given by
1000: $-\gamma h_{\rm p}/[2g_{\parallel}(0)]$. Using the definitions of
1001: reduced quantities given after Eq. (\ref{Gamma3}), we may write the
1002: actual critical bias in volts as
1003: \begin{equation} \label{VPAP}
1004: V_{P\rightarrow AP}=M\gamma M_{\rm s}H_{\rm d}/[2g_{\parallel}(0)|e|],
1005: \end{equation}
1006: where $M$ is the number of atomic planes in the switching magnet,
1007: $M_{\rm s}$ is the average moment ($J/T$) of the switching magnet
1008: per atomic plane per unit area, and $H_{\rm d}=\hbar H_{\rm
1009: p0}/(2\mu_{\rm B})$ is the easy-plane anisotropy field in tesla. As
1010: expressed earlier $g_{\parallel}(0)=({\rm d}T_{\parallel}/{\rm
1011: d}\psi)_{\psi =0}$ where the torque $T_{\parallel}$ is per unit area
1012: in units of $eV_{\rm B}$. (The calculated torques in Figs.
1013: \ref{fig13} and \ref{fig14} of Sec. \ref{cinque} are per surface
1014: atom so that if these are used to determine $g_{\parallel}(0)$ in
1015: Eq. (\ref{VPAP}) $M_{\rm s}$ must be taken per surface atom.)
1016:
1017: An obvious way to reduce the critical bias, and hence the critical
1018: current, is to reduce $M$, the thickness of the switching magnet.
1019: Calculations show \cite{13} (see also Fig. \ref{fig14}) that
1020: $g_{\parallel}$ does not decrease with $M$ and may, in fact,
1021: increase for small values such as $M=2$. Careful design of the
1022: device might also increase $g_{\parallel}(0)$ beyond the values
1023: ($<0.01$ per surface atom) which seem to be obtainable in simple
1024: trilayers \cite{13}. Jiang {\it et al.} \cite{37,38}, have studied
1025: various structures in which the polarising magnet is pinned by an
1026: adjacent antiferromagnet (exchange biasing) and in which a thin Ru
1027: layer is incorporated between the switching layer and the lead.
1028: Critical current densities of $2\times 10^{6}$Acm$^{-2}$ have been
1029: obtained which are substantially lower than those in Co/Cu/Co
1030: trilayers. Such structures can quite easily be investigated
1031: theoretically by the methods of Section \ref{cinque}.
1032:
1033: Decreasing the magnetisation $M_{\rm s}$, and hence the
1034: demagnetising field ($\propto H_{\rm d}$), would be favourable but
1035: $g_{\parallel}$ then tends to decrease also \cite{13}. A possible
1036: way of decreasing $H_{\rm d}$ without decreasing local magnetic
1037: moments in the system is to use a synthetic ferrimagnet as the
1038: switching magnet \cite{39}. The Gilbert damping factor $\gamma$ is
1039: another crucial parameter but it is uncertain whether this can be
1040: decreased significantly. However, the work of Capelle and Gyorffy
1041: \cite{40} is an interesting theoretical development. The search for
1042: structures with critical current densities low enough for use in
1043: spintronic devices ($10^{5}$Acm$^{-2}$ perhaps) \cite{41} is an
1044: enterprise where experiment and quantitative calculations \cite{13}
1045: should complement each other fruitfully.
1046:
1047: \section{Quantitative theory of spin-transfer torque}\label{quattro}
1048: \subsection{General principles}
1049:
1050: To put the phenomenological treatment of Sec. \ref{tre} on a
1051: first-principle quantitative basis we must calculate the
1052: spin-transfer torques (Eqs. (\ref{torque2}) in a steady state for
1053: real systems. For this purpose it is convenient to describe the
1054: magnetic and nonmagnetic layers of Fig. \ref{fig1} by tight-binding
1055: models, in general multiorbital with s, p, and d orbitals whose
1056: one-electron parameters are fitted to first-principle bulk band
1057: structure \cite{42}. The hamiltonian is therefore of the form
1058: \begin{equation} \label{HAM}
1059: H=H_0+H_{\rm int}+H_{\rm anis}
1060: \end{equation}
1061: where the one-electron hopping term $H_0$ is given by
1062: \begin{equation} \label{H0}
1063: H_0=\sum_{k_{\parallel}\sigma}\sum_{m\mu,n\nu}t_{m\mu,n\nu}({\bf
1064: k}_{\parallel})
1065: c^{\dagger}_{{\bf k}_{\parallel}m\mu\sigma}
1066: c_{{\bf k}_{\parallel}n\nu\sigma},
1067: \end{equation}
1068: where $c^{\dagger}_{k_{\parallel}m\mu\sigma}$ creates an electron in
1069: a Bloch state, with in-plane wave vector ${\bf k}_{\parallel}$ and
1070: spin $\sigma$, formed from a given atomic orbital $\mu$ in plane
1071: $m$. Eq. \ref{H0} generalises the single orbital eq.
1072: (\ref{hamiltonian}). $H_{\rm int}$ is an on-site interaction between
1073: electrons in d orbitals which leads to an exchange splitting of the
1074: bands in the ferromagnets and is neglected in the spacer and lead.
1075: Finally, $H_{\rm anis}$ contains anisotropy fields in the switching
1076: magnet and is given by
1077: \begin{equation}
1078: \label{Hanis}
1079: H_{\rm anis}=-\sum_{n}{\bf S}_{n}\cdot{\bf H}_{\rm A},
1080: \end{equation}
1081: where ${\bf S}_n$ is the operator of the total spin angular momentum
1082: of plane $n$ and ${\bf H}_{\rm A}$ is given by Eqs. (\ref{hh})-(\ref{hp})
1083: with the unit vector ${\bf m}$ in the direction of
1084: $\sum_{n}\langle{\bf S}_n\rangle$, where $\langle{\bf S}_n\rangle$
1085: is the thermal average of ${\bf S}_n$. We assume here that the
1086: anisotropy fields $H_{\rm u0}$,$H_{\rm p}$ are uniform throughout
1087: the switching magnet but we could generalise to include, for example,
1088: a surface anisotropy.
1089:
1090: In the tight-binding description, the spin angular momentum operator
1091: ${\bf S}_n$ is given by
1092: \begin{equation}
1093: \label{Sn}
1094: {\bf S}_{n}=\frac12\hbar\sum_{k_{\parallel\mu}}
1095: (c^{\dagger}_{k_{\parallel}n\mu\uparrow},
1096: c^{\dagger}_{k_{\parallel}n\mu\downarrow})
1097: {\bm\sigma}(c_{k_{\parallel}n\mu\uparrow},
1098: c_{k_{\parallel}n\mu\downarrow})^{\rm T}
1099: \end{equation}
1100: and the corresponding operator for the spin angular momentum current
1101: between planes $n-1$ and $n$ is
1102: \begin{equation}
1103: \label{jj}
1104: {\bf j}_{n-1}=-\frac12 \hbar\sum_{{\bf k}_{\parallel}\mu\nu}
1105: t({\bf k}_{\parallel})_{n\nu,n-1\mu}
1106: \left(c^{\dagger}_{k_{\parallel}n\nu\uparrow},
1107: c^{\dagger}_{k_{\parallel}n\nu\downarrow}\right)
1108: {\bm\sigma}
1109: \left(c_{k_{\parallel}n-1\mu\uparrow},
1110: c_{k_{\parallel}n-1\mu\downarrow}\right)^{\rm T}+{\rm h.c.},
1111: \end{equation}
1112: which generalises the single orbital expression (\ref{gamma}). The
1113: rate of change of ${\bf S}_n$ in the switching magnet is given by
1114: \begin{equation}\label{rate}
1115: {\rm i}\hbar\frac{{\rm d}{\bf S}_n}{{\rm d}t}=[{\bf S}_n,H_0]+[{\bf
1116: S}_n,H_{\rm anis}].
1117: \end{equation}
1118: This results holds since the spin operator commutes with the
1119: interaction hamiltonian $H_{\rm int}$.
1120:
1121: It is straightforward to show that
1122: \begin{equation}
1123: \label{commut}
1124: [{\bf S}_n,H_0]={\rm i}\hbar({\bf j}_{n-1}-{\bf j}_{n}),
1125: \end{equation}
1126: and
1127: \begin{equation}\label{commut2}
1128: [{\bf S}_n,H_{\rm anis}]=-{\rm i}\hbar({\bf H}_{\rm A}\times{\bf
1129: S}_n).
1130: \end{equation}
1131: On taking the thermal average, Eq. (\ref{rate}) becomes
1132: \begin{equation}\label{thermal}
1133: \langle\frac{{\rm d}{\bf S}_n}{{\rm d}t}\rangle=\langle
1134: {\bf j}_{n-1}\rangle-\langle {\bf j}_n\rangle-{\bf H}_{\rm
1135: A}\times\langle
1136: {\bf S}_{\rm tot}
1137: \rangle,
1138: \end{equation}
1139: This corresponds to an equation of continuity, stating that the rate
1140: of change of spin angular momentum on plane $n$ is equal to the
1141: difference between the rate of flow of this quantity onto and off
1142: the plane, plus the rate of change due to precession around the
1143: field ${\bf H}_{\rm A}$. When Eq. (\ref{thermal}) is summed over all
1144: planes in the switching magnet we have
1145: \begin{equation}\label{thermal2}
1146: \frac{{\rm d}}{{\rm d}t} \langle{\bf S}_{\rm tot}\rangle=
1147: {\bf T}^{\rm s-t}-{\bf H}_{\rm
1148: A}\times \langle {\bf S}_{\rm tot}\rangle,
1149: \end{equation}
1150: where the total spin-transfer torque ${\bf T}^{\rm s-t}$ is
1151: given by Eq. (\ref{torque}) and $\langle{\bf S}_{\rm tot}\rangle$ is
1152: the total spin angular momentum of the switching magnet. Equation
1153: (\ref{thermal2}) is equivalent to Eq. (\ref{llg}), for zero external
1154: field, in the absence of damping. Equation (\ref{torque}) shows how
1155: ${\bf T}^{\rm s-t}$ required for the phenomenological treatment of Sec.
1156: \ref{tre} is to be determined from the calculated spin currents in
1157: the spacer and lead. As discussed in Sec. \ref{tre}, the
1158: magnetization of a single-domain sample is essentially uniform and
1159: the spin-transfer torque ${\bf T}^{\rm s-t}$ depends on the angle $\psi$
1160: between the magnetisations of the polarising and switching magnets.
1161:
1162: To consider time-dependent solutions of Eq. (\ref{llg}) it is
1163: necessary to calculate ${\bf T}^{\rm s-t}$ for arbitrary angle $\psi$ and
1164: for this purpose ${\bf H}_{\rm A}$ can be neglected. To reduce the
1165: calculation of the spin-transfer torque to effectively a
1166: one-electron problem, we replace $H_{\rm int}$ by a selfconsistent
1167: exchange field term $-\sum_n{\bf S}_n\cdot{\bm\Delta}_n$, where the
1168: exchange field ${\bm\Delta}_n$ should be determined
1169: selfconsistently in the spirit of an unrestricted Hartree-Fock (HF)
1170: or local spin density (LSD), approximation. The essential
1171: selfconsistency condition in any HF or LSD calculation is that the
1172: local moment $\langle{\bf S}_n\rangle$ in a steady state is in the
1173: same direction as ${\bm\Delta}_n$. Thus we require
1174: \begin{equation}\label{delta}
1175: {\bm\Delta}_n\times\langle{\bf S}_n\rangle=0
1176: \end{equation}
1177: for each atomic plane of the switching magnet. It is useful to
1178: consider first the situation when there is no applied bias and the
1179: polarising and switching magnets are separated by a spacer which is
1180: so thick that the zero-bias oscillatory exchange coupling \cite{44}
1181: is negligible. In that case we have two independent magnets and the
1182: selfconsistent exchange field in every atomic plane of the switching
1183: magnet is parallel to its total magnetisation which is uniform and
1184: assumed to be along the $z$-axis. Referring to Fig. \ref{fig1} the
1185: selfconsistent solution therefore corresponds to uniform exchange
1186: fields in the polarising and switching magnets which are at an
1187: assumed angle $\psi=\theta$ with respect to one another.
1188:
1189: When a bias $V_{\rm b}$ is applied , with a uniform exchange field
1190: ${\bm\Delta }=\Delta{\bf e}_z$ in the switching magnet imposed, the
1191: calculated local moments $\langle{\bf S}_n\rangle$ will deviate from
1192: the $z$-direction so that the solution is not selfconsistent. To
1193: prepare a selfconsistent state with ${\bm\Delta}$ and all
1194: $\langle{\bf S}_n\rangle=\langle {\bf S}\rangle$ in the
1195: $z$-direction it is necessary to apply fictitious constraining
1196: fields ${\bf H}_n$ of magnitude proportional to $V_{\rm b}$. The
1197: local field for plane $n$ is thus ${\bm\Delta}+{\bf H}_n$ but to
1198: calculate the spin currents in the spacer and lead, and hence
1199: ${\bf T}^{\rm s-t}$ from Eq. (\ref{torque}), the fields ${\bf H}_n$, of
1200: the order of $V_{\rm b}$, may be neglected compared with ${\bf
1201: \Delta}$. Although the fictitious constraining fields ${\bf H}_n$
1202: need therefore never be calculated, it is interesting to see that
1203: they are in fact related to ${\bf T}^{\rm s-t}$. For the constrained
1204: self-consistent steady state ($\langle{\bf S}_n\rangle=\langle {\bf
1205: S}\rangle$, $\langle{\bf \dot{S}}_n\rangle=0$) in the presence of the
1206: constraining fields, with ${\bf H}_{\rm A}$ neglected as discussed
1207: above, it follows from Eq. (\ref{thermal}) that
1208: \begin{equation}
1209: \label{jjj} \langle{\bf j}_{\bf n-1}\rangle-\langle{\bf j}_{\bf
1210: n}\rangle=(\Delta+{\bf H}_n)\times\langle S\rangle={\bf
1211: H}_n\times\langle{\bf S}\rangle,
1212: \end{equation}
1213: where the local field ${\bm\Delta}+{\bf H}_n$ replaces ${\bf
1214: H}_{\rm A}$. On summing over all atomic planes $n$ in the switching
1215: magnet we have
1216: \begin{equation}
1217: \label{ttt} {\bf T}^{\rm s-t}=\langle{\bf j}_{\rm
1218: spacer}\rangle-\langle{\bf j}_{\rm lead}\rangle=\sum_n{\bf
1219: H}_n\times\langle{\bf S}\rangle.
1220: \end{equation}
1221: Thus, as expected, in the prepared state with a given angle $\psi$
1222: between the magnetisations of the magnetic layers the spin-transfer
1223: torque is balanced by the total torque due to the constraining
1224: fields.
1225:
1226: In the simple model of Section \ref{due}, with infinite exchange
1227: splitting in the magnets, the local moment is constrained to be in
1228: the direction of the exchange field so the question of
1229: selfconsistency is not raised.
1230:
1231: The main conclusion of this Section is that the spin-transfer torque
1232: for a given angle $\psi$ between magnetisations may be calculated
1233: using uniform exchange fields making the same angle with one
1234: another. Such calculations are described in Sec. \ref{due} and
1235: \ref{cinque}. The use of this spin-transfer torque in the LLG
1236: equation of Section \ref{tre} completes what we shall call the
1237: ''standard model" (SM). It underlies the original work of
1238: Slonczewski \cite{3} and most subsequent work. The spin-transfer
1239: torque calculated in this way should be appropriate even for
1240: time-dependent solutions of the LLG equation. This is based on the
1241: reasonable assumption that the time for the electronic system to
1242: attain a ''constrained steady state" with given $\psi$ is short
1243: compared with the time-scale ($\approx$1ns) of the macroscopic
1244: motion of the switching magnet moment.
1245:
1246: Although the SM is a satisfactory way of calculating the
1247: spin-transfer torque its lack of selfconsistency leads to some
1248: non-physical concepts. The first of these is the "transverse spin
1249: accumulation" in the switching magnet \cite{46,47}, This refers to
1250: the deviations of local moments $\langle{\bf S}_n\rangle$ from the
1251: direction of the exchange field, assumed uniform in the SM. In a
1252: self-consistent treatment such deviations do not occur because the
1253: exchange field is always in the direction of the local moment. A
1254: related non-physical concept is the ''spin decoherence length" over
1255: which the spin accumulation is supposed to decay \cite{46,47}, More
1256: detailed critiques of these concepts are given elsewhere
1257: \cite{13,em}.
1258:
1259: \subsection{Keldysh formalism for fully realistic calculations of the
1260: spin-transfer torque}
1261:
1262: The wave-function approach to spin-transfer torque described in
1263: Section \ref{due} is difficult to apply to realistic multiorbital
1264: systems. For this purpose Green functions are much more convenient
1265: and Keldysh \cite{11} developed a Green function approach to the
1266: non-equilibrium problem of electron transport. In this section we
1267: apply this method to calculate spin currents in a magnetic layer
1268: structure, following Edwards {\it et al.} \cite{13}.
1269:
1270: The structure we consider is shown schematically in Fig. \ref{fig1}.
1271: It consists of a thick (semi-infinite) left magnetic layer
1272: (polarising magnet), a nonmagnetic metallic spacer layer of $N$
1273: atomic planes, a thin switching magnet of $M$ atomic planes, and a
1274: semi-infinite lead. The broken line between the atomic planes $n-1$
1275: and $n$ represents a cleavage plane separating the system into two
1276: independent parts so that charge carriers cannot move between the
1277: two surface planes $n-1$ and $n$. It will be seen that our ability
1278: to cleave the whole system in this way is essential for the
1279: implementation of the Keldysh formalism. This can be easily done
1280: with a tight-binding parametrisation of the band structure by simply
1281: switching off the matrix of hopping integrals $t_{n\nu,n-1\mu}$
1282: between atomic orbitals $\nu$, $\mu$ localised in planes $n-1$ and
1283: $n$. We therefore adopt the tight-binding description with the
1284: Hamiltonian defined by Eqs. (\ref{HAM}-\ref{Sn}).
1285:
1286: To use the Keldysh formalism \cite{11,12,53} to calculate the
1287: charge or spin currents flowing between the planes $n-1$ and $n$, we
1288: consider an initial state at time $\tau=-\infty$ in which the
1289: hopping integral $t_{n\nu,n-1\mu}$ between planes $n-1$ and $n$ is
1290: switched off. Then both sides of the system are in equilibrium but
1291: with different chemical potentials $\mu_{\rm L}$ on the left and
1292: $\mu_{\rm R}$ on the right, where $\mu_{\rm L}-\mu_{\rm R}=eV_{\rm
1293: b}$. The interplane hopping is then turned on adiabatically and the
1294: system evolves to a steady state. The cleavage plane, across which
1295: the hopping is initially switched off, may be taken in either the
1296: spacer or in one of the magnets or in the lead. In principle, the
1297: Keldysh method is valid for arbitrary bias $V_{\rm b}$ but here we
1298: restrict ourselves to small bias corresponding to linear response.
1299: This is always reasonable for a metallic system. For larger bias,
1300: which might occur with a semiconductor or insulator as spacer,
1301: electrons would be injected into the right part of the system far
1302: above the Fermi level and many-body processes neglected here would
1303: be important. Following Keldysh \cite{11,12}, we define a two-time
1304: matrix
1305: \begin{equation}\label{kel}
1306: G_{\rm RL}^{+}(\tau,\tau^{\prime})={\rm i}\langle c_{\rm
1307: L}^{\dagger}(\tau^{\prime})c_{\rm R}(\tau)\rangle,
1308: \end{equation}
1309: where $R\equiv(n,\nu,\sigma^{\prime})$ and
1310: $L\equiv(n-1,\mu,\sigma)$, and we suppress the $k_{\parallel}$
1311: label. The thermal average in Eq. (\ref{kel}) is calculated for the
1312: steady state of the coupled system. The matrix $G_{\rm
1313: RL}^{\dagger}$ has dimensions $2m\times 2m$ where $m$ is the number
1314: of orbitals on each atomic site, and is written so that the $m\times
1315: m$ upper diagonal block contains matrix elements between $\uparrow$
1316: spin orbitals and the $m\times m$ lower diagonal block relates to
1317: $\downarrow$ spin. $2m\times 2m$ hopping matrices $t_{\rm LR}$ and
1318: $t_{\rm RL}$ are written similarly and in this case only the
1319: diagonal blocks are nonzero. If we denote $t_{\rm LR}$ by $t$, then
1320: $t_{\rm RL}=t^{\dagger}$. We also generalise the definition of
1321: ${\bm\sigma}$
1322: so that its components are now direct products of the
1323: $2\times 2$ Pauli matrices $\sigma_x$, $\sigma_y$, $\sigma_z$, and
1324: the $m\times m$ unit matrix. The thermal average of the spin current
1325: operator, given by Eq. (\ref{rate}), may now be expressed as
1326: \begin{equation}\label{uffa}
1327: \langle{\bf j}_{n-1}\rangle=\frac12\sum_{{\bf k}_{\parallel}}{\rm
1328: Tr}\left\{\left[G_{\rm RL}^{+}\left(\tau,\tau\right)t-G_{\rm
1329: LR}^{+}(\tau,\tau)t^{\dagger}\right]{\bm\sigma}\right\}.
1330: \end{equation}
1331: Introducing the Fourier transform $G^{+}(\omega)$ of
1332: $G^{+}(\tau,\tau^{\prime})$ , which is a function of
1333: $\tau-\tau^{\prime}$, we have
1334: \begin{equation}\label{uffa2}
1335: \langle{\bf j}_{n-1}\rangle=\frac12\sum_{{\bf k}_{\parallel}}
1336: \int\frac{{\rm d}\omega}{2\pi}{\rm Tr}\left\{\left[G_{\rm
1337: RL}^{+}\left(\omega\right)t-G_{\rm
1338: LR}^{+}(\omega)t^{\dagger}\right]{\bm\sigma}\right\}.
1339: \end{equation}
1340: The charge current is given by Eq. (\ref{uffa2}) with
1341: $\frac12{\bm\sigma}$ replaced by the unit matrix multiplied by
1342: $e/\hbar$.
1343:
1344: Following Keldysh \cite{11,12} we now write
1345: \begin{equation}\label{uffa3}
1346: G_{\rm AB}^{+}(\omega)=\frac12\left(F_{\rm AB}+G_{\rm AB}^{\rm
1347: a}-G_{\rm AB}^{\rm r}\right),
1348: \end{equation}
1349: where the suffices $A$ and $B$ are either $R$ or $L$. $F_{\rm
1350: AB}(\omega)$ is the Fourier transform of
1351: \begin{equation}\label{uffa4}
1352: F_{\rm AB}(\tau,\tau^{\prime})=-{\rm i}\langle[c_{\rm A}(\tau),c_{\rm
1353: B}^{\dagger}(\tau^{\prime})]_{-}\rangle
1354: \end{equation}
1355: and $G^{\rm a}$, $G^{\rm r}$ are the usual advanced and retarded
1356: Green functions \cite{54}. Note that in \cite{11} and \cite{12} the definitions
1357: of $G^{\rm a}$ and $G^{\rm r}$ are interchanged and that in the
1358: Green function matrix defined by these authors $G^{+}$ and $G^{-}$
1359: should be interchanged.
1360:
1361: Charge and spin current are related by Eqs. (\ref{uffa2}) and (\ref{uffa3}) to
1362: the quantities $G^{\rm a}$, $G^{\rm r}$ and $F_{\rm AB}$. The latter
1363: are calculated for the coupled system by starting with decoupled
1364: left and right systems, each in equilibrium, and turning on the
1365: hopping between planes L and R as a perturbation. Hence, we express
1366: $G^{\rm a}$, $G^{\rm r}$ and $F_{\rm AB}$ in terms of retarded
1367: surface Green functions $g_{L}\equiv g_{\rm LL}$, $g_{\rm R}\equiv
1368: g_{\rm RR}$ for the decoupled equilibrium system. It is then found
1369: \cite{13} that the spin current between the planes $n-1$ and $n$ can
1370: be written as the sum $\langle{\bf j}_{n-1}\rangle=\langle{\bf
1371: j}_{n-1}\rangle_1+\langle{\bf j}_{n-1}\rangle_2$, where the two
1372: contributions to the spin current $\langle{\bf j}_{n}\rangle_1$,
1373: $\langle{\bf j}_{n}\rangle_2$ are given by
1374: \begin{equation}\label{nonloso1}
1375: \langle{\rm j}_{n-1}\rangle_1=\frac1{4\pi}\sum_{{\bf
1376: k}_\parallel}\int{\rm d}\omega\Re{\rm Tr}[(B-A){\bm\sigma}]
1377: [f(\omega-\mu_{\rm L})+f(\omega-\mu_{\rm R})].
1378: \end{equation}
1379:
1380: \begin{equation}\label{nonloso2}
1381: \langle{\rm j}_{n-1}\rangle_2=\frac1{2\pi}\sum_{{\bf
1382: k}_\parallel}\int{\rm d}\omega\Re{\rm Tr}\left\{[g_{\rm L}tABg_{\rm
1383: R}^{\dagger}t^{\dagger}-AB+\frac12(A+B)]{\bm\sigma}\right\}
1384: [f(\omega-\mu_{\rm L})-f(\omega-\mu_{\rm R})].
1385: \end{equation}
1386: Here, $A=[1-g_{\rm R}t^{\dagger}g_{\rm L}t]^{-1}$,
1387: $B=[1-g^{\dagger}_{\rm R}t^{\dagger}g^{\dagger}_{\rm L}t]^{-1}$, and
1388: as in Section \ref{due} $f(\omega-\mu)$ is the Fermi function with
1389: chemical potential $\mu$ and $\mu_{\rm L}-\mu_{\rm R}=eV_{\rm b}$.
1390: In the linear-response case of small bias which we are considering,
1391: the Fermi functions in Eq. (\ref{nonloso2}) are expanded to first
1392: order in $V_{\rm b}$. Hence the energy integral is avoided, being
1393: equivalent to multiplying the integrand by $eV_{\rm b}$ and
1394: evaluating it at the common zero-bias chemical potential $\mu_0$.
1395:
1396: It can be seen that Eqs. (\ref{nonloso1}) and (\ref{nonloso2}), which determine
1397: the spin and the charge currents, depend on just two quantities,
1398: {\it i.e.} the surface retarded one-electron Green functions for a
1399: system cleaved between two neighbouring atomic planes. The surface
1400: Green functions can be determined without any approximations by the
1401: standard adlayer method (see {\it e.g.} \cite{42,44}) for a fully
1402: realistic band structure.
1403:
1404: We first note that there is a close correspondence between Eqs.
1405: (\ref{nonloso1}), (\ref{nonloso2}) and the generalised Landauer formula
1406: (\ref{landau33}). The first term in Eq. (\ref{landau33}) corresponds to the
1407: zero-bias spin current $\langle{\bf j}_{n-1}\rangle_1$ given by
1408: Eq. (\ref{nonloso1}). When the cleavage plane is taken in the
1409: spacer, the spin current $\langle{\bf j}_{n-1}\rangle_1$ determines
1410: the oscillatory exchange coupling between the two magnets and it is
1411: easy to verify that the formula for the exchange coupling obtained
1412: from Eq. (\ref{nonloso1}) is equivalent to the formula used in
1413: previous total energy calculations of this effect \cite{42,44}. The
1414: contribution to the transport spin current given by Eq.
1415: (\ref{nonloso2}) clearly corresponds to the second term in the
1416: Landauer formula (\ref{landau33}) which is proportional to the bias
1417: in the linear response limit. Placing the cleavage plane first
1418: between any two neighbouring atomic planes in the spacer and then
1419: between any two neighbouring planes in the lead, we obtain from Eq.
1420: (\ref{nonloso2}) the total spin-transfer torque ${\bf T}^{\rm s-t}$
1421: of Eq. (\ref{torque}) in Section \ref{due}.
1422:
1423: The equivalence of the Keldysh and Landauer methods has been
1424: demonstrated by calculating the currents
1425: (\ref{nonloso1}) and (\ref{nonloso2}) analytically for the simple single
1426: orbital model of Section \ref{due}. The results of that section,
1427: such as Eq. (\ref{torquex}) are reproduced \cite{13}.
1428:
1429: \section{Quantitative results for
1430: C\lowercase{o}/C\lowercase{u}/C\lowercase{o}(111)}\label{cinque}
1431:
1432: We now discuss the application of the Keldysh formalism to a real
1433: system. In particular we consider a realistic multiorbital model of
1434: fcc Co/Cu/Co(111) with tight-binding parameters fitted to the
1435: results of the first-principles band structure calculations, as
1436: described previously \cite{42,44}.
1437:
1438: Referring to Fig. \ref{fig1}, the system considered by Edwards {\it
1439: et al.} \cite{13} consists of a semi-infinite slab of Co (polarising
1440: magnet), the spacer of 20 atomic planes of Cu, the switching magnet
1441: containing $M$ atomic planes of Co, and the lead which is
1442: semi-infinite Cu. The spacer thickness of 20 atomic planes of Cu was
1443: chosen so that the contribution of the oscillatory exchange coupling
1444: term is so small that it can be neglected. The spin currents in the
1445: right lead and in the spacer were determined from Eq.
1446: (\ref{nonloso2}).
1447: Figure \ref{fig13}(a),(b) shows the angular dependences of
1448: $T_{\parallel}$, $T_{\perp}$ for the cases $M=1$ and $M=2$.
1449: respectively.
1450: \begin{figure}
1451: \includegraphics[width=12cm]{Fig13.eps}
1452: \caption{Dependence of the spin-transfer torque $T_{\parallel}$ and
1453: $T_{\perp}$ for Co/Cu/Co(111) on the angle $\psi$. The torques per
1454: surface atom are in units of $eV_{\rm b}$. Figure (a) is for $M=1$,
1455: and (b) for $M=2$ monolayers of Co in the switching magnet. }
1456: \label{fig13}
1457: \end{figure}
1458: For the monolayer switching magnet, the torques $T_{\parallel}$ and
1459: $T_{\perp}$ are equal in magnitude and they have opposite sign.
1460: However, for $M=2$, the torques have the same sign and $T_{\perp}$
1461: is somewhat smaller than $T_{\parallel}$. A negative sign of the
1462: ratio of the two torque components has important and unexpected
1463: consequences for hysteresis loops as already discussed in Section
1464: \ref{tre}. It can be seen that the angular dependence of both torque
1465: components is dominated by a $\sin\psi$ factor but distortions from
1466: this dependence are clearly visible. In particular, the slopes at
1467: $\psi=0$ and $\psi=\pi$ are quite different. As pointed out in
1468: Section \ref{tre}, this is important in the discussion of the
1469: stability of steady states and leads to quite different magnitudes
1470: of the critical biases $V_{\rm P}\rightarrow V_{\rm AP}$ and $V_{\rm
1471: AP}\rightarrow V_{\rm P}$.
1472:
1473: In Fig. \ref{fig14} we reproduce the dependence of $T_{\perp}$ and
1474: $T_{\parallel}$ on the thickness of the Co switching magnet. It can
1475: be seen that the out-of-plane torque $T_{\perp}$ becomes smaller
1476: than $T_{\parallel}$ for thicker switching magnets.
1477: \begin{figure}
1478: \includegraphics[width=8cm]{Fig14.eps}
1479: \caption{Dependence of the spin-transfer torque $T_{\parallel}$ and
1480: $T_{\perp}$ for Co/Cu/Co(111) on the thickness of the switching
1481: magnet $M$ for $\psi=\pi/3$. The torques are in units of $eV_{\rm
1482: b}$.}
1483: \label{fig14}
1484: \end{figure}
1485: However, $T_{\perp}$ is by no means negligible (27$\%$ of
1486: $T_{\parallel}$) even for a typical experimental thickness of the
1487: switching Co layer of ten atomic planes. It is also interesting that
1488: beyond the monolayer thickness, the ratio of the two torques is
1489: positive with the exception of $M=4$.
1490:
1491: The microscopically calculated spin-transfer torques for
1492: Co/Cu/Co(111) were used by Edwards {\it et al.} \cite{13} as an
1493: input into the phenomenological LLG equation. For simplicity the
1494: torques as functions of $\psi$ were approximated by sine curves but
1495: this is not essential. The LLG equation was first solved numerically
1496: to determine all the steady states and then the stability discussion
1497: outlined in the phenomenological section was applied to determine
1498: the critical bias for which instabilities occur. Finally, the
1499: ballistic resistance of the structure was evaluated from the
1500: real-space Kubo formula at every point of the steady state path.
1501: Such a calculation for the realistic Co/Cu system then gives
1502: hysteresis loops of the resistance versus bias which can be compared
1503: with the observed hysteresis loops. The LLG equation was solved
1504: including a strong easy-plane anisotropy with $h_{\rm p}=100$. If we
1505: take $H_{\rm u0}=1.86\times 10^9$sec$^{-1}$, corresponding to a
1506: uniaxial anisotropy field of about 0.01T, this value of $h_{\rm p}$
1507: corresponds to the shape anisotropy for a magnetisation of
1508: $1.6\times 10^6$A/m, similar to that of Co \cite{sun}. Also a
1509: realistic value \cite{sun} of the Gilbert damping parameter
1510: $\gamma=0.01$ was used. Finally, referring to the geometry of Fig.
1511: \ref{fig1}, two different values of the angle $\theta$ were employed
1512: in these calculations: $\theta=2$rad and $\theta=3$rad, the latter
1513: value being close to the value of $\pi$ which is realised in most
1514: experiments.
1515:
1516: We first reproduce in Fig. \ref{fig16} the hysteresis loops for the
1517: case of Co switching magnet consisting of two atomic planes. We note
1518: that the ratio $r=T_{\perp}/T_{\parallel}\approx 0.65$ deduced from
1519: Fig. \ref{fig13} is positive in this case. Fig. \ref{fig16}(a) shows
1520: the hysteresis loop for $\theta=2$ and Fig. \ref{fig16}(b) that
1521: for $\theta=3$.
1522: \begin{figure}
1523: \includegraphics[width=13cm]{Fig16.eps}
1524: \caption{Resistance of the Co/Cu/Co(111) junction as a function of
1525: the applied bias, with $M=2$ monolayers of Co in the switching
1526: magnet. (a) is for $\theta=2$rad and (b) for $\theta=3$rad.}
1527: \label{fig16}
1528: \end{figure}
1529: The hysteresis loop for $\theta=3$ shown in Fig. \ref{fig16}(b)
1530: is an illustration of the stability scenario in zero applied field
1531: with $r>0$ discussed in Section \ref{tre}. As pointed out there the
1532: hysteresis curve is that of Fig. \ref{fig5}(a) which agrees with
1533: Fig. \ref{fig16}(b)
1534: when we remember that the reduced bias used in Fig. \ref{fig5} has
1535: the opposite sign from the bias in volts used in Fig. \ref{fig16}.
1536: It is rather interesting that the critical bias for switching is $\approx
1537: 0.2$mV both for $\theta=2$ and $\theta=3$. When this bias
1538: is converted to the current density using the calculated ballistic
1539: resistance of the junction, it is found \cite{13} that the critical
1540: current for switching is $\approx 10^7$A/cm$^2$, which is in very
1541: good agreement with experiments \cite{albert}.
1542:
1543: The hysteresis loops for the case of the Co switching magnet
1544: consisting of a single atomic plane are reproduced in Fig.
1545: \ref{fig17}. The values of $h_{\rm p}$, $\gamma$, $H_{\rm u0}$, and
1546: $\theta$ are the same as in the previous example.
1547: \begin{figure}
1548: \includegraphics[width=13cm]{Fig17.eps}
1549: \caption{Resistance of the Co/Cu/Co(111) junction as a function of
1550: the applied current, with $M=1$ monolayer of Co in the switching
1551: magnet. (a) is for $\theta=2$rad and (b) for $\theta=3$rad.}
1552: \label{fig17}
1553: \end{figure}
1554: However the ratio $r\approx -1$ is now negative and the hysteresis
1555: loops in Fig. \ref{fig17} illustrate the interesting behaviour
1556: discussed in Section \ref{tre} when the system subjected to a bias
1557: higher than a critical bias moves to the ''both unstable" region
1558: shown in Fig. \ref{fig6}.
1559: As in Fig. \ref{fig7} the points on the hysteresis loop in
1560: Fig. \ref{fig17} corresponding to the critical bias are labelled by asterisks.
1561: Fig. \ref{fig17}(b)
1562: and Fig. \ref{fig7}(a) are in close correspondence because Fig.
1563: \ref{fig7}(a) is for $r_{\rm c}<r<0$ and in the present case $r=-1$,
1564: $r_{\rm c}=-2/(\gamma h_{\rm p})=-2$. Also, from Fig.
1565: \ref{fig13}(a), $g_{\parallel}<0$ so that $vg_{\parallel}$ in Fig.
1566: \ref{fig7}(a) has the same sign as the voltage $V$ in Fig.
1567: \ref{fig17}(b).
1568:
1569: \section{Summary}
1570:
1571: Spin-transfer torque is responsible for current-driven switching of
1572: magnetisation in magnetic layered structures. The simplest
1573: theoretical scheme for calculating spin-transfer torque is a
1574: generalised Landauer method and this is used in Section \ref{due} to
1575: obtain analytical results for a simple model. The general
1576: phenomenological form of spin-transfer torque is deduced in Section
1577: \ref{tre} and this is introduced into the Landau-Lifshitz-Gilbert
1578: equation, together with torques due to anisotropy fields. This
1579: describes the motion of the magnetisation of the switching magnet
1580: and the stability of the steady states (constant current and
1581: stationary magnetisation direction) is studied under different
1582: experimental conditions, with and without external field. This leads
1583: to hysteretic and reversible behaviour in resistance versus bias (or
1584: current) plots in agreement with a wide range of
1585: experimental observations. In Section \ref{quattro} the general
1586: principles of a self-consistent treatment of spin-transfer torque are
1587: discussed and the Keldysh formalism for quantitative calculations is
1588: introduced. This approach to the non-equilibrium problem of electron
1589: transport uses Green functions which are very convenient to
1590: calculate for a realistic multiorbital tight-binding model of the
1591: layered-structure. In Section \ref{cinque} quantitative calculations
1592: for Co/Cu/Co(111) systems are presented which yield switching
1593: currents of the observed magnitude.
1594:
1595: This study of current-driven switching of magnetisation was carried
1596: out in collaboration with J. Mathon and A. Umerski and financial
1597: support was provided by the UK Engineering and Physical Research
1598: Council (EPSRC).
1599:
1600:
1601: \begin{thebibliography}{99}
1602:
1603: \bibitem{1} P. Gr\"{u}nberg, R. Schreiber, Y. Pang, M. B. Brodsky, and
1604: H. Sower, Phys. Rev. Lett. {\textbf 57}, 2442 (1986); M. N. Baibich,
1605: J. M. Broto, A. Fert, Van Dau Nguyen, F. Petroff, P. Etienne,
1606: G. Creuset, A. Friederich, and J. Chazelas, Phys. Rev. Lett. {\bf
1607: 61}, 2472 (1998)
1608: \bibitem{2}S. S. P. Parkin {\it et al.}, J. Appl. Phys. {\bf 85},
1609: 5828 (1999).
1610: \bibitem{3}J. C. Slonczewski, J. Magn. Magn. Mater. {\bf 159}, L1 (1996).
1611: \bibitem{4}X. Waintal, E. B. Myers, P. W. Brouwer, and D. C. Ralph,
1612: Phys. Rev. B {\bf 62}, 12317 (2000)
1613: \bibitem{5}{\it e.g.} ``Transport in Nanostructures'' by D. K. Ferry
1614: and S. M. Goodnick (Cambridge University Press 1997).
1615: \bibitem{sun} J. Z. Sun, Phys. Rev. B {\bf 62}, 570 (2000).
1616: \bibitem{albert} F. J. Albert, J. A. Katine, R. A. Buhrman, and
1617: D. C. Ralph, Appl. Phys. Lett. {\bf 77}, 3809 (2000).
1618: \bibitem{16} J. Grollier, V. Cross, A. Hamzic, J. M. George,
1619: H. Jaffres, A. Fert, G. Faini, J. Ben Youssef, and H. Le Gall,
1620: Appl. Phys. Lett. {\bf 78}, 3663 (2001).
1621: \bibitem{17}E. C. Stoner and E. P. Wohlfarth, Phil, Trans. Roy. Soc. A
1622: {\bf 240}, 599 (1948).
1623: \bibitem{18}J. Grollier, V. Cross, H. Jaffres, A. Hamzic, J. M. George,
1624: G. Faini, J. Ben Youssef, H. Le Gall, and A. Fert,
1625: Phys. Rev. B {\bf 67}, 174402 (2003).
1626: \bibitem{19} F. J. Albert, N. C. Emley, E. B. Myers, D. C. Ralph, and
1627: R. A. Buhrman, Phys. Rev. Lett. {\bf 89}, 226802 (2002).
1628: \bibitem{20} D. W. Jordan and P. Smith, Nonlinear Ordinary
1629: Differential Equations, Clarendon Press, Oxford (1977).
1630: \bibitem{13} D. M. Edwards, F. Federici, G. Mathon, and A. Umerski,
1631: Phys. Rev. B {\bf 71}, 134501 (2005).
1632: \bibitem{21} J.A. Katine, F. J. Albert, R. A. Buhrman, E. B. Myers, and
1633: D. C. Ralph, Phys. Rev. Lett. {\bf 84}, 3149 (2000).
1634: \bibitem{22} S. I. Kiselev, J. C. Sankey, I. N. Krivorotov,
1635: N. C. Emley, R. J. Schoelkopf, R. A. Buhrman, and D. C. Ralph, Nature {\bf 425}, 380 (2003).
1636: \bibitem{23} S. Urazhdin, N. O. Birge, W. P. Pratt, Jr., and J. Bass, Phys. Rev. Lett. {\bf 91}, 146803 (2003)
1637: \bibitem{24} M. R. Pufall, W. H. Rippard, S. Kaka, S. E. Russek,
1638: T. J. Silva, J. Katine, and M. Carey, Phy. Rev. Lett. {\bf 69},
1639: 214409 (2004).
1640: \bibitem{25} E. B. Myers, F. J. Albert, J. C. Sankey, E. Bonet,
1641: R. A. Buhrman, and D. C. Ralph, Phys. Rev. Lett. {\bf 89}, 196801 (2002)
1642: \bibitem{26} M. A. Zimmler, B. \"{O}zyilmaz, W. Chen, A. D. Kent,
1643: J. Z. Sun, M. J. Rooks, and R. H. Koch, Phy. Rev. B {\bf 70},
1644: 184438 (2004).
1645: \bibitem{27}M. Tsoi, A. G. M. Jansen, J. Bass, W. C. Chiang. M. Seck, V.Tsoi,
1646: and P. Wyder, Phys. Rev. Lett. {\bf 80}, 4281 (1998).
1647: \bibitem{28}M. Tsoi, A. G. M. Jansen, J. Bass, W. C. Chiang. V. Tsoi,
1648: and P. Wyder, Nature {\bf 406}, 46 (2000)
1649: \bibitem{29} E. B. Myers, D. C. Ralph, J. A. Katine, R. N. Louie and
1650: R. A. Buhrman, Science {\bf 285}, 867 (1999).
1651: \bibitem{6} J. C. Slonczewski, J. Magn. Magn. Mater. {\bf 195}, L261 (1999);
1652: {\bf 247}, 324 (2002).
1653: \bibitem{30}B. \"{O}zyilmaz, A. D. Kent, D. Monsma, J. Z. Sun,
1654: M. J. Rooks, and R. H. Koch, Phys. Rev. Lett. {\bf 91}, 067203 (2003).
1655: \bibitem{31} M. Tsoi, J. Z. Sun,
1656: M. J. Rooks, R. H. Koch, and S. S. P. Parkin, Phys. Rev. B {\bf 69}, 100406(R) (2004).
1657: \bibitem{32}W. F. Brown, Phys. Rev. B {\bf 130}, 1677 (1963).
1658: \bibitem{33}Z. Li and S. Zhang, Phys. Rev. B {\bf 69}, 134416 (2004).
1659: \bibitem{34}J. Miltat, G. Albuquerque, A. Thiaville, and C. Vouille, J. Appl.
1660: Phys. {\bf 89}, 6982 (2001).
1661: \bibitem{35}Z. Li and S. Zhang, Phys. Rev. B {\bf 68}, 024404 (2003).
1662: \bibitem{36} C. R. P. Cowburn, C. K. Koltsov, A. O. Adeyeye, and M.
1663: E. Welland, Phys. Rev. Lett. {\bf 83}, 1042 (1999).
1664: \bibitem{37} Y. Jiang, S. Abe, T. Ochiai, T. Nozaki, A. Hirohata, N.
1665: Tezuka, and K. Inomata, Phys. Rev. Lett. {\bf 92}, 167204 (2004).
1666: \bibitem{38} Y. Jiang, T. Nozaki, S. Abe, T. Ochiai, A. Hirohata, N.
1667: Tezuka, and K. Inomata, Nature Materials {\bf 3}, 361 (2004).
1668: \bibitem{39} N. Tezuka (private communication)
1669: \bibitem{40} K. Capelle and B. L. Gyorffy, Europhys. Lett. {\bf 61},
1670: 354 (2003).
1671: \bibitem{41} J. Sun, Nature {\bf 424}, 359 (2003).
1672: \bibitem{42} J. Mathon, Murielle Villeret, A. Umerski, R. B. Muniz,
1673: J. d'Albuquerque e Castro, and D. M. Edwards, Phys. Rev B, {\bf 56},
1674: 11797 (1997).
1675: \bibitem{44} J. Mathon, Murielle Villeret, R. B. Muniz, J.
1676: d'Albuquerque e Castro, and D. M. Edwards, Phys. Rev. Lett. {\bf
1677: 74}, 3696 (1995).
1678: \bibitem{46}S. Zhang, P. M. Levy, and A. Fert , Phys. Rev. Lett. {\bf
1679: 88},236601 (2002).
1680: \bibitem{47} A. A. Kovalev, A. Brataas, and G. E. W. Bauer, Phys.
1681: Rev. B {\bf 66 }, 224424 (2002).
1682: \bibitem{em} D. M. Edwards and J. Mathon in ''Nanomagnetism: Multilayers,
1683: Ultrathin Films and Textured Media", eds. J. A. C. Bland and D. L.
1684: Mills (Elsevier, to be published).
1685: \bibitem{11} L. V. Keldysh, Sov. Phys. JETP, {\bf 20}, 1018 (1965).
1686: \bibitem{12} C. Caroli, R. Combescot, P. Nozieres, and D.
1687: Saint-James, J. Phys. C {\bf 4}, 916 (1971).
1688: \bibitem{53}D. M. Edwards in: Exotic states in quantum
1689: nanostructures, ed. by S. Sarkar, Kluwer Academic Press (2002).
1690: \bibitem{54} G. D. Mahan, Many Particle Physics, 2nd Ed., Plenum
1691: Press, New York (1990).
1692: \end{thebibliography}
1693:
1694:
1695:
1696:
1697: \end{document}
1698: