+ All Categories
Home > Documents > Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic...

Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic...

Date post: 06-Mar-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
22
Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modeling D.V. Berkov a, , J. Miltat b a Innovent Technology Development, Pruessingstr. 27B, D-07745 Jena, Germany b Laboratoire de Physique des Solides, Univ. Paris-Sud & CNRS, 91405 Orsay, France Available online 31 December 2007 Abstract In this paper, we present an overview of recent progress made in the understanding of the spin-torque induced magnetization dynamics in nanodevices using mesoscopic micromagnetic simulations. We first specify how a spin-torque term may be added to the usual Landau–Lifshitz–Gilbert equation of magnetization motion and detail its physical meaning. After a brief description of spin-torque driven dynamics in the macrospin approximation, we discuss the validity of this approximation for various experimentally relevant geometries. Next, we perform a detailed comparison between accurate experimental data obtained from nanopillar devices and corresponding numerical modeling. We show that, on one hand, many qualitatively important features of the observed magnetization dynamics (e.g. non-linear frequency shift and frequency jumps with increasing current) can be satisfactory explained by sophisticated micromagnetic models, but on the other hand, understanding of these experiments is still far from being complete. We proceed with the numerical analysis of point-contact experiments, where an even more complicated magnetization dynamics is observed. Simulations reveal that such a rich behavior is due to the formation of several strongly non-linear oscillation modes. In the last part of the paper we emphasize the importance of sample characterization and conclude with some important remarks concerning the relation between micromagnetic modeling and real experiments. r 2007 Elsevier B.V. All rights reserved. PACS: 75.40.Gb; 75.40.Mg; 75.75.+a; 85.75.d Keywords: Spin injection; Spin-torque; Magnetization dynamics; Numerical simulations 1. Introduction An intuitive approach to micromagnetics might consider a ferromagnet as an assembly of localized dipoles governed by Heisenberg exchange interactions, dipole–dipole inter- actions as well as anisotropy, a characteristic of magnetic bodies tightly linked to properties of the orbital moment. The distribution of magnetization within tiny magnetic islands grown on favorable substrates in ultra-high vacuum (e.g. Ref. [1]) may be studied in such a way [2]. Classical micromagnetics [3–5] treats the magnetization in the continuum limit; assuming a constant saturation magneti- zation for the operating temperature, the magnetization distribution within a ferro- or ferri-magnetic body is thus becoming a vector field M(r)=M S m(r), where m(r) may only reside on the unit sphere |m(r)|=1. Dipole–dipole interactions are replaced by magnetostatics, a formal equivalent to electrostatics with the necessary condition that magnetic ‘‘charges’’ sum-up to zero, akin a simple dipole. Magnetic ‘‘charges’’ arise either from divergences of the magnetization vector field within the volume of the ferromagnetic body (rm(r)6¼0), or, assuming so-called ‘‘free boundary’’ conditions to apply, as soon as the magnetization vector fails to be parallel to a free surface or an interface ((m n)6¼0, where n is the outwards normal to the surface or interface). Heisenberg-type exchange inter- actions are taken in the continuous limit and assumed isotropic. Therefore, a single parameter describes the stiffness of the magnetization vector field to distortions whether bending, splay or twist. Anisotropy is described on a purely phenomenological basis. Lastly, starting from a magnetization distribution at equilibrium, the magnetization ARTICLE IN PRESS www.elsevier.com/locate/jmmm 0304-8853/$ - see front matter r 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.jmmm.2007.12.023 Corresponding author. Tel.: +49 3641 282 5357; fax: +49 3641 252 2530. E-mail address: [email protected] (D.V. Berkov).
Transcript
Page 1: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

0304-8853/$

doi:10.1016

�Correspfax: +4936

E-mail a

Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259

www.elsevier.com/locate/jmmm

Current Perspectives

Spin-torque driven magnetization dynamics: Micromagnetic modeling

D.V. Berkova,�, J. Miltatb

aInnovent Technology Development, Pruessingstr. 27B, D-07745 Jena, GermanybLaboratoire de Physique des Solides, Univ. Paris-Sud & CNRS, 91405 Orsay, France

Available online 31 December 2007

Abstract

In this paper, we present an overview of recent progress made in the understanding of the spin-torque induced magnetization dynamics

in nanodevices using mesoscopic micromagnetic simulations. We first specify how a spin-torque term may be added to the usual

Landau–Lifshitz–Gilbert equation of magnetization motion and detail its physical meaning. After a brief description of spin-torque

driven dynamics in the macrospin approximation, we discuss the validity of this approximation for various experimentally relevant

geometries. Next, we perform a detailed comparison between accurate experimental data obtained from nanopillar devices and

corresponding numerical modeling. We show that, on one hand, many qualitatively important features of the observed magnetization

dynamics (e.g. non-linear frequency shift and frequency jumps with increasing current) can be satisfactory explained by sophisticated

micromagnetic models, but on the other hand, understanding of these experiments is still far from being complete. We proceed with the

numerical analysis of point-contact experiments, where an even more complicated magnetization dynamics is observed. Simulations

reveal that such a rich behavior is due to the formation of several strongly non-linear oscillation modes. In the last part of the paper we

emphasize the importance of sample characterization and conclude with some important remarks concerning the relation between

micromagnetic modeling and real experiments.

r 2007 Elsevier B.V. All rights reserved.

PACS: 75.40.Gb; 75.40.Mg; 75.75.+a; 85.75.�d

Keywords: Spin injection; Spin-torque; Magnetization dynamics; Numerical simulations

1. Introduction

An intuitive approach to micromagnetics might considera ferromagnet as an assembly of localized dipoles governedby Heisenberg exchange interactions, dipole–dipole inter-actions as well as anisotropy, a characteristic of magneticbodies tightly linked to properties of the orbital moment.The distribution of magnetization within tiny magneticislands grown on favorable substrates in ultra-high vacuum(e.g. Ref. [1]) may be studied in such a way [2]. Classicalmicromagnetics [3–5] treats the magnetization in thecontinuum limit; assuming a constant saturation magneti-zation for the operating temperature, the magnetizationdistribution within a ferro- or ferri-magnetic body is thus

- see front matter r 2007 Elsevier B.V. All rights reserved.

/j.jmmm.2007.12.023

onding author. Tel.: +493641 282 5357;

41 252 2530.

ddress: [email protected] (D.V. Berkov).

becoming a vector field M(r)=MS �m(r), where m(r) mayonly reside on the unit sphere |m(r)|=1. Dipole–dipoleinteractions are replaced by magnetostatics, a formalequivalent to electrostatics with the necessary conditionthat magnetic ‘‘charges’’ sum-up to zero, akin a simpledipole. Magnetic ‘‘charges’’ arise either from divergences ofthe magnetization vector field within the volume of theferromagnetic body (rm(r) 6¼0), or, assuming so-called‘‘free boundary’’ conditions to apply, as soon as themagnetization vector fails to be parallel to a free surface oran interface ((m � n) 6¼0, where n is the outwards normal tothe surface or interface). Heisenberg-type exchange inter-actions are taken in the continuous limit and assumedisotropic. Therefore, a single parameter describes thestiffness of the magnetization vector field to distortionswhether bending, splay or twist. Anisotropy is described ona purely phenomenological basis. Lastly, starting from amagnetization distribution at equilibrium, the magnetization

Page 2: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1239

may be set into motion under the action of an externalfield, often called the applied field, or a spin-polarizedcurrent [6,7]. Both exert a torque on the magnetization.

An energy density is associated to each type ofinteraction within the magnetic body, spin-torque omitted.A magnetization distribution is thus the result of conflictingrequirements. Exchange interactions promote uniformmagnetization distributions. On the other hand, imposingthe condition (m � n)=0 along the boundaries of a magneticbody necessarily leads to non-uniform distributions.Consider for instance a flat cylindrical micron size platelet:imposing the condition (m � n)=0 states that the magneti-zation should remain in-plane and tangential to the rim ofthe platelet. Imposing the condition rm(r)=0 means that,away from the rim, the magnetization should remainorthogonal to any radius drawn from the cylinder axis tothe rim. When approaching the cylinder axis, however,such a circular magnetization distribution, called a vortex,leads to increasing exchange interactions, and, soon to adivergence of the exchange energy. Because of the cons-traint |m(r)|=1, the magnetization needs to pop-up out ofthe plane. Due to symmetry, the magnetization directionalong the core of the vortex may only be perpendicular tothe plane of the platelet, up or down. The transition fromin-plane to out-of-plane magnetization orientation takesplace over distances not exceeding a few nanometers (seeRef. [8] for a short introduction to vortices). More gene-rally, in the absence of any spin-polarized current, amagnetization distribution reaches equilibrium when theenergy reaches its minimal value. When not submitted to anapplied field or spin-polarized current, any magnetizationdistribution is energetically equivalent to the distributionobtained through the transformation m(r)-�m(r). Addi-tional degeneracy may arise from the geometrical symme-tries of micron or sub-micron size magnetic elements.

Numerical micromagnetics as a tool proves adequate fora fine description of magnetization distributions atremanence within finite magnetic bodies or their transfor-mation under the action of a steady field (see Ref. [9] forrecent reviews of numerical techniques). Just to quote asingle example, magnetization distributions observed(MFM and X-PEEM experiments) in thick, facetted, singlecrystalline Fe islands grown on a Mo(1 1 0) surface [10]although intriguing at first sight appear to closely fitmicromagnetic simulations taking due account of the shapefeatures of these (nearly) perfect crystals. Although notfully revealed by experiments, the internal wall structure isanticipated to still prove complex. In the following we doassume that fundamental material parameters such as thesaturation magnetization MS at the working temperature,anisotropy constants such as Kun (characteristic of auniaxial anisotropy), K1 and K2 (characteristic of a cubicanisotropy), etc., and the exchange constant A, are knownwith an accuracy sufficient not to impair simulation results.Such quantities may, in principle, be known independentlyvia, e.g. SQUID magnetometry, torque magnetometry andspin waves analysis, respectively.

The primary aim of this ‘‘Perspective’’ is the evaluationof the ability of micromagnetics to describe magnetizationdynamics under spin-torque excitation. As a prerequisite,however, it seems worthwhile examining whether numer-ical micromagnetics remains accurately predictive in thefield-driven dynamical regime. On top of material para-meters listed above, one now also needs, as shown below inSection 2, to have at hand a fair evaluation of thegyromagnetic ratio g, as well as the damping constant a.Rare are the experiments where all of the quantities thatenter the Landau–Lifshitz or Landau–Lifshitz–Gilbertequation of magnetization motion (see Section 2) areknown accurately. For instance, in their study of modes inthe vortex state, Buess et al. [11] claim a remarkableagreement between experimental data and micromagneticsimulations, with an experimental accuracy better than10%. However, the authors’ choice of a free electron valuefor g is most likely inappropriate. Arguably, though, theirchoice of a rather high saturation magnetization forPermalloy at room temperature may well compensate forthe likely E5% or more error in the gyromagnetic ratio.An interesting outcome of this study is that when themagnetization is ‘tipped’ away from its original orientationand then allowed to relax towards equilibrium in theabsence of any applied field, then one obtains the samevalue of the damping parameter, irrespective of the activemode (aE0.008 for the samples studied). Similarly,Novosad et al. [12] found an excellent agreement betweenthe vortex gyration frequency and numerical simulations,also in Permalloy disks (under gyration we mean herevortex motion along an essentially circular closed orbit).Here experimental data represent an average over a largenumber of similar magnetic elements. Simulations do relyon a rather depressed (very thin samples) saturationmagnetization measured independently and on a freeelectron value of the gyromagnetic ratio as well as a‘‘standard’’ value of the damping parameter in Permalloy(a ¼ 0.01). If eigenmode frequencies most often amount toa few GHz, vortex gyration frequencies depend, for a giventhickness, on the disk diameter and do not exceed a fewhundreds of MHz. Eigenmodes analysis relies on small-amplitude motion of the magnetization whereas vortexgyration implies larger amplitude motion. Work quotedabove altogether displays a very satisfactory agreementbetween measured and computed frequencies (see alsoRefs. [13,14] for recent studies of vortex motion and vortexcore reversal under field or spin-polarized current).This, however, is not always the case: in their study of

localized spin-wave modes in micron-wide magnetic wires,Park et al. [15] do reach good agreement betweenexperiments and simulations for fields applied in the planeof and normal to the wire edge in excess of E100Oe.Below that value, experimental frequencies prove signifi-cantly lower than predicted by simulations. Similarly, asystematic discrepancy between experiments and simula-tions has been found in the study of eigenmodes in Co rings[16]. In this particular case, however, computed frequencies

Page 3: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591240

prove lower than experimental values. Such facts probablyneed to be borne in mind when analyzing the outcome ofsimulations picturing spin-transfer torque driven magneti-zation dynamics.

Lastly, since sustained precession under the action of asteady spin-polarized current may imply large angleprecession, it would have been extremely useful to compareexperimental data dealing with ballistic switching (alsoimplying large angle precession) (e.g. Refs. [17,18]) anddetailed micromagnetic simulations. Rather unfortunatelyin a sense, magnetization dynamics in the single spin limitwith ad-hoc parameters proved sufficient for a very stun-ning description of phase coherent ballistic switching inGMR or TMR micron size elements. Nevertheless, italtogether appears that whereas micromagnetics provestightly predictive when describing the static behaviorof magnetic elements, it performs less satisfactorilywhen attempting a quantitative analysis of magnetizationdynamics.

Let us now move to the core of this paper. Becausesustained precession in CPP-GMR nanopillars and pointcontacts leads to precise observables, namely a well-definedcurrent- and field-dependent frequency, and an associatedpower spectral density, the present perspective is limited tospin-transfer torque induced magnetization dynamics inthese structures, leaving aside spin pressure effects on wallsin magnetic nanowires. It is organized as follows. Section 2recalls which are the minimal modifications to be broughtto the equation of magnetization motion in order to takeinto account spin-torque effects. In Section 3, we attemptto analyze reasons why a single spin approximation fails toprovide an acceptable picture of observed phenomena inthe structures considered. The following section illustrateshow different working hypotheses may lead to markedlydifferent micromagnetics simulation results when dealingwith nanopillar geometries. Peculiarities of the point-contact geometry are analyzed in Section 5. Finally,we discuss in Section 6 the crucial importance of sam-ple characterization for a meaningful numerical analysisof spin-torque experiments, before reaching concludingremarks.

2. Including spin-torque effects into magnetization dynamics

Almost all of micromagnetic simulations involvingmagnetization dynamics rely on the Landau–Lifshitz–Gil-bert equation of magnetization motion, namely (if using SIunits)

dMðr; tÞ

dt¼ � g0 Mðr; tÞ �Heff ðr; tÞ½ � þ a Mðr; tÞ �

dMðr; tÞ

dt

� �

Heff ðr; tÞ ¼ �1

m0

d�dM

, (1)

where M(r,t) and Heff(r,t) are the magnetization andeffective field, respectively. Both are functions of spaceand time. e denotes the energy density functional,

g0 ¼ m0 � ðgmB=_Þ (g0 ffi 2:211� 105 mA�1 s�1 for a freeelectron), mB is the Bohr magneton and a is the dampingparameter. The effective field is the sum of the applied,anisotropy and demagnetizing fields, supplemented by fieldcomponents arising from exchange interactions. Eq. (1)means that precession around the local effective field is thefundamental magnetization motion. Note, however, thatthe effective field moves together with the magnetization,and, thus, the simple idea of a precession around a fieldwith fixed direction may prove extremely misleading.Damping is required in order to align the magnetizationalong the acting field: the Gilbert form used in Eq. (1) isconsistent with Rayleigh-type dissipation. As noticednumerous times before, Eq. (1) is strictly equal to:

dMðr; tÞ

dt¼ �

g01þ a2

Mðr; tÞ �Heff ðr; tÞ

þa

MSMðr; tÞ � ½Mðr; tÞ �Heff ðr; tÞ�

�, (2)

thus recovering the initial damping formulation of theLandau–Lifshitz equation at the expense of a (minor sincea� 1, usually) renormalization of both the gyromagneticratio and the damping parameter. Eq. (2) may also beenwritten as:

dMðr; tÞ

dt¼ �

g01þ a2

½Mðr; tÞ � ðHeff ðr; tÞ þHdampðr; tÞÞ�

Heff ðr; tÞ ¼ �1

m0

d�dM

; Hdampðr; tÞ ¼a

MS½Mðr; tÞ �Heff ðr; tÞ�

(3)

As noticed by Smith [19], Eq. (3) treats on an equalfooting two field terms: the effective field that isconservative (it derives from an energy density functional)and a field that, per definition, is non-conservative (energyis transferred to an external bath).In the presence of an electric current, an additional

torque may act on the magnetization within a thinferromagnetic layer, arising primarily from the transmis-sion and reflection of incoming electrons with moments atarbitrary angles to the magnetization [6,20–22]. As a netresult, reflected and transmitted spin currents have virtuallyno component transverse to the magnetization. In otherwords, this pure ballistic effect leads to a close to completeabsorption of the transverse spin current, itself the sourceof the spin-transfer torque (see Ref. [23] for a review ofconcepts). In a CPP-GMR stack, the electrons acquire spinpolarization either because they first cross the pinned (orhard) layer of the stack, or because they get reflected fromthe latter. Let p and m ¼M/MS be the unit vectors alongthe magnetization of the pinned and soft layers of thestack, respectively. For reasons outlined above, the spin-transfer torque is proportional to the sine of the anglebetween p and m, or in vector notation to [m� [m� p]].The spin-transfer torque is also proportional to thequantum of angular momentum carried by one electrontimes the density of carriers per unit time weighted by the

Page 4: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1241

electron polarization P, i.e. / Pð_=2ÞðJ=jejÞ½m� ½m� p��,where J is the current density and e the electroncharge (eo0).

In CPP-GMR stacks, the electron polarization at theinterface between the normal metal spacer and theferromagnetic free layer is affected by the spin-dependenttransport characteristics of the whole stack, due to (i) spinrelaxation within the bulk of the layers or at interfaces, and(ii) spin accumulation [24]. Slonczewski proposed in 2002an elegant approach to this problem where spin-flipscattering is not allowed within the spacer layer [25]. This‘‘full acceptance’’ model may be developed into a simplecircuit theory that leads to immediately usable expressionsfor both the CPP-GMR and spin-transfer torque in theparticular case of ‘‘symmetrical’’ spin valves, i.e. stackswhere the two FM layers are made of the same materialwith equal thicknesses, and leads are also made of a uniquematerial and have equal lengths. For these reasons,Slonczewski’s vintage’ 2002 model soon became popular,although it needs to be realized that fulfilling the conditionof two identical ferromagnetic layers with an identicalenvironment does not easily fit with the necessity to pin orharden the magnetization of one of the ferromagneticlayers in order for a CPP-GMR device to be functional.According to this Slonczewski’s model, the spin-transfertorque may be written as1:

dM

dt¼ � g0

_

2

1

m0M2S

1

d

J

e

!

�PSloncz=2

cos2ðW=2Þ þ ð1=1þ waÞsin2ðW=2Þ

� M� ½M� p�½ �,

(4)

(still assuming eo0), whereas the GMR response turns outto be proportional to

1� cos2ðW=2Þ1þ wacos2ðW=2Þ

. (5)

In Eqs. (4) and (5), W is the angle between m and p

(cos W ¼ ðm � pÞ); the asymmetry parameter wa and thepolarization PSloncz do both depend on the stack-circuitcharacteristics and may be expressed as combinations ofspin-dependent bulk and interface resistances. The currentis negative when electrons first cross the ‘‘pinned’’, then the‘‘free’’ layer, and positive otherwise. Finally, it ought to beunderstood that the spin-transfer torque has been dis-tributed over the thickness d of the ‘‘free’’ or ‘‘soft’’ layer inEq. (4). Eqs. (4) and (5) do capture essential ingredients of

1For cgs units addicts, Eq. (4) becomes:

dM

dt¼ �gcgs

_

2

1

4pM2S

1

d

Je

e

!�

Peff

cos2ðW=2Þ þ ð1=1þ waÞsin2ðW=2Þ

� M� ½M� p�½ �,

where gcgs ¼ gmcgsB =_cgs ¼ ð103=4pÞg0 (gcgs ffi 1:76� 107ðOe sÞ�1 for a

free electron) and all quantities are expressed in cgs units.

spin-polarized transport in CPP-GMR stacks, namely anasymmetry of the GMR response as well as an asymmetryof the spin-torque. More detailed calculations [26] did notuncover significant deviations from Slonczewski’s modelfor materials and thicknesses entering typical CPP-GMRnanopillars.Most simulations relying on Slonczewski’s or Xiao/

Zangwill/Stiles’ analysis of spin-torque effects in magneticnanostructures forget about the meaning of the polariza-tion factor and replace ð1=2ÞPSloncz by a simple polariza-tion-like adjustable parameter Peff . Several additionalremarks ought to be made:

(i)

cumbersome physical constants may be avoided via

variable reduction. In SI units, defining reducedvariables as w ¼ ð_=2Þð1=m0M

2SÞð1=dÞðJ=eÞPeff , t ¼ g0

MSt, m ¼M=MS, h ¼ H=MS, leads to the simplemagnetization dynamics equation including spin-torque:

dm

dt¼ �½m� heff � � wgðm � pÞ½m� ½m� p�� þ a m�

dm

dt

� �(6)

heff ¼ �1

m0M2S

d�dm

;

gðm � pÞ ¼ cos2ðW=2Þ þ1

1þ wasin2ðW=2Þ

� ��1.

Note that both w and a are ‘‘small’’ parameters andthat g(m � p) does not depart strongly from 1 forcommon asymmetry parameter values, implying thatmagnetization trajectories will to a good approxima-tion be determined by the energy landscape.

(ii)

As alluded to earlier, the validity of the Gilbertdamping term in the presence of spin-torque is beingdebated [19,27]. Because this perspective is primarilyconcerned with sustained precession in the nanopillarand point-contact geometries, the following simpleargument shows that a Gilbert (embedding the wholeof dm/dt) or Landau–Lifshitz formulation restrictedto the conservative component of the field, heff, maynot lead to significant differences. Starting with Eq. (6)and taking the cross product of both parts with dm/dtyields:

heff �dm

dtþ wgðm � pÞðm� pÞ �

dm

dt� a

dm

dt

� �2¼ 0. (7a)

Expression (7a) means that the work of the effectivefield augmented with the work of the spin-torqueequivalent field, hST ¼ w � gðm � pÞðm� pÞ, is balancedby dissipation, at any arbitrary time and at anylocation within the ferromagnetic body. This is theGilbert picture. Expression (7a) may also, according to

Page 5: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591242

Eq. (2), be written as

fheff þ hST þ a½m� ðheff þ hSTÞ�g �dm

dt¼ 0 (7b)

stating that the work of the total field acting on themagnetization, htot ¼ heff+hST, is again equilibratedby the work of the total equivalent damping field.Relation (7b) is more in tune with the Landau–Lifshitzapproach, which defines the required additionaldamping field in order to relax the magnetizationtowards equilibrium without altering the magnitude ofthe magnetization. In (7b), however, the damping fieldstill arises from both the effective field and the spin-torque equivalent field.

Since the effective field is conservative, for a closedorbit G and in the absence of any time-dependentapplied field, relations (7a) and (7b) reduce to

IG

wgðm � pÞ½m� p� �dm

dt� a

dm

dt

� �2" #

¼ 0, (8a)

IG

wgðm � pÞ½m� p� �dm

dtþ a½m� ðheff þ hSTÞ� �

dm

dt

� �¼ 0.

(8b)

The first expression states that, on a closed orbit, thework of the spin-torque equivalent field is exactlycanceled by a definite-positive damping integral and,thus, allows for the existence of precessional statesunder the sole action of a steady current (i.e. in theabsence of any time-dependent field or current). Thesecond establishes a relation between the work ofthe spin-torque equivalent field and the work of thedamping field without lending itself to a clear physicalinterpretation as noted earlier [19], especially ifdamping is truncated to the sole effective field.Because, however, proceeding from the very samepremises, relations (8a) and (8b) cannot bear differentmeanings.

Incidentally, the sum of the exchange, magnetostaticand anisotropy energy may rise during part of theorbiting motion. If it does so, part of the energy isgiven back to the system at a later stage along the orbitin such a way that when integrated over one cycle, Eq.(8a) or (8b) remains satisfied. Similar situations arecommon, for example, in the changes in domain wallstructures during wall motion under the action of apulsed field. The energy of the domain wall increasesas the wall distorts in response to the changed field.This increase is usually obscured because the micro-magnetic energy as a whole decreases with time,primarily due to the decrease of the applied field (orZeeman) energy during the wall motion. At the sametime, wall distortion usually leads to an increase of the

other energy components. At the end of the pulse, thewall reverses to a more stable configuration andreleases the stored energy, which causes additionalwall motion, called overshoot. When an applied fieldcauses domain wall distortion and propagation, thatfield provides the initial torque; when a spin-polarizedcurrent causes distortion, oscillation and/or propaga-tion, the magnetization distribution is set into motionthanks to the initial torque provided by the current.

Now, precessional states become stable in nanopil-lars for values of w/a close to unity. Therefore, relaxingthe spin-torque contribution to damping is equivalentto the neglect of a term of order �a2 in the dampingprocess. Numerical simulations confirm that, whendealing with precessional states in either of thegeometries considered here, the Gilbert form or theLandau–Lifshitz approach with the conservative fieldas a sole source of damping lead to virtuallyundistinguishable results.

(iii)

The damping process may be affected by ‘‘spin-pumping’’ [28,29], i.e. a transfer of angular momentumvia the inelastic scattering of electrons with energiesclose to the Fermi energy flowing from the ferro-magnet into the normal metal spacer or lead.Practically, this could be taken into account via aspecific surface damping mechanism within numericalmicromagnetics. To the knowledge of the authors,surface damping effects have not yet been includedinto micromagnetic simulations.

(iv)

Temperature is most commonly introduced via astochastic thermal field leading to what is commonlytermed Langevin magnetization dynamics, followingthe pioneering work of Brown [30], where this methodis introduced for a single magnetic moment. Thejustification of this approach within numerical micro-magnetics, where several interactions between elemen-tary moments (discretization cells) do exist, isdiscussed in detail in Ref. [31]. We note also that insystems submitted to spin-transfer torques, an addi-tional source of noise exists, namely spin currentfluctuations.

(v)

The spin-torque equivalent field may be supplementedwith an additional term that appears from firstprinciple calculations [22] and is linked to theimaginary part of the mixing conductance in circuittheory (see Ref. [32] and Refs. therein). This additionalfield may, for obvious geometrical reasons, only beparallel to p. The total spin-torque equivalent fieldthen reads hST ¼ w½gðm � pÞ½m� p� þ bp�, b� 1. Theso-called ‘‘non-adiabatic’’ spin-transfer term intro-duced in order to explain wall motion in nanowiresat low current densities [33,34] is phenomenologicallysimilar. The inclusion of such an additional term willnot be considered below (simulations performed byone of the authors in the past have shown that for theproblems of interest here, addition of a haddST ¼ wbpdoesnot significantly impact simulation results).
Page 6: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1243

3. Validity range of the macrospin approximation

3.1. Spin-torque driven dynamics in the macrospin

approximation

Consider an elliptical soft magnetic element deprived ofany growth-induced anisotropy. Shape anisotropy tells usthat the long axis of the ellipse is the easy magnetizationaxis and that moving the magnetization out of the planewill prove more costly than moving the magnetizationaway from the easy axis in the plane. The energy functionalof such an element may be reduced to:

� ¼ Kð1�m2xÞ þ

1

2m0M

2Sm2

z � m0M �Happ, (9)

where x coincides with the easy axis and z is normal to theplane of the element. The first term in Eq. (9) is the uniaxialanisotropy (K40; K � m0M

2S); the second term describes

demagnetizing effects in the thin film limit, the third is theZeeman energy. If the external field is applied alongthe easy axis with a sole component Happ

x , then theeffective field for this element treated in the single spinlimit reads:

Heff ¼ �1

m0MS

d�dm¼ ½Happ

x þHK mx; 0;�MSmz�, (10)

where HK ¼ 2K=m0MS is the anisotropy field. Let usfurther assume the spin-torque to be symmetrical (wa ¼ 0,i.e. gðm � pÞ ¼ 1) and the electron polarization to be alignedwith the ellipse long axis (p ¼ x). Solving Eq. (1) or Eq. (2)augmented with Eq. (3), that is with spin-torque added,soon leads to a phase diagram such as shown in Fig. 1(T ¼ 0).

As long as Happx remains smaller than the anisotropy

field, there exists a transition between the parallel ‘‘P’’ state(both the ‘‘pinned’’ and ‘‘soft’’ layers are magnetized alongthe þx direction) and the antiparallel ‘‘AP’’ state(‘‘pinned’’ magnetized along þx, ‘‘soft’’ layer along �x)for a positive current (the convention here is that thecurrent is positive if flowing from the ‘‘pinned’’ to the‘‘soft’’ or ‘‘free’’ layer; stated otherwise, the current isnegative if electrons flow from the ‘‘pinned’’ to the ‘‘free’’layer). Conversely, the transition occurs between the ‘‘AP’’and ‘‘P’’ states for negative currents. However, thetransition is not direct, except for pathological pointsw=a ¼ �1=2; Happ

x ¼ þHK and w=a ¼ 1=2; Happx ¼ �HK ;

sustained precession states are expected from the simpletheory with ‘‘clamshell’’-type orbits for current densitiesw1owow2, and out-of-plane orbits for w4w2.

Our main point here is to show that an actually terriblyoversimplified macrospin approach does lead to a phasediagram for an applied field along the easy anisotropy axisthat has much in common with the experimental phasediagram to be found in, e.g. Ref. [35] (except for the fieldoffset observed in the experimental phase diagram due tothe magnetostatic coupling between the ‘‘soft’’ and ‘‘hard’’layers). Three additional remarks here: (i) rounding-up of

the phase diagram boundaries is expected from finitetemperature effects as well as regions of states overlap [36],(ii) precessional states have been observed not only forjHapp

x j4HK , but also in the region Happx

�� ��oHK [37,38] and(iii) the phase diagram in Fig. 1 proves only mildly robustin the presence of a transverse in-plane field componenteven at zero temperature [39].As recalled above, the single spin approximation is able

to capture some fundamental aspects of spin-torqueinduced (STI) magnetization dynamics. In some instanceseven analytical calculations may be pursued far enough soas to define thresholds for the onset of precessional statesand switching. However, recent experiments [40] havedemonstrated that switching in these systems provescomplex, a fact actually predicted by all full micromagneticsimulations previously performed. It ought to be men-tioned, however, that critical currents are particularly highin Ref. [40], resulting in an exalted influence of the Oerstedfield, i.e. the field created by the current flowing across thepillar. In order to understand why the macrospinapproximation may fail, although the size of manysystems where the STI magnetization dynamics has beenobserved is very small, we first recall several importantissues concerning the determination of critical single-domain sizes.

3.2. Critical sizes for the validity of the macrospin

approximation

First we remind the reader that the concept of a ‘strictly’single-domain magnetic particle (body) does indeed exist.We mean here a particle that remains homogeneouslymagnetized independently of external conditions, like thevalue and the direction of the applied field (of course,the temperature is still assumed to be well below Tc). Thequalitative estimation of the critical size for such a ‘strictly’single-domain behavior relies on the comparison between(i) closed-flux magnetization configurations governed by asole exchange energy Eexch and (ii) the collinear statecharacterized by Eexch ¼ 0 and Edem40, where Edem is themagnetostatic energy (see Chapter 3.3 in Ref. [4] andRefs. therein). The exchange energy density of the closedM(r)-configuration obviously increases with decreasingparticle size (magnetization gradients are getting larger),leading to the result that only a collinear magnetizationstate is energetically stable below the critical size lcrwhich

scales with the exchange length lexch ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2A=m0M

2S

q. We

note in passing that the anisotropy energy is not includedinto these considerations; in general, for a single crystalferromagnet, this anisotropy is expected to stabilize thehomogeneous magnetization state, thus increasing lcr. Anexact value of lcr can only be determined by rigorousnumerical simulations. It depends on many physicalfactors, and amongst them, primarily the particle shape.It also depends on the non-collinear state used to computeEexch, and, in most cases, amounts to lcr ð4� 8Þ � lexch [4].

Page 7: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

3

x1

x1

xy

z

x2x1

In-planeprec. states

In-planeprec. states

Out-of-planeprec. states

Out-of-planeprec. states

x2num

x2num

x2anal

x2anal

Hx

HK

2

1

-1

-2

-3-1.5 -1 -0.5 0 0.5 1 1.5

x/a

0

P

P/AP

AP

Fig. 1. Simulated phase diagram in the single spin limit at zero temperature as a function of reduced co-ordinates w/a and Hx/HK. The current is negative

when electrons first cross the ‘‘Pinned’’ and, then, the ‘‘Free’’ layer. The phase diagram separates regions where either the ‘‘P’’ or the ‘‘AP’’ state are stable

from regions where, either bi-stability exists ‘‘P/AP’’, or sustained precession under the sole action of a steady current. The single spin limit establishes a

clear difference between ‘‘in-plane’’ or ‘‘clamshell’’ orbits and ‘‘out-of-plane’’ orbits to be found at larger current densities for fields in excess of the

anisotropy field (upper-right and lower-left regions of the phase diagram delimited by the horizontal segment Hx/HK ¼ 1 or �1 and the computed phase

boundary wnum2 ). Blue and red lines are computed analytically (see Ref. [23]) whereas open symbols are simulation results.

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591244

The exchange length is about lexchE5 nm or less for mostferromagnets, so that an upper bound for the single-domainthreshold is lcrE40 nm. This would mean that for nearly allspin-torque experiments (leaving aside data obtained onpoint contacts to be discussed below) the lateral elementsize is far above lcr, invalidating the macrospin approxima-tion and raising the question why this approximation cangive any reasonable predictions at all. In order to answerthis question in particular and to make further methodo-logical progress in general, several factors need to be takeninto account:

(i)

The lcr-estimate given above is solely based on theenergy comparison between different configurations.Thus, it cannot be used to predict whether thetransition from a single- to a multi-domain state willreally occur during a remagnetization process, becausethis transition often requires overcoming an energybarrier. This means, that an approximately homo-genous (single-domain) magnetization state, being forsome specific external conditions only metastable, canstill exist, because the transition to, e.g. some closedmagnetization configuration with a smaller energyrequires overcoming a prohibitively large energybarrier. Indeed, simulations have shown that forcertain particle shapes almost collinear magnetizationstates persist during the whole remagnetization processfor nanoelements with lateral sizes as large as severalhundred nanometers in a homogeneous external field.

(ii)

Most calculations leading to the estimation lcr ð4�8Þ � lexch given above were performed for particles with

sizes of the same order of magnitude in all threedimensions (cubes, spheres, etc.). For a thin filmelement with thickness much smaller than its lateralsizes, both exchange and stray-field energies mighthave a different size dependency, which, in turn, mightsubstantially affect lcr.

(iii)

Looking at the spin-torque distribution in quasi-singledomain elements proves instructive. For, e.g. a squarenanoelement in a ‘‘flower’’ remanent state and for aspin polarization collinear with its mean magnetizationdirection, the spin-torque (4) has opposite directionsnear adjacent corners of the square as displayed inFig. 2a. In other words, the initial spin-torquedistribution proves highly non-homogeneous [41],impairing easy magnetization reversal. Such a situa-tion is however not unique to spin-torque action on anon uniform magnetization distribution: in a thin filmelement, reversal under the action of a homogeneousfield antiparallel to the mean magnetization directionwould result in an equivalent initial torque distributionsince the precession of the magnetization around theapplied field gives rise to a local demagnetizing fieldHdem, with, as a result, magnetization precessionaround the latter. The cross-product [M� p] actuallyplays the role of Hdem.

The line of arguments presented above means that each

specific experimental situation requires a separate analysisusing full-scale micromagnetic simulations to find outwhether the macrospin approximation is valid for itsdescription. An example of such an analysis can be found

Page 8: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

J/e> 0Ha

M(r)

Hdem

∝−[M×Hdem]dMdt

M(r)

p

[M×p]

NMST∝−[M×[M×p]]

Fig. 2. (a) Torque distribution in a nearly uniformly magnetized particle

(‘‘flower’’ state) for a current promoting parallel alignment (J, eo0 in Eq.

(4)): the torque is strongly inhomogeneous even for a slightly non-collinear

magnetization configuration; (b) torque distribution under the action of a

homogeneous applied field.

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1245

in Ref. [42], where STI precession in a square element withthickness h ¼ 2.5 nm and typical magnetic parameters2

(MSE1000G, A ¼ 2� 10�6 erg/cm) was studied. It hasbeen shown than the magnetization configuration duringSTI precession significantly deviates from single-domainbehavior already for lateral sizes as small as L ¼ 30–40 nm.This value is significantly smaller than the lateral sizes ofnanoelements used in all experiments reported so far innanopillars.

Concluding this discussion, we point out that there existsa class of experimental systems where the macrospinapproximation is invalid for any sizes of the current floodedarea, namely the so-called point-contact setup. In suchexperiments [43–46], a current-carrying wire with a verysmall diameter (20–80 nm) is attached to a system ofextended magnetic layers, usually with lateral extension�10 mm. As mentioned above, the size of the area floodedby a current through the ‘free’ magnetic layer can be assmall as 25 nm [43], making it very tempting to declare thatthe macrospin approximation is applicable to analyze theseexperiments even quantitatively.

However, in contrast to the multilayer nanopillars,where the whole area of magnetic layers is flooded by theelectric current, in the nanocontact setup only themagnetization within the area under the contact ‘feels’ aspin-polarized current and can be directly excited by it. Onthe other hand, there exists a strong exchange interactionbetween this area and the outer film region (which results,in particular, in the necessity for higher current densities inorder to excite steady-state magnetization precession whencompared to the nanopillar geometry). This exchangeinteraction is qualitatively important: first, it should beincluded in order to accurately determine the equilibrium

2Most material parameters are expressed in cgs units because these units

are still used by most experimentalists. Use the following transformations

if necessary: 1000G is equivalent to 106A/m; in other words, 1 kA/m (SI)

is equivalent to 1G (cgs). 1Oe is equivalent to (1000/4p)A/m. A field

equal to 1mT means a 10Oe field. Exchange constant of 10�6 erg/cm is

equivalent to 10�11 J/m. Lastly, the energy density (anisotropy constant)

of 1 erg/cm3 is equivalent to 0.1 J/m3.

magnetization configuration of the system; it may be quitecomplicated due to a large Oersted field arising for thehigh-density currents through the contact. Second, ex-change is responsible for the formation of spin waves thatcarry energy away from the point-contact area, so that theinclusion of this interaction is crucial even for a qualitativeunderstanding of the magnetization dynamics [20,47,48].Unfortunately, within the macrospin formalism there is noadequate method to incorporate exchange. Hence, themacrospin model is, strictly speaking, invalid in thissituation for any size of the point contact.

3.3. Misleading artefacts of the macrospin approximation

As briefly recalled above, the macrospin dynamics is veryrich, so that the macrospin model can successfully reproducemany features of the actual magnetization dynamics of realsystems. The reverse of this coin is the danger that due tothis richness the macrospin approximation can accidentallyreproduce some real dynamics features suggesting acompletely misleading explanation of them.An excellent example of such an artefact can be found in

the famous paper of Kiselev et al. [49]. Trying to explainmany interesting and highly non-trivial features of theexperimentally observed STI-dynamics in Co/Cu/Co na-nopillars, Kiselev et al. performed macrospin simulations.Using several adjustable parameters (like the saturationmagnetization and an homogeneous uniaxial anisotropy),they succeeded in reproducing not only the value of themagnetization oscillation frequency and the correct slopeof its current dependence, but also the frequency jump withincreasing current! Within the framework of macrospinsimulations this jump was elegantly interpreted as thetransition between the small-amplitude elliptical trajectoryand the large-amplitude ‘clamshell’ trajectory (see Fig. 2 inRef. [49]), so that a nearly perfect quantitative agreementbetween experiment and macrospin simulations wasachieved.Unfortunately, full-scale micromagnetic simulations

performed later [50] have revealed that this jump is anartifact of the macrospin approach. For the ellipticalnanoparticle with the geometry given in Ref. [49] andmagnetic parameters typical for Co, transition fromregular magnetization oscillations (where the nanoelementmagnetization remains at least approximately collinear) toa ‘‘chaotic’’ magnetization dynamics (which is a signatureof a highly irregular magnetization configuration duringoscillations) occurs before the transition from ‘small-angle’to ‘large-angle’ oscillations, so that the correspondingfrequency jump cannot be explained correctly by themacrospin model. We shall return to the analysis of thisvery challenging paper later on.

4. Steady-state precession in the nanopillar geometry

In this section we focus our attention on micromagneticsimulations of a steady-state precession in nanopillars,

Page 9: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591246

i.e. persisting magnetization oscillations when a dc-currentflows through a multilayer stack of thin magnetic nanoele-ments separated by non-magnetic spacer(s). We actuallyconcentrate mainly on two such experiments [49,51] forseveral reasons. First, both papers contain highly non-trivialresults (and for this reason are probably the most frequentlycited on this topic). Second, both studies include a reason-able [49] or very detailed [51] sample characterization, sothat attempting a quantitative comparison with simulationdata makes sense. Third, systematic numerical simulationshave been performed in order to understand dynamicsobserved in [49] (see, e.g. Refs. [50,52–54]) and very recentlyRefs. [39,55], also to analyze results from Ref. [51].

Before we present detailed analysis of these twoexperiments, let us point out that, apart from the use ofdifferent magnetic materials, the experimental setups usedin Refs. [49,51] differ in two fundamental aspects. In Ref.[49], both the applied field and the conduction electronpolarization are nominally collinear and pointing approxi-mately along the long axis of the elliptical soft element. Weshall refer here to a ‘‘longitudinal’’ geometry. In Ref. [51],the angle between the field and the electron polarization isclose to 901 and neither the field nor the electronpolarization direction coincide with the major ellipse axis.We shall call this setup a ‘‘skewed’’ geometry. Asdemonstrated below, the external field direction as well asthe electron polarization direction are qualitatively im-portant for the STI-dynamics, so that we analyze these twoexperimental situations separately.

4.1. Longitudinal geometry: STI-dynamics features which

can be explained by simulations

We start with the analysis of the famous paper of theCornell group [49], which was historically the first oneallowing for a quantitative analysis of experimental resultson STI-dynamics. In this work, the authors used a Co/Cu/Co nanopillar structure with an extended bottom thick Colayer (40 nm) and elliptical thin Co (3 nm) nanoelementwith nominal lateral size 130� 70 nm2 on top of it.Classical magnetostatic coupling between the ‘‘hard’’ and‘‘soft’’ layers was thus presumably moderate if not absent.Furthermore, since the non-magnetic Cu spacer was 10 nmthick, any indirect exchange interaction between the twoCo layers could be excluded. When the electron flow fromthe dc-current was directed from the thin to the thick Colayer, Kiselev at al. could detect microwave-frequencyoscillations of the sample resistance in a large interval ofcurrents and external magnetic fields. As mentioned above,in the Kiselev experimental geometry the magnetizationdirection of the ‘fixed’ layer and the applied field areparallel and nearly aligned with the long axis of theelliptical ‘soft’ (‘free’) element. The observed dynamics hadthe following main features:

(1)

Small-amplitude signal at low currents with a (high)frequency virtually independent from the current.

(2)

Huge frequency drop (from E17 to E6.5GHz) whenthe current was increased, accompanied by a dramaticgrowth of the signal amplitude (‘large-amplitude’signal) at first, followed by a vanishing-out of the rfpower for the largest currents (E6mA).

(3)

Continuous frequency decrease with increasing currentin the large-amplitude regime.

(4)

The large-amplitude signal showed several equidistantspectral bands, where upper bands were obviouslyhigher harmonics of the basic frequency f0 and asignificant power contribution in the regime of very low(compared to f0!) frequencies (0–1GHz).

(5)

Broad spectral lines (�1GHz), especially in the large-amplitude regime.

A quantitative analysis of the system magnetizationdynamics using micromagnetic simulations proved farfrom straightforward, due first to the large sample-to-sample variations of experimental data and, second, torather controversial specifications of the Co magneticparameters given in [49]. However, many importantfeatures listed above could be successfully reproduced atleast qualitatively [50], namely:

(i)

Small-amplitude signal with the current-independent

frequency: According to simulations, the ‘small-ampli-tude’ signal with a nearly constant (current-indepen-dent) frequency corresponds to the small-amplitude(linear) magnetization oscillations excited for currentsonly slightly different from the threshold current, jcr,for the steady-sate precession onset at zero tempera-ture. When the current is not much higher than jcr andincreases, then the oscillation amplitude increases also,but remains still small enough to ensure a quasi-independence of the oscillation frequency from oscilla-tion amplitude, as expected from standard classicalmechanics. An important point here is that as long asthermal fluctuations were neglected in simulations, theoscillation amplitude was found to grow very rapidlywith the current j4jcr (so called ‘stiff’ oscillationgeneration), so that the current interval where theamplitude remained small enough to keep the fre-quency almost constant, was negligibly small. Onlyinclusion of thermal fluctuations allowed for a gradual(relatively slow) increase of the oscillation amplitude,so that the ‘small-amplitude’ regime could be satisfac-tory explained [50].

(ii)

Downwards frequency drop with increasing current: Thisjump was interpreted by Kiselev et al. within theframework of macrospin simulations as the transitionbetween the small-angle elliptical precession and large-angle ‘clamshell’ orbit motion of the macrospin.Numerical simulations [50] revealed that the coherenceof the magnetization configuration is completely lostbefore the transition to the ‘clamshell’ orbit, so that themacrospin picture is invalid for such large currents.These simulations have shown that the transition to
Page 10: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1247

fully incoherent magnetization oscillations (differentparts of a nanoelement oscillate with differentfrequencies and amplitudes, leading to the so calledquasichaotic regime) is accompanied by the abruptdecrease of the oscillation frequency. This abruptdecrease could be interpreted as a frequency jump inthe experiment. However, the amplitude of the jumpfound in simulations (DfexpE3GHz) proved by far notas large as in the experiment (DfexpE10GHz).

(iii)

Frequency decrease with increasing current in the ‘large-

amplitude’ regime: Increasing the current strengthnormally leads to an increase of the oscillationamplitude, simply because more energy is ‘pumped’into the system. In principle, the reasons why thefrequency may strongly depend on the oscillationamplitude in the case of non-linear oscillations—andoscillations in the large-amplitude regime are ob-viously non-linear-is well known in mechanical sys-tems. A detailed explanation of this phenomenon formagnetization dynamics in the macrospin approxima-tion can be found, e.g. in Ref. [47]. Roughly speaking,the oscillation frequency f is proportional to theproduct f /

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiHðH þNdMeqÞ

p, where Meq is the

magnetization projection on the equilibrium directionof the magnetization, around which precession takesplace, and Nd is the demagnetizing factor. Foroscillations around the in-plane orientation of themagnetization (which was the case in experimentsunder discussion, due to the in-plane anisotropy of athin ellipsoidal plate and the in-plane external field),this factor is positive, and the average projection of themagnetization on its equilibrium (in-plane) directionobviously decreases with increasing amplitude.Hence, the second factor under the square rootffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

HðH þNdMeqÞp

decreases, pushing the frequencydown. In the quasichaotic regime, when the macrospinapproximation cannot be applied any longer,one can still follow the same line of argu-ments, saying that qualitatively the relation f /ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

HðH þNdMeqÞp

remains valid, and the averagemagnetization drops with increasing current simplybecause in the quasichaotic regime the magnetizationbecomes more inhomogeneous when the currentstrength grows, thus leading to the same effect (fdecreases with I). However, one should keep in mind,that for oscillations with the characteristic wavelengthsas small as in the experiments under consideration(�10–20 nm) the contribution of the exchange inter-action to the dynamical system behavior is verysignificant, so that a theory of quasichaotic dynamicswhich includes also the exchange stiffness of thesystem, is actually required for a thorough under-standing of its behavior.

(iv)

Several equidistant spectral bands: The presence ofsignificant spectral peaks with frequencies at 2f0, 3f0,etc., where f0 denotes the ‘basic’ oscillation frequency,is usually considered in classical mechanics as an

evidence that the oscillations are strongly non-linear,so that the time dependence of the oscillating co-ordinate is no longer simply sinusoidal and hence itsFourier expansion contains substantial contributionsat frequencies corresponding to higher harmonics.When we consider magnetization dynamics, thisargument remains valid; however, an additionalcomplication arises due to the fact that magnetizationoscillations actually arise from the precession of themagnetization vector. This precessional character ofthe magnetization oscillation leads to the followingpicture.

When the magnetization vector M with constantmagnitude M0 oscillates around, say, the x-axis withfrequency f0, then its y- and z-projections oscillatewith the same frequency: My ¼M0 sin y0 cosð2pf 0tÞ; Mz ¼M0 sin y0 sinð2pf 0tÞ, where y0 is theprecession angle, i.e. angle between M and the x-axis.In the simplest case, assuming circular precession, y0 isconstant (time-independent) and so is the x-projectionMx ¼M0 cos y0. However, in the experimental situa-tions under consideration, the shape anisotropy of themagnetic nanoelement gives rise to an elliptical orbitand Mx is also time-dependent: Mx ¼M0 cos y0ðtÞ.Now, because the magnetization magnitude should bekept constant, the elliptical precession of m ¼M/MS

takes place on the unit sphere and the angle y0 reachesits minimal (and its maximal) value twice during oneoscillation period. This means that the x-projection ofthe magnetization oscillates with the frequency 2f0even when the precession angle is small, i.e. well in thelinear regime.

To proceed further, we note that magnetizationoscillations are detected experimentally using somekind of a MR-effect (here the GMR), which isproportional to the scalar product of magnetizationsof the ‘fixed’ and ‘free’ layers: DR(t)pmfree(t) �mfixed.For the in-plane external field (the case studied in Ref.[49]), the fixed layer moment lies in the nanoelementplane, which we denote as the xy-plane. Then the MRtime dependence is given by DRðtÞ / m

ðxÞfreeðtÞ �m

ðxÞfixedþ

mðyÞfreeðtÞ �m

ðyÞfixed, thus containing both the basic fre-

quency f0 coming from my-oscillations, and the nextharmonic 2f0 due to mx-oscillations. This logic is validfor arbitrary small oscillation amplitude, so that in thegeometry with the in-plane orientation of the externalfield and the fixed layer magnetization the secondharmonics should be present in the linear regime also.An important exception is the case when equilibriumorientations of the ‘free’ and fixed layers coincide,which may be the case when either the external field isvery strong or the experiment geometry has a specialsymmetry—like an elliptical (in-plane) multilayer stackwith the external field oriented exactly along the majorellipse axis. In this case one should not detect the signalwith the basic frequency f0. To understand why this isso, it is enough to choose the x-axis along the

Page 11: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591248

equilibrium orientations of the magnetizations of bothlayers. Then m

ðyÞfixed ¼ 0 and we are left with the signal

due to oscillations of mðxÞfreeðtÞ only, which have the

frequency 2f0. The basic frequency is then notobserved at all!

(v)

Large spectral linewidth in the large-amplitude regime:Such broad spectral lines, observed for magnetizationoscillations of so small elements, are normallyexplained as a thermal fluctuation effect. However,micromagnetic simulations have shown, that evenwithout including these fluctuations into the equationof motion, we still observe very broad spectral linesdue to the quasichaotic nature of magnetizationoscillations in the ‘large-amplitude’ regime. On theother hand, the explanation of still relatively largelinewidths in the regular regime (for current onlyslightly above the critical current for oscillation onset)definitely requires the introduction of thermal fluctua-tions [50].

(vi)

Absence of the out-of-plane regime due to the loss of

coherence: The so called ‘out-of-plane’ oscillationregime when the magnetization of a nanoelement isoscillating around an axis pointing out of the elementplane, is an inherent feature of the steady-stateprecession in the macrospin approximation (seeabove). The signature of this regime is an increase ofthe oscillation frequency with increasing current.Spectral bands with the corresponding f(I)-dependencehave been observed neither in Ref. [49], nor in mostother experiments carried out in the nanopillargeometry when the external field was oriented in thefilm plane. Simulations have explained why these out-of-plane oscillations are almost never observed:macrospin theory shows that ‘out-of-plane’ precessionshould occur for large currents and values of theexternal field large enough for a single energyminimum to exist. However, for such large currentsthe coherence of the magnetization configurationduring the steady-state precession is largely lost, sothat conditions for the existence of this regime are notfulfilled anymore. Modeling the situation studied inRef. [49] has shown that for the currents such that out-of-plane precession would be expected from themacrospin model, relatively small fluctuations of therandom crystal grain anisotropy of Co and thermalfluctuations are strong enough to significantly disturbthe collinearity of the magnetization configuration.Thus the oscillation amplitude of the average magne-tization is strongly reduced, so that the out-of-planeregime could hardly be detected experimentally [50].

Simulations based on a similar geometry, but withmagnetic parameters typical of Permalloy (saturationmagnetization similar to those used in Ref. [50] MS ¼

800Oe, but with a much smaller exchange constantA ¼ 1.0� 10�6 erg/cm) and a slightly reduced element size[39,53], have revealed another possibility to explain the

features of the ‘small-angle’ regime found in Ref. [49].Montigny et al. argued that the almost constant frequencyregime with a low oscillation power should correspond tothe first eigenmode of the elliptical element. For theexchange constant used in Ref. [53], it was shown that thismode is localized near the edges of the elliptical nanoele-ment. Upon current increase, the red-shift mode involvinga quasi-uniform precession within the whole element wasseen to develop with a fast initial rise in rf power. It wasfurther found [39] that for Permalloy-like parameters,‘‘clamshell’’-type orbits do indeed exist in the micromag-netic regime although rather irregular, resulting in a broadlinewidth. Unlike the single spin picture, the largestopening of ‘‘clamshell’’ orbits was not larger than E1801,in agreement with Ref. [50].The frequency drop at the onset of ‘‘clamshell’’ orbits is

of the order of 2GHz for the material parameterconsidered, which is similar to the value obtained inRef. [50], but much smaller than observed in Ref. [49].Montigny’s results seem more in tune with recent experi-ments of Mistral et al. [56], in spite of the line width inMistral’s experiment proving significantly narrower thanever uncovered in simulations (note also that the red-shifting region in Mistral’s experiments proves extremelynarrow).The existence of spectral bands is quite generic as soon as

the angle between the magnetization at rest within the hardand soft layers departs from zero. Lastly, in the sameelements, it was found that out-of-plane orbits also exist,with as a consequence the existence of a blue-shift regime,even in full micromagnetic simulations. However, the out-of-plane trajectories prove notoriously unstable so that thesystem may spend a few nanoseconds precessing out of theplane, then decay in some kind of chaotic trajectory of themean magnetization, before eventually turning back to anout-of-plane orbit. Telegraphic noise type hopping betweenout-of-the plane orbits above and below the plane of thesample is not excluded.Further discussion of simulation results is deferred until

Section 4.4. At this stage, however, it appears that the mostmeaningful difference between simulations outputs in Refs.[50,39,53] occurs in the low current regime. At highercurrents, both authors conclude to the existence of highlyperturbed and limited extent orbits before a ‘‘weak’’ out-of-plane precession regime takes over.

4.2. Skewed geometry (isolated free element): features of

the STI-dynamics which can be reproduced and explained by

simulations [55]

There are two main differences between the multilayerednanopillars used in Refs. [49,51]. First, the stack inRef. [51] contains Py layers instead of Co as magneticcomponents and second, the lower Py layer was sputteredonto IrMn, an antiferromagnet (AFM). It is well knownthat when a FM layer is deposited onto a high-quality layerof some specific AFMs (including IrMn), the exchange

Page 12: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1249

interaction between the FM and the upper atomic layer ofthe AFM, a phenomenon known as ‘exchange bias’,stabilizes the FM magnetization in the direction parallelto the orientation of the upper AFM magnetic moments.Krivorotov et al. have used this phenomenon to ‘pin’ themagnetization of the lower Py layer in a directionsignificantly tilted away from the long axis of the ‘soft’elliptical element. Because the equilibrium magnetizationdirections of the ‘free’ and ‘fixed’ layers are clearly not

collinear (look at geometry in Fig. 5), the initial torqueexerted by the spin-polarized current is anticipated to bealmost homogeneous. For this very reason, the wholestructure is expected to display a markedly more coherentbehavior [41].

An additional advantage of the experiment reported inRef. [51] was the careful characterization of the Py layers.Not only the saturation magnetization MS (which turnedout to be quite low-MSE640G), but also the magnetiza-tion damping parameter a( ¼ 0.025) entering the LLGequation of motion for the system magnetization weremeasured independently—see Refs. [51,55] for details.Another simplification occurs due to a very low value ofthe Py magnetocrystalline anisotropy (KE5� 103 erg/cm3),so that this anisotropy and thus—the grain structure of Pycan be safely neglected. Finally, the experiments underdiscussion were carried out at low temperatures (nominallyat liquid He, but due to the Joule heating the actualtemperature was estimated to be T�10–50K), so thatthermal fluctuations should be less important than for thedata from Ref. [49], hopefully leading to an easier analysis.

As for the previous experiment, we first list the mostimportant features of the experimental results [51]:

(1)

Strong decrease of the oscillation frequency withincreasing current.

(2)

Existence of several frequency jumps with increasingcurrent.

(3)

Evidence for extremely narrow line widths(Df�10–100MHz), which vary non-monotonically withcurrent.

Now we turn our attention to the analysis of thesefeatures based on the insight offered by numericalsimulations.

(i)

Decrease of the oscillation frequency with increasing

current strength: The frequency drop observed inRef. [51] proves quite different from the ‘red’frequency shift discussed previously: the frequencydecreases very rapidly from the very beginning of theoscillation onset, exhibiting several jumps. We remindthat the frequency of ‘small-amplitude’ oscillationsobserved in Ref. [49] was nearly current-independentup to the transition to the ‘large-amplitude’ regime(i.e. in the region 1.7–2.4mA), after which thefrequency started to decrease continuously withincreasing current (without further jumps). Numerical

simulations have shown, as expected, that the rapidfrequency decrease immediately after the oscillationonset is a non-linear effect due to the fast growth of theoscillation amplitude with increasing current. Thisincrease of the oscillation amplitude could also bedetected as a fast increase of the measured microwaveoscillation power emitted by the device [55]. Due to thecareful experimental characterization of the samplesnumerical simulations could reproduce this fast initialfrequency decrease not only qualitatively, but alsoquantitatively without any adjustable parameters (seeFig. 3).

(ii)

Origin of the frequency jumps: The nature of thefrequency jumps in Ref. [51] turned out to be also verydifferent from that of the jump observed in Ref. [49].The jump in Ref. [49] signaled the transition from theregular (with a relatively homogenous magnetizationconfiguration) to the quasichaotic oscillation regimeowing to Berkov’s analysis. This reason could be safelyexcluded for the system studied in Ref. [51], becauseboth before and after the jumps, the oscillationlinewidth was extremely narrow (becoming somewhatlarger in the immediate vicinity of the jumps). Wecould show that the frequency jumps in this case weredue to the transitions between different types ofregular (non-chaotic) oscillation modes, namely—between modes with different spatial localizations.

Simulations have revealed the following picture. Atthe beginning of the magnetization oscillations spatialpower is relatively homogeneously distributed acrossthe nanoelement. When the current is so large that theoscillation amplitude of these homogeneous oscilla-tions is nearly maximal, the transition to a modelocalized in the central region of the ellipticalnanoelement occurs. This transition results in the firstfrequency jump, which was observed for many samplesstudied by the Cornell group. For some samples—including that studied in detail in Ref. [55] a secondfrequency jump could also be observed. Again,simulations have shown that this jump is due to amore subtle transition—namely, from the modelocalized only in the direction along the major ellipseaxis to the mode localized in all directions (seecorresponding detailed explanation with spatial mapsof the oscillation power for various localization typesin [55]). This second transition was observed experi-mentally for only a few samples; taking into accountthat simulations were performed on the structure withperfect borders, we can assume that samples where thistransition was found, have an especially neatly shapededge surface.

(iii)

Very small spectral linewidth: A linewidth as small as10MHz for ranges of current values is a reliableevidence that magnetization oscillations observed inRef. [51] remain regular up to the currents about atleast �8mA. This is also in sharp contrast with resultsin Ref. [49], where the transition to the quasichaotic
Page 13: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

Current I (mA)0 4 10

Inte

rgr.

sign

al a

mpl

. Sin

t

0.0

0.1

0.2

0.3

0.4

0.5

Freq

uenc

y f

(GH

z)

0

2

4

6

8

simulationsexperiment

2 6 8Current I (mA)

0 4 102 6 8

Fig. 3. (a) Comparison of experimentally measured (triangles) and simulated (circles) dependencies of the oscillation frequency of on current strength. (b)

The same for the integrated spectral density Sint.

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591250

behavior occurred already for IE2mA for thenanoelement with approximately the same lateral sizes(130� 70 nm2). This is especially astonishing, takinginto account that Co used in Ref. [49] as a material forFM layers, has a much higher exchange stiffnessconstant (ACoE3� 10�6 erg/cm) than Py (APyE1.3�10�6 erg/cm), and a higher exchange stiffness shouldobviously stabilize a more uniform magnetizationstructure, thus preventing the system from ‘sliding tochaos’ (we remind the reader that according to themeasurements reported in Refs. [49,51], we have usedfor Co (MSE800G) and Py (MSE650G) values of thesaturation magnetizations which are quite close). Asomewhat higher thickness of the FM layer used inRef. [51] (hPyE4 nm) compared to Ref. [49] (hCoE2.5 nm) can hardly overcompensate such a differencein the exchange constants.

We assume [55] that the preservation of the regularoscillation regime up to such high currents as observed inRef. [51] is due to the large angle between the polarizationdirection of electrons responsible for the spin-torque effectand the equilibrium magnetization of the ‘free’ layer.Indeed, if the electron polarization vector p is tilted relativeto the average magnetization M of the free layer strongenough, the spin-torque Nst has roughly the same directionfor the magnetization across the whole elliptical nanoele-ment, thus supporting—and not destroying, as shown inFig. 2a, the homogeneous magnetization state. This is mostprobably the reason why the regular oscillation regime‘survives’ up to pretty high currents in nanopillars with thehard layer pinned at some fair angle from the easy axis.

4.3. Skewed geometry: simulations including the

magnetodipolar interlayer interaction

Up to this point, the free layer has been treated asisolated. However, in a stack such as described in Ref. [51],the etching process is anticipated to embed both the freeand the pinned layer. It ensues that the latter exerts a stray-

field on the former. That field is by no means small. Fig. 4shows the values of the x and y components of the stray orbiasing field averaged over the volume of the free layer aswell as the extremum value of the z component as afunction of a hypothetical cone angle of the pillar (sucheffects are rather common in nano-fabrication). Clearly,the average in-plane stray-field components prove littlesensitive to the cone angle. In contradistinction, the maxi-mum (and minimum) value of the out-of-plane stray-fieldcomponent decreases sharply with increasing cone anglealthough in all cases remaining particularly large. Beinghighly inhomogeneous, the stray-field potentially affectsmode localization. Moreover, this field is in no way muchsmaller than the applied field or the demagnetizing fieldfrom the free layer itself.The stray-field in Fig. 4 has been corrected for stair-case

effects that arise as soon as the charge distribution alongthe rim of the element becomes non-monotonic and piece-wise constant due to the tile decomposition of the curvedrim boundary. The stray-field distribution indeed becomesirregular in the immediate vicinity of a stair-case boundary,a clearly unphysical result (see Refs. [57–60] for differentapproaches to this problem). The same remark applies todemagnetizing field computations for all simulation resultsin this section. Dispersion curves computed in the singlespin limit for a perfectly cylindrical pillar, including thestray-field (hHstr

x i ¼ �8:63 mT, hHstry i ¼ �11:77 mT), ap-

plied field (Happx ¼ þ48 mT, Happ

y ¼ �48 mT, Happ ¼

68 mT) and anisotropy field representative of a 4 nm thickfree layer (MS=650 kA/m) with elliptical cross-section anddimensions 2a=130 nm, 2b=60nm (i.e. HK ffi 50 mT) isshown in Fig. 5 for two values of the damping constant:a ¼ 0.010 and 0.025. The magnetization direction withinthe fixed layer is still supposed to be pinned at +301 fromthe long ellipse axis.For the small damping constant, red-shifting proves slow

after the fast initial decrease because of the competitionbetween orbit opening as a function of increasing currentdensity and a high mobility along the trajectory due to thelow damping constant. The existence of a large region with

Page 14: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

-130

-120

-110

-100

-90

-80

310

320

330

340

350

360

370

380

390

0 10 20 30 40

<Hx>

HzMAX

Pillar Angle (°)

<Hy>

Fig. 4. Variation of the average in-plane and maximum out-of-plane components of the stray-field due to the fixed layer if pinned at 301 from the ellipse

long axis. The extremum values of the out-of-plane stray-field component are to be found in the dark blue and red regions of the field map on the right.

x

yp+30°

HaHb

–45°

3

4

5

6

7

8

0 10 15

0 0.5 1 1.5 2 2.5J (A/µm2)

I (mA)

f (G

Hz)

α=0.010

α=0.025

≈–126°

5

2.5 mA

Fig. 5. Dispersion relation in the macrospin approximation for a

‘‘symmetrical’’ spin valve Py (4 nm)/Cu (8 nm)/Py (4 nm) for an

asymmetry parameter wa ¼ 0.5. Simulation parameters: MS ¼ 650 kA/m,

HK ¼ 50.2mT, Happ¼ 68mT, HstrE14.5mT The effective polarization

has been adjusted so that the critical current is in both cases close to

2.7mA, i.e. Peff ¼ 37:5% for a ¼ 0.025 and Peff ¼ 15% for a ¼ 0.010.

The temperature is assumed equal to T ¼ 40K, independently of current

amplitude.

3

4

5

6

7

8

0 10 15

0 0.5 1 1.5 2 2.5J (A/µm2)

I (mA)

f (G

Hz)

Mode Hopping

α=0.010

Mode Hopping

? α=0.025

5

3

4

5

6

7

8

0 10 15

0 0.5 1 1.5 2 2.5J (A/µm2)

I (mA)f (

GH

z)

5

2.5 mA 2.5 mA

Fig. 6. Comparison of dispersion relations between macrospin and full

micromagnetic simulations for two values of the damping parameter.

Simulation parameters: wa ¼ 0.5, MS ¼ 650 kA/m, Happ¼ 68mT,

HbE14.5mT. The effective polarization has been adjusted so that the

critical current is in both cases close to 2.7mA, i.e. Peff ¼ 37:5% for

a ¼ 0.025 and Peff ¼ 15% for a ¼ 0.010. Temperature T ¼ 40K is

independent of current.

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1251

reduced df/dJ is reminiscent of similar features insystems with p ¼ x when a transverse in-plane field adds-up onto the usual longitudinal field [39] and is consistentwith the modification of the energy landscape: thecombination of the applied and biasing fields indeed actsas a field with a large longitudinal but even strongertransverse component. For the larger damping constant,there exists a critical current above which precession shiftsinto the out-of-plane regime, as also observed in themicromagnetic regime in the absence of Oersted field [55].Experimental current values do not exceed 10mA, valuesfor which the precession frequency owing to the single spinapproximation remains particularly large w.r.t. experimen-tal values.

Moving to the micromagnetic regime leads to thedispersion curves shown in Fig. 6. From these figures,several conclusions may be drawn:

(i)

The thermal frequency of the order of E6.5GHzremains close to experimental values.

(ii)

Red-shifting always occurs faster for simulationsoperated in the full micromagnetic regime as comparedto macrospin simulations, which, as noticed pre-viously, are unable to predict mode hopping.

(iii)

Contrary to simulations performed in the absence ofany biasing field, true mode hopping is only observedfor the low damping constant, whereas, for the highervalue, frequency jumps merely decay into slopechanges in the dispersion curve.

(iv)

Frequency jumps or local changes in the slope of thedispersion curve occur at well-defined frequencies.
Page 15: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSf(

GH

z)

Fig.

Fig. I

No i

comp

Some

harm

agree

10.45

Happ

are th

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591252

Characteristic frequencies amount here to about 5.5and 4.5GHz. They differ little from experimentalvalues. If these frequencies are to be viewed asrepresentative of the energy landscape under thepresence of both a biasing and an applied field, thenit needs being admitted that they differ little from jumpfrequencies calculated in the absence of any biasingfield.

(v)

Very unfortunately, red-shifting still proves muchslower than in experiments. The second transition,especially, takes place at current densities almost twiceas large as the experimental value and the lastcomputed frequency is for the lowest dampingconstant also almost twice as large as the experimentalfrequency for I ¼ 10mA.

On the other hand, the power spectral density remainsnicely concentrated within extremely narrow line widths inthese simulations as evidenced in Fig. 7, where nosubstantial line broadening may be observed, even for thelargest current densities, except for the largest dampingparameter value.

Summarizing at this point, it appears that simulationsincluding the biasing field do only remotely describe thecharacteristics of experimental dispersion curves. Bestagreement is found for low damping that preserves theexistence of marked frequency jumps occurring at well-defined frequencies. If the effective polarization is chosensuch as to match the current threshold for the onset ofprecessional states, then the average red-shifting hdf =dJi

proves smaller than expected by a factor close to 2. Even if

2.5

5.0

7.5

10

J(A /µm2)0 2

J(A /µm2)0 2

α α = 0.010 α α = 0.025

7. Power spectral density (PSD) maps (log scale) corresponding to

II.4-3. Low damping map on the left, higher damping on the right.

nterpolation has been performed: the power spectral density is

uted at the center of each visible frequency segment in the images.

current domains display clear overlapping frequencies. A non-

onic spreading of the PSD may also be observed. Horizontal scale in

ment with current ranges in Fig. 6; vertical axis ranging from 0.67 to

GHz. Simulation parameters: wa ¼ 0.5, MS ¼ 650 kA/m,

¼ 68mT, HbE14.5mT. The effective polarization and temperature

e same as in Fig. 6.

taking into account some of the worse discrepancies foundin either mode analysis or vortex gyration experimentsquoted earlier, such a result may only appear as extremelypoor.More generally, it may be stated that

(1)

for a given damping constant, increasing the effectiveelectron polarization leads to faster red-shifting. How-ever, (i) the critical current for the onset of precessionalstates decreases accordingly, (ii) frequency jumpsbecome gradually smeared-out,

(2)

for a given damping parameter and a given effectiveelectron polarization, increasing the asymmetry para-meter wa tends to translate the dispersion curvestowards lower current densities without significantalteration of the average red-shifting velocity.

4.4. Nanopillar geometry: STI-dynamics features which

could not be reproduced by simulations

Although a fair number of features of spin-torque drivenmagnetization dynamics could be successfully explainedeither by means of macrospin or full-scale micromagneticsimulations, these achievements should not be mistaken fora full understanding of this phenomenon. Several impor-tant issues must be clarified before the physical model ofthe spin-torque driven magnetization motion can bedeclared at least as reliable as standard magnetizationdynamics studies based on the LLG equation. Here aresome of the open problems related to experiments in thenanopillar geometry.

(1)

A very serious discrepancy is the large difference

between simulated [50] and measured oscillation frequen-

cies in the Co/Cu/Co nanostack in Ref. [49]. To startwith the analysis of this problem, we first point out thatthe frequency is the observable that can be measuredmost reliably, with a very high precision, indeed. Allother features of the spectral lines—line widths, peakheights, total power, etc.—require a much moresophisticated analysis to enable a quantitative compar-ison with theory, but the frequency is usually deter-mined unambiguously. Keeping this argument in mind,the inability of simulations to reproduce experimentallymeasured oscillation frequency of the Co/Cu/Conanopillar is alarming.

The first attempt to reproduce the experimentalfrequency was made by Lee et al. [52], who used thestandard Co saturation magnetization MS ¼ 1400Gand a lower value of the Co exchange stiffness(ACo ¼ 2� 10�6 erg/cm instead of the more usual valueACo ¼ 3� 10�6 erg/cm). The choice for a lower ex-change constant was justified by the authors [52] as amean to take into account thermal fluctuations effectswithout having to solve a stochastic equation of motion(we note in passing that the validity of this tactics is

Page 16: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1253

presently a subject of debate, but well beyond thescope of this paper). Comparison of simulation resultsfrom Ref. [52] with measured frequencies is discoura-ging: in the quasichaotic regime the simulated frequen-cies were in the range fsimE10–12GHz (see Fig. 3d in[52]) for an applied field value equal to H0 ¼ 400Oe.The measured frequencies are about fexp�5.5GHz forH0 ¼ 2000Oe, which means that for H0 ¼ 400Oe fexpwould be even lower. We have simulated the quasi-chaotic dynamics of the Kiselev’s geometry with thesame magnetic parameters as in Ref. [52] and foundthat the quasichaotic regime starts with fsimE15GHzfor H0 ¼ 2000Oe, i.e. nearly three times the measuredvalue!

One possible reason for such a discrepancy could benon-standard values of Co magnetic parameters (firstof all, the saturation magnetization MS) when Co ispresent as a thin film. In an attempt to determine theseparameters experimentally, Kiselev et al. have mea-sured the saturation magnetization of a stack consistingof several 3 nm thick extended (not patterned-pro-cessed!) Co layers and have found the value4pMS ¼ 1 0:1 T, so that MSE800G [49]. Oursimulations [50] with this reduced MS value yielded(for H0 ¼ 2000Oe) an oscillation frequency atthe beginning of the quasichaotic regime equal tofsimE11GHz (for the exchange stiffness ACo ¼ 2�10�6 erg/cm) and fsimE9GHz (for ACo ¼ 3� 10�6 erg/cm), still a much too high value. Apart from theunknown crystalline structure of the Co films used inRef. [49], surface anisotropy and edge oxidation effectsmay play an important role in such small systems.Whether one or both of these reasons play a significantrole requires careful magnetic and spatially resolvedstructural (and chemical) measurements performed onmultilayers identical to those intended for use in spin-torque experiments. In addition, increasing the torqueasymmetry can reduce the oscillation frequency in thelarge-angle regime (see results in Ref. [55]), but again,one needs independent experimental measurements ofthe corresponding asymmetry parameter to drawmeaningful conclusions.

Lastly, rather unexpectedly, a better fit betweenmicromagnetic simulations and experimental resultswas obtained in the red-shift regime for materialparameters classical for a Ni80Fe20 alloy and a fieldequal to 3HK [39,53]. The frequency observed in thevery low current regime could certainly not bereproduced, unless assuming that the GMR signalobserved at low current densities only displays the 2f

signature of a collinear system whereas the funda-mental frequency appears at larger densities due,perhaps, to current induced magnetization oscillationswithin the hard layer. Such an assumption, previouslyunnoticed, potentially offers a credible explanation forthe huge frequency drop occurring at the onset ofprecessional states in Ref. [49].

(2)

The second important problem in nanopillar experi-ments concerns the dependence of the oscillation power

on the current strength. We note first, that obtainingquantitatively accurate values of the oscillation power(in order to enable a meaningful comparison withsimulations) is a non-trivial experimental task, as canbe seen from the corresponding discussion in Ref. [55].However, this task was successfully solved for theIrMn/Py/Cu/Py nanopillar in the experiment analyzedin detail in Section 4.2. Comparison of experimentaland numerical results (together with the frequency vscurrent dependence f(I)) is shown in Fig. 3b. Thequalitative discrepancy between simulations and ex-periments is evident. Simulated power grows veryrapidly for currents slightly above the critical value(due to the rapid growth of the oscillation amplitudeassociated to a nearly coherent magnetization preces-sion), reaches its maximal value just after the firstfrequency jump and then slowly decreases, as aconsequence of the increasing inhomogeneity of themagnetization configuration at higher currents. Incontrast to this behavior, the experimentally measuredpower grows much slower for currents just above thethreshold and increases monotonically with current(except the dips for the current values where thefrequency jumps take place).

This disagreement is especially surprising taking intoaccount the rather good coincidence of measured andsimulated frequencies (see panel (a) of the same figure).As explained above, frequency decrease with increasingcurrent strength is a non-linear effect arising from thedependence of the oscillation amplitude on currentvalue. The oscillation power is directly related to theamplitude, so that the large (partly qualitative!)disagreement between simulated and measured powersfor the situation where the corresponding frequenciesagree well, is a puzzling phenomenon, clearly requiringfurther investigation.

(3)

Turning back to simulations including the bias fieldarising from the pinned layer, the dispersion relationf(J) is clearly unsatisfactory, but, on the other hand, theoverall behavior of the line width and power spectraldensity vs. current (see Fig. 7) does not appear that faraway from experimental data. In order to reach a betterfit in the dispersion relation, unless admitting that thebiasing field proves, for an unknown reason, muchsmaller than estimated from micromagnetics, is thereany parameter that could still be tuned in thesimulations? It seems indeed rather paradoxical thatnone of the simulations we know of correctly predictsthe right current for the second frequency jump.Heuristically, beyond surface anisotropy and potentialexchange biasing along the edges, all simulationparameters could be made current density dependent.For instance, both the damping parameter a and thetorque asymmetry parameter wa could be thought ofincreasing with increasing current density J. Although
Page 17: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591254

highly speculative, it could also be assumed that thesaturation magnetization MS decreases with increasingJ. Test simulations seem to indicate that, at least forsimulations avoiding the intrinsic deleterious effects ofstair-case magnetostatics, assuming current densitydependent parameters such as wa or MS does not leadto line width broadening, even for the largest currentdensities. However, before moving ahead, theoreticalinsights would here be more than appreciated.

5. Simulations of steady-state precession in the point-contact

geometry

5.1. Methodological problems

Before we start with the discussion of experiments andsimulations obtained in the point-contact geometry, wewould like to point out that simulations of this geometryencounter several complicated methodological problems[48,61,62], which make this kind of simulations much morechallenging than modeling of the nanopillar devices.

We remind that the corresponding experimental setupconsists of a nanowire in contact with an upper layer of amultilayered ferromagnetic system. The current flowsthrough this nanowire and magnetic layers contacted byit. If we assume for simplicity that the system is placed intoan external field large enough to saturate magnetic layers,then the lower layer will preferably reflect those currentelectrons, which magnetic moments are oriented antipar-allel to its magnetization. According to the general conceptof the spin-torque, such electrons create a torque acting onthe upper layer magnetization, which can lead to theinstability, spin-wave excitations and even switching in theupper layer, exactly as in the nanopillar geometry.

However, the point-contact geometry and hence—corresponding physics—is quite different from that of thenanopillar device. In the point-contact geometry, the non-magnetic spacer thickness (5–10 nm) is much smaller thanthe contact diameter (25–80 nm). Thus, the transversaldiffusion of current electrons is negligible and reflectedelectrons act mainly on the magnetization in the upperlayer region under the contact. The rest of the layer is (in afairly good approximation) not affected by the spin-torque.This means that magnetization excitations created by thespin-torque in the region under the point contact, canpropagate in the rest of the magnetic layer as waves emittedby a small source (but not a point source!). Hence, themagnetization dynamics in this case is expected to bequalitatively different from the nanopillar case where spinwaves are excited ‘simultaneously’ over the whole area of amagnetic nanoelement.

This difference immediately leads to several seriouscomplications when modeling magnetization dynamics.Namely, in order to describe adequately the wavepropagation and the influence of the magnetization of therest of the layer on the point area, we need now to simulatenot only the thin film region where the magnetization is

excited by the spin-torque (region under the contact), butalso the surrounding thin film. The problem is that actuallateral sizes of a multilayer in such experiments are about�10–20 mm, making the simulation time for the wholesystem prohibitively long.The obvious solution seems to simulate a smaller area,

say 1� 1 mm2, in the hope that this area is still much largerthan the point-contact size, so that our simulations willcapture all the necessary physics. This hope is in principlecorrect, but only in principle, because the naive straightfor-ward implementation of this idea fails due to improperboundary conditions. In a real experiment the emittedwave propagates in the laterally wide thin film until it diesdue to natural energy dissipation. In simulations relying ona much smaller lateral system size, two situations arepossible: (i) the emitted wave of a significant amplitude isreflected from the thin film border—for open boundaryconditions, or (ii) the wave coming from the systemreplica(s)—for periodic boundary conditions—penetratesinto the simulated area. In both cases we deal with anartificial ‘incoming’ wave, which propagates from theborders of the simulated region towards the point-contactarea. The interference of this second wave with themagnetization oscillations within the point-contact regionmay lead to unpredictable and physically completelymeaningless results, especially taking into account thatthis interference is particularly strong—both waves havethe same frequency.The ‘clean’ solution of this problem requires the

analytical derivation of perfectly absorbing boundaryconditions which should be imposed on the magnetizationat the system borders in order to completely absorb theincident wave. However, this derivation has not been doneyet and it is not clear whether it can be done at all. For thisreason we have proposed a numerical trick basing on theintroduction of a space-dependent dissipation [61,62].Namely, dissipation constant at and near the point-contactarea is set to real physical dissipation value, so thatoscillations of the point-contact area and the emitted waveare not affected. With the growing distance to the contactcenter the dissipation increases, so that near the simulationarea borders it is so large that the wave passing throughthis area loses all its energy and ‘disappears’. The spatialprofile of this dissipation increase and the maximumdissipation value at the area border should be chosencarefully: if the profile is too steep, wave reflection can stilloccur due to too rapid a change of the medium properties.The rigorous theory how to choose such a profile is alsonot available (although numerical criteria whether theprofile is chosen correctly, do exist [62,63]), so thecorresponding operation is still more art than science.

5.2. Various dynamic modes and their excitation threshold

Experimental results on point-contact systems, whichcan be used for a meaningful comparison with simulations,are very rare [43,44,64,65], what is surely one of the major

Page 18: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1255

reasons why the corresponding dynamics is understoodmuch poorer that for nanopillar devices.

After the methodical difficulties mentioned in theprevious section were resolved and the first rigoroussimulations of STI-dynamics in point-contact devices couldbe performed, it has turned out that for this kind ofexperiments not even a qualitative agreement betweensimulations [66] and ‘real’ data [43] could be claimed. Twomajor qualitative discrepancies between theory and experi-ments were: (i) for the current just after the oscillationonset, the simulated frequency was almost two times largerthan measured experimentally and (ii) even when theOersted field of the electric current was neglected, twosteady-state precession regimes were found in simulations(with nearly opposite directions of the precession axes, verydifferent oscillation frequencies and different frequency vs.current characteristics), when only one regime wasobserved in the experiment.

The discrepancy between the oscillation frequenciesshould indeed be considered as a qualitative, and not aquantitative one. Namely, the frequency measured experi-mentally fexpE7.6GHz (see Fig. 1 in Ref. [43]) wasnot only nearly two times smaller than the valuefsimE12–13GHz found numerically; the major problemwas that the experimental frequency was below thehomogeneous FMR frequency fFMR ¼ 8.4GHz for thesystem studied in Ref. [43], i.e. a thin film made of Py withthe saturation magnetization MS ¼ 640G, placed in anexternal field H0 ¼ 1000Oe. This means that the modedetected experimentally could not be the propagatingwave, as it was the case in numerical simulations for theregime observed directly after the oscillation onset. In thisfirst numerical study [66], a regime with magnetizationoscillations localized under the point-contact area was alsofound—this was the second regime mentioned above.However, the nature of this localized oscillation modewas not clarified in Ref. [66].

Detailed theoretical studies of magnetization oscillationsinduced by a point-contact spin-torque were performed bySlavin et al. [47,67] using analytical methods suitable fornon-linear dynamics. In Refs. [47,67], it was shown thatthere exist at least two solutions for the equations of themagnetization motion under the influence of the Slonc-zewski torque. The first solution which was already foundby Slonczewski [20] is the standard solution of thelinearized equation of motion and represents the propagatingwave with the wave vector k�1/Rc (Rc being the radiusof the point contact). The frequency f(k) for this regimeis much higher than the FMR frequency, because thecorresponding wave vector is relatively large. As a solutionof the linearized equation, it could have an arbitrarily smallamplitude. Oscillations in the first regime—just after thecritical current—found in Ref. [66] do correspond to thispropagating solution.

Slavin and Tiberkevich [67] have found out, that the‘native’ (non-linearized) equations of motion in the point-contact geometry possess also another solution—the

so-called non-linear ‘bullet’ mode. Magnetization oscilla-tions in this mode are localized within the point-contactarea, i.e. their amplitude b0 decreases exponentially withdistance r from the point-contact center (b0�exp(�r/Rdec),decay radius Rdec�Rc); we recall that b0�1/r

1/2 for thelinear mode in 2D. The frequency of such a localized modeis smaller than the lowest possible frequency of thepropagating mode fFMR; this is an inherent feature of alocalized solution. The amplitude of this mode cannot bearbitrarily small: as often the case for non-linear oscilla-tions, the amplitude acquires a finite value immediatelyabove the critical current threshold.Another interesting feature of this mode is also due to its

localization. Namely, being localized, the ‘bullet’ does notlose energy due to energy radiation in form of apropagating wave and thus can have a smaller excitationthreshold (smaller critical current required for its excita-tion), despite the fact that right after the excitation itshould have a significant amplitude. This analyticalprediction was the subject of a controversial debate.Namely, the localized mode reported in Ref. [66] anddescribed in more detail in Ref. [61] had many commonfeatures with the ‘bullet’ described by Slavin et al.: spatiallocalization near the point-contact area, frequency lowerthan fFMR, nearly homogeneous structure of the modekernel. On the other hand, in simulations the localizedmode was excited after the linear (Slonczewski) mode, i.e.for currents larger than the linear mode threshold, so thatthe question about which mode was observed in a realexperiment, still remained unclear.This controversy was resolved in the work of Consolo et

al. [63], who could show that in order to observenumerically the non-linear bullet with the excitationthreshold smaller than that for the linear mode, one shouldnot increase the current, but rather move from largecurrents—where the ‘bullet’ does already exist—to smallerones. In this case the non-linear solution persisted down tocurrents smaller than the linear mode threshold, confirmingthe prediction from Ref. [67]. Stated otherwise, magnetiza-tion excitation dependence on current strength for thepoint-contact geometry demonstrates a hysteretic behaviorat T ¼ 0.Essentially the same explanation for the discrepancy

discussed above was suggested in Ref. [48], where theenergy behavior for corresponding modes was studied as afunction of increasing current. It was found, that althoughthe non-linear mode was excited later than the linear onewith increasing current, the average system energy for thisnon-linear mode was significantly lower than for theSlonczewski solution (propagating wave), so that theSlonczewski mode was only metastable. It was excited firstbecause simulations done for T ¼ 0 could not providefluctuations required to excite a non-linear mode with afinite excitation energy threshold. So at the present stateof knowledge the most likely explanation of the resultsdescribed in Ref. [43] is that due to thermal fluctuations thenon-linear ‘bullet’ mode was observed there. Experiments

Page 19: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESS

Pow

er a

s <mz2 (t)

>

0.0

0.1

0.2

0.3

0.4

0.5

aJ

aJ

0 4

Freq

uenc

y f(m

z), G

Hz

0

4

8

12

16

W L1 L2

2D Graph 1

aJ

4.5 5.0 5.5 6.0

7.27.37.47.5

150 x 150 nm2 100 x 100 nm2

765

0 4 765

f, G

Hz

a

b

c d

Fig. 8. Dependencies of the oscillation frequency f (panel (a)) and power

P (panel (b)) on the spin-torque magnitude aJ (proportional to the current

strength I) for the point-contact setup. Transitions from the extended wave

mode W to the localized mode L1 and from L1 to the second localized

mode L2 are clearly seen. Snapshots of the in-plane magnetization

projection perpendicular to the applied field direction for all mode types

are shown as gray-scale maps. Panels (c) and (d) display the in-plane

magnetization arrow images of the localized mode cores: (c) L1 mode

(every second moment is shown), (d) L2 mode.

D.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591256

from Ref. [43] were performed at room temperature, sothat this mode could be excited although observations weremade under increasing current.

As a matter of facts, it quickly turned out that from thetheoretical point of view the situation proves even morecomplicated. Already in Ref. [62] another non-linearlocalized mode was detected. In contrast to the solutionfound in Ref. [67], this additional mode had a kernel with ahighly complicated magnetization structure consisting oftwo vortex–antivortex pairs (see Fig. 8d). More detailedstudies have shown [48] that for currents where both non-linear modes (‘bullet’ and vortex–antivortex) can exist, theenergy of the vortex mode in significantly lower, so thatthis mode should be actually observed. Moreover, we havealso shown, that some perturbations present in real system(e.g. the dipolar field from the ‘fixed’ layer) strongly favorthe transition from the ‘bullet’ to the vortex mode. Thereason why Rippard et al. have found the ‘bullet’ mode,thus remains unclear.

Experimentally this second (vortex–antivortex) modetype can be distinguished from the ‘bullet’-like oscillationsusing the following criteria.

First, the frequency of the vortex mode is muchlower than that of the ‘bullet’ mode; e.g. for con-ditions corresponding to the experiment of Rippardet al. (MS ¼ 640G, H0 ¼ 1000Oe) [43] we have foundfbullE2fvort [48]. In smaller external fields, the frequency ofthis mode can be as low as �200–300MHz. The reason forsuch a low frequency is that the oscillations of the averagemagnetization in this mode occur due to the creation–an-nihilation of vortex–antivortex pairs, which is a rather slowprocess. As mentioned above, frequency of the microwaveoutput signal can be measured experimentally and deter-mined in simulations very reliably, so that this criterion isrelatively easy to apply, if the real system is characterizedcarefully enough.

Second, the total oscillation power in the vortex mode(averaged over the point-contact area) is significantlysmaller—up to several times—than that of the ‘bullet’ mode.This is due to a more complicated magnetization configura-tion of the vortex mode kernel. Usage of this criterionrequires reliable quantitative measurements of the microwaveoscillation power produced by the point-contact device.

Third, although the frequency of both localized modes islower than fFMR, so that they cannot emit power in form of‘normal’ propagating waves, they still radiate energy in formof spatially localized solitons. This radiation is highlyanisotropic and the anisotropy patterns are very differentfor ‘bullet’ and vortex modes. However, to apply thiscriterion, one needs to perform measurements of themagnetization oscillations for frequencies in the GHz rangeand with the spatial resolution �50nm, which is extremelydifficult. On the other hand, any attempt to obtain thesynchronized point-contact oscillators (keyword ‘frequencylocking’) [45,46] in the field-in-plane geometry requiresdetailed studies of the energy radiated from point contacts,so that further progress on this topic is quasi unavoidable.

Concluding this subsection, we would like to mention,that our preliminary simulation results indicate that thevortex–antivortex mode could be responsible for the low-frequency oscillations reported very recently in Ref. [68],where experimental data for the point-contact geometrywith a somewhat higher contact radius (RE40–50 nm) arediscussed.

6. Importance of a sample characterization

It was known from the very beginning of research on STImagnetization dynamics that one would need extremely

Page 20: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1257

small sizes of the corresponding structural units (devices) inorder to separate the STI-induced phenomena from effectsarising from the Oersted field generated by the large electriccurrent required to induce steady-state precession or evenswitching of the sample magnetization. Indeed, most of thehigh-quality experiments used to compare theoreticalpredictions with experimental findings were performed onvery small nanopillars (lateral sizes �50–200 nm) or point-contact devices with contact diameter 25–80 nm.

Characterization of magnetic structures with such smallsizes presents a new challenge to the experimentalists: oneshould not only know the exact magnetic parameters of theextended films which are used to produce, e.g. nanopillardevices, but also be able to measure structural, chemicaland magnetic properties of the nanoelements themselves.

First such a knowledge is indeed crucial if a realtheoretical understanding of spin-torque physics is soughtfor, and not only a qualitative description making use ofanalytical perturbation theories or the macrospin approx-imation. Micromagnetics is credited with a high predictivepower, because for ‘standard’ (not involving spin-torque)magnetization dynamics is it able to describe and explainexperimental data with a high precision. On top ofexamples quoted in the introduction, a good recentexample can be found, e.g. in Ref. [69], where non-trivialdiscrete spectra of magnetization oscillations in rectangularnanoelements (measured via the Brillouin light scattering)were reproduced with a very satisfactory quantitativeagreement.

Hence, if aiming at such a quantitative agreement for theSTI-dynamics also, we actually need a precise character-ization of each particular device the dynamics of whichneeds to be modeled. The reason why we really needcharacteristics of a particular sample is the large sample-to-sample variation of experimental data. For example,looking at the f(I) curves in Fig. 1 from Ref. [49] andcomparing them with the phase diagram of the oscillationpower in (I–H0) co-ordinates (Fig. 2 from Ref. [49]), we canimmediately recognize that the two devices which data areshown on these two figures, produce significantly differentdynamical output. The same conclusion naturally resultsfrom the wide region of fitting parameters which, accordingto Ref. [49], was required to fit the data of various samples.In Ref. [55], frequency vs. current curves are explicitlyshown (see Fig. 6 in Ref. [55]) and clearly demonstratesignificant variations from sample-to-sample.

The most important features of the nanodevices (leavingaside evidently required characteristics like the saturationmagnetization, exchange constant, average bulk anisotro-py, etc.) which one definitely needs for a quantitative- andoften even for a qualitative-data analysis using micro-magnetics, are the following:

(1)

Polycrystalline structure of the magnetic films used toproduce nanopillars or point-contact devices. Here weneed the average grain size, grain texture (if present),and the type of crystallites (crystal lattice). These data

are especially important when analyzing data fromdevices based on Co or CoFe films, because themagnetocrystalline anisotropy of these materials canbe high and there are even two possible crystallographicphases for Co-hcp and fcc. In our detailed analysis wehave shown [50] that the magnetization dynamics of Conanoellipses can qualitatively depend on the type of Cocrystal grains.

(2)

Surface and edge anisotropy are also very important forthe devices under discussion, because, due to the verysmall element sizes, surface and edge effects play a veryimportant role. We strongly suspect that the largesample-to-sample variations of the observed dynamicsfound, e.g. in Refs. [49,55] (and many other papers) isat least partly due to an imperfect element shape andedge anisotropy.

(3)

Irregularities of the chemical composition: In particular,it is known that for some specific patterning techniquesnanostructure edges are oxidized to some extent. It isalso known, that oxidation may affect not only thesaturation magnetization, but also the damping con-stant of a magnetic material, which is also a veryimportant parameter when performing simulations.

(4)

In exchange biased spin valve systems the orientation of

the exchange bias direction is an essential information.It is hardly possible to change the exchange biasorientation directly, so one should perform themeasurements of quasistatic hysteresis loops on suchspin valves using, e.g. the GMR effect and then try toobtain the required direction by fitting this loop.

7. Conclusion

Keeping in mind that this series of ‘Current Perspectives’papers should present the very recent state of the art andsimultaneously ‘reflect the personal point of view’ of theauthors, it is almost impossible to write a conclusion in thesense known from ‘standard’ scientific contributions. Forthis reason we would like to confine ourselves to thefollowing final remarks.First, nowadays everybody is used to numerical simula-

tion analysis as an extremely useful tool for interpretingexperimental results and gaining new insights into thephysics of various processes. On the other hand, formicromagnetics in particular, this acceptance of numericalsimulations coexists (in a very strange way) with the pointof view, that when full-scale simulations give a poorer

agreement with experiment than oversimplified models(first of all the macrospin approximation), then one shouldabandon any attempt to obtain a good agreement usingfull-scale micromagnetics and return to these simplemodels. So, especially in connection to micromagneticsimulations of the STI-dynamics, which still demonstratemany important discrepancies with experimental results,we would like to emphasize the following. Full-scalesimulations are supposed, first, to take into account more

Page 21: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–12591258

(and not fewer) known features of the studied systems thansimplified (semi-)analytical models, and second, to usefewer (if any) adjustable parameters. This means, that iffull-scale simulations demonstrate a poorer agreement withexperiment than the macrospin analytics and/or simula-tions, then we should realize that we do not understand

some crucial physical properties of the system under study.Hence, the disagreement between simulations and experi-ment obviously calls for further research, and not forreturning to oversimplified models, where a good fit toexperimental data can still be achieved by adjusting severalparameters without having a real physical justification forsuch an adjustment.

Second, meaningful full-scale simulations require anaccurate knowledge of the underlying system parameters,obtained from independent sources (preferably high-quality experiments). Being still very time-consuming,full-scale simulations usually do not allow for an explora-tion of the whole parameter space, so that the correspond-ing experimental input is really mandatory. Applied tosimulations of the STI-dynamics, this statement means, ofcourse, that we need independent analytical calculationsand experimental measurements of the STI-specific char-acteristics like the current polarization degree for variousFM/NM interfaces. However, probably even more im-portant is the understanding, that we need not only theaverage values of the saturation magnetization, exchangestiffness, etc. of nanomagnets under study, but also theinformation how the processing of corresponding devicesmay change magnetic parameters locally, especially nearthose edges of the nanostructures which are influenced (oreven created) by this processing. As shown above, suchinformation is crucially important for the modeling ofnanodevices due to their extremely small sizes, resulting inthe increasing influence of the structure imperfections andlocal parameter variations.

And last, but not least, recent simulations haveshown that the influence of the Oersted field mayhave a qualitative impact on the observed dynamics. Thisinsight raises the question how to correctly calculatethis field in the state-of-the-art spin-transfer devices.Being in principle straightforward and purely tech-nical, such calculations still require an exact knowledgeof the 3D distribution of an electric current in suchsystems, a task which is definitely out of the scope of themicromagnetics and requires a significant separate effortfor its solution.

Acknowledgments

JM gratefully acknowledges support from the EuropeanCommunity programme ‘‘Structuring the ERA’’, undercontract MRTN-CT-2006-035327 SPINSWITCH, DBthanks the German Research Society for its support inframes of the SPP ‘‘Ultrafast magnetization dynamics’’(project BE 2464/4).

References

[1] P. Gambardella, et al., Science 300 (2003) 1130.

[2] S. Rohart, V. Repain, A. Thiaville, S. Rousset, Phys. Rev. B 76 (2007)

104401.

[3] W.F. Brown, Micromagnetics, Interscience Publishers, Gouda, 1963.

[4] A. Hubert, R. Schafer, Magnetic Domains: The Analysis of Magnetic

Microstructures, Springer, Heidelberg, 1998.

[5] J. Miltat, G. Albuquerque, A. Thiaville, An introduction to

micromagnetics in the dynamical regime, in: B. Hillebrands,

K. Ounadjela (Eds.), Spin Dynamics In Confined Magnetic

Structures I, Springer, Berlin, Heidelberg, 2002.

[6] J.C. Slonczewski, J. Magn. Magn. Mater. 159 (1996) L1.

[7] L. Berger, Phys. Rev. B 54 (1996) 9353.

[8] J. Miltat, A. Thiaville, Science 298 (2002) 555.

[9] H. Kronmuller, S. Parkin (Eds.), Handbook of Magnetism, vol. 2,

Wiley, 2007.

[10] R. Hertel, O. Fruchart, S. Cherifi, P.-O. Jubert, S. Heun, A. Locatelli,

J. Kirschner, Phys. Rev. B 72 (2005) 214409.

[11] M. Buess, T. Haug, M.R. Scheinfein, C.H. Back, Phys. Rev. Lett. 94

(2005) 127205.

[12] V. Novosad, F.Y. Fradin, P.E. Roy, K.S. Buchanan, K. Yu.

Guslienko, S.D. Bader, Phys. Rev. B 72 (2005) 024455.

[13] B. Van Waeyenberge, A. Puzic, H. Stoll, K.W. Chou, T. Tyliszczak,

R. Hertel, M. Fahnle, H. Bruckl, K. Rott, G. Reiss, I. Neudecker,

D. Weiss, C.H. Back, G. Schutz, Nature 444 (2006) 461.

[14] K. Yamada, et al., Nat. Mater. 6 (2007) 269.

[15] J.P. Park, P. Eames, D.M. Engebretson, J. Berezovsky, P.A. Crowell,

Phys. Rev. Lett. 89 (2002) 277201.

[16] I. Neudecker, M. Klaui, K. Perzlmaier, D. Backes, L.J. Heydermann,

C.A.F. Vaz, J.A.C. Bland, U. Rudiger, C.H. Back, Phys. Rev. Lett.

96 (2006) 057207.

[17] H.W. Schumacher, C. Chappert, P. Crozat, R.C. Sousa, P.P. Freitas,

J. Miltat, J. Fassbender, B. Hillebrands, Phys. Rev. Lett. 90 (2003)

017201.

[18] H.W. Schumacher, C. Chappert, R.C. Sousa, P.P. Freitas, J. Miltat,

Phys. Rev. Lett. 90 (2003) 017204.

[19] N. Smith, cond-mat arXiv:0706.1736.

[20] J.C. Slonczewski, J. Magn. Magn. Mater. 195 (1999) L261.

[21] X. Waintal, E.B. Myers, P.W. Brouwer, D.C. Ralph, Phys. Rev. B 62

(2000) 12317.

[22] M.D. Stiles, A. Zangwill, Phys. Rev. B 66 (2002) 014407.

[23] M.D. Stiles, J. Miltat, Spin transfer torque and dynamics, in: B.

Hillebrands, A. Thiaville (Eds.), Spin Dynamics in Confined

Magnetic Structures III, Springer, Berlin, Heidelberg, 2006.

[24] T. Valet, A. Fert, Phys. Rev. B 48 (1993) 7099.

[25] J.C. Slonczewski, J. Magn. Magn. Mater. 247 (2002) 324.

[26] J. Xiao, A. Zangwill, M.D. Stiles, Eur. Phys. J. B (2007), doi:10.1140/

epjb/e2007-00004-0.

[27] M.D. Stiles, W.M. Saslow, M.J. Donahue, A. Zangwill, arXiv:cond-

mat/0702020.

[28] Y. Tserkovnyak, A. Brataas, G.E.W. Bauer, B.I. Halperin, Rev.

Mod. Phys. 77 (2005) 1375.

[29] M.D. Stiles, J. Xiao, A. Zangwill, Phys. Rev. B 69 (2004) 054408.

[30] W.F. Brown, Phys. Rev. 130 (1963) 1677.

[31] D.V. Berkov, Magnetization dynamics including thermal fluctua-

tions, in: H. Kronmuller, S. Parkin (Eds.), Handbook of Magnetism

and Advanced Magnetic Materials, vol. 2, Wiley, Chichester, UK,

2007, pp. 795–823.

[32] A. Brataas, G.E.W. Bauer, P. Kelly, Phys. Rep. 427 (2006) 157.

[33] S. Zhang, Z. Li, Phys. Rev. Lett. 93 (2004) 127204.

[34] A. Thiaville, Y. Nakatani, J. Miltat, Y. Suzuki, Europhys. Lett. 69

(2005) 990.

[35] J.C. Sankey, I.N. Krivorotov, S.I. Kiselev, P.M. Braganca, N.C.

Emley, R.A. Buhrman, D.C. Ralph, Phys. Rev. B 72 (2005)

224427.

[36] J. Xiao, A. Zangwill, M.D. Stiles, Phys. Rev. B 72 (2005) 014446.

Page 22: Current Perspectives Spin-torque driven magnetization dynamics: Micromagnetic modelinglayer.uci.agh.edu.pl/~mczapkie/Research/symul/articles/e... · 2013. 5. 2. · steady spin-polarized

ARTICLE IN PRESSD.V. Berkov, J. Miltat / Journal of Magnetism and Magnetic Materials 320 (2008) 1238–1259 1259

[37] I.N. Krivorotov, N.C. Emley, A.G.F. Garcia, J.C. Sankey, S.I.

Kiselev, D.C. Ralph, R.A. Buhrman, Phys. Rev. Lett. 93 (2004)

166603.

[38] T. Devolder, P. Crozat, C. Chappert, J. Miltat, A. Tulapurkar,

Y. Suzuki, K. Yagami, Phys. Rev. B 71 (2005) 184401.

[39] B. Montigny, Tranfert de Spin et Dynamique de l’Aimantation,

Ph.D. Thesis, Orsay, 2007, /http://tel.archives-ouvertes.frS.

[40] Y. Acremann, J.P. Strachan, V. Chembrolu, S.D. Andrews,

T. Tyliszczak, J.A. Katine, M.J. Carey, B.M. Clemens, H.C.

Siegmann, J. Stohr, Phys. Rev. Lett. 96 (2006) 217202.

[41] J. Miltat, G. Albuquerque, A. Thiaville, C. Vouille, J. Appl. Phys. 89

(2001) 6982.

[42] D.V. Berkov, N.L. Gorn, Phys. Rev. B 71 (2005) 052403.

[43] W.H. Rippard, M.R. Pufall, S. Kaka, S.E. Russek, T.J. Silva, Phys.

Rev. Lett. 92 (2004) 027201.

[44] W.H. Rippard, M.R. Pufall, S. Kaka, T.J. Silva, S.E. Russek, Phys.

Rev. B 70 (2004) 100406(R).

[45] F.B. Mancoff, N.D. Rizzo, B.N. Engel, S. Tehrani, Nature 437 (2005)

393.

[46] S. Kaka, M.R. Pufall1, W.H. Rippard, T.J. Silva, S.E. Russek,

J.A. Katine, Nature 437 (2005) 389.

[47] A.N. Slavin, P. Kabos, IEEE Trans. Magn. MAG-41 (2005) 1264.

[48] D.V. Berkov, N.L. Gorn, cond-mat arxiv:0705.1515.

[49] S.I. Kiselev, J.C. Sankey, I.N. Krivorotov, N.C. Emley,

R.J. Schoelkopf, R.A. Buhrman, D.C. Ralph, Nature 425 (2003) 380.

[50] D.V. Berkov, N.L. Gorn, Phys. Rev. B 72 (2005) 094401.

[51] I.N. Krivorotov, N.C. Emley, J.C. Sankey, S.I. Kiselev, D.C. Ralph,

R.A. Buhrman, Science 307 (2005) 228.

[52] K.J. Lee, A. Deac, O. Redon, J.P. Nozieres, B. Dieny, Nat. Mater. 3

(2004) 877.

[53] B. Montigny, J. Miltat, J. Appl. Phys. 97 (2005) 10C708.

[54] V. Tiberkevich, A. Slavin, Phys. Rev. B 75 (2007) 014440.

[55] I.N. Krivorotov, D.V. Berkov, N.L. Gorn, N.C. Emley, J.C. Sankey,

D.C. Ralph, R.A. Buhrman, Phys. Rev. B 76 (2007) 024418.

[56] Q. Mistral, J.-V. Kim, T. Devolder, P. Crozat, C. Chappert,

J.A. Katine, M.J. Carey, K. Ito, Appl. Phys. Lett. 88 (2006) 192507.

[57] G.J. Parker, C. Cerjan, D.W. Hewett, J. Magn. Magn. Mater. 214

(2000) 130.

[58] C.J. Garcıa-Cervera, Z. Gimbutas, E. Weinan, J. Comput. Phys. 184

(2003) 37.

[59] M.J. Donahue, R.D. McMichael, IEEE Trans. Magn. 43 (2007)

2878.

[60] J. Miltat, B. Montigny, in preparation.

[61] D.V. Berkov, J. Magn. Magn. Mater. 300 (2006) 159.

[62] D.V. Berkov, N.L. Gorn, J. Appl. Phys. 99 (2006) 08Q701.

[63] G. Consolo, B. Azzerboni, G. Gerhart, G.A. Melkov, V. Tiberkevich,

A.N. Slavin, Phys. Rev. B 76 (2007) 144410.

[64] W.H. Rippard, M.R. Pufall, S.E. Russek, Phys. Rev. B 74 (2006)

224409.

[65] M.R. Pufall, W.H. Rippard, M.L. Schneider, S.E. Russek, Phys. Rev.

B 75 (2007) 140404.

[66] D.V. Berkov, N.L. Gorn, Talk 053-HP (invited) on the Joint

European Magnetic Symposium, Dresden Germany, September

2004;

D.V. Berkov, N.L. Gorn, Talk 28TL-D-6 (invited) on the Moscow

International Symposium on Magnetism, June 2005, Moscow, Russia.

[67] A. Slavin, V. Tiberkevich, Phys. Rev. Lett. 95 (2005) 237201.

[68] M.R. Pufall, W.H. Rippard, M.L. Schneider, S.E. Russek, Phys. Rev.

B 75 (2007) 140404(R).

[69] C. Bayer, J. Jorzick, B. Hillebrands, S.O. Demokritov, R. Kouba,

R. Bozinoski, A.N. Slavin, K. Guslienko, D.V. Berkov, N.L. Gorn,

M.P. Kostylev, Phys. Rev. B 72 (2005) 064427.


Recommended