24.02.2013 Views

Read Back Signals in Magnetic Recording - Research Group Fidler

Read Back Signals in Magnetic Recording - Research Group Fidler

Read Back Signals in Magnetic Recording - Research Group Fidler

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

DIPLOMARBEIT<br />

<strong>Read</strong> <strong>Back</strong> <strong>Signals</strong> <strong>in</strong><br />

<strong>Magnetic</strong> Record<strong>in</strong>g<br />

ausgeführt am<br />

Institut für Festkörperphysik<br />

der Technischen Universität Wien<br />

unter der Anleitung von<br />

Univ.Doz. Dipl.Ing. Dr.techn. Thomas Schrefl<br />

durch<br />

Otmar Ertl<br />

Re<strong>in</strong>prechtsdorferstraße 20/16<br />

1050 Wien<br />

Mat. Nr. 0026810<br />

Wien, am 15. Juli 2005 ________________


Kurzfassung<br />

Diese Arbeit widmet sich der Berechnung des <strong>Signals</strong> von Leseköpfen, wie sie <strong>in</strong><br />

herkömmlichen Festplatten Verwendung f<strong>in</strong>den. Dabei wird zuerst auf analytische Methoden<br />

zurückgegriffen, die es erlauben, den Verlauf des <strong>Signals</strong> für e<strong>in</strong> vere<strong>in</strong>fachtes 2-<br />

dimensionales Modell zu ermitteln. Insbesondere wird das Reziprozitätstheorem verwendet.<br />

Der Großteil dieser Arbeit beschäftigt sich dann mit der Simulation des Lesevorgangs mittels<br />

der Methode der f<strong>in</strong>iten Elemente. Dabei wird die Landau-Lifshitz-Gilbert Gleichung gelöst,<br />

die das dynamische mikromagnetische Verhalten e<strong>in</strong>es magnetischen Modells beschreibt. Um<br />

die Auswirkungen des Sensorstroms auf den Lesekopf zu berücksichtigen, muss die<br />

Stromverteilung für <strong>in</strong>homogene Leiter berechnet werden. Da es sich bei e<strong>in</strong>em modernen<br />

Lesekopf um e<strong>in</strong>en magnetoresistiven Sensor handelt, hängt die Leitfähigkeit von der<br />

Magnetisierung des Lesekopfes ab. Die berechnete Stromverteilung erzeugt wiederum e<strong>in</strong><br />

Magnetfeld, das auf die Magnetisierung des Lesekopfs e<strong>in</strong>wirkt. Die programmtechnische<br />

Umsetzung und die Realisierung solcher Simulationen von Leseköpfen waren die primären<br />

Ziele dieser Diplomarbeit.<br />

Im letzten Teil dieser Arbeit wird anhand e<strong>in</strong>es 3-dimensionalen Modells e<strong>in</strong>es Lesekopfes<br />

die Funktionalität des Computerprogramms erprobt. Es werden der Gleichgewichtszustand,<br />

die Sensorkennl<strong>in</strong>ie und das Relaxationsverhalten berechnet. Weiters wird der E<strong>in</strong>fluss der<br />

magnetischen Schilde des Lesekopfes untersucht. Zum Schluss wird das Lesesignal von<br />

perfekten magnetischen Übergängen beziehungsweise von periodischen Bitmustern bestimmt.<br />

Mit dem im Zuge der Diplomarbeit erstellten Computerprogramm ist e<strong>in</strong> Werkzeug<br />

geschaffen worden, das die komplette Simulation des Lesevorganges, und auch die<br />

Entwicklung von Leseköpfen am Computer erlaubt. Dabei können die dynamischen und<br />

statischen Eigenschaften von Leseköpfen bestimmt werden.<br />

2


Abstract<br />

This thesis concentrates on calculat<strong>in</strong>g the read back signals of read heads, as used <strong>in</strong><br />

conventional hard disks. First the read back signal of a simplified 2-dimensional model is<br />

determ<strong>in</strong>ed by analytical means. Here the reciprocity pr<strong>in</strong>ciple is used.<br />

The ma<strong>in</strong> part of this work deals with the simulation of the read back process with the F<strong>in</strong>ite<br />

Element Method. The Landau-Lifshitz-Gilbert equation is solved, which describes the<br />

dynamic micromagnetic behavior of a magnetic model. To take effects of the sense current of<br />

the read head <strong>in</strong>to account, the current distribution for <strong>in</strong>homogeneous conductors has to be<br />

evaluated. Modern read heads use magnetoresistive effects. Thus the conductivities depend on<br />

the magnetization of the read head. The calculated current distribution generates a magnetic<br />

field, act<strong>in</strong>g on the magnetization of the read head itself. The programm<strong>in</strong>g and the<br />

implementation to make such simulations possible were the primary targets of this work.<br />

In the last part of this thesis the functionality of the program is tested with a simple 3dimensional<br />

model of a read head. The equilibrium state, the sensor curve, and the relaxation<br />

behavior are calculated. Moreover the <strong>in</strong>fluence of the magnetic shields is exam<strong>in</strong>ed. At last<br />

the read back signal is calculated for perfect transitions and for periodic bit patterns.<br />

This work resulted <strong>in</strong> a tool for simulat<strong>in</strong>g the whole read back process, which also allows the<br />

computer-based development of read heads. Furthermore the dynamic and static<br />

micromagnetic properties of read heads can be calculated.<br />

3


Contents<br />

Contents<br />

KURZFASSUNG................................................................................................................................................... 2<br />

ABSTRACT........................................................................................................................................................... 3<br />

CONTENTS........................................................................................................................................................... 4<br />

1 INTRODUCTION ....................................................................................................................................... 7<br />

2 BASICS....................................................................................................................................................... 11<br />

2.1 OHM’S LAW........................................................................................................................................ 11<br />

2.2 MAXWELL’S EQUATIONS .................................................................................................................... 11<br />

2.3 BASIC EQUATIONS OF MAGNETOSTATICS ........................................................................................... 12<br />

2.4 INHOMOGENEOUS CONDUCTORS......................................................................................................... 14<br />

2.5 GIBBS’ FREE ENERGY ......................................................................................................................... 15<br />

2.5.1 Magnetostatic Energy.................................................................................................................... 16<br />

2.5.2 Stray Field Energy......................................................................................................................... 16<br />

2.5.3 Exchange Energy........................................................................................................................... 16<br />

2.5.4 Magnetocrystall<strong>in</strong>e Anisotropy Energy ......................................................................................... 17<br />

2.6 EFFECTIVE FIELD ................................................................................................................................ 18<br />

2.7 LANDAU-LIFSHITZ-GILBERT EQUATION ............................................................................................. 18<br />

2.8 MAGNETORESISTANCE........................................................................................................................ 20<br />

2.8.1 Magnetoresistance of Ferromagnets............................................................................................. 20<br />

2.8.2 Anisotropic Magnetoresistance Effect........................................................................................... 21<br />

2.8.3 Giant Magnetoresistance Effect .................................................................................................... 22<br />

2.8.4 Tunnel<strong>in</strong>g Magnetoresistive Effect ................................................................................................ 26<br />

2.9 RECORDING SCHEMES......................................................................................................................... 26<br />

2.9.1 Longitud<strong>in</strong>al Record<strong>in</strong>g................................................................................................................. 27<br />

2.9.2 Perpendicular Record<strong>in</strong>g .............................................................................................................. 27<br />

3 ANALYTICAL CALCULATIONS ......................................................................................................... 29<br />

3.1 RECIPROCITY PRINCIPLE..................................................................................................................... 29<br />

3.2 LONGITUDINAL RECORDING ............................................................................................................... 29<br />

3.2.1 Head Field..................................................................................................................................... 30<br />

3.2.2 Signal Field ................................................................................................................................... 34<br />

3.2.3 Signal of a Perfect Transition........................................................................................................ 34<br />

4


Contents<br />

3.2.4 Signal of Written Bits..................................................................................................................... 35<br />

3.3 PERPENDICULAR RECORDING ............................................................................................................. 36<br />

3.4 SIGNAL FIELD IN GAP ......................................................................................................................... 38<br />

4 NUMERICAL METHODS....................................................................................................................... 41<br />

4.1 FINITE ELEMENT METHOD.................................................................................................................. 42<br />

4.1.1 Dirichlet Boundary Conditions ..................................................................................................... 44<br />

4.1.2 Neumann Boundary Conditions..................................................................................................... 44<br />

4.1.3 L<strong>in</strong>ear equation solver................................................................................................................... 45<br />

4.2 STRAY FIELD CALCULATION............................................................................................................... 46<br />

4.3 ADAPTIVE CROSS-APPROXIMATION TECHNIQUE ................................................................................ 48<br />

4.4 CONDUCTOR MODEL .......................................................................................................................... 49<br />

4.4.1 Conductivity .................................................................................................................................. 49<br />

4.4.2 Electric Potential........................................................................................................................... 50<br />

4.4.3 Current Distribution...................................................................................................................... 50<br />

4.4.4 <strong>Magnetic</strong> Field of Current............................................................................................................. 51<br />

4.4.5 Hybrid FEM/BEM for Current Field............................................................................................. 52<br />

4.4.6 Example: Long Wire...................................................................................................................... 54<br />

4.4.7 Example: Coil................................................................................................................................ 55<br />

4.5 TIME INTEGRATION ............................................................................................................................. 57<br />

5 READ HEAD DESIGN ............................................................................................................................. 59<br />

5.1 BIAS SCHEMES.................................................................................................................................... 59<br />

5.1.1 Demagnetiz<strong>in</strong>g Factor................................................................................................................... 59<br />

5.1.2 Hard Bias ...................................................................................................................................... 60<br />

5.1.3 Exchange Bias ............................................................................................................................... 60<br />

5.1.4 Synthetic Antiferromagnet............................................................................................................. 61<br />

5.1.5 Sense Current ................................................................................................................................ 62<br />

5.1.6 Crystall<strong>in</strong>e Anisotropy................................................................................................................... 64<br />

5.2 SHIELDING .......................................................................................................................................... 64<br />

5.3 THERMAL STABILITY .......................................................................................................................... 64<br />

6 FEM SIMULATIONS............................................................................................................................... 65<br />

6.1 MODEL................................................................................................................................................ 65<br />

6.1.1 <strong>Magnetic</strong> Model............................................................................................................................. 65<br />

6.1.2 Conductor Model........................................................................................................................... 67<br />

6.2 RESULTS ............................................................................................................................................. 69<br />

6.2.1 Equilibrium State........................................................................................................................... 69<br />

6.2.2 Transfer Curve .............................................................................................................................. 70<br />

6.2.3 Influence of Shields ....................................................................................................................... 72<br />

5


Contents<br />

6.2.4 Relaxation...................................................................................................................................... 73<br />

6.2.5 Signal of a Perfect Transition........................................................................................................ 74<br />

6.2.6 <strong>Read</strong> <strong>Back</strong> Signal........................................................................................................................... 75<br />

7 OUTLOOK................................................................................................................................................. 77<br />

APPENDIX A ...................................................................................................................................................... 78<br />

APPENDIX B ...................................................................................................................................................... 80<br />

LIST OF FIGURES ............................................................................................................................................ 83<br />

BIBLIOGRAPHY ............................................................................................................................................... 86<br />

DANKSAGUNG.................................................................................................................................................. 90<br />

LEBENSLAUF .................................................................................................................................................... 91<br />

6


Introduction<br />

1 IntroductionEquation Section 1<br />

The idea to use magnetism for record<strong>in</strong>g orig<strong>in</strong>ates from Oberl<strong>in</strong> Smith (1840-1926), an<br />

<strong>in</strong>ventor, <strong>in</strong>dustrialist, and mechanical eng<strong>in</strong>eer. His idea based on a visit to Thomas Edison’s<br />

laboratory <strong>in</strong> 1878. Lack<strong>in</strong>g time to pursue further development - it is not known whether he<br />

built a work<strong>in</strong>g device - he published a description of magnetic record<strong>in</strong>g <strong>in</strong> Electrical World,<br />

Sep. 8, 1888 [1]. Figure 1.1 shows the orig<strong>in</strong>al draw<strong>in</strong>g of Oberl<strong>in</strong> Smith. It demonstrates the<br />

basic pr<strong>in</strong>ciple of magnetic audio record<strong>in</strong>g used up to now. The audio waves are transformed<br />

by the microphone (A) <strong>in</strong>to an electrical signal and are recorded <strong>in</strong> the form of magnetization<br />

patterns on a str<strong>in</strong>g covered with iron fil<strong>in</strong>gs (C), pass<strong>in</strong>g through an electromagnet (B). To<br />

reduce noise a battery (F) supplies bias current. The str<strong>in</strong>g is moved from the supply reel (D)<br />

to the take up reel (E).<br />

Probably <strong>in</strong>spired by Oberl<strong>in</strong> Smith’s article Valdemar Poulson built the first work<strong>in</strong>g<br />

magnetic record<strong>in</strong>g device (see Figure 1.1). He used an <strong>in</strong>ductive head mov<strong>in</strong>g over a wire<br />

wrapped drum. In the follow<strong>in</strong>g decades a lot of improvements were <strong>in</strong>vented, but as<br />

<strong>in</strong>formation carrier ma<strong>in</strong>ly wire wrapped drums were used. Later tapes coated with magnetic<br />

powder replaced them.<br />

Figure 1.1: The orig<strong>in</strong>al draw<strong>in</strong>g of Oberl<strong>in</strong> Smith (1888) and Valdemar Poulson’s<br />

telegraphone [2].<br />

7


Introduction<br />

In the late forties and the early fifties the first comput<strong>in</strong>g mach<strong>in</strong>es where developed. To store<br />

program and data <strong>in</strong>formation punched paper cards or punched paper tapes where used. Later<br />

magnetic tape record<strong>in</strong>g was used, which had a big disadvantage. The random data request of<br />

tapes was very time consum<strong>in</strong>g, because the non-sequential access required long rew<strong>in</strong>d and<br />

fast-forward times. That is why there was the demand for a fast random access storage<br />

system.<br />

In the late 1940s a new computer design program started at the University at Berkeley to<br />

develop an <strong>in</strong>termediate sized computer based on a magnetic drum memory. To get rapid data<br />

access, high rotational speeds were used, which generated <strong>in</strong>terest <strong>in</strong> non-contact record<strong>in</strong>g<br />

heads. Among the tested head designs, there was also a head with a wide gap and a drum with<br />

a soft magnetic layer beneath the usual magnetic layer, the so called Soft-Under-Layer (SUL).<br />

The result was that the magnetite dipoles were oriented perpendicular to the surface and not<br />

along the surface circumferentially [3] (1949).<br />

In 1956 IBM announced the first commercial magnetic disk drive, the Random Access<br />

Method of Account<strong>in</strong>g and Control (RAMAC) 350. This drive had a storage capacity of 5 MB<br />

and a data density of 2 kbit/<strong>in</strong> 2 . 50 24-<strong>in</strong> alum<strong>in</strong>um disks coated with iron oxide pa<strong>in</strong>t acted as<br />

media. The whole stack of disks had a weight of 250 kg. Air pressure held the distance of<br />

25 µm between head and disk. The complex air pressure mechanism was very expensive. That<br />

is why only one actuator-arm assembly with a head pair - one for the upper surface and one<br />

for the lower surface of a disk - was used (see Figure 1.2). Nevertheless the access time was<br />

less than a second.<br />

Figure 1.2: The first computer disk drive RAMAC. The pneumatically driven actuator arm<br />

and some of the 50 disks are visible [4].<br />

8


Introduction<br />

In 1955, before the RAMAC was announced, IBM decided to start the development of the<br />

next generation hard disks. The targets of the so called Advanced Disk File (ADF) project<br />

were a 10 times larger capacity and a 1/10 the access time compared to RAMAC. The faster<br />

access time could only be reached by us<strong>in</strong>g one head for each disk surface. It was clear that<br />

the uneconomic air pressure system had to be replaced by fly<strong>in</strong>g heads (air bear<strong>in</strong>g sliders).<br />

Fly<strong>in</strong>g heads allow a smaller head-disk spac<strong>in</strong>g, which f<strong>in</strong>ally allows higher l<strong>in</strong>ear data<br />

densities. The experiences of the RAMAC lead to the decision to use a harder media. They<br />

chose soft magnetic steel disks. Oxidiz<strong>in</strong>g <strong>in</strong> a steam atmosphere leads to a hard th<strong>in</strong> layer on<br />

their surface. In addition the manufactur<strong>in</strong>g costs were less than that of RAMAC’s alum<strong>in</strong>um<br />

disks.<br />

The project was very risky, because of lack<strong>in</strong>g knowledge about air bear<strong>in</strong>g heads and about<br />

perpendicular record<strong>in</strong>g. In 1960, due to several failures and major problems (e.g. the surface<br />

quality of the steel disks was unacceptable), IBM turned away from perpendicular record<strong>in</strong>g<br />

and towards longitud<strong>in</strong>al record<strong>in</strong>g. Due to major achievements <strong>in</strong> longitud<strong>in</strong>al record<strong>in</strong>g<br />

perpendicular record<strong>in</strong>g became un<strong>in</strong>terest<strong>in</strong>g for many decades.<br />

All magnetic storage devices were us<strong>in</strong>g coils to read out magnetic <strong>in</strong>formation. In 1971<br />

however, a new type of transducer us<strong>in</strong>g the anisotropy magneto-resistance effect (AMR) was<br />

reported by R. B. Hunt [5]. AMR heads generate larger read back signals than <strong>in</strong>ductive heads<br />

for common relative velocities between disk and head. Another advantage of AMR heads<br />

over <strong>in</strong>ductive heads is their lower resistance and <strong>in</strong>ductivity result<strong>in</strong>g <strong>in</strong> lower Johnson noise.<br />

In 1991 the first commercial hard disks us<strong>in</strong>g AMR read heads were available and broke the<br />

100 Mbit/<strong>in</strong> 2 record<strong>in</strong>g density barrier.<br />

In 1988 Baibich et al. [6] discovered a new magneto-resistance effect. Due to its large change<br />

<strong>in</strong> resistance compared to AMR this effect was called Giant Magneto-Resistance (GMR).<br />

GMR read heads were first <strong>in</strong>troduced <strong>in</strong> commercial hard disks by IBM <strong>in</strong> 1997. The<br />

<strong>in</strong>vention of AMR and GMR read heads has accelerated the growth <strong>in</strong> data density (see Figure<br />

1.3).<br />

9


Introduction<br />

Figure 1.3: The development of bit sizes and storage densities over the last years [7].<br />

Before the <strong>in</strong>troduction of magnetoresistive (MR) sensors <strong>in</strong> read heads the density doubled<br />

every three years. When AMR read heads reached the market the doubl<strong>in</strong>g time was reduced<br />

to two years. And when the GMR head was <strong>in</strong>troduced, the density has been doubled each<br />

year from 1998 to 2001. In the last years the growth rate decreased to 25-40%.<br />

It is clear that this exponential growth cannot last forever. The major limitation, which<br />

worries disk-drive builders most, is the so called superparamagnetic limit.<br />

Superparamagnetism occurs for very small bits. Then the energy barrier is very small, so that<br />

the magnetization can switch due to thermal fluctuations. The limit actually depends on the<br />

media.<br />

There are several different attempts to <strong>in</strong>crease these limit with special media. One possibility<br />

is to use antiferromagnetically coupled media [8] [9]. So the media exists of a second<br />

ferromagnetic layer, which is antiferromagnetically coupled with the actual data layer. Both<br />

layers are always magnetized <strong>in</strong> opposite directions, lead<strong>in</strong>g to a better thermal stability.<br />

Other approaches are the re<strong>in</strong>troduction of perpendicular record<strong>in</strong>g, the usage of patterned<br />

media [10], where the bit cells are predef<strong>in</strong>ed as isolated magnetic doma<strong>in</strong>s <strong>in</strong> a nonmagnetic<br />

matrix, or thermally or optically assisted magnetic record<strong>in</strong>g [11]. At last to reach a maximum<br />

data density there are experiments adapt<strong>in</strong>g an atomic force microscope to store <strong>in</strong>formation<br />

at the scale of <strong>in</strong>dividual atoms [12].<br />

10


2 BasicsEquation<br />

Section (Next)<br />

2.1 Ohm’s Law<br />

Basics<br />

Georg Simon Ohm (1789-1854) found that the current I through most materials is<br />

proportional to the applied voltage V.<br />

V = R⋅ I<br />

(2.1)<br />

The proportional coefficient is called resistance R. The more general form of this equation is<br />

shown below.<br />

σ⋅ E = j (2.2)<br />

E [V/m] denotes the electric field, and j [A/m 2 ] the current density. σ [A/Vm] is a property<br />

of matter named conductivity. Generally the conductivity is a tensor, but here all conductors<br />

will be assumed to be isotropic.<br />

2.2 Maxwell’s Equations<br />

James Clerk Maxwell (1831-1879) was the first who jo<strong>in</strong>ed the theories of electricity and<br />

magnetism. He proposed follow<strong>in</strong>g four equations to describe both theories (<strong>in</strong> SI).<br />

� ∫ DdA= ∫ρdV<br />

(2.3)<br />

∂V<br />

V<br />

� ∫ BdA= 0<br />

(2.4)<br />

∂V<br />

∂<br />

� ∫Eds=− d<br />

∂t<br />

∫B<br />

A<br />

(2.5)<br />

∂A<br />

A<br />

∂<br />

� ∫ Hds= ∫jdA+ d<br />

∂t<br />

∫D<br />

A<br />

(2.6)<br />

∂A<br />

A A<br />

These four <strong>in</strong>tegral equations can also be written as equivalent differential equations.<br />

11


Basics<br />

div D =ρ<br />

(2.7)<br />

div B = 0<br />

(2.8)<br />

∂B<br />

curlE<br />

=− (2.9)<br />

∂t<br />

∂D<br />

curl H = j + (2.10)<br />

∂t<br />

D [C/m 2 ] is the electric displacement, B [T] is the magnetic <strong>in</strong>duction, E [V/m] is the electric<br />

field, H [A/m] is the magnetic field and j [A/m 2 ] denotes the current density. The third<br />

equation (2.9) is the Law of Induction found by Michael Faraday (1791-1867), who first<br />

recognized that an alternat<strong>in</strong>g magnetic field produces a current. The last equation (2.10) is<br />

the generalization of Ampere’s Law:<br />

� ∫ Hds= ∫jdA.<br />

(2.11)<br />

∂A<br />

A<br />

The magnetic <strong>in</strong>duction B and the magnetic field H are related by<br />

( )<br />

B=μ H+ M =μ H+ J, (2.12)<br />

0 0<br />

and the displacement D and the electric field E by<br />

Here<br />

D=ε 0E+<br />

P. (2.13)<br />

μ = π× is the permeability of vacuum, and<br />

−7<br />

0 4 10 Vs/Am<br />

ε = × is<br />

−12<br />

0 8.8542 10 As/Vm<br />

the absolute dielectric constant of vacuum. M [A/m] denotes the total magnetization, J [T] the<br />

magnetic polarization, and P the electric polarization [C/m 2 ].<br />

Tak<strong>in</strong>g the divergence of both sides <strong>in</strong> (2.10) and us<strong>in</strong>g (2.7) lead to the cont<strong>in</strong>uum equation<br />

∂ρ<br />

div j + = 0 . (2.14)<br />

∂t<br />

2.3 Basic Equations of Magnetostatics<br />

In case of vanish<strong>in</strong>g field E = 0 , non-dielectric materials ( P = 0 ), and vanish<strong>in</strong>g current<br />

density j = 0 , (2.10) simplifies to<br />

12


Basics<br />

curl H = 0 . (2.15)<br />

So H is a irrotational field and can therefore be expressed as gradient of a scalar field Ψ ,<br />

called magnetic potential.<br />

H =−grad Ψ<br />

(2.16)<br />

The magnetic potential of a volume V with a given magnetic polarization J can be calculated<br />

by solv<strong>in</strong>g the Poisson equation<br />

ρm<br />

ΔΨ = − μ<br />

with the jump condition<br />

0<br />

out <strong>in</strong><br />

∂Ψ ∂Ψ σ<br />

− =−<br />

∂n ∂n μ<br />

∂V<br />

m<br />

0<br />

(2.17)<br />

(2.18)<br />

at the boundary ∂ V . Here ρ m and σ m are magnetic charge densities <strong>in</strong> the volume V and at<br />

its boundary ∂ V . These charges are only virtual, because there are no magnetic monopoles <strong>in</strong><br />

reality. These virtual charges are useful to calculate the magnetic potential of a polarization<br />

distribution. They can be obta<strong>in</strong>ed by tak<strong>in</strong>g the divergence and the face divergence of the<br />

magnetic polarization J.<br />

ρ =−div J, σ =−Div<br />

J (2.19)<br />

m m<br />

The face divergence is def<strong>in</strong>ed as difference of the normal components of outer and <strong>in</strong>ner<br />

field.<br />

( )<br />

out <strong>in</strong> out <strong>in</strong><br />

Div J = J −J ⋅ n = J −J<br />

(2.20)<br />

n n<br />

Assum<strong>in</strong>g that the magnetic potential is regular at <strong>in</strong>f<strong>in</strong>ity,<br />

1<br />

Ψ→ for r →∞, (2.21)<br />

r<br />

the solution of (2.17) and (2.18) is<br />

1 ⎛ ρ σ ⎞<br />

Ψ= dV + dA<br />

4<br />

∫ ∫� . (2.22)<br />

m m<br />

⎜ ⎟<br />

πμ0 ⎝ r r<br />

V ∂V<br />

⎠<br />

13


Basics<br />

2.4 Inhomogeneous Conductors<br />

To calculate the read back signal of magnetoresistive heads, it is necessary to calculate the<br />

current distribution, because the conductivity does not only depend on locus, but also on the<br />

magnetization:<br />

σ=σrMr (, (,)) t . (2.23)<br />

Consequently a general treatment is necessary to calculate the electric potential and the<br />

current density. The electrical field can be splitted <strong>in</strong>to two parts, an irrotational and a<br />

solenoidal part:<br />

with<br />

and<br />

E= Efree + E eddy<br />

(2.24)<br />

curlE 0<br />

(2.25)<br />

free =<br />

divE 0 . (2.26)<br />

eddy =<br />

It can be shown, that the irrotational part of the electrical field, denoted with E free , is caused<br />

by external applied voltage or current. The solenoidal part denoted by E eddy is due to magnetic<br />

<strong>in</strong>duction described by Faraday’s law (2.9) and orig<strong>in</strong>ates from eddy currents. For the moment<br />

we can neglect eddy currents. Later they can be partially taken <strong>in</strong>to account by choos<strong>in</strong>g<br />

higher damp<strong>in</strong>g constants <strong>in</strong> the LLG equation (see Section 2.7) [41] [42]. This implies, that<br />

the total electrical field is irrotational<br />

curlE = 0<br />

(2.27)<br />

and can be written as gradient of a scalar potential function Φ , also called electric potential<br />

E =−grad Φ.<br />

(2.28)<br />

Furthermore the whole problem is considered to be quasi-static. In other words the<br />

displacement current term <strong>in</strong> (2.10) can be neglected<br />

curl H = j. (2.29)<br />

14


Basics<br />

This is justified, because the sense current through MR sensors is normally held constant and<br />

the current density is quite large for the small dimensions of a MR read<strong>in</strong>g head. (The current<br />

density is only limited due to the heat dissipation. Typical values for the critical current<br />

density are 10 7 -10 8 A/cm 2 .) So local charge concentrations are immediately compensated.<br />

Apply<strong>in</strong>g the divergence operator on both sides of equation (2.29) gives<br />

div j = 0 . (2.30)<br />

Thus the source term of the cont<strong>in</strong>uum equation (2.14) vanishes. Comb<strong>in</strong><strong>in</strong>g (2.2), (2.28), and<br />

(2.30) leads to a generalized form of Laplace equation<br />

div( σgrad Φ ) = 0.<br />

(2.31)<br />

If the electric potential is known, it is possible to evaluate the current density us<strong>in</strong>g<br />

j =−σgrad Φ.<br />

(2.32)<br />

2.5 Gibbs’ Free Energy<br />

The equilibrium state of an isothermal and isobaric thermodynamic system is characterized by<br />

the m<strong>in</strong>imum of its Gibbs’ free energy E.<br />

∫<br />

E = F − J⋅H dV<br />

(2.33)<br />

ext<br />

F = U −T⋅ S+ p⋅ V is the free energy, where U is the <strong>in</strong>ner energy, T the absolute<br />

temperature, S the entropy, and V the volume of the magnetic system. For constant<br />

temperatures and neglected magnetostrictive effects, which means keep<strong>in</strong>g the volume and<br />

the temperature constant, the free energy is reduced to U. The <strong>in</strong>ner energy U can be<br />

calculated as the sum of the stray field energy E S , the exchange energy E exch , and the<br />

magnetocrystall<strong>in</strong>e anisotropy energy E ani , so the total Gibbs’ free energy can be written as<br />

E = ES+ Eexch + Eani+ EH.<br />

(2.34)<br />

Here E H is the magnetostatic energy and arises from the external field H ext . The equilibrium<br />

state of such a magnetic system is determ<strong>in</strong>ed by<br />

15


Basics<br />

δ E<br />

= 0 . (2.35)<br />

δJ<br />

2.5.1 Magnetostatic Energy<br />

The energy associated with a magnetized body <strong>in</strong> the presence of an external field H ext is<br />

called magnetic field energy and can be calculated via<br />

∫<br />

E =− J⋅H dV . (2.36)<br />

H ext<br />

V<br />

This energy will cause the magnetization to align parallel to the external field <strong>in</strong> order to<br />

m<strong>in</strong>imize the energy.<br />

2.5.2 Stray Field Energy<br />

The stray field energy term writes similar to the magnetic field energy.<br />

1<br />

ES =− ⋅ S dV<br />

2 ∫ J H (2.37)<br />

V<br />

The additional factor ½ takes <strong>in</strong>to account, that the stray field H S , which is also called<br />

demagnetiz<strong>in</strong>g field, is caused by the magnetization itself. This is <strong>in</strong> contrast to the magnetic<br />

energy where an external field is given, which is not <strong>in</strong>fluenced by the magnetization. The<br />

stray field can be determ<strong>in</strong>ed by calculat<strong>in</strong>g the magnetic scalar potential Ψ S as described <strong>in</strong><br />

Section 2.3. Then the stray field can be calculated us<strong>in</strong>g (2.16):<br />

H =−grad Ψ . (2.38)<br />

S S<br />

2.5.3 Exchange Energy<br />

In the approach of micromagnetics the quantum mechanical sp<strong>in</strong> operators are substituted by<br />

classical magnetization vectors. Thus, the exchange energy <strong>in</strong> the Heisenberg model [13] can<br />

be expressed as<br />

N N<br />

exch =−∑∑ i j<br />

ij ⋅S ⋅S<br />

. (2.39)<br />

i= 1 j≠i E J<br />

16


Basics<br />

Here J ij represents the exchange <strong>in</strong>tegral. In case of ferromagnetism the exchange <strong>in</strong>tegral is<br />

positive and leads to the observed strong coupl<strong>in</strong>g and parallel alignment of sp<strong>in</strong>s.<br />

Consider<strong>in</strong>g only nearest neighbor (n.n.) angular momentums and assum<strong>in</strong>g a constant<br />

exchange <strong>in</strong>tegral Jij � J lead to<br />

N n. n.<br />

i j<br />

Eexch =−J ∑∑S ⋅S<br />

. (2.40)<br />

i= 1 j≠i The assumption of a constant exchange <strong>in</strong>tegral is justified, because experiments show only<br />

small differences when measur<strong>in</strong>g along different crystallographic directions.<br />

As angular moments are described by a cont<strong>in</strong>uous function J = J() r <strong>in</strong> micromagnetics, the<br />

sum can be approached by a volume <strong>in</strong>tegral,<br />

A<br />

E = ( ∇ J ) + ( ∇ J ) + ( ∇J<br />

) dV<br />

2 ∫ . (2.41)<br />

2 2 2<br />

exch<br />

J s V<br />

x y z<br />

Here J s is the spontaneous polarization, and the exchange constant A [J/m] is proportional to<br />

the exchange <strong>in</strong>tegral J. The value of A has been found experimentally for a lot of magnetic<br />

materials.<br />

2.5.4 Magnetocrystall<strong>in</strong>e Anisotropy Energy<br />

In general a ferromagnetic material does not show isotropic magnetization behavior.<br />

Experimentally certa<strong>in</strong> directions can be found, which favor magnetization (easy axes) and<br />

others which do not (hard axes). This leads to a new energy term.<br />

ani ani<br />

V<br />

( )<br />

E = ∫ f J dV<br />

(2.42)<br />

Here f ani denotes a direction dependent function. In case of hexagonal systems (e.g. Co) a<br />

uniaxial anisotropy can be assumed. Therefore f ani can be expressed as series expansion<br />

f ( θ ) = K + K s<strong>in</strong> θ+ K s<strong>in</strong> θ+ ... . (2.43)<br />

ani<br />

2 4<br />

0 1 2<br />

θ denotes the angle between J and the c-axis of the hexagonal system and 0 K , 1 K , 2<br />

are the so called anisotropy constants. In practice all higher terms are neglected.<br />

K , etc.<br />

In case of cubic systems, their cubic symmetry can be taken <strong>in</strong>to account, if f ani is written as<br />

17


Basics<br />

2 2 2 2 2 2 2 2 2<br />

( ) ( )<br />

f ( α , α , α ) = K + K α α +α α +α α + K α α α + ... . (2.44)<br />

ani<br />

1 2 3 0 1 1 2 2 3 3 1 2 1 2 3<br />

α 1 , 2 α , and α 3 denote the direction cos<strong>in</strong>es of the cubic lattice vectors to the direction of the<br />

magnetic polarization J.<br />

In both cases the first term K 0 can be neglected, because it does not depend on angle and is<br />

therefore a constant term.<br />

2.6 Effective Field<br />

The effective magnetic field H eff which acts on the magnetization can be calculated by the<br />

variation of Gibbs’ Free Energy.<br />

δE<br />

2A δfani<br />

( J)<br />

H =− = ΔJ− + H + H + H<br />

eff 2<br />

S ext curr<br />

δJ J s δJ<br />

(2.45)<br />

Here the first term corresponds to the exchange field H exch , and the second one to the<br />

anisotropy field H ani . The additional contribution H curr is the field due to currents. In fact this<br />

field is a contribution to the external field H ext , but here it is written separately, because from<br />

now on ext H is used to describe homogeneous fields, whereas H curr are fields due to currents,<br />

and can be calculated us<strong>in</strong>g the Biot-Savart Law<br />

1 r−r′ Hcurr () r = ×<br />

4π<br />

∫ j<br />

r−r′ V<br />

3<br />

dV ′ , (2.46)<br />

which was empirically found by Jean Baptiste Biot and Félix Savart <strong>in</strong> 1820.<br />

2.7 Landau-Lifshitz-Gilbert Equation<br />

To f<strong>in</strong>d out the equilibrium state of a magnetic system equation (2.35) can be used. To get<br />

<strong>in</strong>formation about the magnetization dynamics, i.e. about the dynamic behavior of the<br />

magnetic polarization J, it is necessary to apply a full 3-dimensional treatment of the time<br />

evolution of the magnetic polarization, which is described by the Gilbert equation of motion<br />

18


s<br />

Basics<br />

∂J α ∂J<br />

=−γ J× Heff + J × . (2.47)<br />

∂t J ∂t<br />

This equation is mathematically equivalent to the Landau-Lifshitz-Gilbert equation (LLG)<br />

∂Jγ γ α<br />

=− J× Heff − J× ( J× H eff ) . (2.48)<br />

∂ t +α J +α<br />

2 2<br />

( 1 ) s ( 1 )<br />

Here J s is the spontaneous polarization and γ is the gyromagnetic ratio<br />

5<br />

γ= 2.210175× 10 m/As . The first term <strong>in</strong> this equation describes the precessional movement<br />

of the magnetic polarization vector J around the effective field H eff . The second<br />

phenomenological term stands for the energy dissipation dur<strong>in</strong>g the movement and causes a<br />

movement towards the direction of the effective field. The motion of the magnetization is<br />

visualized <strong>in</strong> Figure 2.1a. The strength of the damp<strong>in</strong>g is characterized by the dimensionless<br />

Gilbert damp<strong>in</strong>g constant α . Large α means strong damp<strong>in</strong>g. There are many processes<br />

which contribute to the damp<strong>in</strong>g <strong>in</strong> magnetic solids, for example magnon-magnon, magnon-<br />

phonon <strong>in</strong>teractions, or eddy currents. So the damp<strong>in</strong>g constant is an <strong>in</strong>tr<strong>in</strong>sic property of<br />

matter and can also be varied <strong>in</strong> practice, for <strong>in</strong>stance by dop<strong>in</strong>g [14]. In general, except for<br />

th<strong>in</strong> layers (see Section 6.2.4), high damp<strong>in</strong>g constants lead to coherent motion of the<br />

magnetization and therefore to faster reversal processes. Figure 2.1b shows the field of a<br />

simulated write head as function of time for different Gilbert damp<strong>in</strong>g constants. It can be<br />

seen that for damp<strong>in</strong>g constants of about 0.5 the fastest switch<strong>in</strong>g time is achieved [43].<br />

∂ J<br />

In equilibrium, thus if = 0 , equation (2.47) simplifies to<br />

∂t<br />

J× H eff = 0,<br />

(2.49)<br />

which means that <strong>in</strong> equilibrium the magnetic polarization J is aligned either parallel or<br />

antiparallel to the effective field H eff . The unrealistic case of antiparallel alignment is<br />

irrelevant, because this solution is not stable, and <strong>in</strong> this case small deviations lead to a<br />

movement <strong>in</strong> direction of the effective field due to the damp<strong>in</strong>g term <strong>in</strong> (2.48) (see also Figure<br />

2.1a).<br />

19


a)<br />

Heff<br />

J<br />

−J x Jx Heff<br />

−J x Heff<br />

b)<br />

Basics<br />

Figure 2.1: a) Illustration of the damped precession of the magnetic polarization J around<br />

the effective field Heff, taken from [15]. b) Influence of the Gilbert damp<strong>in</strong>g constant on the<br />

field rise time <strong>in</strong> magnetic read heads. Head field strength as a function of time for different<br />

Gilbert damp<strong>in</strong>g constants with<strong>in</strong> the head [43]. The dashed l<strong>in</strong>e depicts the current profile.<br />

The solid l<strong>in</strong>es give the head field for α = 1, 0.5, 0.1, and 0.02. The fastest field rise time is<br />

achieved with <strong>in</strong>termediate damp<strong>in</strong>g (α = 0.5).<br />

2.8 Magnetoresistance<br />

2.8.1 Magnetoresistance of Ferromagnets<br />

A change <strong>in</strong> resistance of a material under an applied field is known as magnetoresistance.<br />

This effect was first discovered <strong>in</strong> 1856 by Lord Kelv<strong>in</strong> who was exam<strong>in</strong><strong>in</strong>g the resistance of<br />

an iron sample. He found a 0.2% <strong>in</strong>crease of resistance when he applied a field <strong>in</strong> longitud<strong>in</strong>al<br />

direction, and a 0.4% decrease for the transversal direction.<br />

Thus, a positive and a negative change of resistance is possible, although <strong>in</strong>tuitively one<br />

expects only positive magnetoresistance, because electrons are forced to take longer paths due<br />

to the magnetic field, which leads to more scatter<strong>in</strong>g. Nevertheless negative<br />

magnetoresistance occurs, which has to be expla<strong>in</strong>ed <strong>in</strong> a different way. Mott [16] [17] could<br />

give a first explanation for this behavior. Figure 2.2 shows the typical band configuration of a<br />

transition metal. The bands are spontaneously split without external field. This phenomenon is<br />

also called band ferromagnetism. Scatter<strong>in</strong>g of electrons from the s- to the d-bands occurs<br />

only, if there are any unoccupied states available. Figure 2.2 shows the electron configuration<br />

at low temperatures. Here the sp<strong>in</strong> up d-band is fully occupied. Thus scatter<strong>in</strong>g can only occur<br />

20


Basics<br />

for sp<strong>in</strong> down electrons. An applied magnetic field may <strong>in</strong>crease the sp<strong>in</strong> polarization and<br />

allows fewer s to d transitions, result<strong>in</strong>g <strong>in</strong> lower resistance.<br />

Figure 2.2: The sp<strong>in</strong>-split bands of a ferromagnet [18].<br />

2.8.2 Anisotropic Magnetoresistance Effect<br />

The resistance of several ferromagnetic materials shows a dependence on the angle between<br />

magnetization and current. This effect is known as anisotropic magnetoresistance (AMR). Its<br />

orig<strong>in</strong> is connected with the sp<strong>in</strong>-orbit <strong>in</strong>teraction and its <strong>in</strong>fluence on s-d scatter<strong>in</strong>g.<br />

This phenomenon can be described by assum<strong>in</strong>g different resistivities for currents parallel and<br />

perpendicular to the magnetic field and leads to follow<strong>in</strong>g relation:<br />

0<br />

2<br />

( 1 RAMR<br />

cos )<br />

ρ=ρ + ⋅ θ . (2.50)<br />

Here θ is the angle between the current and the magnetization, and R AMR is the<br />

magnetoresistance ratio. Typical values of R AMR are about 2%.<br />

The concept of us<strong>in</strong>g the AMR effect for read<strong>in</strong>g heads was first proposed by Hunt <strong>in</strong> 1971<br />

[5]. AMR read heads were used from 1992 to 1998 <strong>in</strong> commercial hard disks. Their advantage<br />

was the much higher signal amplitude for smaller bits (higher data densities) compared to<br />

<strong>in</strong>ductive read heads. Additionally magnetoresistive heads have reduced noise and <strong>in</strong>ductance,<br />

and the output voltage is not dependent on the relative velocity of the head and the data layer.<br />

21


2.8.3 Giant Magnetoresistance Effect<br />

Basics<br />

Giant magnetoresistance was first reported by Baibich et al. <strong>in</strong> 1988 [6]. They discovered that<br />

the resistance of Fe/Cr multilayers decreased 50% when they applied a magnetic field (see<br />

Figure 2.3).<br />

Figure 2.3: Giant magnetoresistance <strong>in</strong> Fe/Cr multilayers [18].<br />

Without apply<strong>in</strong>g an external field the magnetic layers are antiparallel coupled. If a strong<br />

enough field is applied to align the magnetization of adjacent magnetic layers, the resistivity<br />

of the multilayers drops significantly. Thus resistivity is lower <strong>in</strong> case of parallel alignment<br />

and higher for the antiparallel configuration. Multilayers require very high saturation fields to<br />

overcome the strong antiferromagnetic coupl<strong>in</strong>g. Therefore the field sensitivity of such<br />

multilayers is actually lower compared to sensors us<strong>in</strong>g the AMR effect. These considerations<br />

motivated the <strong>in</strong>vention of an alternative GMR structure for sensors shown <strong>in</strong> Figure 2.4. It<br />

consists of four ma<strong>in</strong> layers. There are two ferromagnetic layers which are separated by a<br />

nonmagnetic layer, called spacer layer (mostly Cu). All these layers are only few nanometers<br />

thick. The magnetization of one ferromagnetic layer is fixed via the exchange anisotropy of an<br />

antiferromagnetic layer (see Section 5.1.2). The magnetization of the other ferromagnetic<br />

layer can rotate freely and is therefore called free layer. In contrast to multilayer structures<br />

this structure does not show any hysteresis.<br />

22


Basics<br />

Figure 2.4: Schematic model of a sp<strong>in</strong>-valve GMR head (CIP mode) with exchange-p<strong>in</strong>ned<br />

layer and longitud<strong>in</strong>al hard bias [19].<br />

The change <strong>in</strong> resistance of the GMR effect can be mathematically approximated by<br />

follow<strong>in</strong>g formula:<br />

( θ θ )<br />

⎛ 1−cos 1− 2 ⎞<br />

ρ = ρ0⎜1+ RGMR<br />

⋅<br />

⎟<br />

⎝ 2 ⎠<br />

(2.51)<br />

Figure 2.5 shows the dependence of the resistivity as function of the angle between the<br />

magnetization orientations of the two ferromagnetic layers. To achieve highest sensitivity, the<br />

work<strong>in</strong>g po<strong>in</strong>t of the GMR sp<strong>in</strong> valve should be chosen at maximum slope. If there is no<br />

external field, the magnetization of the free layer should be ideally perpendicular to the<br />

p<strong>in</strong>ned layer magnetization. To adjust the magnetization of the free layer, hard bias magnets<br />

are used as shown <strong>in</strong> Figure 2.4.<br />

Figure 2.5: Change <strong>in</strong> resistance of a sp<strong>in</strong> valve compared to AMR (NiFe) vs.<br />

magnetization orientation [19].<br />

23


Basics<br />

There are two different modes a sp<strong>in</strong> valve can be operated, the current-<strong>in</strong>-plane (CIP) mode,<br />

and the current-perpendicular-to-plane (CPP) mode. The difference between these two modes<br />

is the current direction. In the first case the sense current is parallel to the layers of the GMR<br />

element, while for the second case the current is applied perpendicular to the layers. The two<br />

different modes are illustrated <strong>in</strong> Figure 2.6.<br />

Figure 2.6: The two different modes of a sp<strong>in</strong> valve. The current is applied either<br />

longitud<strong>in</strong>al (CIP) or perpendicular (CPP) to the layer structure [7].<br />

Up to now only the CIP configuration is used <strong>in</strong> commercial hard disks. For smaller read head<br />

designs the CPP mode becomes more and more <strong>in</strong>terest<strong>in</strong>g, because the output voltage for<br />

CPP heads is roughly <strong>in</strong>versely proportional to the square root of the sensor area. In addition<br />

the change <strong>in</strong> resistance is higher, because the sense current is not shunted by the conductive<br />

spacer layer, which does not contribute to magneto-resistance. Besides the CPP structure<br />

allows smaller read gaps, thus smaller distances between the two shields, because the shields<br />

can be used as leads. This design allows read gaps down to 20 nm, compared to 60 nm for<br />

CIP heads. CPP heads are thermally more stable, because the layer structure is <strong>in</strong> contact to<br />

the relative large shields lead<strong>in</strong>g to a better heat dissipation.<br />

The orig<strong>in</strong> of the GMR effect is the sp<strong>in</strong> dependent scatter<strong>in</strong>g of FM/NM/FM multilayers.<br />

The sp<strong>in</strong> dependent scatter<strong>in</strong>g at <strong>in</strong>terfaces is the ma<strong>in</strong> contribution. Figure 2.7 shows a<br />

schematic picture of the potential landscape of a GMR layer structure for the two sp<strong>in</strong><br />

24


Basics<br />

channels <strong>in</strong> the parallel and <strong>in</strong> the antiparallel state, <strong>in</strong>dicat<strong>in</strong>g the different orig<strong>in</strong>s of<br />

scatter<strong>in</strong>g. Only deviations from the periodic atomic potentials are shown, as no scatter<strong>in</strong>g<br />

takes place <strong>in</strong> a regular periodic lattice. The size of the potential step and of the scatter<strong>in</strong>g<br />

potentials at imperfections of the <strong>in</strong>terfaces depends on the degree of match<strong>in</strong>g of the<br />

electronic structures of the ferromagnetic layers and the nonmagnetic spac<strong>in</strong>g layer. In order<br />

to obta<strong>in</strong> a large GMR ratio, good electronic structure match<strong>in</strong>g is required for one sp<strong>in</strong><br />

direction and bad match<strong>in</strong>g for the opposite sp<strong>in</strong> direction. It is clear that giant<br />

magnetoresistance only occurs <strong>in</strong> layers with thicknesses much smaller than the electron<br />

mean-free path.<br />

Figure 2.7: Potential landscape <strong>in</strong>side the FM/NM/FM active part of a sp<strong>in</strong>-valve, for the<br />

two sp<strong>in</strong> channels <strong>in</strong> the parallel state (a,b) and <strong>in</strong> the antiparallel state (c), <strong>in</strong>dicat<strong>in</strong>g the<br />

different orig<strong>in</strong>s of scatter<strong>in</strong>g [20].<br />

The GMR effect can be qualitatively described by a very simplified two current model. The<br />

current densities are assumed to be homogenous for both sp<strong>in</strong> directions. If the resistances for<br />

the ‘+’ and ‘-‘ channels <strong>in</strong> the parallel state are denoted with R + and R − , then the total<br />

resistance for the parallel state is<br />

25


R<br />

↑↑<br />

R+ R−<br />

=<br />

R + R<br />

+ −<br />

Basics<br />

. (2.52)<br />

R+ + R−<br />

The resistance of both channels for the antiparallel configuration is set to result<strong>in</strong>g <strong>in</strong><br />

2<br />

a total resistance of<br />

R + R<br />

= . (2.53)<br />

4<br />

+ −<br />

R ↑↓<br />

For the GMR ratio follows<br />

( ) 2<br />

R − R<br />

Δ R R − R ↑↓ ↑↑ + −<br />

= = . (2.54)<br />

R R 4R<br />

R<br />

↑↑<br />

+ −<br />

This value is always positive, which means that the resistance <strong>in</strong> the anti-parallel<br />

configuration is higher than <strong>in</strong> the parallel configuration, which is consistent with (2.51).<br />

2.8.4 Tunnel<strong>in</strong>g Magnetoresistive Effect<br />

Another magnetoresistive effect is the tunnel<strong>in</strong>g magnetoresistive (TMR) effect. The design<br />

of a TMR element is quite similar to a GMR element <strong>in</strong> CPP mode. The ma<strong>in</strong> difference is the<br />

<strong>in</strong>sulat<strong>in</strong>g layer between p<strong>in</strong>ned and free layer. The orig<strong>in</strong> is also different, because TMR is<br />

due to sp<strong>in</strong> dependent tunnel<strong>in</strong>g and not due to sp<strong>in</strong> dependent scatter<strong>in</strong>g. With TMR<br />

elements changes <strong>in</strong> resistance up to 25% can be achieved at room temperature. The<br />

application <strong>in</strong> read heads is restricted because of the high resistance, which limits the<br />

operat<strong>in</strong>g frequency and causes Johnson and Shot noise.<br />

2.9 Record<strong>in</strong>g Schemes<br />

There are three possibilities to align the written bits <strong>in</strong> the data layer. Longitud<strong>in</strong>al to track,<br />

perpendicular to track, or transversal to the data layer called longitud<strong>in</strong>al, perpendicular and<br />

transversal record<strong>in</strong>g respectively. Only first both record<strong>in</strong>g schemes are of technical <strong>in</strong>terest,<br />

because transversal written bits are hard to detect, because of their small stray fields.<br />

Although there were also attempts to use the perpendicular scheme <strong>in</strong> the early years, the<br />

longitud<strong>in</strong>al mode is well established and is used <strong>in</strong> all commercial hard disk drives.<br />

26


Basics<br />

Nowadays there are aga<strong>in</strong> efforts to use perpendicular record<strong>in</strong>g because of the theoretical<br />

higher data density limit.<br />

2.9.1 Longitud<strong>in</strong>al Record<strong>in</strong>g<br />

The pr<strong>in</strong>ciple of longitud<strong>in</strong>al record<strong>in</strong>g is illustrated <strong>in</strong> Figure 2.8. For the writ<strong>in</strong>g process a<br />

r<strong>in</strong>g head is used, which generates a stray field at the gap above the record<strong>in</strong>g layer, so that it<br />

is magnetized <strong>in</strong> longitud<strong>in</strong>al direction. To retrieve the written <strong>in</strong>formation, the read head<br />

detects the stray field of the data layer, which has its maxima at the transitions of two opposite<br />

magnetized bits. In longitud<strong>in</strong>al record<strong>in</strong>g the read back signal is roughly proportional to the<br />

first derivation of the magnetization.<br />

Figure 2.8: A comparison of longitud<strong>in</strong>al record<strong>in</strong>g and perpendicular record<strong>in</strong>g [7].<br />

2.9.2 Perpendicular Record<strong>in</strong>g<br />

As shown <strong>in</strong> Figure 2.8 an additional soft magnetic underlayer is necessary for perpendicular<br />

record<strong>in</strong>g. This soft under layer (SUL) closes the magnetic circuit. The yoke of the write head<br />

has a small write pole, which bundles the flux to achieve high writ<strong>in</strong>g fields, and a wide return<br />

27


Basics<br />

pole, so that the flux density is small and can not <strong>in</strong>fluence the magnetization of the record<strong>in</strong>g<br />

layer. Here the read back signal is directly proportional to the magnetization of the data layer.<br />

28


Analytical Calculations<br />

3 Analytical CalculationsEquation Section (Next)<br />

3.1 Reciprocity Pr<strong>in</strong>ciple<br />

For any two volum<strong>in</strong>a V 1 and V 2 with given magnetization the follow<strong>in</strong>g equation is valid.<br />

∫ ∫<br />

Mr ( ) ⋅ H( r) dV = Mr ( ) ⋅H(<br />

r ) dV<br />

(3.1)<br />

1 2 1 1 2 1 2 2<br />

V V<br />

1 2<br />

Here 1 H and 2<br />

H denote the magnetic fields caused by the magnetizations <strong>in</strong> volumes 1 and 2<br />

respectively. This fact is called reciprocity pr<strong>in</strong>ciple, and its proof is illustrated below.<br />

Partial <strong>in</strong>tegration of (3.1) and tak<strong>in</strong>g equations (2.16) and (2.19) <strong>in</strong>to consideration lead to<br />

�∫<br />

∫<br />

σ ( r) ⋅Ψ ( r) dA − ρ ( r) ⋅Ψ ( r)<br />

dV =<br />

m 1 2 1 1 m 1 2 1 1<br />

∂V1<br />

V1<br />

�∫<br />

∫<br />

σ ( r ) ⋅Ψ ( r ) dA − ρ ( r ) ⋅Ψ ( r ) dV .<br />

m 2 1 2 2 m 2 1 2 2<br />

∂V2<br />

V2<br />

The f<strong>in</strong>al substitution of both magnetic potentials us<strong>in</strong>g (2.22) shows equality.<br />

3.2 Longitud<strong>in</strong>al Record<strong>in</strong>g<br />

Longitud<strong>in</strong>al record<strong>in</strong>g means that the magnetization of written bits shows <strong>in</strong> track direction.<br />

To read <strong>in</strong>formation it is necessary to detect the stray field of the stored bits. For longitud<strong>in</strong>al<br />

record<strong>in</strong>g the stray field reaches its maximum at the transitions, where opposite oriented<br />

magnetizations come together.<br />

(3.2)<br />

Figure 3.1 shows a schematic model of a read head. In the middle of two shields there is the<br />

sens<strong>in</strong>g layer. In case of a GMR sensor this layer is equal to the free layer. The shields are<br />

assumed to have <strong>in</strong>f<strong>in</strong>ite permeability and <strong>in</strong>f<strong>in</strong>ite extension. The plane, which conta<strong>in</strong>s the<br />

bottom sides of the two shields and the sens<strong>in</strong>g layer, is called air bear<strong>in</strong>g surface (ABS).<br />

29


Analytical Calculations<br />

With the help of the reciprocity pr<strong>in</strong>ciple we are now able to calculate the effective signal<br />

field<br />

eff<br />

H sig caused by the data layer at the sens<strong>in</strong>g layer. Here only the z-component of the field<br />

will be considered. The field components <strong>in</strong> the other direction can be neglected s<strong>in</strong>ce the<br />

sensor only registers the z-component anyway. The effective field<br />

field H = H ( r ) , which acts on the magnetization of the sens<strong>in</strong>g layer (SL).<br />

sig sig<br />

eff<br />

H sig is the average signal<br />

eff<br />

1<br />

Hsig = Hsig= Hsig() dV<br />

V ∫ r (3.3)<br />

SL VSL<br />

We assume that the magnetization of the sens<strong>in</strong>g layer is uniform. With this condition the<br />

reciprocity pr<strong>in</strong>ciple is applied on the volume of the sens<strong>in</strong>g layer V SL and the volume of the<br />

data layer.<br />

eff<br />

V M H = Mr () ⋅H()<br />

r dV<br />

(3.4)<br />

SL SL sig H<br />

V<br />

∫<br />

DL<br />

Here H H is the so called head field. It is the field which is generated by the magnetization of<br />

the sens<strong>in</strong>g layer.<br />

3.2.1 Head Field<br />

z<br />

x<br />

Shield<br />

Shield<br />

t<br />

µ → ∞ µ → ∞<br />

g g<br />

Figure 3.1: A schematic 2-dimensional model of a read head for longitud<strong>in</strong>al record<strong>in</strong>g.<br />

The head field of the read<strong>in</strong>g head due to the magnetization of the sense layer can be<br />

calculated us<strong>in</strong>g the Karlquist approximation [21]. This approximation assumes a l<strong>in</strong>ear<br />

MR<br />

d<br />

Data Layer<br />

30


Analytical Calculations<br />

develop<strong>in</strong>g of the magnetic potential Ψ between the shields and the sens<strong>in</strong>g layer at the ABS.<br />

The magnetic potential is assumed to vanish at the ABS of the shields due to their high<br />

permeability.<br />

⎧ 1 x < t/2<br />

⎪<br />

⎪ g+ t/2+ x<br />

Ψ ( x,0) = C⋅MSL ⋅⎨ t/2 ≤ x < g+ t/2<br />

⎪ g<br />

⎪<br />

⎩<br />

0 g+ t/2≤ x<br />

Figure 3.2: The magnetic potential at the ABS under Karlquist approximation.<br />

This function is shown <strong>in</strong> Figure 3.2. The amplitude of this function is proportional to the<br />

magnetization of the sens<strong>in</strong>g layer. The magnetization of the sens<strong>in</strong>g layer can be replaced by<br />

(3.5)<br />

top<br />

face charges m 0 SL M σ =μ at the bottom and bottom<br />

m 0 SL M<br />

σ =−μ at the top side under the terms of<br />

(2.19). The top face charge can be neglected, because the sens<strong>in</strong>g layer height is usually much<br />

larger than the distance between ABS and data layer surface. Common fly<strong>in</strong>g heights are<br />

about 10 nm. The magnetic charge at the bottom side leads to a jump of the magnetic field<br />

H − H =− M . Due to the symmetry of the surface charge <strong>in</strong> the near region<br />

<strong>in</strong> out<br />

z z SL<br />

z<br />

H (0,0) = M / 2<br />

(3.6)<br />

H SL<br />

Ψ(x,0)<br />

C⋅MSL<br />

-t/2-g -t/2 t/2 t/2+g x<br />

can be assumed. This is the necessary condition, which allows the determ<strong>in</strong>ation of the<br />

unknown proportional factor C <strong>in</strong> (3.5).<br />

31


Analytical Calculations<br />

To calculate the magnetic potential for the whole half space with the boundary condition<br />

given <strong>in</strong> (3.5) one has to solve the Laplace equation<br />

ΔΨ ( xz , ) = 0.<br />

(3.7)<br />

This can be done analytically by a Fourier transformation <strong>in</strong> x [19]. The general solution of<br />

this boundary value problem is<br />

∞<br />

z Ψ(<br />

x,0)<br />

Ψ ( x, z) =− dx′<br />

π ∫ . (3.8)<br />

′<br />

−∞<br />

( ) 2 2<br />

x− x + z<br />

Apply<strong>in</strong>g (3.8) on the boundary condition (3.5) yields<br />

with<br />

C⋅MSL Ψ ( x, z) = Γ ( x+ g, z) −Γ ( x, z) +Γ( − x+ g, z) −Γ( −x,<br />

z)<br />

g ⋅π<br />

( )<br />

( ( ) ) ( )<br />

2 2 + /2<br />

(3.9)<br />

z ⎛ x t ⎞<br />

Γ ( xz , ): = ln x+ t/2 + z − x+ t/2<br />

arctan⎜<br />

⎟.<br />

(3.10)<br />

2<br />

⎝ z ⎠<br />

Now we are able to evaluate the head field by tak<strong>in</strong>g the gradient of the magnetic potential.<br />

⎛γ x( x + gz , ) −γx( xz , ) −γx( − x+ gz , ) +γx( −xz<br />

, ) ⎞<br />

CMSL<br />

⎜ ⎟<br />

H H ( xz , ) = ⋅ 0<br />

g ⎜ ⎟<br />

(3.11)<br />

π ⎜γ z( x + gz , ) −γ z( xz , ) +γz( − x+ gz , ) −γz( −xz<br />

, ) ⎟<br />

⎝ ⎠<br />

Here x γ and γ z are the derivatives of Γ .<br />

∂ ⎛ x+ t/2⎞<br />

γ x ( xz , ): = Γ ( xz , ) =−arctan⎜ ⎟<br />

∂x ⎝ z ⎠ (3.12)<br />

2<br />

2<br />

( ( ) )<br />

∂<br />

γ z ( xz , ): = Γ ( xz , ) = 1+ 1/2⋅ ln z + x+ t/2<br />

∂z<br />

(3.13)<br />

Now we have to determ<strong>in</strong>e the proportional constant C. Here we assume that 2g � t,<br />

which<br />

yields<br />

H<br />

2CM ⎛ g+ t/2⎞<br />

4CM<br />

(0,0) = ln ⎜ ⎟�<br />

. (3.14)<br />

πg ⎝ t/2 ⎠ πt<br />

z<br />

H<br />

SL SL<br />

Compar<strong>in</strong>g with (3.6) gives<br />

32


Analytical Calculations<br />

t<br />

C<br />

8<br />

π<br />

= . (3.15)<br />

Insert<strong>in</strong>g <strong>in</strong> (3.11) leads to the f<strong>in</strong>al result for the head field<br />

⎛γ x( x + gz , ) −γx( xz , ) −γx( − x+ gz , ) +γx( −xz<br />

, ) ⎞<br />

MSL t ⎜ ⎟<br />

H H ( xz , ) = ⋅ 0<br />

8 g ⎜ ⎟<br />

. (3.16)<br />

⎜γ z( x + gz , ) −γ z( xz , ) +γz( − x+ gz , ) −γz( −xz<br />

, ) ⎟<br />

⎝ ⎠<br />

Now a normalized head field H H can be <strong>in</strong>troduced, which is <strong>in</strong>dependent from the sensor<br />

magnetization<br />

H<br />

n<br />

H<br />

⎛γ x( x + gz , ) −γ x( xz , ) −γ x( − x+ gz , ) + γ x(<br />

−xz<br />

, ) ⎞<br />

t ⎜ ⎟<br />

( xz , ) = ⋅ 0<br />

8 g ⎜ ⎟<br />

. (3.17)<br />

⎜<br />

z( x + gz , ) − z( xz , ) + z( − x+ gz , ) − z(<br />

−xz<br />

, ) ⎟<br />

⎝γ γ γ γ ⎠<br />

Figure 3.3 shows the normalized head field at z = − 10 nm and for t = 5 nm and g = 27.5 nm ,<br />

so that the total gap is 60 nm. For longitud<strong>in</strong>al record<strong>in</strong>g only the x-component contributes,<br />

because the magnetization po<strong>in</strong>ts ma<strong>in</strong>ly <strong>in</strong> this direction.<br />

H x<br />

0.06<br />

0.04<br />

0.02<br />

0.00<br />

-0.02<br />

-0.04<br />

-0.06<br />

-100 -50 0 50 100<br />

x [nm]<br />

H z<br />

-0.02<br />

-100 -50 0 50 100<br />

x [nm]<br />

Figure 3.3: The x- and the z-component of the normalized head field of a read head with<br />

60 nm gap, and sens<strong>in</strong>g layer thickness 5 nm, 10 nm below the ABS.<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

-0.01<br />

33


3.2.2 Signal Field<br />

Analytical Calculations<br />

Now we are able to calculate the effective signal field from (3.4)<br />

eff 1<br />

n<br />

Hsig = () ⋅ H () dV<br />

V ∫ Mr H r . (3.18)<br />

SL V<br />

DL<br />

In case of a mov<strong>in</strong>g head with velocity v the effective signal field can be written as function<br />

of time<br />

eff 1<br />

n<br />

Hsig () t = ( ) ⋅ H ( − t) dV<br />

V ∫ Mr H r v . (3.19)<br />

SL V<br />

DL<br />

The advantage of us<strong>in</strong>g the reciprocity pr<strong>in</strong>ciple is that the head field has to be calculated only<br />

once. Then the signal field can be determ<strong>in</strong>ed by fold<strong>in</strong>g the head field with the magnetization<br />

of the record<strong>in</strong>g layer.<br />

3.2.3 Signal of a Perfect Transition<br />

The maximum signal field is achieved for perfect transitions, which are def<strong>in</strong>ed <strong>in</strong> case of<br />

longitud<strong>in</strong>al record<strong>in</strong>g by<br />

J ( x, z) =± J ⋅ sgn( x)<br />

. (3.20)<br />

x s<br />

Here J s is the spontaneous polarization, which is <strong>in</strong> the range of 0.3 T to 0.5 T for common<br />

record<strong>in</strong>g layers. For the head field shown <strong>in</strong> Figure 3.3 we get a pulse for the effective signal<br />

field, which is shown <strong>in</strong> Figure 3.4 for a 10 nm thick data layer positioned 10 nm below the<br />

ABS. The maximum signal field is reached whenever the center of the head is exactly over<br />

the transition. The full width at half maximum is about 50 nm and is therefore <strong>in</strong> the order of<br />

the distance between the two shields (60 nm).<br />

34


µ 0 Hsig [T]<br />

0.030<br />

0.025<br />

0.020<br />

0.015<br />

0.010<br />

0.005<br />

Analytical Calculations<br />

0.000<br />

-200 -100 0 100 200<br />

x [nm]<br />

Figure 3.4: The effective signal field as function of the relative position of the read head <strong>in</strong><br />

respect to a perfect transition (Js = 0.5 T).<br />

3.2.4 Signal of Written Bits<br />

The maximum read back signal is achieved for perfect transitions. In practice written bits do<br />

not have perfect transitions. Figure 3.5 shows the results of two record<strong>in</strong>g simulations with a<br />

head velocity of 10 m/s and 20 m/s.<br />

Figure 3.5: These two figures show the x-component of the magnetization (Js = 0.5 T) of<br />

the data layer after a write process with head velocity 10 m/s (left) and 20 m/s (right). Red<br />

color represents magnetization <strong>in</strong> positive x-direction and blue color magnetization <strong>in</strong><br />

negative x-direction. The scale is <strong>in</strong> µm, thus the bit length is approximately 40 nm. The<br />

data layer thickness is 12 nm.<br />

35


Analytical Calculations<br />

The head velocity <strong>in</strong>fluences the bit quality. The higher the velocity, the worse the quality of<br />

the bit transitions is.<br />

For both data layer structures the read back signals were calculated. Aga<strong>in</strong> a head with 60 nm<br />

gap, 5 nm free layer, and a fly<strong>in</strong>g height of 10 nm was assumed. The spontaneous<br />

magnetization of the data layer was aga<strong>in</strong> J s = 0.5 T , but the data layer thickness was 12 nm.<br />

Figure 3.6 show smaller amplitudes for the effective field compared to the amplitude of the<br />

perfect transition. The worse transition quality of the faster written bits leads to a smaller<br />

signal.<br />

µ 0 Hsig [T]<br />

0.020<br />

0.015<br />

0.010<br />

0.005<br />

0.000<br />

-0.005<br />

-0.010<br />

-0.015<br />

-0.020<br />

0 50 100 150<br />

x [nm]<br />

200 250 300<br />

µ 0 Hsig [T]<br />

0 50 100 150 200 250 300<br />

x [nm]<br />

Figure 3.6: The effective fields of the record<strong>in</strong>g layers (shown <strong>in</strong> Figure 3.5) as function of<br />

the head’s position. The dashed l<strong>in</strong>es <strong>in</strong>dicate the boundaries of the record<strong>in</strong>g layer model.<br />

3.3 Perpendicular Record<strong>in</strong>g<br />

For perpendicular record<strong>in</strong>g we can use the reciprocity pr<strong>in</strong>ciple aga<strong>in</strong>. We only have to take<br />

the SUL <strong>in</strong>to account (see Figure 3.7). Analogically to the shields the SUL is also assumed to<br />

have <strong>in</strong>f<strong>in</strong>ite permeability. Then the problem is equivalent to assum<strong>in</strong>g a second read head,<br />

mirrored at the SUL surface as shown <strong>in</strong> Figure 3.7 [22]. In addition the two shields with their<br />

high permeability act as second mirror. Therefore the head field for perpendicular record<strong>in</strong>g<br />

can be approximated by a series expansion of head fields of the orig<strong>in</strong>al head and its<br />

reflections at the SUL surface and the ABS. If the distance ABS to SUL is denoted with L, the<br />

series expansion of the head field for the perpendicular case can be written us<strong>in</strong>g (3.17)<br />

0.020<br />

0.015<br />

0.010<br />

0.005<br />

0.000<br />

-0.005<br />

-0.010<br />

-0.015<br />

-0.020<br />

36


Analytical Calculations<br />

n n<br />

⎛HH, x( x, z− 2 iL) + HH, x(<br />

x, −z− 2( i+ 1) L)<br />

⎞<br />

∞ ⎜ ⎟<br />

n<br />

H H ( xz , ) = ∑⎜0<br />

⎟.<br />

(3.21)<br />

k = 0 ⎜ n n<br />

HH, z( x, z−2 iL) −HH, z(<br />

x, −z− 2( i+ 1) L)<br />

⎟<br />

⎝ ⎠<br />

Figure 3.7: The simplified read head model (left) for perpendicular record<strong>in</strong>g. In the right<br />

picture the mirror heads are <strong>in</strong>dicated, which replace the SUL with its high permeability<br />

[22].<br />

The headfield for perpendicular record<strong>in</strong>g is shown <strong>in</strong> Figure 3.8. Here the gap was aga<strong>in</strong><br />

60 nm and the sens<strong>in</strong>g layer thickness 5 nm. The fly<strong>in</strong>g height was assumed to be 10 nm and<br />

the distance ABS to SUL 20 nm. The series expansion was aborted after the first 10 terms.<br />

For perpendicular record<strong>in</strong>g, where the data layer is ma<strong>in</strong>ly magnetized <strong>in</strong> z-direction, only<br />

H contributes to the signal field. So for perpendicular record<strong>in</strong>g the signal field is<br />

n<br />

H , z<br />

approximately proportional to the magnetization of the data layer. In contrast to longitud<strong>in</strong>al<br />

record<strong>in</strong>g the bits are detected and not their transitions.<br />

37


H x<br />

0.06<br />

0.04<br />

0.02<br />

0.00<br />

-0.02<br />

-0.04<br />

-0.06<br />

-100 -50 0 50 100<br />

x [nm]<br />

Analytical Calculations<br />

H z<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

-0.01<br />

-0.02<br />

-100 -50 0 50 100<br />

x [nm]<br />

Figure 3.8: The x- and the z-component of the normalized head field for perpendicular<br />

record<strong>in</strong>g.<br />

3.4 Signal Field <strong>in</strong> Gap<br />

After calculat<strong>in</strong>g the signal field H sig of the data layer at the air-bear<strong>in</strong>g surface we are now<br />

able to estimate the run of the magnetic field between the two shields. By averag<strong>in</strong>g H sig over<br />

the MR sensor’s height h we can determ<strong>in</strong>e the effective field<br />

eff<br />

h<br />

1<br />

sig sig<br />

h 0<br />

sig<br />

eff<br />

H sig act<strong>in</strong>g on the sensor.<br />

H = H = ∫ H ( z) dz<br />

(3.22)<br />

To get Hsig ( z ) we assume that the field <strong>in</strong> gap- (x-) direction Hg( z ) is constant between<br />

sensor and shield and only depends on z.<br />

As shown <strong>in</strong> Figure 3.9 we use Ampere’s Law on the dashed marked path, which gives<br />

and f<strong>in</strong>ally<br />

H ( z) ⋅ dz + H ( z + dz) ⋅g−H( z) ⋅ g = 0<br />

(3.23)<br />

sig g g<br />

dH g<br />

Hsig ( z) =− g . (3.24)<br />

dz<br />

38


z<br />

Analytical Calculations<br />

Figure 3.9: a) Application of Ampere’s Law on the dashed path, b) application of Gauss’s<br />

Law on the dashed box.<br />

In addition we use Gauss’ Law for a box with <strong>in</strong>f<strong>in</strong>itesimal small height dz over the sens<strong>in</strong>g<br />

layer:<br />

x<br />

a) g g<br />

b)<br />

−μ ⋅μ ⋅H ( z) ⋅ t+ 2 ⋅μ ⋅H ( z) ⋅ dz+μ ⋅μ ⋅ H ( z+ dz) ⋅ t = 0.<br />

(3.25)<br />

r 0 sig 0 g r 0 sig<br />

Here μ r is the relative permeability of the sens<strong>in</strong>g layer. The last equation is equivalent to<br />

1 dH sig<br />

Hg( z) =− ⋅μr ⋅t⋅ (3.26)<br />

2 dz<br />

Comb<strong>in</strong><strong>in</strong>g (3.24) and (3.26) gives a differential equation.<br />

2<br />

dHsig<br />

2<br />

1<br />

Hsig ( z) = ⋅μr⋅g⋅t⋅ (3.27)<br />

2 dz<br />

Assum<strong>in</strong>g the boundary conditions<br />

ABS<br />

H (0) = H and H ( h)<br />

= 0<br />

(3.28)<br />

sig sig sig<br />

leads to the solution<br />

⎛h−z⎞ s<strong>in</strong>h ⎜ ⎟<br />

λ<br />

Hsig ( z) = Hsig⋅<br />

⎝ ⎠<br />

with λ=<br />

⎛h⎞ s<strong>in</strong>h ⎜ ⎟<br />

⎝λ⎠ ABS r<br />

μ ⋅g⋅t . (3.29)<br />

2<br />

Here λ is the flux propagation length. Now the effective field can be calculated under the<br />

terms of (3.22):<br />

Hsig<br />

t<br />

Hg Hg<br />

Hg<br />

dz<br />

z<br />

x<br />

Hsig<br />

t<br />

dz<br />

g g<br />

39


H = H<br />

eff ABS<br />

sig sig<br />

λ cosh( h / λ) −1<br />

h s<strong>in</strong>h( h/<br />

λ)<br />

Analytical Calculations<br />

For large flux propagation lengths λ � h , which is normally the case, (3.30) simplifies to<br />

(3.30)<br />

h<br />

eff ABS λ ⎛ − ⎞<br />

λ<br />

Hsig = Hsig⎜1−e ⎟.<br />

(3.31)<br />

h ⎝ ⎠<br />

So the signal field roughly decreases exponentially between the shields from the ABS<br />

upwards.<br />

40


4 Numerical MethodsEquation<br />

Numerical Methods<br />

Section (Next)<br />

Micromagnetic simulations require a discrete representation of the cont<strong>in</strong>uous magnetization<br />

distribution on a computer. Due to the strong ferromagnetic coupl<strong>in</strong>g the magnetization does<br />

not vary rapidly over short distances. So the discretization size can be much larger than the<br />

atomic lattice constant. Usual lengths of significant magnetization changes (widths of doma<strong>in</strong><br />

walls) are <strong>in</strong> the order of some nanometers for hard magnetic materials, for soft magnetic<br />

materials even one or two orders of magnitude larger. To keep numerical errors small, the<br />

discretization size should be smaller than the doma<strong>in</strong> wall widths.<br />

Dynamic micromagnetic calculations require two major tasks. The first part is to calculate the<br />

stray field. For this purpose a partial differential equation (PDE) of the form<br />

Lu [ ( r)] = s(<br />

r )<br />

(4.1)<br />

has to be solved. Here L denotes the differential operator, s is called source term and u is the<br />

unknown function, which fulfills this PDE. For stray field calculations u represents the<br />

magnetic potential. We also have to solve an equation of type (4.1) for the electric potential of<br />

an <strong>in</strong>homogeneous conductor (see (2.31)). For this purpose the F<strong>in</strong>ite Element Method is<br />

used.<br />

As second task we have solve an <strong>in</strong>itial-value problem (IVP) of the form:<br />

y� = f(, t y), y( t ) = y . (4.2)<br />

0 0<br />

In case of the LLG equation y represents the magnetization. Therefore y 0 would be the <strong>in</strong>itial<br />

magnetization. Here the time has to be discretized. We use a time <strong>in</strong>tegration rout<strong>in</strong>e, which<br />

automatically adapts the time steps. So for fast physical processes a smaller time step is<br />

chosen to <strong>in</strong>crease accuracy.<br />

41


4.1 F<strong>in</strong>ite Element Method<br />

Numerical Methods<br />

The F<strong>in</strong>ite Element Method (FEM) is a well established tool for solv<strong>in</strong>g partial differential<br />

equations (PDE). The ma<strong>in</strong> advantage of FEM is the good applicability on arbitrary shaped<br />

geometries. Usually tetrahedrons are used to fractionalize the model. In addition the<br />

discretization size, i.e. the size of the tetrahedrons, can be easily varied over the whole model.<br />

The general form of a PDE is<br />

Lu [ � ( r)] = f(<br />

r)<br />

. (4.3)<br />

Here L denotes the differential operator, f is called source term and u� is the unknown exact<br />

solution, which fulfills this PDE. The ma<strong>in</strong> idea of FEM is to discretize the solution. So a set<br />

of N basis functions ϕi ( r) is assumed. Now all l<strong>in</strong>ear comb<strong>in</strong>ations of these basis functions<br />

are considered:<br />

N<br />

∑<br />

u() r = u ⋅ϕ () r . (4.4)<br />

i=<br />

1<br />

i i<br />

Then the numerical solution u is that l<strong>in</strong>ear comb<strong>in</strong>ation, which fits best to the true solution<br />

u� . Thus the coefficients u i have to be determ<strong>in</strong>ed.<br />

To measure the quality of different solutions the residual function R( r) is <strong>in</strong>troduced<br />

R(): r = Lu [()] r − f()<br />

r . (4.5)<br />

The smaller the residual the smaller the error of the solution is. Now we can make different<br />

demands on the residual <strong>in</strong> order to get a good solution. One popular method is the Galerk<strong>in</strong><br />

Method, whose conditions for the residual are<br />

∫ R() r wi() r dr = 0 1≤i≤<br />

N . (4.6)<br />

V<br />

In other words, the solution has to fulfill the PDE only <strong>in</strong> average. Here wi ( r ) are N weight<br />

functions, which have to satisfy certa<strong>in</strong> conditions to lead actually to a small residual. One<br />

possibility is to set these weight functions equal to the basis functions.<br />

w() r =ϕ() r 1≤i≤N<br />

. (4.7)<br />

i i<br />

The basis functions usually are piecewise l<strong>in</strong>ear functions with the property<br />

42


Numerical Methods<br />

ϕ ( x ) =δ . (4.8)<br />

i j ij<br />

Each basis function ϕ i refers to the i-th node of the tetrahedral mesh. Figure 4.1 shows such<br />

functions for a 2-dimensional mesh, which are also called “hat functions” due to their<br />

appearance.<br />

Figure 4.1: Typical basis functions of two neighbor<strong>in</strong>g nodes of a 2-dimensional mesh.<br />

For our simulations we need to solve Laplace and Poisson equations of the form<br />

( )<br />

∇ σ() r ∇ u() r = s()<br />

r . (4.9)<br />

Here s( r ) is the source term, which vanishes <strong>in</strong> case of a Laplace equation. The additional<br />

factor σ( r ) also <strong>in</strong>cludes the generalized Laplace equation for <strong>in</strong>homogeneous conductors<br />

(see (2.31)). Now we apply the Galerk<strong>in</strong> Method, which leads to<br />

( ( ) )<br />

∫ ∇ σ() r ∇u() r −s() r ϕ i () r dV = 0,<br />

(4.10)<br />

V<br />

Partial <strong>in</strong>tegration yields<br />

∫ ∫ ∫<br />

σ() r ∇u() r ∇ϕ i() r dV = � σ() r ∇u() r ϕi() r dA− s() r ϕi()<br />

r dV . (4.11)<br />

V ∂V<br />

V<br />

We take <strong>in</strong>to account that the solution u is a l<strong>in</strong>ear comb<strong>in</strong>ation of all basis functions ϕ i :<br />

N<br />

∑ ∫ ∫ ∫<br />

u σ() r ∇ϕ () r ∇ϕ () r dV = � σ() r ∇u() r ϕ () r dA− s() r ϕ () r dV . (4.12)<br />

j j i i i<br />

j= 1 V ∂V<br />

V<br />

With the def<strong>in</strong>itions<br />

43


Numerical Methods<br />

Aij = ∫ σ() r ∇ϕ j () r ∇ϕi()<br />

r dV , (4.13)<br />

V<br />

si =−∫s() r ϕi()<br />

r dV , (4.14)<br />

V<br />

∂u()<br />

r<br />

bi = ∫ σ() r ∇u() r ϕ i() r dA= ∫ σ() r ϕi()<br />

r dA<br />

∂n<br />

� � (4.15)<br />

∂V ∂V<br />

(4.12) can be written as<br />

N<br />

∑ Au ij j<br />

j=<br />

1<br />

= bi + si = ri<br />

. (4.16)<br />

Here the system matrix A = ( Aij<br />

) is symmetric and positive def<strong>in</strong>ite. s i is the source term,<br />

and b i is due to boundary conditions and only contributes, if the i-th node is situated at the<br />

surface ∂ V ( r i ∈∂V<br />

). In this case we have to determ<strong>in</strong>e the <strong>in</strong>tegral <strong>in</strong> (4.15) by tak<strong>in</strong>g<br />

boundary conditions <strong>in</strong>to account. Here we have to dist<strong>in</strong>guish between Dirichlet and<br />

Neumann boundary conditions. r i denotes the right hand side and is the sum of the boundary<br />

and the source term.<br />

4.1.1 Dirichlet Boundary Conditions<br />

In case of Dirichlet boundary conditions the value of the i-th node u( r ) = b is known. So the<br />

whole i-th equation can be replaced by<br />

i i<br />

ui = bi.<br />

(4.17)<br />

This would destroy the symmetry of the matrix A ij . To ma<strong>in</strong>ta<strong>in</strong> its symmetry, which is<br />

convenient for solv<strong>in</strong>g a system of l<strong>in</strong>ear equations, the matrix entry A ii is just multiplied by a<br />

6<br />

large number ( ~10 ). Then the right hand side is set to b i multiplied with the new coefficient<br />

A ii .<br />

4.1.2 Neumann Boundary Conditions<br />

In case of Neumann boundary conditions the normal component of σ() r ∇u()<br />

r , that is<br />

n⋅σ() r ∇ u() r =− j()<br />

r , is given at the boundary of the mesh. Here j() r can be <strong>in</strong>terpreted as a<br />

44


Numerical Methods<br />

current flow<strong>in</strong>g out of the volume. Normally the Neumann boundary conditions are given<br />

either for each boundary triangle<br />

j() r = j , ∀r ∈Δ<br />

(4.18)<br />

ijk ijk<br />

or for each boundary node of the mesh<br />

j( r) = j , ∀r ∈∂V<br />

. (4.19)<br />

i i i<br />

For the first case we get for the boundary <strong>in</strong>tegral term <strong>in</strong> (4.15)<br />

1<br />

σ() r ∇u() r ϕ () r dA=− j() r ϕ () r dA=−∑ j ϕ () r dA=− ∑A<br />

j<br />

3<br />

�∫ � ∫ ∫<br />

. (4.20)<br />

i i ijk i ijk ijk<br />

∂V ∂V Δijk Δijk<br />

j< k<br />

Δijk<br />

j< k<br />

For the second case we get for the current j by l<strong>in</strong>ear <strong>in</strong>terpolation between the sampl<strong>in</strong>g<br />

po<strong>in</strong>ts (po<strong>in</strong>ts at the boundary) us<strong>in</strong>g the test functions i ϕ<br />

∑<br />

j() r = j ⋅ϕ () r . (4.21)<br />

i<br />

i i<br />

Then the result of the boundary <strong>in</strong>tegral is<br />

�∫ �∫<br />

σ() r ∇u() r ϕ () r dA=− j() r ϕ () r dA=<br />

i i<br />

∂V ∂V<br />

1<br />

−∑ ∫ ( jiϕ i() r + jjϕ j() r + jkϕk() r ) ϕ i() r dA=− ∑Aijk<br />

( 2 ji + jj + jk)<br />

.<br />

12<br />

Δijk Δ<br />

Δ<br />

ijk<br />

ijk<br />

j< k j< k<br />

4.1.3 L<strong>in</strong>ear equation solver<br />

The FEM method leads to a system of l<strong>in</strong>ear equations (4.16)<br />

(4.22)<br />

Au ij j = ri<br />

⇔ Au ⋅ = r (4.23)<br />

with a sparse symmetric and positive def<strong>in</strong>ite matrix A. To solve this system of l<strong>in</strong>ear<br />

equations a Cholesky factorization method for sparse matrices seemed to be practicable [23].<br />

The advantage of this method over iterative methods is the <strong>in</strong>sensibility of “bad” meshes. The<br />

convergence fails with iterative methods if the tetrahedrons show much difference from<br />

regular tetrahedrons, so that they have small angles. This is problematic for meshes of th<strong>in</strong><br />

layers, as used <strong>in</strong> GMR elements.<br />

45


4.2 Stray Field Calculation<br />

Numerical Methods<br />

The numerical calculation of the stray field is not very efficient, if it is calculated us<strong>in</strong>g (2.22),<br />

because for each po<strong>in</strong>t <strong>in</strong> space one has to calculate the whole volume <strong>in</strong>tegral <strong>in</strong> (2.22). So<br />

the calculation requires comput<strong>in</strong>g of<br />

calculate the stray field more efficiently with numerical means.<br />

2<br />

ON ( ) . Fredk<strong>in</strong> and Koehler [24] proposed a way to<br />

As described <strong>in</strong> Section 2.3 the field of a given magnetization, <strong>in</strong> this case the stray field, can<br />

be represented as gradient of a magnetic scalar potential.<br />

H =−grad Ψ . (4.24)<br />

S S<br />

The magnetic potential of a given magnetization <strong>in</strong> a volume V can be calculated by solv<strong>in</strong>g<br />

the Poisson equation<br />

ΔΨ = divM<br />

(4.25)<br />

S<br />

with the jump condition at the boundary ∂ V<br />

out <strong>in</strong><br />

∂Ψ S ∂Ψ S − =−Mn ⋅<br />

∂n ∂n<br />

∂V<br />

. (4.26)<br />

Now the idea of Fredk<strong>in</strong> and Koehler was to split the magnetic scalar potential <strong>in</strong>to two parts<br />

Ψ=Ψ S,1 +Ψ S,2.<br />

(4.27)<br />

Inside the volume V Ψ S ,1 is the solution of a Poisson equation<br />

ΔΨ = M (4.28)<br />

<strong>in</strong><br />

S ,1 div<br />

with Neumann boundary conditions<br />

<strong>in</strong><br />

∂Ψ S ,1<br />

= ⋅<br />

∂n<br />

Mn<br />

∂V<br />

. (4.29)<br />

Outside of V this part of the magnetic potential is set to zero:<br />

Ψ = . (4.30)<br />

out<br />

S ,1 0<br />

Ψ S ,1 can be solved now as performed <strong>in</strong> Section 4.1. The boundary term and the source term<br />

are calculated as follows:<br />

46


Numerical Methods<br />

∫ ∫ ∫<br />

si =− div Mϕ i( r) dV = M⋅grad ϕi( r) dV −� Mϕi( r) dA,<br />

(4.31)<br />

∫<br />

V V ∂V<br />

bi = � Mϕi() r dA.<br />

(4.32)<br />

∂V<br />

Both terms contribute to the right hand side, result<strong>in</strong>g <strong>in</strong><br />

N<br />

r = M⋅grad ϕ ( r) dV = M ϕ ( r)grad ϕ ( r ) dV . (4.33)<br />

∫ ∑ ∫<br />

i i j j i<br />

V j=<br />

1 V<br />

Here the <strong>in</strong>tegral can be calculated analytically for each tetrahedron of the mesh.<br />

S<strong>in</strong>ce the total potential S S,1 S,2<br />

satisfy the Laplace equation<br />

<strong>in</strong> out<br />

S,2 S,2<br />

Ψ =Ψ +Ψ must fulfill equations (4.25) and (4.26), Ψ ,2 must<br />

ΔΨ = 0, ΔΨ = 0<br />

(4.34)<br />

with boundary conditions<br />

and<br />

out <strong>in</strong><br />

∂Ψ S,2 ∂Ψ S,2<br />

− = 0<br />

∂n ∂n<br />

∂V<br />

out <strong>in</strong> <strong>in</strong><br />

S,2 S,2 S,1<br />

∂V<br />

S<br />

(4.35)<br />

Ψ −Ψ =Ψ . (4.36)<br />

Fortunately the last equations also describe the magnetic scalar potential of a dipole layer with<br />

moment Ψ S ,1 at the surface ∂ V . The scalar potential of such a dipole layer is known:<br />

1<br />

r−r′ Ψ () r = � Ψ ( r′ ) dA′<br />

. (4.37)<br />

∫<br />

S,2 4π<br />

∂V<br />

<strong>in</strong><br />

S,13<br />

r−r′ i<br />

In practice the magnetic potential Ψ S ,2 can be evaluated at each node S,2 S,2 i<br />

i<br />

with FEM. Generally the potential vector ( S ,2 )<br />

j<br />

with the vector ( Ψ S ,1 ) ,<br />

i j<br />

S,2 ij S,1<br />

Ψ =Ψ ( r ) , if used<br />

Ψ can be calculated by a matrix multiplication<br />

Ψ = M Ψ . (4.38)<br />

47


Here the matrix = ( M ij )<br />

Numerical Methods<br />

M describes a dipole <strong>in</strong>teraction of all boundary nodes of V, with all<br />

nodes of V, and is normally fully populated. This matrix has to be calculated only once,<br />

because it does not depend on the boundary conditions. For large models this matrix requires<br />

a lot of memory. So it is more convenient to calculate only the <strong>in</strong>teraction between boundary<br />

and boundary nodes, and evaluate the magnetic potential Ψ S ,2 by solv<strong>in</strong>g the Laplace<br />

equation <strong>in</strong> (4.34) with Dirichlet boundary conditions. In addition the <strong>in</strong>teraction matrix is<br />

compressed as described <strong>in</strong> next section.<br />

4.3 Adaptive Cross-Approximation Technique<br />

The discretization of boundary <strong>in</strong>tegral equations leads to large dense matrices that have no<br />

explicit structure <strong>in</strong> general. However, by suitable renumber<strong>in</strong>g and permut<strong>in</strong>g the boundary<br />

nodes, the dense matrix can be written <strong>in</strong> a form that conta<strong>in</strong>s blocks that can be approximated<br />

by low-rank matrices [25]. The renumber<strong>in</strong>g of the nodes is done by geometrical criterions.<br />

Consecutive boundary nodes should be located close to each other. A simple example of the<br />

cluster<strong>in</strong>g is given <strong>in</strong> Figure 4.2. Boundary nodes that are located close to each other are<br />

comb<strong>in</strong>ed <strong>in</strong> a cluster. A large distance between cluster results <strong>in</strong> entries <strong>in</strong> the renumbered<br />

boundary matrix that are well separated from each other. For example the cluster with the<br />

nodes {6,7,8,9,10} and the cluster {1,2,3,4,5} are admissible cluster pairs. This pair of<br />

clusters leads to a block <strong>in</strong> the boundary matrix which can be approximated by a low-rank<br />

matrix. The boundary matrix correspond<strong>in</strong>g to the cluster<strong>in</strong>g is also shown <strong>in</strong> Figure 4.2. The<br />

off-diagonal blocks of the matrix represent far-field <strong>in</strong>teractions.<br />

This technique leads to compressed <strong>in</strong>teraction matrices, which save a lot of memory.<br />

Additionally the algorithmic complexity for sett<strong>in</strong>g up the boundary matrix and for matrix-by-<br />

vector products is reduced from<br />

2<br />

ON ( ) to ON ( ) .<br />

48


Numerical Methods<br />

Figure 4.2: The left figure shows a set of 10 nodes after renumber<strong>in</strong>g and the cluster<strong>in</strong>g.<br />

The right figure shows the associated boundary matrix structure. The large off-diagonal<br />

blocks, correspond<strong>in</strong>g to far field <strong>in</strong>teractions of two clusters, can be approximated by lowrank<br />

matrices [25].<br />

4.4 Conductor Model<br />

The sense current through the MR sensor generates a magnetic field, which has to be taken<br />

<strong>in</strong>to account. The current is even used for bias<strong>in</strong>g <strong>in</strong> AMR heads (see Section 6.1.2).<br />

Therefore the sense current cannot be neglected.<br />

So we have a conductor model <strong>in</strong> addition to the magnetic model. As there are materials<br />

which are magnetic and conductive these two models can overlap. The current model consists<br />

of materials with different conductivities and magnetoresistive effects as described <strong>in</strong> Section<br />

2.8. So the current density must be assumed as function of locus and as function of<br />

magnetization.<br />

σ=σrM (, )<br />

(4.39)<br />

Aga<strong>in</strong> we use a tetrahedral mesh and the F<strong>in</strong>ite Element Method to calculate the electric<br />

potential and further the current distribution.<br />

4.4.1 Conductivity<br />

In case of non-magnetoresistive materials we have constant conductivities. In<br />

magnetoresistive materials, however, the conductivity has to be evaluated from the<br />

magnetization (see (2.50) or (2.51)). To specify conductivity and magnetoresistance there are<br />

49


Numerical Methods<br />

several material parameters, listed and described <strong>in</strong> Table 4.1. The listed parameters are<br />

assigned for each material. The tetrahedrons <strong>in</strong>herit the properties of the material, which they<br />

are assigned to.<br />

If a magnetoresistive material is specified, the conductivity has to be calculated first, <strong>in</strong> order<br />

to be able to calculate the current density. In case of a AMR (Type = 1) or GMR element<br />

(Type = 2) the magnetic material or two magnetic materials, respectively, have to be<br />

specified, which <strong>in</strong>fluence the conductivity. To calculate the conductivity for a tetrahedron of<br />

a GMR material, the nearest po<strong>in</strong>ts to the midpo<strong>in</strong>t of the tetrahedron <strong>in</strong> both materials are<br />

searched first. This is done us<strong>in</strong>g the Approximate Nearest Neighbor (ANN) Search<br />

Algorithm [26]. Then the effective conductivity can be obta<strong>in</strong>ed us<strong>in</strong>g (2.50) or (2.51).<br />

Parameter Description<br />

σ Conductivity <strong>in</strong> Ω -1 m -1<br />

Type 0 Normal conductor, 1 AMR, 2 GMR<br />

MR-Ratio The maximum change <strong>in</strong> resistance.<br />

Only relevant if Type is not equal to 0.<br />

<strong>Magnetic</strong> Material 1 Only relevant for AMR or GMR, def<strong>in</strong>es number of (1 st )<br />

magnetic material, which <strong>in</strong>fluences the conductivity.<br />

<strong>Magnetic</strong> Material 2 Only relevant for GMR, def<strong>in</strong>es number of 2 nd material,<br />

which <strong>in</strong>fluences conductivity.<br />

Table 4.1: Material parameters for the conductor model.<br />

4.4.2 Electric Potential<br />

To get the electric potential Φ Equation (2.31)<br />

div( σgrad Φ ) = 0<br />

(4.40)<br />

has to be solved. In addition we have to specify boundary conditions. There are two<br />

possibilities: Either the potential or the out-/<strong>in</strong>flow<strong>in</strong>g current has to be declared. F<strong>in</strong>ally this<br />

problem can be solved by apply<strong>in</strong>g the FEM approach as described <strong>in</strong> Section 4.1.<br />

N<br />

Φ () r = Φ ⋅ϕ () r (4.41)<br />

∑<br />

i= 1<br />

i i<br />

The result gives the values of the electric potential at the nodal po<strong>in</strong>ts.<br />

4.4.3 Current Distribution<br />

Then the current distribution can be evaluated us<strong>in</strong>g (2.32)<br />

50


Numerical Methods<br />

j =−σgrad Φ.<br />

(4.42)<br />

With the calculated electric potential (4.41) we are able to determ<strong>in</strong>e the current distribution<br />

for each tetrahedron ijkl. Insert<strong>in</strong>g <strong>in</strong> (4.42) gives<br />

( grad grad grad grad )<br />

j =−σ Φ ⋅ ϕ +Φ ⋅ ϕ +Φ ⋅ ϕ +Φ ⋅ ϕ . (4.43)<br />

ijkl ijkl i i j j k k l l<br />

4.4.4 <strong>Magnetic</strong> Field of Current<br />

When the current distribution is known for each element, the next task is to determ<strong>in</strong>e the<br />

additional magnetic field due to the current. The current is uniform for each tetrahedron. So<br />

the Biot-Savart Law (2.46) can be written as<br />

or<br />

1<br />

Hijkl () r = jijkl<br />

×<br />

4π<br />

∫<br />

Vijkl<br />

r−r′ r−r′ 3<br />

dV ′<br />

(4.44)<br />

1 1<br />

Hijkl () r = j ′ ijkl × ∇ dV ′<br />

4 ∫<br />

. (4.45)<br />

π r−r′ Vijkl<br />

Integration delivers (see Appendix A)<br />

1 1<br />

Hijkl () r = jijkl × d ′<br />

4 �∫<br />

A . (4.46)<br />

π r−r′ ∂Vijkl<br />

So the field of a tetrahedron with uniform current density can be calculated by <strong>in</strong>tegrat<strong>in</strong>g<br />

over its surface triangles:<br />

1 ⎛ 1 1 1 1 ⎞<br />

Hijkl () r = jijkl × ⎜ d ′ + d ′ + d ′ + d ′ ⎟<br />

4 ⎜ ∫ A<br />

− ′ ∫ A<br />

− ′ ∫ A<br />

− ′ ∫ A . (4.47)<br />

π<br />

− ′ ⎟<br />

⎝<br />

r r r r r r r r<br />

Δijk Δjlk Δilk Δijl<br />

⎠<br />

Fortunately the <strong>in</strong>tegrals <strong>in</strong> (4.47) can be solved analytically for triangle surfaces, result<strong>in</strong>g <strong>in</strong><br />

quite complicated formulas [27] [28]. The total field due to the current can be obta<strong>in</strong>ed by<br />

summation over all tetrahedral elements ijkl. Now we are able to determ<strong>in</strong>e the magnetic field<br />

at all nodes r i of the magnetic model. Generally the effective field can be written as<br />

multiplication of the <strong>in</strong>teraction matrix and the current density vector.<br />

∑<br />

Hr ( ) = A ⋅j<br />

. (4.48)<br />

i ij j<br />

jl<br />

51


Numerical Methods<br />

Here A ij are 3× 3-submatrices<br />

describ<strong>in</strong>g the <strong>in</strong>teraction of the j-th tetrahedron of the<br />

conductor model with the i-th node of the magnetic model. Thus the matrix A= ( A ) is a<br />

( 3n) ( 3m)<br />

× -Matrix, where n denotes the number of nodes of the magnetic model and m the<br />

number of tetrahedrons <strong>in</strong> the conductor model. Usually the matrix A is fully populated and<br />

requires a lot of memory. The multiplication with such a matrix is very time consum<strong>in</strong>g. To<br />

speed up the calculation and save memory an adaptive cross-approximation technique can be<br />

used (see chapter 4.3). Unfortunately this technique needs a lot of memory when calculat<strong>in</strong>g<br />

the approximation of the <strong>in</strong>teraction matrix A for large magnetic and conductor models.<br />

4.4.5 Hybrid FEM/BEM for Current Field<br />

To overcome the storage problem we proposed an alternative way to calculate the magnetic<br />

field <strong>in</strong> the magnetic model. Here we use a common mesh for the magnet and the conductor<br />

model. Similar to the stray field calculation (see Section 2.5.2) we can reduce the <strong>in</strong>teraction<br />

matrix to a boundary <strong>in</strong>teraction matrix.<br />

Start<strong>in</strong>g from Equation (2.29)<br />

curl H = j, (4.49)<br />

apply<strong>in</strong>g the curl-operator on both sides, and consider<strong>in</strong>g that div H = 0 lead to a Poisson<br />

equation for each component of H.<br />

Δ H =−curl<br />

j (4.50)<br />

with boundary condition (see Appendix B)<br />

out <strong>in</strong><br />

∂H ∂H<br />

− = n× j<br />

∂n ∂n<br />

∂V<br />

. (4.51)<br />

Now the magnetic field can be calculated quite similar to the stray field, described <strong>in</strong> Section<br />

4.2. We split the magnetic field <strong>in</strong>to two parts<br />

H = H1+ H 2,<br />

(4.52)<br />

where H 1 is the solution of<br />

Δ H =− j (4.53)<br />

<strong>in</strong><br />

1 curl<br />

ij<br />

52


Numerical Methods<br />

with<strong>in</strong> the volume V. Outside the magnetic field H 1 is set to zero<br />

out<br />

H 1 = 0 . (4.54)<br />

Then the boundary condition writes as<br />

∂H<br />

∂n<br />

<strong>in</strong><br />

1<br />

=− n× j<br />

∂V<br />

. (4.55)<br />

Now H 1 can be calculated with<strong>in</strong> the volume V as described <strong>in</strong> Section 4.1. The source term<br />

is determ<strong>in</strong>ed by<br />

∫<br />

s = curl jϕ ( r ) dV . (4.56)<br />

i i<br />

V<br />

Here the source term is a vector. Each component refers to one of the three Poisson equations<br />

<strong>in</strong> (4.53). Partial <strong>in</strong>tegration gives<br />

si = � ∫ n× jϕi() r dA − ∫j×<br />

grad ϕi()<br />

r dV . (4.57)<br />

∂V<br />

V<br />

The boundary <strong>in</strong>tegral becomes<br />

<strong>in</strong><br />

∂H1<br />

bi = ∫ ϕ i() r dA =− × ϕi()<br />

dA<br />

∂ ∫ n j r<br />

n<br />

� � . (4.58)<br />

∂V ∂V<br />

So the first <strong>in</strong>tegral <strong>in</strong> the source term and the boundary term cancel each other. The right<br />

hand side is therefore<br />

∫<br />

r =− j× grad ϕ(<br />

r ) dV . (4.59)<br />

i i<br />

V<br />

The gradient of the test function and the current density are constant for each tetrahedron, so<br />

the <strong>in</strong>tegral (4.59) can be calculated as sum over all tetrahedrons, which are enclosed to the<br />

i-th node.<br />

The second part H 2 has to fulfill<br />

and further<br />

Δ H = 0, Δ H = 0<br />

(4.60)<br />

<strong>in</strong> out<br />

2 2<br />

53


out <strong>in</strong> <strong>in</strong><br />

2 2 1<br />

∂V<br />

Numerical Methods<br />

H − H = H (4.61)<br />

to guarantee (4.50) and (4.53). Aga<strong>in</strong> we have an equivalent problem as <strong>in</strong> Section 4.2, but for<br />

three dimensions. Analogously the field H 2 can be evaluated by<br />

( − ′ )<br />

1 r r n<br />

H () r = ( ′ ) dA<br />

π � ∫ H r<br />

. (4.62)<br />

′<br />

2<br />

4 ∂V<br />

<strong>in</strong><br />

1<br />

r−r 3<br />

To save memory we only create an <strong>in</strong>teraction matrix from boundary to boundary, and solve<br />

the Laplace equation (4.60) aga<strong>in</strong> to calculate H 2 .<br />

4.4.6 Example: Long Wire<br />

To verify the program and the calculated fields, some relative simple models are simulated,<br />

which can also be calculated analytically. The first example is a long straight Cu-wire with<br />

length l = 200 mm and wire radius r = 2 mm . The conductivity of Cu is<br />

7 -1 -1<br />

σ= 5.961⋅10 Ω m . At least we have to specify the boundary conditions of the wire. At one<br />

side the electric potential was set to V 1 = 0 . The boundary condition of the opposite end was<br />

chosen <strong>in</strong> a way, that the total <strong>in</strong>flow<strong>in</strong>g current has a value of I = 1 mA . With these<br />

parameters the total resistance of the Cu-wire can be calculated: R = 267 µ Ω . Us<strong>in</strong>g Ohm’s<br />

Law (2.1) we get for the electric potential V 2 = 0.267 µV . This analytically calculated value<br />

matches well with that of the FEM calculation 2<br />

0.278 µV<br />

FEM<br />

FEM<br />

V = . The error ( 2<br />

V is about<br />

4.2% larger than V 2 ) is traced back to the mesh. The mesh has to approximate a cyl<strong>in</strong>drical<br />

surface with triangles. So the cross sectional area of the mesh model is smaller than that of the<br />

realistic model. This leads to a larger resistance and further to a larger voltage.<br />

I<br />

2<br />

The current density of the realistic model is j = = 79.58 A/m . The magnetic field of an<br />

2<br />

r π<br />

<strong>in</strong>f<strong>in</strong>ite long straight wire with constant current can be easily calculated due to the symmetry<br />

<strong>in</strong> respect of rotations around the wire’s axis. The only non-zero field component is that <strong>in</strong><br />

azimuthal direction and can be determ<strong>in</strong>ed us<strong>in</strong>g Ampere’s Law (2.6).<br />

� ∫ H⋅ ds= ∫j×<br />

dA<br />

(4.63)<br />

∂A<br />

A<br />

54


Numerical Methods<br />

Integration of Ampere’s law gives for the radial field<br />

H<br />

( )<br />

⎧ ρ<br />

⎪<br />

j : 0≤ρ<<br />

r<br />

⎪ 2<br />

. (4.64)<br />

ϕ ρ =⎨ 2<br />

⎪<br />

r<br />

j : r ≤ρ<br />

⎪⎩ 2ρ<br />

This function is plotted <strong>in</strong> Figure 4.3, and shows a l<strong>in</strong>ear <strong>in</strong>crease, reach<strong>in</strong>g its maximum at<br />

the wire’s surface and decl<strong>in</strong>es proportional to 1/ρ outside.<br />

H ϕ [A/m]<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.00<br />

0<br />

0 10 20 30 40 50<br />

ρ [mm]<br />

Figure 4.3: Hϕ for an <strong>in</strong>f<strong>in</strong>ite long wire is plotted over the distance from the middle axis.<br />

The second curve shows the error of the numerical calculation.<br />

The second curve <strong>in</strong> Figure 4.3 shows the relative error of the FEM calculation compared to<br />

(4.64). The error is negative and <strong>in</strong>creases accord<strong>in</strong>g to amount with <strong>in</strong>creas<strong>in</strong>g ρ . The model<br />

with f<strong>in</strong>ite length used for the FEM calculation is the reason for that. The f<strong>in</strong>ite length has a<br />

greater effect for larger radial distances. The discont<strong>in</strong>uities of the error curve are due to the<br />

Adaptive Cross-Approximation Technique (see Section 4.3). For small ρ we have a very<br />

good match of the numerical and analytical calculated fields.<br />

4.4.7 Example: Coil<br />

A coil as shown <strong>in</strong> Figure 4.4 is the second example. For the coil wire we have aga<strong>in</strong> the same<br />

7 -1 -1<br />

parameters as for the straight wire ( σ= 5.961⋅10 Ω m , I = 1 mA , r = 2 mm ). The radius<br />

of the coil is r′ = 20 mm , the length is l = 100 mm , and it has N = 20 w<strong>in</strong>d<strong>in</strong>gs. The model<br />

additionally has two straight leads with lengthl L = 80 mm . Similar to the straight wire we can<br />

now calculate the difference of the electric potential by analytical means:<br />

-10<br />

-8<br />

-6<br />

-4<br />

-2<br />

Error [%]<br />

55


V<br />

2l + N2πr′ ≈ = 3.57 µV<br />

σrπ L<br />

2 2<br />

Numerical Methods<br />

(4.65)<br />

With the FEM we get V 2 = 3.81µV . The difference is aga<strong>in</strong> due to the smaller cross sectional<br />

area of the meshed model.<br />

H z (r=0) [A/m]<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0<br />

-100 -50 0 50 100 150 200<br />

z [mm]<br />

Figure 4.4: The model of the coil (left). The color code shows the l<strong>in</strong>ear decrease of the<br />

electric potential <strong>in</strong>side the wire. The chart (right) shows the magnetic field of a solenoid <strong>in</strong><br />

z-direction at the middle axis. The two dotted l<strong>in</strong>es identify the position of the solenoid.<br />

The second graph shows the relative error of the coil’s magnetic field calculated with FEM<br />

calculation compared to the solenoid’s field.<br />

The magnetic field for a perfect solenoid (hollow cyl<strong>in</strong>der with <strong>in</strong>f<strong>in</strong>itesimal th<strong>in</strong> nappe) with<br />

length L, face current density J and radius r′ can be calculated exactly along the middle axis<br />

(z-axis):<br />

⎛ ⎞<br />

J L−z z<br />

Hz( z)<br />

= ⋅ ⎜ + ⎟.<br />

(4.66)<br />

2 ⎜ 2 2 2 2<br />

( L− z) + r′<br />

z + r′<br />

⎟<br />

⎝ ⎠<br />

When we now approximate the coil by such a solenoid and set the face current density equal<br />

to<br />

NI<br />

J = , we have the possibility to compare the field of the solenoid H z with that of the<br />

L<br />

calculated field of our model. The run of the magnetic field along the middle axis of the<br />

solenoid is shown <strong>in</strong> Figure 4.4. In addition the relative error of the FEM calculated coil<br />

compared with the solenoid is given. So the quantitative values are aga<strong>in</strong> consistent.<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Error [%]<br />

56


4.5 Time <strong>in</strong>tegration<br />

Numerical Methods<br />

Us<strong>in</strong>g the LLG equation, the time evolution of the magnetization can be calculated. The<br />

discretization of realistic problems leads to an <strong>in</strong>itial-value problem (4.2) with up to a million<br />

unknowns.<br />

To solve such problems the free solver CVODE is used [29] [30], which offers iterative and<br />

direct multistep methods. Multistep methods make use of the past values of the solution.<br />

For our purposes the iterative backward differentiation formula (BDF) seemed to fit best to<br />

solve our problem, <strong>in</strong> order to achieve fastest convergence times. The nonl<strong>in</strong>ear system is<br />

solved us<strong>in</strong>g a Newton method, which requires only a few iterations. In order to improve the<br />

performance, CVODE allows precondition<strong>in</strong>g. By declar<strong>in</strong>g an estimated Jacobian matrix the<br />

convergence time can be drastically decreased.<br />

CVODE is capable of choos<strong>in</strong>g the time step size automatically. For rapidly chang<strong>in</strong>g<br />

processes a smaller time step is chosen to <strong>in</strong>crease accuracy.<br />

Figure 4.5 schematically shows the complete simulation cycle. First the mesh model and the<br />

material parameters are read. Then the boundary matrices for the FEM/BEM methods for the<br />

stray field and the current field are set up. Afterwards the <strong>in</strong>itial magnetization is read and set,<br />

and the time <strong>in</strong>tegration starts.<br />

CVODE solves the LLG equation, and requests the right hand side of the LLG equation for<br />

certa<strong>in</strong> po<strong>in</strong>ts of time. CVODE passes the actual time and the magnetic polarization J i of<br />

each node to the LLG calculation rout<strong>in</strong>e, which calculates the right hand side of the LLG<br />

equation. All magnetic fields which contribute to the total effective magnetic field H eff are<br />

calculated for each nodal po<strong>in</strong>t. Then the effective field is <strong>in</strong>serted <strong>in</strong> the right hand side of the<br />

LLG equation, which leads to new functions f i of time and magnetic polarization of the i-th<br />

node. These functions are transmitted back to the solver, which decides to modify the<br />

magnetic polarization. If the desired precision is not fulfilled the new magnetic polarizations<br />

are passed to the LLG calculation rout<strong>in</strong>e aga<strong>in</strong>, and the cycle is repeated. Otherwise the<br />

solver chooses the next time step and the system time is <strong>in</strong>creased. The whole procedure is<br />

repeated until a user chosen f<strong>in</strong>al time is reached.<br />

57


Time Integration<br />

CVODE<br />

∂J<br />

∂t<br />

i<br />

= f ( t,<br />

J , J ,..., J )<br />

i 1 2 N<br />

Numerical Methods<br />

START<br />

<strong>Read</strong> geometry (FE mesh)<br />

and material parameters<br />

Calculate the B matrix for<br />

the hybrid FE/BE method<br />

Set the <strong>in</strong>itial magnetization<br />

t, J1, J2,..., J N<br />

f1, f2,..., f N<br />

Increase time (t)<br />

t < tf<strong>in</strong>al<br />

END<br />

Hexch + Hani + HS+ Hcurr + Hext<br />

�������������<br />

H<br />

Figure 4.5: Schematic diagram of the simulation cycle. Adapted from [31].<br />

eff<br />

LLG<br />

f1, f2,..., f N<br />

σ ( J )<br />

Φ, j<br />

H ext () t<br />

58


5 <strong>Read</strong> Head DesignEquation<br />

5.1 Bias Schemes<br />

<strong>Read</strong> Head Design<br />

Section (Next)<br />

For AMR and GMR read<strong>in</strong>g heads, magnetic layers with a uniform predef<strong>in</strong>ed magnetization<br />

direction are necessary. Besides a s<strong>in</strong>gle doma<strong>in</strong> state is desirable, to avoid Barkhausen noise<br />

due to mov<strong>in</strong>g doma<strong>in</strong> walls. Therefore the magnetizations of these layers have to be forced<br />

<strong>in</strong>to a certa<strong>in</strong> direction by bias<strong>in</strong>g. The most important bias schemes used <strong>in</strong> MR read heads<br />

are expla<strong>in</strong>ed below.<br />

5.1.1 Demagnetiz<strong>in</strong>g Factor<br />

The field of a magnetized body, the stray field, or also called demagnetiz<strong>in</strong>g field H S , acts on<br />

the magnetization itself. This field is proportional to the magnetization M and opposite<br />

directed with<strong>in</strong> the body. So the demagnetiz<strong>in</strong>g field can be written as<br />

N =− H M (5.1)<br />

S d<br />

Here N d is the demagnetiz<strong>in</strong>g factor, which is generally a tensor and only depends on the<br />

shape of the body. Pr<strong>in</strong>cipal axis transformation leads to three demagnetiz<strong>in</strong>g factors a N , N b ,<br />

N c for each axis one respectively. For their sum<br />

Na + Nb + Nc<br />

= 1<br />

(5.2)<br />

always holds. The longest axis corresponds to the smallest demagnetiz<strong>in</strong>g factor. It is also the<br />

preferred direction for the magnetization, because then the stray field energy has its m<strong>in</strong>imum<br />

(see (2.37)).<br />

A GMR element basically exists of two ferromagnetic layers, which should be magnetized<br />

perpendicular <strong>in</strong> respect to each other <strong>in</strong> ground state. Moreover the free layer should be able<br />

to rotate freely. Therefore the shape of these layers has to be almost quadratic (with an aspect<br />

ratio close to 1) <strong>in</strong> order not to discrim<strong>in</strong>ate a magnetization direction.<br />

59


<strong>Read</strong> Head Design<br />

On the other hand the signal field decays with distance from the record<strong>in</strong>g layer (see Section<br />

3.4) and therefore a higher effective (average) signal field is reached for smaller sensor<br />

heights. Thus a compromise must be found. Usual width-to-height ratios are between 1.2 and<br />

1.5.<br />

5.1.2 Hard Bias<br />

To align the magnetization of the free layer hard magnets are used (see Figure 2.4). Usually<br />

the GMR element is placed between two hard magnets, whose static magnetic field effects the<br />

perpendicular alignment <strong>in</strong> respect to the p<strong>in</strong>ned layer. The permanent magnetic fields also<br />

<strong>in</strong>fluence the magnetization of the p<strong>in</strong>ned layer. Therefore the bias<strong>in</strong>g field for the p<strong>in</strong>ned<br />

layer has to overcome the hard bias field.<br />

5.1.3 Exchange Bias<br />

Exchange bias is based on the exchange anisotropy, which was discovered by Meiklejohn and<br />

Bean 50 years ago [32]. They <strong>in</strong>vestigated small Co particles coated with its<br />

antiferromagnetic oxide CoO. A field cool<strong>in</strong>g process, i.e. heat<strong>in</strong>g up above the Néel<br />

temperature and cool<strong>in</strong>g under apply<strong>in</strong>g a magnetic field, leads to a shift of the hysteresis<br />

loop, which is shown <strong>in</strong> Figure 5.1. The exchange bias effect can be characterized by the<br />

displacement of the hysteresis loop, denoted by H ex .<br />

Figure 5.1: The shift of the hysteresis loop after a field cool<strong>in</strong>g treatment of a<br />

ferromagnetic/antiferromagnetic system [19] and the temperature dependence of the<br />

exchange bias field for NiFe(4-8nm)/Cu(2.2nm)/CoFe(2nm)/IrMn(15nm) sp<strong>in</strong> valves with<br />

different free layer thicknesses [33].<br />

60


<strong>Read</strong> Head Design<br />

The exchange bias field depends on the thicknesses of the antiferromagnet and the<br />

ferromagnetic layer. If the antiferromagnet is thicker than the ferromagnetic layer the<br />

exchange bias field decreases with the thickness of the ferromagnet t p like<br />

H<br />

J<br />

1<br />

t p<br />

− [34].<br />

ex<br />

ex = . (5.3)<br />

µ 0 Ms, ptp Here J ex is the exchange bias <strong>in</strong>teraction parameter [J/m 2 ]. This k<strong>in</strong>d of dependence leads to<br />

the conclusion, that the exchange anisotropy is an <strong>in</strong>terface effect. The maximum exchange<br />

bias field, which can be achieved for usual layer thicknesses, is about 0.05 T [20]. The<br />

exchange bias field decreases l<strong>in</strong>early with temperature [33]. Figure 5.1 shows the<br />

temperature dependence of the exchange bias field for NiFe(4-8nm)/Cu(2.2nm)/CoFe(2nm)/<br />

IrMn(15nm) sp<strong>in</strong> valves with different free layer thicknesses.<br />

The exchange bias effect is used to align the magnetization of the p<strong>in</strong>ned layer perpendicular<br />

to the record<strong>in</strong>g layer. The advantage of this bias scheme is that the free layer is not<br />

<strong>in</strong>fluenced by the antiferromagnet. Unfortunately the stray field of the p<strong>in</strong>ned layer causes an<br />

unwanted torsion of the free layer’s magnetization, which leads to a displacement of the<br />

equilibrium state towards the antiparallel state. The result is an asymmetric sensor curve.<br />

5.1.4 Synthetic Antiferromagnet<br />

To reduce the asymmetry of the transfer curve of the sp<strong>in</strong> valve, a stronger exchange bias is<br />

required. The limitation of the exchange bias field leads to an alternative sp<strong>in</strong> valve design<br />

[35]. Instead of the direct coupl<strong>in</strong>g of the antiferromagnet with the p<strong>in</strong>ned layer, an additional<br />

ferromagnetic layer is <strong>in</strong>troduced which is separated by a very th<strong>in</strong> (0.7-0.9 nm) nonmagnetic<br />

layer (mostly Ru) as shown <strong>in</strong> Figure 5.2. This layer structure is also called a synthetic<br />

antiferromagnet, because the p<strong>in</strong>ned layer is antiferromagnetically coupled with the additional<br />

ferromagnetic layer by <strong>in</strong>terlayer exchange coupl<strong>in</strong>g (IEC).<br />

61


<strong>Read</strong> Head Design<br />

Figure 5.2: A comparison (left) of a conventional sp<strong>in</strong> valve and a sp<strong>in</strong> valve us<strong>in</strong>g a<br />

synthetic ferromagnet to fix the p<strong>in</strong>ned layer [20]. The <strong>in</strong>terlayer exchange constant (right)<br />

for two Ni80Co20 layers as function of the Ru spacer layer thickness [32].<br />

The IEC effect has its orig<strong>in</strong> <strong>in</strong> quantum <strong>in</strong>terference of electron states [36]. The exchange<br />

coupl<strong>in</strong>g constant J IEC [J/m 2 ] depends on the nonmagnetic layer thickness t NM like<br />

⎛ tNM<br />

⎞<br />

s<strong>in</strong> ⎜2π +Φ<br />

Λ<br />

⎟<br />

J = J<br />

⎝ ⎠<br />

(5.4)<br />

IEC<br />

0 2<br />

tNM<br />

as shown <strong>in</strong> Figure 5.2. Λ , Φ , and J 0 are constants, depend<strong>in</strong>g on the composition of the<br />

layers. The advantage of sp<strong>in</strong>-valves with a synthetic antiferromagnetic p<strong>in</strong>ned layer is the<br />

very large exchange field of IEC up to 0.5 T. This is an <strong>in</strong>crease of a factor 10 compared to<br />

the direct exchange coupl<strong>in</strong>g with an antiferromagnet. So the stability aga<strong>in</strong>st external fields<br />

is much better due to the higher anisotropy. Moreover the demagnetiz<strong>in</strong>g fields of the two<br />

ferromagnetic layers Fp,1 and Fp,2 cancel each other partially, which leads to a reduced<br />

<strong>in</strong>fluence on the free layer.<br />

5.1.5 Sense Current<br />

The sense current flow<strong>in</strong>g through the GMR sensor generates a magnetic field. This field also<br />

<strong>in</strong>fluences the p<strong>in</strong>ned layer magnetization. For conventional simple sp<strong>in</strong> valves there is one<br />

current direction which is more favorable. The ma<strong>in</strong> part of the current flows through the Cu<br />

layer, due to its higher conductivity compared to the other layers. So the current field either<br />

adds up with the demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer or weakens the field act<strong>in</strong>g on the<br />

62


<strong>Read</strong> Head Design<br />

free layer (see Figure 5.3). Unfortunately the more favorable current direction causes a<br />

decl<strong>in</strong>e of the exchange bias field.<br />

Figure 5.3: The effect of the sense current direction on the free layer (FL). For the first<br />

case (left) the current field Hcurr partially cancels the stray field HS of the p<strong>in</strong>ned layer (PL)<br />

and weakens the exchange bias field Hexch. For the other current direction (right) the stray<br />

field and the exchange bias field are supported by the current field.<br />

Shield<br />

µ → ∞<br />

Hcurr<br />

I<br />

HS<br />

FL Cu PL<br />

Hexch<br />

Hcurr<br />

I I<br />

I<br />

Hmirr<br />

FL Cu PL<br />

Shield<br />

µ → ∞<br />

Figure 5.4: The layers of the GMR element also see mirror currents due to the high<br />

permeability of the shields. The field of these mirror currents is able to reduce the field of<br />

the orig<strong>in</strong>al current at the free layer and the p<strong>in</strong>ned layer.<br />

Hcurr<br />

Hcurr<br />

FL Cu PL<br />

Hmirr<br />

I<br />

HS<br />

Hexch<br />

63


<strong>Read</strong> Head Design<br />

In presence of shields with high permeability the free layer sees an additional mirror current,<br />

which must also be considered, especially for small gap distances. These mirror currents<br />

reduce the current field as shown <strong>in</strong> Figure 5.4.<br />

5.1.6 Crystall<strong>in</strong>e Anisotropy<br />

Similar to the shape anisotropy the crystall<strong>in</strong>e anisotropy can be used for bias<strong>in</strong>g. For<br />

example a field cool<strong>in</strong>g treatment of a material can generate an <strong>in</strong>tr<strong>in</strong>sic anisotropy.<br />

Nevertheless this bias scheme is not often used <strong>in</strong> GMR sensors, because the free layer should<br />

be able to rotate freely and the p<strong>in</strong>ned layer is normally biased by exchange anisotropy.<br />

5.2 Shield<strong>in</strong>g<br />

To avoid <strong>in</strong>fluence of external fields and neighbor<strong>in</strong>g bit transitions, shields with high<br />

permeability enclose the GMR element. Shields normally consist of Permalloy (Ni80Fe20).<br />

The relative permeability of Permalloy is <strong>in</strong> the order of<br />

4<br />

10 .<br />

The gap distance between both shields is limited to 60 nm for CIP sp<strong>in</strong> valves and 20 nm for<br />

CPP sp<strong>in</strong> valves (see Section 2.8.3). In order to have an improvement of the read back signal<br />

the gap distance should not exceed the bit length.<br />

5.3 Thermal Stability<br />

The sense current leads to a heat<strong>in</strong>g of the GMR element. The properties of the GMR sensor<br />

are temperature dependent. For <strong>in</strong>stance the exchange bias field decreases with temperature<br />

(see Figure 5.2). Moreover the noise <strong>in</strong>creases. To achieve thermal stability a constant sens<strong>in</strong>g<br />

current is applied <strong>in</strong>stead of a constant voltage. Then the change <strong>in</strong> resistance is determ<strong>in</strong>ed<br />

by measur<strong>in</strong>g the change <strong>in</strong> voltage. The heat production of the current limits the current<br />

density. The heat dissipation is especially a problem for GMR sensors work<strong>in</strong>g <strong>in</strong> CIP mode,<br />

because for the CPP mode the shields are used as leads. Therefore the electric conductance<br />

between GMR sensor and shields is larger, which normally also leads to a larger thermal<br />

conductance due to the Wiedemann-Franz Law.<br />

64


FEM Simulations<br />

6 FEM SimulationsEquation Section (Next)<br />

6.1 Model<br />

The model of the read head used <strong>in</strong> the follow<strong>in</strong>g calculations is very simplified, because of a<br />

shortage of detailed <strong>in</strong>formation. Despite cooperation with hard disk manufacturers, their<br />

<strong>in</strong>formation was very restricted. Nevertheless we were able to simulate the qualitative<br />

behavior and to prove the functionality of our code. The whole read head model is separated<br />

<strong>in</strong>to two parts: A conductor model and a magnetic model.<br />

6.1.1 <strong>Magnetic</strong> Model<br />

A simple GMR read head basically exists of five parts, which def<strong>in</strong>e the magnetic behavior:<br />

The free layer, the p<strong>in</strong>ned layer, the hard bias, the shields, and f<strong>in</strong>ally the antiferromagnet for<br />

fix<strong>in</strong>g the p<strong>in</strong>ned layer. The magnetic net moment of an antiferromagnetic material is zero.<br />

Apart from exchange coupl<strong>in</strong>g the antiferromagnet does not contribute to the magnetic<br />

behavior. Therefore the antiferromagnetic part can be neglected, if the exchange coupl<strong>in</strong>g is<br />

taken <strong>in</strong>to account by add<strong>in</strong>g the exchange bias field to the effective field with<strong>in</strong> the p<strong>in</strong>ned<br />

layer.<br />

The geometry of the FEM model for the calculations is shown <strong>in</strong> Figure 6.1. The free layer<br />

and the p<strong>in</strong>ned layer dimensions are h GMR = 80 nm <strong>in</strong> height and w GMR = 100 nm <strong>in</strong> lateral<br />

FL<br />

PL<br />

direction. Their thicknesses are t = 5 nm and t = 2 nm , and their spontaneous<br />

FL<br />

polarizations are J = 1.2 T , which is the average of a composite free layer (NiFe/CoFe) and<br />

s<br />

J = 1.885 T (CoFe) respectively. Both layers are assumed to have no anisotropy and are<br />

PL<br />

s<br />

Cu<br />

separated from each other by a t = 2.7 nm th<strong>in</strong> nonmagnetic Cu layer. The two hard bias<br />

magnets have uniaxial anisotropy <strong>in</strong> lateral direction<br />

HB<br />

5 3<br />

K 1 = 2.5⋅ 10 J/m and a spontaneous<br />

HB<br />

polarization of J = 1T.<br />

Their lateral distance from the GMR layer structure is 10 nm. The<br />

s<br />

exchange bias field is always set to H ex = 0.05 T , unless otherwise noted. This corresponds<br />

65


FEM Simulations<br />

approximately to the exchange coupl<strong>in</strong>g with the p<strong>in</strong>ned layer for a 8 nm IrMn<br />

antiferromagnet. The two shields are assumed to be made of Permalloy (NiFe) with<br />

Shields<br />

spontaneous polarization J = 1T,<br />

and their dimensions are 50 nm× 100 nm× 50 nm . The<br />

s<br />

gap between the two shields is 60 nm. The Gilbert damp<strong>in</strong>g constant for the whole model is<br />

always 0.1 unless otherwise noted. The exchange constants are<br />

PL<br />

A<br />

−11<br />

= 1.2⋅ 10 J/m ,<br />

HB<br />

A<br />

are summarized <strong>in</strong> Table 6.1.<br />

−11<br />

= 1.0⋅ 10 J/m , and<br />

Shields<br />

A<br />

FL<br />

A<br />

−11<br />

= 1.3⋅ 10 J/m ,<br />

−11<br />

= 1.3⋅ 10 J/m . All magnetic properties<br />

The GMR element of the model is meshed with a size of 5 nm. The mesh of the shields has a<br />

general size of 50 nm, but it is ref<strong>in</strong>ed near the gap and near the track region.<br />

Figure 6.1: The magnetic model used for the FEM calculations. The mesh size for the<br />

shields (green) is 50 nm, and is ref<strong>in</strong>ed to 5 nm near the track and near the sp<strong>in</strong> valve<br />

layers. The hard bias magnets (purple), the free layer (dark blue), and the p<strong>in</strong>ned layer<br />

(light blue) are also meshed with triangle size of 5 nm.<br />

z<br />

y<br />

x<br />

66


FEM Simulations<br />

<strong>Magnetic</strong> Material Js [T] K1 [J/m 3 ] A [J/m] Hex [T] Α<br />

Free Layer 1 0 1.3E-11 0 0.1 *<br />

P<strong>in</strong>ned Layer 1.885 0 1.2E-11 0.05 *<br />

0.1 *<br />

Hard Bias 1 2.5E5 1.3E-11 0 0.1 *<br />

Shields 1 0 1.0E-11 0 0.1 *<br />

Table 6.1: The values of the spontaneous polarization, the anisotropy constant, the<br />

exchange constant, and the antiferromagnetic exchange coupl<strong>in</strong>g field, and the Gilbert<br />

damp<strong>in</strong>g constant, which are assumed for our magnetic model. The asterisk-marked values<br />

are valid unless otherwise noted.<br />

6.1.2 Conductor Model<br />

For the current model we have to consider all conductive parts of the model. This also<br />

<strong>in</strong>cludes the antiferromagnet or capp<strong>in</strong>g layers. The conductor model is shown <strong>in</strong> Figure 6.2.<br />

However, only the most important layers are taken <strong>in</strong>to account for our simulations. So we<br />

neglect the th<strong>in</strong> capp<strong>in</strong>g layers, although they make a m<strong>in</strong>or contribution to the conductance.<br />

Thus our conductor model conta<strong>in</strong>s the free layer, p<strong>in</strong>ned layer, nonmagnetic spacer layer<br />

(Cu), the antiferromagnetic layer, and parts of the leads.<br />

Figure 6.2: The FEM conductor model is shown. This simplified model shows the four<br />

ma<strong>in</strong> layers of a GMR element, which contribute most to the conductance: The free layer<br />

(dark blue) the nonmagnetic Cu layer (orange), the p<strong>in</strong>ned layer (light blue), and the<br />

antiferromagnet (yellow). Additionally the high conductive leads (p<strong>in</strong>k) are shown.<br />

x<br />

y<br />

z<br />

67


FEM Simulations<br />

The sense current through the sp<strong>in</strong> valve is applied at the lateral surfaces of the conductor<br />

model and has a value of 3 mA.<br />

The effective conductivities of the <strong>in</strong>dividual parts are estimated from the measured<br />

conductivities <strong>in</strong> Figure 6.3 [37]. These conductivities differ much from the bulk values,<br />

because the electron mean-free path is larger than the layer thicknesses.<br />

Figure 6.3: The effective conductivities for a typical sp<strong>in</strong> valve taken from [37].<br />

In our simulation the free layer has a conductivity of<br />

PL 6 −1 −1<br />

σ = 3.5⋅10 Ω m , the Cu layer<br />

FL 6 −1 −1<br />

σ = 310 ⋅ Ω m , the p<strong>in</strong>ned layer<br />

Cu 6 −1 −1<br />

σ = 4.5⋅10 Ω m , the antiferromagnet<br />

AF 6 −1 −1<br />

Lead 7 −1 −1<br />

σ = 110 ⋅ Ω m and f<strong>in</strong>ally the leads σ = 110 ⋅ Ω m . Here the conductivities of free<br />

layer, p<strong>in</strong>ned layer and the Cu layer are not constant, because it is assumed that these layers<br />

contribute to the GMR effect. Their maximum change <strong>in</strong> conductivity, the GMR ratio, was<br />

assumed to be 10%. So the effective GMR ratio of the whole layer structure is smaller,<br />

because the sense current is shunted by the antiferromagnet. All values assumed for the<br />

conductor model are listed <strong>in</strong> Table 6.2.<br />

Conductor σ [Ω -1 m -1 ] Type MR-Ratio <strong>Magnetic</strong> Material 1 <strong>Magnetic</strong> Material 2<br />

Free Layer 3.0E6 2 (GMR) 0.1 Free Layer P<strong>in</strong>ned Layer<br />

Cu Layer 4.5E6 2 (GMR) 0.1 Free Layer P<strong>in</strong>ned Layer<br />

P<strong>in</strong>ned Layer 3.5E6 2 (GMR) 0.1 Free Layer P<strong>in</strong>ned Layer<br />

Antiferromagnet 1.0E6 0 - - -<br />

Lead 1.0E7 0 - - -<br />

Table 6.2: The properties for all parts of the conductor model. Here the conductivities, the<br />

type of magnetoresistance, and the magnetic materials, which <strong>in</strong>fluence the conductivity,<br />

are given. (Compare with Table 4.1).<br />

68


6.2 Results<br />

FEM Simulations<br />

In the follow<strong>in</strong>g the results of the simulations are presented. Naturally the model described<br />

above of a read head is not perfect at all. It was not the target to design a perfect read head,<br />

but to be able to simulate the read back process. Although the used read head model is not<br />

technically mature, the results of the simulations show qualitative accordance with reality.<br />

6.2.1 Equilibrium State<br />

Ideally the magnetizations of the p<strong>in</strong>ned and the free layer should be perpendicular to each<br />

other <strong>in</strong> equilibrium state, so that the magnetization of the p<strong>in</strong>ned layer shows <strong>in</strong> z-direction,<br />

and that of the free layer <strong>in</strong> y-direction. Then the total resistance would be 32.87 Ω and the<br />

potential difference 98.6 mV. In reality such a perfect configuration is not achievable.<br />

The equilibrium states of our model for the two different current directions are shown <strong>in</strong><br />

Figure 6.4. For the more favorable current direction the magnetizations are almost<br />

perpendicular to each other <strong>in</strong> the center of the GMR element. Unfortunately the free layer<br />

magnetization is twisted downwards due to the demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer.<br />

Similarly the p<strong>in</strong>ned layer is rotated <strong>in</strong> the same direction due to the hard bias. Both effects<br />

lead to an asymmetric transfer curve (see Section 6.2.2).<br />

Figure 6.4: The normalized magnetization of the free layer and the p<strong>in</strong>ned layer <strong>in</strong><br />

equilibrium state for the two different current directions. The color code gives the zcomponent<br />

of the normalized magnetization, red corresponds to the positive and blue to the<br />

negative z-direction. Thus the red arrows represent the p<strong>in</strong>ned layer magnetization. The<br />

right hand side gives the configuration for the more favorable current direction, because<br />

here the current field cancels partially the demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer.<br />

69


FEM Simulations<br />

If the current flows <strong>in</strong> the more favorable direction the exchange bias field is weakened (see<br />

Section 5.1.5), which can also be seen <strong>in</strong> Figure 6.4, when consider<strong>in</strong>g the torsion of the<br />

magnetization of the p<strong>in</strong>ned layer.<br />

6.2.2 Transfer Curve<br />

To get the transfer curve of the sp<strong>in</strong> valve, a triangle pulsed external field as shown <strong>in</strong> Figure<br />

6.5 is applied. Its amplitude is 0.1 T and its period time is 8 ns. As we will see later <strong>in</strong> Section<br />

6.2.4, the chang<strong>in</strong>g of the external field is slow enough, so that the GMR sensor is always<br />

near equilibrium.<br />

H z [T]<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

-0.05<br />

-0.10<br />

-0.15<br />

0 2 4 6 8 10<br />

Time [ns]<br />

Figure 6.5: The external applied field, which is applied <strong>in</strong> z-direction to evaluate the<br />

transfer curve of the GMR sensor.<br />

70


Output [V]<br />

0.108<br />

0.107<br />

0.106<br />

0.105<br />

0.104<br />

0.103<br />

0.102<br />

0.101<br />

-0.10 -0.05 0.00 0.05 0.10<br />

H z [T]<br />

FEM Simulations<br />

ΔR/R [%]<br />

Current <strong>in</strong> pos. y-direction<br />

Current <strong>in</strong> neg. y-direction<br />

Current <strong>in</strong> pos. y-direction, Hex = 0.1T<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-0.10 -0.05 0.00 0.05 0.10<br />

Hz [T]<br />

Figure 6.6: The output voltage (left) and the relative change <strong>in</strong> resistance (right, <strong>in</strong> respect<br />

to the equilibrium state) over the external applied field <strong>in</strong> z-direction for both current<br />

direction and a larger exchange bias field.<br />

The correspond<strong>in</strong>g outputs for both current directions and a stronger exchange bias are given<br />

<strong>in</strong> Figure 6.8. A positive external field leads to a parallel alignment of the magnetizations of<br />

both layers, result<strong>in</strong>g <strong>in</strong> a low resistance. If the external field is applied <strong>in</strong> negative z-<br />

direction, the resistance <strong>in</strong>creases, because the magnetic configuration approaches the<br />

antiparallel state. However, if the magnetic field overcomes the exchange bias field, the<br />

p<strong>in</strong>ned layer magnetization is not fixed anymore and rotates towards negative z-direction. So<br />

for large negative external fields we have aga<strong>in</strong> a parallel state and therefore low resistance.<br />

As expected the favorable current direction shows less asymmetry <strong>in</strong> change of resistance<br />

around the equilibrium state than the opposite current direction. Moreover the GMR sensor is<br />

more sensitive, because the slope of the transfer curve at zero field is larger. The output<br />

voltage for this case is also higher, because here the magnetization vectors are more likely<br />

aligned antiparallel (see equilibrium state <strong>in</strong> Figure 6.4), lead<strong>in</strong>g to larger resistance.<br />

In case of a stronger exchange bias field, H ex = 0.1T , we have a much better transfer curve<br />

for higher external fields. (Unfortunately such high exchange fields are rather unrealistic. To<br />

<strong>in</strong>crease the bias field for the p<strong>in</strong>ned layer, synthetic antiferromagnets, as described <strong>in</strong> Section<br />

71


FEM Simulations<br />

5.1.4, are used <strong>in</strong> modern sp<strong>in</strong> valves.) The transfer curve and the slope at zero po<strong>in</strong>t do not<br />

differ much compared to the low exchange bias field. However, for large fields we have a<br />

much better sensor performance <strong>in</strong> respect to sensitivity and symmetry. The higher voltage <strong>in</strong><br />

equilibrium is due to the larger angle between both magnetizations. The large exchange bias<br />

field reduces the deflection due to the hard bias, while the free layer magnetization stays<br />

unmodified, because of the widely unchanged demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer.<br />

6.2.3 Influence of Shields<br />

Now the shields are taken <strong>in</strong>to account. It is expected that the shields have an <strong>in</strong>fluence on the<br />

transfer curve due to the mirror currents (see Section 5.1.5). Indeed the shields have a large<br />

<strong>in</strong>fluence. Aga<strong>in</strong> the transfer curve is evaluated by apply<strong>in</strong>g the external field shown <strong>in</strong> Figure<br />

6.5 on the GMR element. The shields are not exposed to the external field. The result is given<br />

<strong>in</strong> Figure 6.7.<br />

Output Voltage [V]<br />

0.109<br />

0.108<br />

0.107<br />

0.106<br />

0.105<br />

0.104<br />

0.103<br />

0.102<br />

-0.10 -0.05 0.00 0.05 0.10<br />

H z [T]<br />

ΔR/R [%]<br />

without shields<br />

with shields<br />

-0.10 -0.05 0.00 0.05 0.10<br />

H z [T]<br />

Figure 6.7: The transfer curves for the sp<strong>in</strong> valve with and without shields (for both<br />

Hexch = 0.1 T).<br />

The presence of shields generally leads to a greater output voltage, particularly at zero field.<br />

The orig<strong>in</strong> of this effect is based on mirror currents. They reduce the current field which<br />

works aga<strong>in</strong>st the demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer. Therefore the demagnetiz<strong>in</strong>g field<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

72


FEM Simulations<br />

has a greater <strong>in</strong>fluence. The result is a more likely antiparallel state, which is equivalent to a<br />

higher total resistance.<br />

This proves that the shield also have an <strong>in</strong>fluence on the read head behavior and cannot be<br />

neglected <strong>in</strong> micromagnetic simulations.<br />

6.2.4 Relaxation<br />

The micromagnetic behavior of the GMR sensor depends on the Gilbert damp<strong>in</strong>g constant.<br />

Start<strong>in</strong>g from an equilibrium state with an external field of H ext = 0.03 T <strong>in</strong> positive z-<br />

direction, the relaxation to the new equilibrium state when suddenly switch<strong>in</strong>g off the field, is<br />

shown <strong>in</strong> Figure 6.8.<br />

Output Voltage [V]<br />

0.1060<br />

0.1055<br />

0.1050<br />

0.1045<br />

0.1040<br />

0.1035<br />

0.1030<br />

0.0 0.2 0.4 0.6 0.8<br />

Time [ns]<br />

α = 0.1<br />

α = 0.2<br />

α = 0.3<br />

α = 0.5<br />

α = 1<br />

Figure 6.8: The output signal over time for a relaxation from equilibrium state with<br />

Hext = 0.3 T to the equilibrium state without field for different Gilbert damp<strong>in</strong>g constants.<br />

For large Gilbert damp<strong>in</strong>g constants we have strong damp<strong>in</strong>g, result<strong>in</strong>g <strong>in</strong> a quite large<br />

relaxation time (larger than 0.4 ns). New read heads work with a read frequency up to 2 GHz.<br />

Therefore such large relaxation times would weaken the read back signal. Fortunately realistic<br />

damp<strong>in</strong>g constants are about α ∼ 0.1,<br />

which lead to an undercritical damp<strong>in</strong>g, result<strong>in</strong>g <strong>in</strong><br />

damped oscillations. Smallest relaxation times are achieved for critical damp<strong>in</strong>g with<br />

damp<strong>in</strong>g constants around 0.3.<br />

73


6.2.5 Signal of a Perfect Transition<br />

FEM Simulations<br />

To calculate the signal of a perfect transition a track width of 120 nm, a data layer thickness<br />

of 12 nm, and a spontaneous polarization of J s = 0.44 T was assumed. The read back signal<br />

was <strong>in</strong>vestigated for both k<strong>in</strong>d of transitions ( →← and ←→).<br />

Figure 6.9 and Figure 6.10 show the read back signal of the read head with and without<br />

shields for the different k<strong>in</strong>d of transitions. The head velocity was 20 m/s. For this calculation<br />

the Gilbert damp<strong>in</strong>g constant was 0.3 and 1.0 for the GMR sensor and the shield respectively.<br />

The signal peaks for the full read head simulation (with shields) are more localized compared<br />

to the simulation without shields. The asymmetric transfer curve leads to a large difference <strong>in</strong><br />

amplitude height for both cases. Aga<strong>in</strong> the output voltage is shifted up for the read back signal<br />

with shields due to mirror currents, as we have already seen <strong>in</strong> Figure 6.7.<br />

Output Voltage [V]<br />

0.1070<br />

0.1065<br />

0.1060<br />

0.1055<br />

0.1050<br />

0.1045<br />

0.1040<br />

-300 -200 -100 0 100 200 300<br />

x [nm]<br />

ΔR/R [%]<br />

0.2<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

-1.2<br />

-1.4<br />

-1.6<br />

-300 -200 -100 0 100 200 300<br />

without shields<br />

with shields<br />

x [nm]<br />

Figure 6.9: The read back signal and the relative change <strong>in</strong> resistance for a perfect →←<br />

transition.<br />

74


Output Voltage [V]<br />

0.1074<br />

0.1072<br />

0.1070<br />

0.1068<br />

0.1066<br />

0.1064<br />

0.1062<br />

0.1060<br />

x [nm]<br />

FEM Simulations<br />

0.1058<br />

-300 -200 -100 0 100 200 300<br />

ΔR/R [%]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

-300 -200 -100 0 100 200 300<br />

without shields<br />

with shields<br />

x [nm]<br />

Figure 6.10: The read back signal and the relative change <strong>in</strong> resistance for a perfect ←→<br />

transition.<br />

6.2.6 <strong>Read</strong> <strong>Back</strong> Signal<br />

F<strong>in</strong>ally the read back signal was calculated. As replacement for the data layer, a bit pattern as<br />

shown <strong>in</strong> Figure 6.11 was used. It consists of 30 bits with perfect transitions, which are<br />

alternately magnetized <strong>in</strong> positive and negative x-direction. The assumed bit length was<br />

60 nm, which is equal to the gap of the read head model. The track width was 120 nm, which<br />

is larger than the free layer width. The thickness was 12 nm and the fly<strong>in</strong>g height of the head,<br />

i.e. the distance between ABS and the data layer was 10 nm. The spontaneous polarization<br />

was 0.44 T.<br />

Figure 6.12 shows the read back signal for a read head with and without shields. Accord<strong>in</strong>g to<br />

the transfer curve the output signal for the whole read head (with shields) is aga<strong>in</strong> shifted up,<br />

due to mirror current effects. The shields lead to an <strong>in</strong>crease of the amplitude. Therefore the<br />

read back signal is much better, although the signal for s<strong>in</strong>gle perfect transitions was lower<br />

than that without shields (see previous section). This fact can be traced back to the wider<br />

peaks <strong>in</strong> the absence of the shields, so that the signals of consecutive transitions partially<br />

cancel each other.<br />

75


Output Voltage [V]<br />

0.1080<br />

0.1075<br />

0.1070<br />

0.1065<br />

0.1060<br />

0.1055<br />

0.1050<br />

FEM Simulations<br />

Figure 6.11: The model of a perfect bit pattern. It consists of 30 perfectly written bits and<br />

is used for our read back signal calculations. The track width is 120 nm and the bit length<br />

60 nm. The layer thickness is 12 nm and the spontaneous polarization is 0.44 T. The bits<br />

are alternately magnetized <strong>in</strong> positive and negative x-direction.<br />

0.1045<br />

0 100 200 300 400<br />

x [nm]<br />

ΔR/R [%]<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

0 100 200 300 400<br />

without shields<br />

with shields<br />

x [nm]<br />

Figure 6.12: The read back signal for the periodic bit pattern calculated for the GMR<br />

sensor and for the whole read head (with shields).<br />

76


7 OutlookEquation<br />

Section (Next)<br />

Outlook<br />

This work was a first step to take arbitrary current distributions <strong>in</strong>to account <strong>in</strong> micromagnetic<br />

simulations. The program solves the LLG equation and simultaneously calculates the current<br />

distribution, which can be <strong>in</strong>fluenced by the magnetization. So it can be used for read back<br />

signal calculations. Of course it is also possible to calculate the read back signal of<br />

perpendicular record<strong>in</strong>g layers. Here the additional SUL has to be taken <strong>in</strong>to account. Sp<strong>in</strong><br />

valves with synthetic antiferromagnet can also be calculated, if the <strong>in</strong>terlayer exchange<br />

coupl<strong>in</strong>g is <strong>in</strong>troduced.<br />

Possible enhancements of this program are the consideration of eddy currents [41] [42] or<br />

f<strong>in</strong>ite temperatures. To take fluctuations at f<strong>in</strong>ite temperatures <strong>in</strong>to account, a stochastic,<br />

thermal field H th is added to the effective field <strong>in</strong> the LLG equation [38]<br />

∂J α ∂J<br />

=−γ J× ( Heff + Hth) + J × . (7.1)<br />

∂t J ∂t<br />

s<br />

Future hard disks will make use of sp<strong>in</strong> valves operat<strong>in</strong>g <strong>in</strong> CPP mode. Therefore simulations<br />

of CPP sp<strong>in</strong> valves become more and more <strong>in</strong>terest<strong>in</strong>g. A current perpendicular to the layer<br />

structure leads to sp<strong>in</strong> transfer, which has to be taken <strong>in</strong>to account by an additional term,<br />

called sp<strong>in</strong> torque term [39]<br />

∂J α ∂J<br />

Cj<br />

=−γ J× H ′<br />

eff + J× + J× J × J. (7.2)<br />

∂t J ∂t<br />

J J′<br />

2<br />

s s s<br />

Here J ′ is the magnetization of a reference layer, which polarizes the electrons. The polarized<br />

current transfers the sp<strong>in</strong> to the actual magnetic polarization J and causes an additional<br />

motion of J. j is the local current density. For CPP sp<strong>in</strong> valves the magnetization is transferred<br />

from the p<strong>in</strong>ned layer to the free layer or vice versa.<br />

The application of this program is not restricted to read head simulations. It can also be<br />

applied, for example, on magnetic tunnel junctions (MTJs), which are used <strong>in</strong> MRAMs<br />

(<strong>Magnetic</strong> Random Access Memory).<br />

77


Appendix A<br />

The theorem<br />

Appendix A<br />

∫grad f dV = � ∫ f dA<br />

(A.1)<br />

V ∂V<br />

is often used <strong>in</strong> literature, but it is normally limited to cont<strong>in</strong>uously differentiable functions.<br />

Nevertheless this relation is also frivolously applied for functions with first order<br />

s<strong>in</strong>gularities, as for example<br />

1<br />

f () r = with V<br />

r−r ′<br />

′∈ r . (A.2)<br />

We were not able to f<strong>in</strong>d any proof, which shows the validity of the above theorem for this<br />

function. Thus it is given here by our own.<br />

3<br />

Let us assume any constant vectorc<br />

∈ � and start from the equality<br />

divcf= c grad f . (A.3)<br />

Integration over the whole volume V gives<br />

divcfdV = � cgrad<br />

f dV . (A.4)<br />

∫ ∫<br />

V V<br />

If f has cont<strong>in</strong>uous first order partial derivatives we can apply the Divergence Theorem<br />

result<strong>in</strong>g <strong>in</strong><br />

� ∫ c f dA= ∫cgrad<br />

f dV . (A.5)<br />

∂V<br />

V<br />

c is constant and therefore it can be brought <strong>in</strong> front of the <strong>in</strong>tegral.<br />

c� ∫ f dA= c∫grad<br />

f dV . (A.6)<br />

∂V<br />

V<br />

Last equation holds for any c, therefore c can be cancelled, and the theorem (A.1) is verified.<br />

78


Appendix A<br />

In case we have the function as given <strong>in</strong> (A.2), we have to divide the volume V <strong>in</strong>to a small<br />

sphere with radius ε around the s<strong>in</strong>gularity r ′ denoted with Kε ( r ′ ) and the rest of V. Then<br />

the theorem (A.1) can be written as<br />

∫ grad f dV + ∫ grad f dV = �∫ f dA− �∫ f dA+ � ∫ f dA.<br />

(A.7)<br />

V / Kε( r′ ) Kε( r′ ) ∂V ∂Kε( r′ ) ∂Kε(<br />

r′<br />

)<br />

For the volume V / Kε ( r ′ ) with the outer surface ∂ V and the <strong>in</strong>ner surface ∂Kε ( r ) the<br />

theorem holds, because the s<strong>in</strong>gularity is excluded. To prove the theorem, we only have to<br />

show, that<br />

∫ ∫<br />

grad f dV = � f dA<br />

. (A.8)<br />

Kε( r′ ) ∂Kε(<br />

r′<br />

)<br />

The equality is valid, if both sides vanish for ε → 0 . Without loss of generality, it is sufficient<br />

to show the equality for the first vector component. Then the <strong>in</strong>tegrand of the left hand side is<br />

( f ) = 3<br />

grad x<br />

x − x′<br />

, (A.9)<br />

r−r′ which is asymmetric around r ′ . Therefore the left hand side <strong>in</strong> (A.8) vanishes. For the right<br />

hand side we have<br />

1<br />

⎛ ⎞<br />

4πε<br />

= ∫ dA = f dA ≥ ⎜ f d ⎟<br />

Kε( ′ − ′ ∫ ⎜ ∫ A<br />

r r<br />

⎟<br />

∂ r ) ∂Kε( r′ ) ⎝∂Kε( r′<br />

) ⎠<br />

� � � . (A.10)<br />

So this side also vanishes, if ε→ 0 , q.e.d.<br />

x<br />

79


Appendix B<br />

Appendix B<br />

To calculate the magnetic field of conductive materials (4.50) has to be solved:<br />

Δ H =−curl<br />

j. (B.1)<br />

In addition boundary conditions are needed to make this problem solvable. The derivation of<br />

the boundary conditions for a surface po<strong>in</strong>t r is given here. We have to calculate<br />

∂ ∂<br />

∂n ∂n<br />

out <strong>in</strong><br />

( H () r − H () r ) = lim ( H( r+ nε) −H( r−nε) )<br />

ε→0<br />

. (B.2)<br />

n denotes the unit vector normal to the boundary face <strong>in</strong> po<strong>in</strong>t r. We start from the Biot-<br />

Savart Law (2.46)<br />

1 r± nε−r′ Hr ( ± nε ) = ×<br />

4π ∫ j<br />

V r± nε−r′ Now we separate a half sphere<br />

Volume V.<br />

3<br />

dV ′ . (B.3)<br />

1/2<br />

Kη () r with radius η (η � ε ) and midpo<strong>in</strong>t r from the<br />

1 r± nε− r′ 1 r± nε−r′ Hr ( ± nε ) = × dV ′ + × dV ′<br />

4π ∫ j<br />

4π<br />

∫ j<br />

3 3<br />

1/2 1/2<br />

V / Kη ( r) r± nε− r′ Kη<br />

( r)<br />

r± nε−r′ (B.4)<br />

In this half sphere the current density j is assumed to be constant. If j is a cont<strong>in</strong>uous function,<br />

1/2<br />

the mean value theorem tells us that there exists a po<strong>in</strong>t r ′′ ∈ K for which<br />

± η<br />

1 r± nε− r′ 1 r± nε−r′ Hr ( ± nε ) = × dV ′ + ( ′′ ) ×<br />

dV ′<br />

4π ∫ j jr<br />

4π<br />

∫<br />

. (B.5)<br />

3 ±<br />

3<br />

1/2 1/2<br />

V / Kη ( r) r± nε− r′ Kη<br />

( r)<br />

r± nε−r′ The <strong>in</strong>tegrand of the second <strong>in</strong>tegral can be expressed as gradient of a scalar term<br />

1 r± nε−r′ 1 1<br />

Hr ( ± nε ) = × dV ′ + ( ′′ ) × ∇′<br />

dV ′<br />

4π ∫ j jr<br />

4π<br />

∫<br />

. (B.6)<br />

r± nε−r′ 3<br />

±<br />

1/2 1/2<br />

V / Kη ( r) r± nε−r′ Kη<br />

( r)<br />

80


Appendix B<br />

Moreover the second <strong>in</strong>tegral can be transformed <strong>in</strong>to a surface <strong>in</strong>tegral by apply<strong>in</strong>g the<br />

theorem, proved <strong>in</strong> Appendix A<br />

1 r± nε−r′ 1 1<br />

Hr ( ± nε ) = × dV ′ + ( ′′ ) ×<br />

d ′<br />

4π ∫ j j r<br />

4π<br />

� ∫<br />

A . (B.7)<br />

r n r r± nε−r′ 3<br />

±<br />

1/2 1/2<br />

V / Kη ( r) ± ε− ′<br />

∂Kη<br />

( r)<br />

Now we are able to calculate the difference <strong>in</strong> (B.2). Due to the difference and ε→ 0 the first<br />

<strong>in</strong>tegral and the rotund surface of the half sphere do not contribute ( η � ε ), which leads to<br />

1 dA′ 1 dA′<br />

Hr ( + nε) −Hr ( −nε ) = jr ( ′′ ) × − ( ′′ ) ×<br />

π ∫ jr<br />

+ ε− ′ π∫ . (B.8)<br />

r n r r−nε−r ′<br />

+ −<br />

4 4<br />

Cη( r) Cη(<br />

r)<br />

Here Cη( r ) denotes a circle with radius η and radius r. Derivation <strong>in</strong> direction n gives<br />

∂ 1 εdA′<br />

∂n4π C r+ nε−r′ ( Hr ( + nε) −Hr ( −nε ) ) =− ( jr ( ′′ + ) + jr ( ′′ −)<br />

) ×<br />

3<br />

∫<br />

η ( r)<br />

. (B.9)<br />

The substitution ρ= r−r ′ and the simultaneous consideration that n is perpendicular to<br />

( − ′ )<br />

r r lead to<br />

∂ 1 ε<br />

( Hr ( + nε) −Hr ( −nε ) ) =− ( jr ( ′′ + ) + jr ( ′′ −)<br />

) × n<br />

ρdρdϕ. (B.10)<br />

3/2<br />

∂n4π ρ +ε<br />

Calculat<strong>in</strong>g the <strong>in</strong>tegral yields<br />

2π<br />

η<br />

∫∫ 2 2<br />

0 0(<br />

)<br />

∂ out <strong>in</strong> jr ( ′′ + ) + jr ( ′′ ) ⎛<br />

− ε ⎞<br />

( H () r − H () r ) = n×<br />

⎜1− ⎟.<br />

(B.11)<br />

∂n 2 ⎜ 2 2<br />

η +ε ⎟<br />

⎝ ⎠<br />

Now the limit can be calculated <strong>in</strong> (B.2)<br />

∂<br />

∂n<br />

out <strong>in</strong> j( r′′ + ) + j( r′′<br />

−)<br />

H r − H r = n×<br />

. (B.12)<br />

2<br />

( () () )<br />

And f<strong>in</strong>ally, for η→ 0 we get the result<br />

∂<br />

∂n out <strong>in</strong><br />

( )<br />

H () r − H () r = n× j() r . (B.13)<br />

In case of an <strong>in</strong>terface between two volumes V 1 and 2<br />

<strong>in</strong>terface j 1 and j 2 respectively, we have a jump for the normal derivation of H<br />

V with different current densities at the<br />

81


∂<br />

∂n 2 1 2 1<br />

( H () r − H () r ) =− n× ( j () r −j()<br />

r )<br />

Appendix B<br />

. (B.14)<br />

82


List of Figures<br />

List of Figures<br />

Figure 1.1: The orig<strong>in</strong>al draw<strong>in</strong>g of Oberl<strong>in</strong> Smith (1888) and Valdemar Poulson’s<br />

telegraphone [2].................................................................................................................7<br />

Figure 1.2: The first computer disk drive RAMAC. The pneumatically driven actuator<br />

arm and some of the 50 disks are visible [4].....................................................................8<br />

Figure 1.3: The development of bit sizes and storage densities over the last years [7]. ........10<br />

Figure 2.1: a) Illustration of the damped precession of the magnetic polarization J<br />

around the effective field Heff, taken from [15]. b) Influence of the Gilbert<br />

damp<strong>in</strong>g constant on the field rise time <strong>in</strong> magnetic read heads. Head field<br />

strength as a function of time for different Gilbert damp<strong>in</strong>g constants with<strong>in</strong> the<br />

head [43]. The dashed l<strong>in</strong>e depicts the current profile. The solid l<strong>in</strong>es give the<br />

head field for α = 1, 0.5, 0.1, and 0.02. The fastest field rise time is achieved with<br />

<strong>in</strong>termediate damp<strong>in</strong>g (α = 0.5).......................................................................................20<br />

Figure 2.2: The sp<strong>in</strong>-split bands of a ferromagnet [18]..........................................................21<br />

Figure 2.3: Giant magnetoresistance <strong>in</strong> Fe/Cr multilayers [18]. ............................................22<br />

Figure 2.4: Schematic model of a sp<strong>in</strong>-valve GMR head (CIP mode) with exchangep<strong>in</strong>ned<br />

layer and longitud<strong>in</strong>al hard bias [19]...................................................................23<br />

Figure 2.5: Change <strong>in</strong> resistance of a sp<strong>in</strong> valve compared to AMR (NiFe) vs.<br />

magnetization orientation [19]. .......................................................................................23<br />

Figure 2.6: The two different modes of a sp<strong>in</strong> valve. The current is applied either<br />

longitud<strong>in</strong>al (CIP) or perpendicular (CPP) to the layer structure [7]. .............................24<br />

Figure 2.7: Potential landscape <strong>in</strong>side the FM/NM/FM active part of a sp<strong>in</strong>-valve, for<br />

the two sp<strong>in</strong> channels <strong>in</strong> the parallel state (a,b) and <strong>in</strong> the antiparallel state (c),<br />

<strong>in</strong>dicat<strong>in</strong>g the different orig<strong>in</strong>s of scatter<strong>in</strong>g [20]............................................................25<br />

Figure 2.8: A comparison of longitud<strong>in</strong>al record<strong>in</strong>g and perpendicular record<strong>in</strong>g [7]. .........27<br />

Figure 3.1: A schematic 2-dimensional model of a read head for longitud<strong>in</strong>al<br />

record<strong>in</strong>g..........................................................................................................................30<br />

Figure 3.2: The magnetic potential at the ABS under Karlquist approximation....................31<br />

Figure 3.3: The x- and the z-component of the normalized head field of a read head<br />

with 60 nm gap, and sens<strong>in</strong>g layer thickness 5 nm, 10 nm below the ABS....................33<br />

Figure 3.4: The effective signal field as function of the relative position of the read<br />

head <strong>in</strong> respect to a perfect transition (Js = 0.5 T)...........................................................35<br />

Figure 3.5: These two figures show the x-component of the magnetization (Js = 0.5 T)<br />

of the data layer after a write process with head velocity 10 m/s (left) and 20 m/s<br />

83


List of Figures<br />

(right). Red color represents magnetization <strong>in</strong> positive x-direction and blue color<br />

magnetization <strong>in</strong> negative x-direction. The scale is <strong>in</strong> µm, thus the bit length is<br />

approximately 40 nm. The data layer thickness is 12 nm. ..............................................35<br />

Figure 3.6: The effective fields of the record<strong>in</strong>g layers (shown <strong>in</strong> Figure 3.5) as<br />

function of the head’s position. The dashed l<strong>in</strong>es <strong>in</strong>dicate the boundaries of the<br />

record<strong>in</strong>g layer model......................................................................................................36<br />

Figure 3.7: The simplified read head model (left) for perpendicular record<strong>in</strong>g. In the<br />

right picture the mirror heads are <strong>in</strong>dicated, which replace the SUL with its high<br />

permeability [22]. ............................................................................................................37<br />

Figure 3.8: The x- and the z-component of the normalized head field for perpendicular<br />

record<strong>in</strong>g..........................................................................................................................38<br />

Figure 3.9: a) Application of Ampere’s Law on the dashed path, b) application of<br />

Gauss’s Law on the dashed box. .....................................................................................39<br />

Figure 4.1: Typical basis functions of two neighbor<strong>in</strong>g nodes of a 2-dimensional mesh......43<br />

Figure 4.2: The left figure shows a set of 10 nodes after renumber<strong>in</strong>g and the<br />

cluster<strong>in</strong>g. The right figure shows the associated boundary matrix structure. The<br />

large off-diagonal blocks, correspond<strong>in</strong>g to far field <strong>in</strong>teractions of two clusters,<br />

can be approximated by low-rank matrices [25]. ............................................................49<br />

Figure 4.3: Hϕ for an <strong>in</strong>f<strong>in</strong>ite long wire is plotted over the distance from the middle<br />

axis. The second curve shows the error of the numerical calculation.............................55<br />

Figure 4.4: The model of the coil (left). The color code shows the l<strong>in</strong>ear decrease of<br />

the electric potential <strong>in</strong>side the wire. The chart (right) shows the magnetic field of<br />

a solenoid <strong>in</strong> z-direction at the middle axis. The two dotted l<strong>in</strong>es identify the<br />

position of the solenoid. The second graph shows the relative error of the coil’s<br />

magnetic field calculated with FEM calculation compared to the solenoid’s field. .......56<br />

Figure 4.5: Schematic diagram of the simulation cycle. Adapted from [31]. ........................58<br />

Figure 5.1: The shift of the hysteresis loop after a field cool<strong>in</strong>g treatment of a<br />

ferromagnetic/antiferromagnetic system [19] and the temperature dependence of<br />

the exchange bias field for NiFe(4-8nm)/Cu(2.2nm)/CoFe(2nm)/IrMn(15nm) sp<strong>in</strong><br />

valves with different free layer thicknesses [33].............................................................60<br />

Figure 5.2: A comparison (left) of a conventional sp<strong>in</strong> valve and a sp<strong>in</strong> valve us<strong>in</strong>g a<br />

synthetic ferromagnet to fix the p<strong>in</strong>ned layer [20]. The <strong>in</strong>terlayer exchange<br />

constant (right) for two Ni80Co20 layers as function of the Ru spacer layer<br />

thickness [32]...................................................................................................................62<br />

Figure 5.3: The effect of the sense current direction on the free layer (FL). For the first<br />

case (left) the current field Hcurr partially cancels the stray field HS of the p<strong>in</strong>ned<br />

layer (PL) and weakens the exchange bias field Hexch. For the other current<br />

direction (right) the stray field and the exchange bias field are supported by the<br />

current field. ....................................................................................................................63<br />

Figure 5.4: The layers of the GMR element also see mirror currents due to the high<br />

permeability of the shields. The field of these mirror currents is able to reduce the<br />

field of the orig<strong>in</strong>al current at the free layer and the p<strong>in</strong>ned layer. .................................63<br />

84


List of Figures<br />

Figure 6.1: The magnetic model used for the FEM calculations. The mesh size for the<br />

shields (green) is 50 nm, and is ref<strong>in</strong>ed to 5 nm near the track and near the sp<strong>in</strong><br />

valve layers. The hard bias magnets (purple), the free layer (dark blue), and the<br />

p<strong>in</strong>ned layer (light blue) are also meshed with triangle size of 5 nm..............................66<br />

Figure 6.2: The FEM conductor model is shown. This simplified model shows the four<br />

ma<strong>in</strong> layers of a GMR element, which contribute most to the conductance: The<br />

free layer (dark blue) the nonmagnetic Cu layer (orange), the p<strong>in</strong>ned layer (light<br />

blue), and the antiferromagnet (yellow). Additionally the high conductive leads<br />

(p<strong>in</strong>k) are shown..............................................................................................................67<br />

Figure 6.3: The effective conductivities for a typical sp<strong>in</strong> valve taken from [37]. ................68<br />

Figure 6.4: The normalized magnetization of the free layer and the p<strong>in</strong>ned layer <strong>in</strong><br />

equilibrium state for the two different current directions. The color code gives the<br />

z-component of the normalized magnetization, red corresponds to the positive and<br />

blue to the negative z-direction. Thus the red arrows represent the p<strong>in</strong>ned layer<br />

magnetization. The right hand side gives the configuration for the more favorable<br />

current direction, because here the current field cancels partially the<br />

demagnetiz<strong>in</strong>g field of the p<strong>in</strong>ned layer. .........................................................................69<br />

Figure 6.5: The external applied field, which is applied <strong>in</strong> z-direction to evaluate the<br />

transfer curve of the GMR sensor. ..................................................................................70<br />

Figure 6.6: The output voltage (left) and the relative change <strong>in</strong> resistance (right, <strong>in</strong><br />

respect to the equilibrium state) over the external applied field <strong>in</strong> z-direction for<br />

both current direction and a larger exchange bias field...................................................71<br />

Figure 6.7: The transfer curves for the sp<strong>in</strong> valve with and without shields (for both<br />

Hexch = 0.1 T). ..................................................................................................................72<br />

Figure 6.8: The output signal over time for a relaxation from equilibrium state with<br />

Hext = 0.3 T to the equilibrium state without field for different Gilbert damp<strong>in</strong>g<br />

constants. .........................................................................................................................73<br />

Figure 6.9: The read back signal and the relative change <strong>in</strong> resistance for a perfect<br />

→← transition................................................................................................................74<br />

Figure 6.10: The read back signal and the relative change <strong>in</strong> resistance for a perfect<br />

←→ transition................................................................................................................75<br />

Figure 6.11: The model of a perfect bit pattern. It consists of 30 perfectly written bits<br />

and is used for our read back signal calculations. The track width is 120 nm and<br />

the bit length 60 nm. The layer thickness is 12 nm and the spontaneous<br />

polarization is 0.44 T. The bits are alternately magnetized <strong>in</strong> positive and negative<br />

x-direction........................................................................................................................76<br />

Figure 6.12: The read back signal for the periodic bit pattern calculated for the GMR<br />

sensor and for the whole read head (with shields). .........................................................76<br />

85


Bibliography<br />

Bibliography<br />

[1] O. Smith, Electrical World, September 8, 1888.<br />

[2] S. E. Schoenherr, Record<strong>in</strong>g Technology History, http://history.acusd.edu/gen/<br />

record<strong>in</strong>g/notes.html, University of San Diego, January 27, 2005.<br />

[3] A. S. Hoagland, “History of <strong>Magnetic</strong> Disc Storage Based on Perpendicular<br />

Record<strong>in</strong>g”, IEEE Trans. Magn., vol. 39, no. 4, pp. 1871-1875, July 2003.<br />

[4] B. Hayes, “Terabyte Territory”, American Scientist, vol. 90, no. 3, pp. 212-216,<br />

May/June 2002.<br />

[5] R. Hunt, “A magnetoresistive readout transducer”, IEEE Trans. Magn., vol. MAG-7,<br />

no. 1, pp. 150-154, March 1971.<br />

[6] M. N. Baibich, J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P. Eitenne, G.<br />

Creuzet, A. Friederich, J. Chazelas, “Giant Magnetoresistance of (001)Fe/(001)Cr<br />

<strong>Magnetic</strong> Superlattices“, Phys. Rev. Lett., vol. 61, no. 21, pp. 2472-2475, November<br />

1988.<br />

[7] Hitachi Global Storage Technologies, Hitachi <strong>Research</strong> and Technology, http://www.<br />

hitachigst.com/hdd/research/, April 22, 2005.<br />

[8] E. E. Fullerton, D. T. Margulies, M. E. Schabes, M. Carey, B. Gurney, A. Moser.,<br />

M. Best, G. Zeltzer, K. Rub<strong>in</strong>, H. Rosen, M. Doerner, “Antiferromagnetically<br />

coupled magnetic media layers for thermally stable high-density record<strong>in</strong>g”, Appl. Phys.<br />

Lett., vol. 77, no. 23, pp. 3806-3808, December 2000.<br />

[9] E. E. Fullerton, D. T, Margulies, N. Supper, H. Do, M. Schabes, A. Berger, A.<br />

Moser, “Antiferromagnetically Coupled <strong>Magnetic</strong> Record<strong>in</strong>g Media”, IEEE Trans.<br />

Magn., vol. 39, no. 2, pp. 639-644, March 2003.<br />

[10] C. T. Rettner, S. Anders, T. Thomson, M. Albrecht, Y. Ikeda, M. E. Best, B. D.<br />

Terris, “<strong>Magnetic</strong> Characterization and Record<strong>in</strong>g Properties of Patterned Co70Cr18Pt12<br />

Perpendicular Media”, IEEE Trans. Magn., vol. 38, no. 4, pp. 1725-1730, July 2002.<br />

86


Bibliography<br />

[11] H. Katayama, M. Hamamoto, J. Sato, Y. Murakami, K. Kojima, “New<br />

Developments <strong>in</strong> Laser-Assisted <strong>Magnetic</strong> Record<strong>in</strong>g”, IEEE Trans. Magn., vol. 36, no.<br />

1, pp. 195-199, January 2000<br />

[12] P. Vettiger, M. Despont, U. Drechsler, U. Dürig, W. Häberle, M. I. Lutwyche, H. E.<br />

Rothuizen, R. Stutz, R. Widmer, G. K. B<strong>in</strong>nig, “The ‘Millipede’ – more than one<br />

thousand tips for future AFM data storage”, IBM J. Res. Dev., vol. 44, no. 3, pp. 323-<br />

340, May 2000.<br />

[13] P. Mohn, Magnetism <strong>in</strong> the Solid State: An Introduction, Spr<strong>in</strong>ger Verlag, Berl<strong>in</strong>, 2003.<br />

[14] W. Bailey, P. Kabos, F. Mancoff, S. Russek, “Control of Magnetization Dynamics <strong>in</strong><br />

Ni81Fe19 Th<strong>in</strong> Films Through the Use of Rare-Earth Dopants”, IEEE Trans. Magn., vol.<br />

37, pp. 1749-1754, July 2001.<br />

[15] M. Kirschner, Exchange Bias <strong>in</strong> Ferro-/Antiferromagnetic Bilayers, Diploma Thesis,<br />

Work<strong>in</strong>g <strong>Group</strong> <strong>Magnetic</strong> Materials and Micromagnetics, Vienna University of<br />

Technology, 2003.<br />

[16] N. F. Mott, “The Electrical Conductivity of Transition Metals“, Proc. Roy. Soc. (Lond.)<br />

A, no. 153, pp. 699-717, 1935.<br />

[17] N. F. Mott, “Electrons <strong>in</strong> Transition Metals“, Adv. Phys., no. 13, pp. 325-422, 1964.<br />

[18] S. Blundell, Magnetism <strong>in</strong> condensed matter, Oxford Univ. Press, New York, 2001.<br />

[19] S. X. Wang, A. M. Tarator<strong>in</strong>, <strong>Magnetic</strong> Information Storage Technology, Academic<br />

Press, San Diego, 1999.<br />

[20] R. Coehoorn, Novel Magnetoelectronic Materials and Devices, Lecture Notes, 2003.<br />

[21] M. M. Aziz, Signal and Noise Properties of Longitud<strong>in</strong>al Th<strong>in</strong>-Film Disk Media,<br />

Master Thesis, Electronic and Information Storage Systems <strong>Research</strong> <strong>Group</strong>, University<br />

of Manchester, 1999.<br />

[22] H. Muraoka, Y.Sugita, Y. Nakamura, “Simplified Expression of Shielded MR Head<br />

Response for Double-Layer Perpendicular Medium“, IEEE Trans. Magn., vol. 35, no. 5,<br />

pp. 2235-2237, September 1999.<br />

[23] T. A. Davis, Algorithm 8xx: a concise sparse Cholesky factorization package,<br />

http://www.cise.ufl.edu/research/sparse/ldl/, January 5, 2004.<br />

[24] D. R. Fredk<strong>in</strong>, T. R. Koehler, “Hybrid Method for Comput<strong>in</strong>g Demagnetiz<strong>in</strong>g Fields“,<br />

IEEE Trans. Magn., vol. 26, no. 2, pp. 415-417, March 1990.<br />

87


Bibliography<br />

[25] S. Kurz, O. Ra<strong>in</strong>, S. Rjasanow, “The adaptive cross-approximation technique for the<br />

3D boundary-element method”, IEEE Trans. Magn., vol. 38, no. 2, pp. 421-424, March<br />

2002.<br />

[26] D. M. Mount, ANN Programm<strong>in</strong>g Manual, http://www.cs.umd.edu/~mount/ANN/,<br />

University of Maryland, 2005.<br />

[27] D. A. L<strong>in</strong>dholm, “Three-Dimensional Magnetostatic Fields from Po<strong>in</strong>t-Matched<br />

Integral Equations with L<strong>in</strong>early Vary<strong>in</strong>g Scalar Sources”, IEEE Trans. Magn. , vol. 20,<br />

no. 5, pp. 2025-2032, September 1984.<br />

[28] S. Pissanetzky, “Comments on ‘Analytical Computation of <strong>Magnetic</strong> Vector Potential<br />

for Tetrahedral Conductors’”, IEEE Trans. Magn., vol. 29, no. 2, pp. 1282-1283, March<br />

1993.<br />

[29] S. D. Cohen, A. C. H<strong>in</strong>dmarsh, “CVODE User Guide“, Lawrence Livermore National<br />

Laboratory report, UCRL-MA-118618, 1994.<br />

[30] S. D. Cohen, A. C. H<strong>in</strong>dmarsh, “CVODE, a stiff/nonstiff ODE solver <strong>in</strong> C”,<br />

Computers <strong>in</strong> Physics, vol. 10, no. 2, pp. 138-143, March/April 1996.<br />

[31] P. Speckmayer, Electron Microscopy and Micromagnetic Simulation of FePt and CoPt<br />

Nanoparticles, Diploma Thesis, Work<strong>in</strong>g <strong>Group</strong> <strong>Magnetic</strong> Materials and<br />

Micromagnetics, Vienna University of Technology, 2004.<br />

[32] W. H. Bean and C. P. Meiklejohn, “New <strong>Magnetic</strong> Anisotropy”, Phys. Rev., vol. 102,<br />

pp. 1413, 1956.<br />

[33] S. Tanoue, K. Tabuchi, “Magnetoresistance characteristics of NiFe/Cu/CoFe/IrMn<br />

sp<strong>in</strong> valves”, J. Vac. Sci. Technol. B, vol. 19, no. 2, pp. 563-566, March 2001.<br />

[34] J. Nogués, C. Leighton, I. K. Schuller, “Correlation between antiferromagnetic<br />

<strong>in</strong>terface coupl<strong>in</strong>g and exchange bias“, Phys. Rev. B, vol. 61, pp. 1315, 2000.<br />

[35] H. A. M. Van den Berg, W. Clemens, G. Gieres, G. Rupp, W. Schelter, M. Veith,<br />

“GMR sensor scheme with artificial antiferromagnetic subsystem”, IEEE Trans. Magn.,<br />

vol. 32, no. 5, pp. 4624-4626, September 1996.<br />

[36] P. Bruno, “Theory of <strong>in</strong>terlayer exchange coupl<strong>in</strong>g”, Phys. Rev. B, vol. 52, no. 1, pp.<br />

411-439, July 1995.<br />

[37] A. T. McCallum, S. E. Russek, “Current density distribution <strong>in</strong> a sp<strong>in</strong> valve<br />

determ<strong>in</strong>ed through <strong>in</strong> sito conductance measurement”, Appl. Phys. Lett., vol. 84, no.<br />

17, pp. 3340-3342, April 2004.<br />

88


Bibliography<br />

[38] W. F. Brown Jr., Micromagnetics, Wiley, New York, 1963.<br />

[39] S. J. Slonczewski, “Current-driven excitation of magnetic multilayers”, J. Magn. Magn.<br />

Mater., vol. 159, no. 1-2, pp. L1-L7, June 1996.<br />

[40] R. L. Comstock, Introduction to Magnetism and <strong>Magnetic</strong> Record<strong>in</strong>g, Wiley, New<br />

York, 1999.<br />

Parts of this work have been published:<br />

[41] G. Hrkac, T. Schrefl, M. Kirschner, F. Dorfbauer, D. Suess, O. Ertl, J. <strong>Fidler</strong>,<br />

“Three-dimenisonal micromagnetic f<strong>in</strong>ite element simulations <strong>in</strong>clud<strong>in</strong>g eddy currents“,<br />

J. Appl. Phys., vol. 97, no. 10, May 2005, <strong>in</strong> press.<br />

[42] G. Hrkac, T. Schrefl, O. Ertl, M. Kirschner, D. Suess, J. <strong>Fidler</strong>, “Influence of Eddy<br />

Currents on Magnetization Processes <strong>in</strong> Sub-Micron Permalloy Structures”, IEEE<br />

Trans. Magn., <strong>in</strong> press.<br />

[43] O. Ertl, T. Schrefl, D. Suess, M. E. Schabes, “Influence of the Gilbert damp<strong>in</strong>g<br />

constants on the flux rise time of write head fields“, J. Magn. Magn. Mater., vol. 290-<br />

291, part 1, pp. 518-521, April 2005.<br />

[44] M. E. Schabes, T. Schrefl, D. Suess, O. Ertl, “Dynamic Micromagnetic Studies of<br />

Anisotropy Effects <strong>in</strong> Perpendicular Write Heads”, IEEE Trans. Magn., <strong>in</strong> press.<br />

[45] T. Schrefl, M. E. Schabes, D. Suess, O. Ertl, M. Kirschner, F. Dorfbauer, G.<br />

Hrkac, J. <strong>Fidler</strong>, “Partition<strong>in</strong>g of the Perpendicular Write Field <strong>in</strong>to Head and SUL<br />

Contributions”, IEEE Trans. Magn., <strong>in</strong> press.<br />

89


Danksagung<br />

Ich möchte all jenen me<strong>in</strong>en Dank aussprechen, die mich bei der Diplomarbeit unterstützt<br />

oder motivierend auf mich e<strong>in</strong>gewirkt haben.<br />

Insbesondere möchte ich mich bei me<strong>in</strong>em Betreuer Doz. Thomas Schrefl für se<strong>in</strong>e ständige<br />

Bereitschaft, mir bei Fragen oder Anliegen weiterzuhelfen, bedanken. Weiters seien Florian<br />

Dorfbauer, G<strong>in</strong>o Hrkac, Markus Kirschner und Dieter Süss erwähnt, die mir viele hilfreiche<br />

Tipps zur Diplomarbeit gaben. Generell möchte ich allen danken, die zum ausgezeichneten<br />

Arbeitsklima <strong>in</strong>nerhalb der Arbeitsgruppe Mikromagnetismus beitrugen.<br />

Chee Lim vom Samsung Advanced Institute of Technology möchte ich für die (wenigen)<br />

Informationen zu aktuellen Leseköpfen danken.<br />

Me<strong>in</strong>en Eltern möchte ich danken, dass sie mir durch ihre f<strong>in</strong>anzielle Unterstützung das<br />

Studium ermöglicht haben.<br />

Me<strong>in</strong>er Freund<strong>in</strong> Katr<strong>in</strong> danke ich für ihre Unterstützung während me<strong>in</strong>er Diplomarbeit.<br />

90


Lebenslauf<br />

Otmar Ertl<br />

Re<strong>in</strong>prechtsdorferstraße 20/16<br />

1050 Wien<br />

Tel.: +43/650/9461912<br />

PERSÖNLICHE ANGABEN<br />

AUSBILDUNG<br />

AUSZEICHNUNGEN<br />

� Staatsangehörigkeit: Österreich<br />

� Geburtsdatum: 22. Februar 1982<br />

� Geburtsort: Vöcklabruck<br />

� Eltern: Mag. Karl Ertl und Renate Ertl<br />

� 1988 bis 1992 Volksschule Vöcklabruck<br />

� 1992 bis 2000 BRG Schloss Wagra<strong>in</strong>, Vöcklabruck<br />

� Juni 2000 Ablegung der Reifeprüfung mit ausgezeichnetem<br />

Erfolg<br />

� seit Okt. 2000 Studium der technischen Physik an der TU Wien<br />

� Juni 2002 Abschluss des 1. Studienabschnitts mit<br />

ausgezeichnetem Erfolg<br />

� 2. Platz (1999) und 1. Platz (2000) beim Bundeswettbewerb der<br />

österreichischen Mathematikolympiade<br />

� Bronzemedaille (1999) und Silbermedaille (2000) bei der<br />

<strong>in</strong>ternationalen Mathematikolympiade<br />

91

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!