Development of Fe-Pd-based ferromagnetic shape memory alloys ...

189
Development of Fe-Pd-based ferromagnetic shape memory alloys by using combinatorial materials science Dissertation zur Erlangung des Grades Doktor-Ingenieur der Fakultät für Maschinenbau der Ruhr-Universität Bochum von Sven Hamann aus Essen Bochum 2013

Transcript of Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Page 1: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Development of Fe-Pd-based ferromagnetic shape memory

alloys by using combinatorial materials science

Dissertation

zur

Erlangung des Grades

Doktor-Ingenieur

der

Fakultät für Maschinenbau

der Ruhr-Universität Bochum

von

Sven Hamann

aus Essen

Bochum 2013

Page 2: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Dissertation eingereicht am: 08.01.2013

Tag der Verteidigung: 20.03.2013

Erster Referent: Prof. Dr.-Ing. Alfred Ludwig

Zweiter Referent: Prof. Dr.-Ing. Gunther Eggeler

Page 3: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Gold is for the mistress - silver for the maid -

Copper for the craftsman cunning at his trade.

"Good!" said the Baron, sitting in his hall,"

But Iron - Cold Iron - is master of them all."

From “Cold Iron” by Rudyard Kipling (1865 - 1936)

“Bella veniunt et abeunt, sed gregarii mei pro aeternitate sunt.”

Page 4: Development of Fe-Pd-based ferromagnetic shape memory alloys ...
Page 5: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Executive Summary / Kurzfassung

Im Rahmen der vorliegenden Promotionsarbeit wurde in einem ersten Schritt die binäre

ferromagnetische Formgedächtnislegierung Fe70Pd30 hinsichtlich ihrer strukturellen

Eigenschaften und in Abhängigkeit verschiedener Herstellungsmethoden sowie der Probenart

untersucht. In einem zweiten Schritt wurden die Methoden der kombinatorischen

Materialforschung genutzt, um neue ternäre Fe-Pd-X ferromagnetische

Formgedächtnissysteme mit verbesserten Eigenschaften zu entwickeln. Dazu wurden dritte

Elemente wie Mn und Cu zum Fe-Pd Ursprungssystem hinzulegiert und deren Einfluss auf

das martensitische Umwandlungsverhalten untersucht. Zur Herstellung wurden Fe-Pd(-X)

Dünnschicht-Materialbibliotheken durch Magnetron-Kathodenzerstäubung auf oxidierten Si

Substraten abgeschieden. Neben simultaner Deposition wurden auch keilförmige

Viellagenschichten zur Herstellung vollständiger binärer sowie partieller ternärer

Materialbibliotheken verwendet. Automatisierte Hochdurchsatz-Charakterisierungsmethoden

wurden genutzt, um die Materialbibliotheken hinsichtlich ihrer Zusammensetzung

(energiedispersive Röntgenanalyse), der Struktur (Röntgenbeugung), des Phasen-

umwandlungsverhaltens (temperaturabhängige Widerstandsmessung) und der magnetischen

Eigenschaften (Magneto-optischer Kerr Effekt) zu untersuchen. Vielversprechende

Zusammensetzungsbereiche wurden dabei mit zusätzlichen Untersuchungsmethoden

hinsichtlich ihres strukturellen Transformationsverhaltens (temperaturabhängige

Röntgenbeugung und Transmission-Elektronen-Mikroskopie) und hinsichtlich der

Sättigungspolarisation, Curie-Temperatur und magnetokristallinen Anisotropie analysiert.

In binären polykristallinen Fe-Pd Materialbibliotheken wurde das zusammensetzungs-

abhängige Transformationsverhalten in Abhängigkeit verschiedener Kathodenzerstäubungs-

methoden untersucht und die damit verbundenen strukturellen Unterschiede mit den

Materialeigenschaften korreliert. Im nächsten Schritt wurde eine neue Methode zur

Herstellung von massiven Fe70Pd30 Proben erprobt und die optimalen

Prozessierungsparameter ermittelt. Das Kapitel der binären Proben schließt mit der

Herstellung von einkristallinen Fe-Pd Dünnschichten und deren Eigenschaften sowie

Umwandlungsverhalten hinsichtlich struktureller Modifikationen. Zunächst wurden die

strukturellen und magnetischen Eigenschaften der verschiedenen Phasen in Abhängigkeit der

Zusammensetzung untersucht. Daran schließt sich die Herstellung freistehender einkristalliner

Fe70Pd30 Dünnschichten an, bei denen eine bisher nicht bekannte reversible strukturelle

Transformation von einer kubisch-raumzentrierten über eine tetragonal-flächenzentrierten hin

Page 6: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

zu einer kubisch-flächenzentrierten Struktur beobachtet wurde. Der letzte Teil des Kapitels

beschreibt die Herstellung von 1.2 µm dicken freistehenden Fe70Pd30 Schichten und deren

strukturelle Eigenschaften.

Im anschließenden Kapitel wurden zwei neue ferromagnetische Formgedächtnis Systeme mit

verbesserten Materialeigenschaften entwickelt und untersucht. Das Fe-Pd-Mn System zeigte

dabei eine Vergrößerung des Zusammensetzungsbereichs, in dem eine reversible strukturelle

Transformation erfolgt. Im Vergleich zum binären Fe70Pd30 System wurden die bisher

höchsten Transformationstemperaturen beobachtet. Diese Erhöhung der

Transformationstemperaturen wurde auf Basis von atomistischen Berechnungen, welche in

Kooperation mit M. E. Gruner (Universität Duisburg-Essen) durchgeführt wurden, mit dem

Effekt der magnetischen Anregung von antiferromagnetischen Elementen in einer

ferromagnetischen Fe-Pd Umgebung korreliert. Eine Verringerung der Curie-Temperatur und

der Sättigungspolarisation wurde für diesen Zusammensetzungsbereich beobachtet.

Das letzte Kapitel behandelt die Entwicklung und Untersuchung des Fe-Pd-Cu Systems.

Neben einer Erweiterung des Zusammensetzungsbereichs, in dem eine reversible strukturelle

Transformation erfolgte, wurde auch hier eine Erhöhung der Transformationstemperaturen

beobachtet und mit der erhöhten Löslichkeit von Fe in der transformierenden Phase im Fe-Pd-

Cu System begründet. Neben einer Verringerung der Curie-Temperatur und der

Sättigungspolarisation für diesen Zusammensetzungsbereich wurde eine Erhöhung der

spontanen Volumenmagnetostriktion entdeckt und in Fe-Pd-Cu Massivmaterialien bestätigt.

Strukturell modifizierte, einkristalline Fe70Pd30-XCuX (X = 3, 7 at.%) Schichten wurden

hergestellt und hinsichtlich ihrer magnetischen Eigenschaften, in Abhängigkeit der

Tetragonalität/Kubizität der Elementarzelle, untersucht. Dabei wurde festgestellt, dass 3 at.%

Cu, im Vergleich zu binärem Fe70Pd30, eine Erhöhung der magnetokristallinen

Anisotropiekonstante K1 um bis zu 40% erzeugt. Die Fe70Pd27Cu3 Zusammensetzung zeigt

somit den höchsten Wert für die magnetokristalline Anisotropiekonstante K1, welcher jemals

für eine ferromagnetische Formgedächtnislegierung gemessen wurde.

Page 7: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

Contents

1. Introduction ..................................................................................................................1

2. Fundamentals................................................................................................................3

2.1 The martensitic transformation................................................................................3

2.2 The thermal shape memory effect............................................................................6

2.3 The ferromagnetic shape memory effects ................................................................8

2.4 Magnetostriction ...................................................................................................10

2.5 The Invar effect.....................................................................................................12

2.6 The Fe-Pd system..................................................................................................13

2.7 Ternary Fe-Pd-X systems......................................................................................23

2.8 Routes for the development of novel Fe-Pd-X alloys.............................................25

2.9 Combinatorial materials science............................................................................28

2.10 Thin film nucleation and growth ...........................................................................31

3. Experimental methods................................................................................................35

3.1 Fabrication and Processing....................................................................................35

3.1.1 Thin film materials libraries...........................................................................35

3.1.2 Epitaxial thin films ........................................................................................40

3.1.3 Bulk samples / Splats.....................................................................................41

3.2 Characterization ....................................................................................................43

3.2.1 Energy-dispersive X-ray analysis (EDX) .......................................................43

3.2.2 Structural analysis by X-ray diffraction .........................................................45

3.2.3 Microstructural analysis by Transmissionen-Electron-Microscopy ................46

3.2.4 Temperature-dependent resistance measurements..........................................47

3.2.5 Magnetic properties, screening and high-resolution measurements ................49

3.2.6 Mechanical properties investigated by nanoindentation .................................50

4. Results and Discussion................................................................................................51

4.1 Binary Fe-Pd Ferromagnetic Shape Memory Alloys..............................................51

4.1.1 Polycrystalline Fe-Pd thin films.....................................................................51

4.1.2 Bulk / Splat Samples .....................................................................................66

4.1.3 Epitaxial Fe-Pd thin films ..............................................................................74

4.2 Ternary Fe-Pd-X Ferromagnetic Shape Memory Alloys......................................100

4.2.1 The Fe-Pd-Mn System.................................................................................100

4.2.1.1 Polycrystalline Fe-Pd-Mn thin films ............................................................100

4.2.2 The Fe-Pd-Cu System..................................................................................119

Page 8: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

4.2.3.1 Polycrystalline Fe-Pd-Cu thin films .............................................................119

4.2.3.2 Fe-Pd-Cu Splats ..........................................................................................135

4.2.3.3 Epitaxial Fe-Pd-Cu thin films ......................................................................142

5. Conclusion and Outlook ...........................................................................................160

6. References .................................................................................................................163

Acknowledgments.............................................................................................................174

Curriculum Vitae .............................................................................................................176

Page 9: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

List of abbreviations and symbols

a [nm] lattice parameter

Af [K] austenite finish temperature

As [K] austenite start temperature at.% [-] atomic percent bcc [-] body centered cubic bct [-] body centered tetragonal c [nm] lattice parameter c/a [-] tetragonality of unit cell CPA [-] coherent potential approximation CSM [-] continuous stiffness method d [nm] lattice spacing DC [A] direct current DFT [-] density functional theory ∆Η [J] latent heat DSC [-] differential scanning calorimetry E [J] energy e/a [-] valence-electron ratio EBSD [-] electron backscatter diffraction EDX [-] energy-dispersive X-ray analysis EELS [-] electron energy loss spectroscopy

EK [J] magnetocrystalline anisotropy energy fcc [-] face centered cubic fct [-] face centered tetragonal FIB [-] focussed ion beam FSMA [-] ferromagnetic shape memory alloy FWHM [-] full-width-at-half-maximum GGA [-] generalized gradient approximations GGA [J] Gibb's free energy H [Oe] magnetic field

HA [Τ] anisotropy field HAADF [-] high-angle annular dark field

HC [Τ] magnetic coercivity HRTM [-] high-resolution TEM

HS [T] saturation field ICDD [-] international centre for diffraction data ICSD [-] inorganic crystal structures database

JS [T] saturation polarization K [kJ/m³] magnetocrystalline anisotropy constant

kB [J/K] Boltzmann constant KKR [-] Korringa-Kohn-Rostoker

Page 10: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

L10 [-] tetragonal phase

L12 [-] cubic austenite phase M [T] magnetization of a martensitic variant MAE [-] magnetocrystalline anisotropy energy

M f [K] martensit finish temperature MF [-] misfit MFIS [-] magnetic field induced strain MIM [-] magnetic field induced martensite MOKE [-] magneto-optical Kerr effect

MS [K] martensit start temperature N [-] demagnetization factor NRA [-] nuclear reaction analysis PCA [-] principal component analysis PLD [-] pulsed laser deposition PPMS [-] physical property measurement system PVD [-] physical vapour deposition Q [J] activation energy for grain boundary movement R(T) [Ohm] temperature-dependent resistance

R0 [nm] initial grain size RBS [-] Rutherford backscatter diffraction Ref. [-] reference RF [Hz] radio frequency

rFe [nm] covalent atomic radius of Fe

rMn [nm] covalent atomic radius of Mn

rcr [nm] critical nucleus radius

s [-] shape parameter in Kuz’min’s equation SAED [-] selected area electron diffraction SEM [-] scanning electron microscope SMA [-] shape memory alloy SPR-KKR [-] spin polarized relativistic Korringa-Kohn-Rostoker code SQUID [-] superconduction quantum interference device STEM [-] scanning transmission electron microscopy T [K] temperature t [s] time

TC [K] Curie temperature TEM [-] transmission electron microscopy URQ [-] ultra-rapid-quenching VASP [-] Vienna Ab-initio Simulation Package VSM [-] vibrating sample magnetometer WDX [-] wavelength-dispersive X-ray analysis XRD [-] X-ray diffraction XRD(T) [-] temperature-dependent X-ray diffraction Z1 [-] tetragonal phase

Page 11: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

α [°] angle α [-] proportional factor β [°] angle ∆Τ [K] thermal hysteresis ε [%] strain φ [°] angle γ [J] interface energy ϕ [°] angle ϑ [°] angle λ [nm] wavelength

λ(hkl) [-] saturation magnetostriction in (hkl) direction

λS [-] saturation magnetostriction θ [°] angle σ [MPa] stress

σbl [MPa] blocking stress

σmag [MPa] magnetic field induced stress

σtbm [MPa] stress for twin boundary movement ψ [°] angle Γ [-] center of the Brillouin zone

Page 12: Development of Fe-Pd-based ferromagnetic shape memory alloys ...
Page 13: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

1

1. Introduction

Over the last two decades the development of new miniaturized actuators and sensors

increased rapidly.1 This progress is significantly caused by the ongoing discovery of new and

the enhancement of already known material systems that exhibit functional effects. Especially

the miniaturization benefits from these functional effects by using the material itself as

actuating or sensing device. Beside the well-known class of conventional shape memory

alloys (SMAs), such as Ni-Ti2, ferromagnetic shape memory alloys (FSMAs), such as Ni-Mn-

Ga3 and Fe70Pd304, became recently the object of increased scientific interest, due to their

advantageous properties.

The Fe-Pd system is a very promising candidate to be implemented as a functional material

into miniaturized systems, since it exhibits many physical effects that can be exploited for

functional application. Because of its high magnetocrystalline anisotropy, that is oriented

perpendicular to the sample surface, when fabricated in a thin film design, it is interesting as a

magnetic recording material.5,6 In contrast to non-ferromagnetic materials, Fe-Pd shows an

anomalous low thermal expansion over a defined temperature range, what is commonly

known as Invar effect.7,8 Furthermore, magnetostricition is observed for Fe-Pd, showing a

significant length change when exposed to a magnetic field.9 When the Fe70Pd30 composition

is quenched from high temperatures, a face-centred cubic austenite phase can be stabilized at

ambient temperature. Upon cooling, the Fe70Pd30 composition undergoes a reversible

martensitic transformation from a face-centred cubic austenite phase to a face-centred

tetragonal martensite phase.4 This martensitic transformation is highly composition dependent

and occurs for Fe70Pd30 slightly below 293 K. With increasing Fe content the martensitic

transformation is shifted to higher temperatures. Due to a twinned microstructure of the face-

centred tetragonal phase, martensitic variants are formed. These variants are separated by twin

boundaries and have a preferred direction of magnetization. This gives rise to a macroscopic

strain up to 3% originated by the reorientation of martensitic variants, when an external

magnetic field is applied. This effect is known as magnetic-field induced strain and can be

utilized in microsystems to work for example as micro-gas valve, as reported in literature.10

Thus, the magnetic-field induced strain effect could be implemented in several micro- and

nano-electro-mechanical-systems to enable friction free and non-vibrating movements.

Most of these effects were observed for single-crystal samples, as reported in literature.11,12

Since the fabrication of single crystals is time and cost consuming, alternative fabrication

Page 14: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

2

methods have to be established to open the way for implementing Fe-Pd as active material

into miniaturized application. Especially thin films are promising for miniaturized application

and allow a rather cost effective fabrication. In contrast to bulk samples, thin films are

strongly influenced by substrate and surface constraints. This can alter significantly the

materials properties when transferred from bulk to thin film geometry. Nevertheless, several

issues still hamper a broad implementation of binary Fe-Pd into technical devices. Especially

the low transformation temperatures, the metastability of the transforming austenite phase and

the insufficiently high magnetocrystalline anisotropy need to be enhanced. Further the high

amount of a cost-expensive element like Pd constrains any mass production. Thus, new

ferromagnetic shape memory alloys based on the Fe-Pd system are of high interest, if they

overcome the previously mentioned constraints. This can be accomplished by the addition of

third elements into Fe-Pd in order to develop novel and enhanced materials. Nevertheless

there is only sparse information about ternary Fe-Pd-X materials so far, and no defined rules

for developing were identified which would allow an alloy design. The impact of Fe content

and valence-electron number e/a as dominating parameters, controlling the temperature shift

of the martensitic transformation, is still not clarified and needs a comprehensive

investigation. This can be achieved by investigating ternary Fe-Pd-X alloys that allow to

separate the different parameters and to gain a deeper insight into the composition-structure-

property relationship.

Page 15: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

3

2. Fundamentals

Within the following chapter, the fundamentals of the Fe70Pd30 FSMA are presented and the

different properties in this alloy, fundamental for the magnetic field-induced strain, are

pointed out. In the first part the martensitic transformation is presented, followed by a

description of the conventional shape memory effect. Then the actuation effects occurring in

FSMAs are shown and the prerequisites for showing these effects are pointed out. The next

subsection emphasizes on magnetostriction and explains the fundamentals of this effect. The

causation of the anomalous reduction of thermal expansion coefficients, also known as Invar

effect, which is occurring in Fe-based alloys, is presented in the following subsection. The

next section introduces the Fe-Pd system with the different structural phases that can occur

and their intrinsic properties. A further part examines the change of intrinsic properties of

ternary Fe-Pd-X alloys and gives an overview of the state-of-the-art literature. Subsequently a

route for the successful development of new Fe-Pd-based FSMAs with enhanced intrinsic

properties is presented. The last section presents the procedure that was used to develop new

Fe-Pd-based FSMAs.

2.1 The martensitic transformation

A martensitic transformation is defined as a displacive, first-order diffusionless

transformation from a high-temperature phase (austenite) having a highly symmetric structure

to a low temperature phase (martensite) with a lowered symmetric structure.13 This

transformation was observed for the first time by Adolf Martens in 1887, when he

investigated hardening mechanisms in steel.14 During a martensitic transformation, the atoms

move only fractions of the lattice distance without any local changes in chemical composition

when the structure transforms from the austenite to the martensite phase. Due to the small

displacement of the atoms in the lattice, well-defined correlations exist between the austenite

and the martensite phase.15 The martensitic transformation is a first order exothermic

transformation.16 A change from the austenite to the martensite phase is induced, when the

difference in Gibb’s free energy between both phases (∆GA-M) is higher than the energy

needed for elastic deformation of the crystal lattice (∆GD), formation of new bounding

surfaces (∆GS) and the amount of internal friction (∆GF):

Page 16: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

4

0)GGG()GG(G FSDMartensiteAusteniteMA >∆+∆+∆+−=∆ − (1)

In Figure 1 the martensitic transformation is depicted in terms of Gibb’s free energy as a

function of temperature for the austenite and the martensite phase. Upon cooling the

transformation starts from the high-temperature austenite phase at the martensite start

temperature (Ms). This temperature is shifted to lower values (∆T) than the equilibrium

temperature T0, since a driving force is needed to overcome ∆GD, ∆GS and ∆GF. The thermal

hysteresis between these two states is originated by the amount of ∆GD, ∆GS and ∆GF.

Figure 1: Temperature-dependence of Gibb’s free energy for both the austenite as well the martensite phase

(adapted from Ref. 17). To induce the transformation from the high-temperature (austenite) to the low-

temperature (martensite) phase an undercooling is needed to overcome ∆GD, ∆GS and ∆GF.

In dependence on the crystallographic compatibility between austenite and martensite phase,

the energy for elastic deformation ∆GD can become negligible, leading to a decrease of the

undercooling ∆T and thus of the thermal hysteresis. When the temperature reaches the

martensite finish temperature (Mf), the structure has changed fully to the martensite phase. If

the temperature is increased again, the structure changes back from the martensite to the

austenite phase. Due to the reversibility of this process, this kind of transformation is defined

as a thermoelastic transformation.

The change of the crystal lattice within a martensitic transformation can be described by three

stages.18 When a sample undergoes a martensitic transformation, the atoms are sheared along

a neutral habit plane (Figure 2 a) to b)) distorting the austenitic matrix.19 By lattice invariant

shearing the martensite adapts the austenite matrix in order to compensate these shearing

stresses. To reduce these stresses the lattice undergoes either a gliding (c)) or a twinning (d))

Page 17: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

5

movement. When the lattice undergoes a gliding (b) to c)), atomic bindings are broken and

new bindings are formed, leading to an irreversible change of the structure.16

Figure 2: Martensitic transformation from the austenite (a)) to the martensite phase by a shearing mechanism

(b)) that can be accompanied by gliding and twinning (c) and d)) for stress relaxation in the martensite (adapted

from Ref. 15).

When a lattice is twinned to reduce shearing stresses, crystal variants (martensitic variants)

are adjusted along a boundary (twin boundary) in a reversed imaged manner. The amount and

kind of bindings of a twin boundary are equal to atoms that do not belong to this boundary.

This is why twin boundaries have a low energy, are highly mobile and can be easily moved

through a crystal. In ordered alloys twinning is the energetically favoured mechanism for

stress relaxation, because all atomic bindings and additionally the degree of order are retained.

Deformation of the lateral dimensions of a sample by twin formation is reversible and

therefore the prerequisite for the shape memory effect. General to all shearing mechanisms a

rotation of the lattice is needed to keep an invariant habit plane. Concluding this, the

martensitic transformation can be defined as a shear-dominated reversible phase

transformation where the nearest atomic neighbours are preserved. Due to the shear

mechanisms, a weakening of the lattice near the martensitic transformation temperature can

Page 18: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

6

be observed giving rise to unusual behaviour of a sample’s intrinsic properties like Young´s

modulus, resistivity and thermal expansion.20

Due to the lower symmetry of the martensite lattice, external macroscopic strains can be

compensated better than in the austenite state. Thus besides temperature also strain can be

used as the driving force to induce a martensitic transformation. When a sample in the

austenite state is strained, the lattice starts to transform into the martensite state at a critical

and temperature-dependent stress. The differential change of transformation temperature dMS

by application of a differential stress dσ is described by the Clausius-Clapeyron

equation:13,21,22

ε⋅∆−=σT

H

dM

d

S

(2)

In this equation ∆H defines the latent heat, which is released within the first-order

transformation. The strain of a sample is represented by ε and the temperature by T.

Materials undergoing a martensitic transformation exhibit an anomalous mechanical

behaviour at temperatures near the transformation like lattice softening.23 When a sample is

just above Af it has an austenitic structure. On the application of an external stress the

martensite phase can form because of a stress-induced martensitic transformation following

the Clausius-Clapeyron relation. In the martensite phase external stresses are compensated by

the development of martensitic variants having their long axis aligned to the external stress

field. This effect allows for high strains up to 10% without significant increase of the external

stress and is defined as superelasticity.24 Since the temperature is above Af the martensite

phase is unstable and will transform back into the austenite phase when the external stress is

removed.

2.2 The thermal shape memory effect

The thermal shape memory effect is defined as the ability of a material to recover its original

shape after deformation, when heated to a certain temperature. This is correlated to the

thermoelastic martensitic transformation. At the beginning the material is in the high-

temperature austenitic state (Figure 3 a)). Upon decreasing the temperature it transforms into

the low-temperature martensite state while keeping its overall shape. Due to spatial

restrictions the lattice change is compensated by developing martensitic twins (Figure 3 b)).

These twins are connected by twin boundaries that are highly mobile and allow the material to

Page 19: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

7

easily deform when being in the martensite state. If the material is deformed by external

forces martensitic variants compensating most of the stress are preferably formed (Figure

3 c)).

Figure 3: Conventional shape memory effect where a cubic austenite phase a) is transformed into a tetragonal

martensite phase showing a twinned structure b) upon cooling. The different variants are connected by twin

boundaries that allow for an easy deformation of the crystal c). When heated, the crystal transforms back to the

austenite phase - c) to a) - recovering its primary shape.25

This is conducted without any breaking of atomic bonds, leading to a defined relationship for

all atoms within both phases. This gives rise to the shape recovery when the material is heated

to its high symmetric austenite state subsequently. When the temperature is lowered the

material transforms again into the martensite phase without any further shape change. Thus

this effect is defined as one-way-effect. By using appropriate thermal processes in accordance

with mechanical treatment, the material can be trained by inducing dislocations within the

structure. This provides a further shape change within the material when undergoing a

martensitic transformation by heating and cooling. This effect is defined as two-way-effect

and can be used as actuating mechanism for conventional shape memory alloys.

Page 20: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

8

2.3 The ferromagnetic shape memory effects

In general two different actuating effects occur within FSMAs. Both effects exhibit a

reversible straining of a sample upon applying an external magnetic field accompanied by a

change in crystal structure. The straining can reach values up to 10% making this more than

ninety-times higher than conventional magnetostriction (TbDyFe26 λS= 1100⋅10-6). Compared

to conventional SMAs, where the actuation is induced by temperature changes, magnetic

fields can be switched faster up into the kHz range. This is only limited by the sonic speed

that restricts the movement of twin boundaries in an alloy.27

For the first effect a martensitic transformation from a weak to a strong magnetic moment

phase is induced by applying an external magnetic field at temperatures near the

transformation. This is correlated to a decrease of free energy during transformation

originated by the magnetic contribution to Gibb’s free enthalpy.28 The effect is known for Ni-

Mn-Ga and Ni-Mn-In alloys, where the martensite phase is induced by magnetic fields and is

defined as magnetic field induced martensite (MIM).29

Figure 4: Schematic of the magnetic field induced strain effect where the orientation of martensitic variants is

changed by an external magnetic field. a) Without a magnetic field a random distribution of martensitic variants

appears. The direction of magnetic moments in the different variants is depicted by a small red arrow. By

applying an external magnetic field (large red arrow in the lower left in b)) the martensitic variants having their

magnetic moment aligned along the external magnetic field are favoured and grow at the expense of others. (the

angle between the easy axes of different variants was chosen to be < 90°). c) With increasing external magnetic

field more variants are aligning along by twin boundary movement (green arrows). d) When the external

magnetic field reaches a critical value all variants are aligned into one direction giving rise to an overall straining

of a sample (adapted from Ref. 30).

In magnetic anisotropic alloys the magnetic moment within the lattice cell is aligned along

certain crystallographic axes. Thus especially in the low symmetry martensite phase the

different variants have a magnetic moment that is oriented to a specific-axis in the lattice (in

Page 21: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

9

the case of Fe70Pd30 it is the shorter c-axis of the fct unit cell).57 Without an external magnetic

field, the different variants are oriented to the crystal lattice. If an external magnetic field is

applied, the martensitic variants and thus the crystal lattice start to align along the field

direction. This gives rise to a magnetic field induced strain (MFIS) effect as depicted in

Figure 4. The MFIS appears only in magnetic materials that have a martensite phase like Ni-

Mn-Ga, Fe70Pd30 and Fe3Pt. The main prerequisite for this effect is a sufficient

magnetocrystalline anisotropy that couples the magnetic moment to a specific-axis of the

crystal lattice. The magnetocrystalline anisotropy generates a direction-dependent

magnetization behaviour of the crystal lattice where the material can be magnetized easily in

certain crystallographic directions (easy axis) while other directions are hard to magnetize

(hard axis). In the following all aspects for magnetic field induced twin boundary motion is

declared for a tetragonal structure where c describes the short and easy axis and a defines the

long axis that is hard to magnetize. In general the magnetocrystalline anisotropy energy EK in

alloys having an uniaxial symmetry can be written as:31

)(cos)(cosK)(sinK)(sinKK),,(E 223

42

210K β⋅α+φ+φ+=βαφ (3)

The parameters φ, α, β define the angles of the magnetization to the c-axis ([001]-direction)

and the a-axes ([100]-and [010]-direction) within the unit cell. The maximum

magnetocrystalline anisotropy energy for uniaxial structures is defined as KU = K1 + K2. This

is the case when sin2φ = 1, which describes the orientation of magnetization aligned to the

hard a-axis.32 When an external magnetic field is applied to two adjacent martensitic variants

having their easy axis oriented at an angle of ϑ = 90° to each other, the maximum energy

difference ∆Emag between is defined as:33

U2

U21max,mag K)(sinKH)MM(E =ϑ⋅+⋅−=∆ (4)

The maximum strain that can be reached is defined by the lattice parameters a and c of the

tetragonal unit cell and can be described by ε0 = 1 − c/a. Further the energy difference

∆Emag,max is proportional to the magnetic field induced stress σmag by:34

0magmagmag0

magEE ε⋅∆=σ⇔∆=

εσ

(5)

Page 22: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

10

If the maximum magnetic field induced stress σmag applied to the martensitic variant is higher

than the stress σtbm (twin boundary movement) required for moving the twin boundary

between two adjacent variants, a reorientation of the variants appears. This is expressed by:35

tbm0

Umag

K σ≥ε

=σ (6)

This gives rise to a full reorientation of all martensitic variants having their easy c-axis

aligned to the external magnetic field.36 If the c-axis is the short axis in the tetragonal crystal

this effect decreases the outer dimensions of a sample giving rise to the actuation effect. In

order to achieve this MFIS easy movable twin boundaries are one prerequisite for FSMAs.

Materials exhibiting the MFIS have to combine a sufficiently high magnetocrystalline

anisotropy KU with easily movable twin boundaries. Mostly known alloys are: Ni2MnGa3,

Fe3Pt37, Fe70Pd304 and La1.99Sr0.01CuO4

38 but also pure elements like Tb and Dy39 exhibit the

MFIS at very low temperatures of about 4.2 K. The MFIS is highly sensitive to structural

defects and grain boundaries that can pin the twin boundary movement and therefore prevent

any straining. Thus, the maximum achievable straining in polycrystalline materials by the

MFIS is reduced. Up to now single crystals37,40 (Ni2MnGa: 9.5 %; Fe70Pd30: 3 %) and foams41

(Ni2MnGa: 8.7 %) show the highest straining up to several percent for the MFIS. The two

main systems are Ni2MnGa and Fe70Pd30 that both combine a high magnetocrystalline

anisotropy (Ni2MnGa: KU= 160 kJ/m3 at 300 K; Fe70Pd30: KU= 180 kJ/m3 at 77 K) with easy

movable twin boundaries (Ni2MnGa: ε = 0.06; Fe70Pd30: ε = 0.07) at a maximum magnetic

field induced stress (Ni2MnGa: σmag = 2.6 MPa; Fe70Pd30: σmag = 2.3 MPa).34,42

2.4 Magnetostriction

In general magnetostriction is defined as the elastic straining of a sample’s volume or shape

originated by magnetism. The effect is based on the spin-orbit coupling in magnetic materials,

which efficiently couples the magnetism directly to the electron spin and lattice orientation. It

is well known that the magnetic properties of metals and alloys depend on the arrangement

and separation of atoms in the lattice. Therefore, the volume of a material in the magnetically

ordered state is different from that it would have, if no spontaneous ordering of spins would

occur. This ordering of spins provides a repulsive force that results in an increase of the

crystal unit cell volume.43

Page 23: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

11

Such an effect is defined as spontaneous volume magnetostriction where the volume of a

material increases without changing the shape. Following Kakehashi et al.50 the spontaneous

volume magnetostriction is a prerequisite for the Invar effect, which influences strongly

thermal expansion and magnetic properties of a material system.

When the electron spins are aligned by an external magnetic field, the electron orbitals change

their shape due to the spin-orbit coupling. Because of the direct coupling of magnetism to the

electron spin and lattice orientation, the crystal lattice changes its distances. This gives rise to

an overall shape deformation and straining of a sample. Such a shape change at constant

volume is defined as Joule’s magnetostriction.

Figure 5: Macroscopic model of the magnetostrictive effect: a) The material is in a paramagnetic state with a

random orientation of magnetic dipoles. b) For temperatures below TC the material changes to a ferromagnetic

state accompanied by an alignment of the magnetic dipoles into areas with uniform magnetic moment (magnetic

domains). c) By applying a magnetic field all areas with uniform magnetic moment align, giving rise to an

overall shape change. (adapted from Ref. 44)

In Figure 5 the principle of Joule’s magnetostriction for an isotropic sample is presented.

When a magnetic material is above its transition temperature (Curie temperature) it is in a

paramagnetic state, with a random distribution of magnetic dipoles (Figure 5 a), dipoles are

depicted by coloured bean-shaped symbols). By lowering the temperature the paramagnetic

state changes to a ferromagnetic state, leading to the formation of regions with aligned

magnetic moments (magnetic domains). In order to minimize the stray field energy the

magnetic domains are aligned in a way where the magnetic flux lines are closed (Figure 5 b)).

If an external magnetic field is applied the magnetic domains align according to this field

leading to a straining of the crystal lattice and thus giving rise to a straining of the sample

(Figure 5 c)). The shape change of the sample in a certain direction is defined by λ(hkl) = dl/l

Page 24: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

12

with a positive algebraic sign “+” for elongation and a negative “−“ for contraction. The

averaged magnetostriction constant for a polycrystalline cubic material is defined by:

111100S 5

3

5

2 λ+λ=λ (7)

Here, λ100 and λ111 are the saturation magnetostriction in (100) and (111) direction of a cubic

lattice. In general nearly all magnetic elements and materials show magnetostriction, although

this effect is rather small for pure Fe (λS = -3.6⋅10-6) it reaches values for lanthanides like Tb

and Dy (λS = -3.6⋅10-6 and λS = -3.6⋅10-6) at cryogenic temperatures.

Like all ferromagnetic materials, disordered Fe70Pd30 exhibits a Joule’s magnetostriction

effect. The averaged magnetostriction constant λS = 5⋅10-5 is relatively high in comparison to

the values for pure Fe.9

2.5 The Invar effect

The Invar effect was first observed in 1896 by the Swiss physician Charles Édouard

Guillaume in a Fe65Ni35 alloy, when he tried to develop new materials for length and mass

standard application.45 In 1920 Guillaume was awarded the Nobel price for this discovery.

This effect denotes the anomalous behaviour of thermal expansion coefficients around room

temperature where the volume expansion of an alloy can be zero or even become negative.

This is caused by a superposition of normal thermal expansion due to anharmonic terms in the

interatomic potential and a volume contraction due to the change of the magnetic state. This

effect is well known to appear in Fe-rich alloys, such as Fe-Ni, Fe-Pt and Fe-Pd but it was

also found in several other alloys.46 This effect is not fully understood so far, although

different approaches have been made to explain it. In general, the main theories correlate this

behaviour to two concurrent magnetic states.47 The two-γ-state model was developed by

Weiss et al. in 1963 for the face-centred cubic γ-Fe and reports on the coexistence of two

different ferromagnetic states: The γ1 state is a low energetic state having a small volume

(a = 0.354 nm) and a weak magnetic moment (0.5 µB); the γ2 state is a high energetic state

with a large volume (a = 0.364 nm) and a strong magnetic moment (2.8 µB).48 If the energy

difference between these two states is small enough, the γ1 state can be activated at moderate

temperatures. Due to the smaller volume of the γ1 state the occupation of this state counteracts

to the normal thermal expansion and thus leads to an overall anomalous decrease of the

Page 25: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

13

thermal expansion coefficients. Another explanation of the Invar effect in Fe-rich alloys was

suggested by Kondorsky and Sedov.49 They suggested a latent antiferromagnetism where an

antiparallel alignment for the Fe atoms within a ferromagnetic matrix is energetically

favoured. If the Fe content is increased to a certain amount, a break down of the exchange

coupling and thus of the spontaneous magnetization appears. These magneto-elastic

anomalies are correlated to changes of the exchange coupling in dependence on volume and

pressure. Therefore it is energetically favoured to increase the volume at low temperatures

when the exchange coupling energy is minimized. Cooling this alloy leads then to an increase

of volume. When it is heated the magnetic exchange coupling is weakened due to

spontaneous spin fluctuations and accordingly the volume decreases again. This is

superimposed by thermal expansion and leads to anomalous behaviour of thermal expansion

coefficients. Kakehashi et al. correlate this to a gradual transition from a strong to a weak

magnetic state in caused by thermal excitations.50

The Invar effect was found to appear in disordered Fe70Pd30 alloys as reported by Schlosser.51

Matsui et al.52 found an anomalous behaviour of thermal expansion of Fe68Pd32 over a

temperature range from 4.2 K to 800 K indicating a reduction of the thermal expansion

coefficient α [K -1] when increasing the temperature > 400 K. This anomalous behaviour of

thermal expansion becomes maximal when a sample is near a structural instability like a

martensitic transformation as reported by Nakayama et al..53

2.6 The Fe-Pd system

The thermodynamic stable structural phases of the Fe-Pd system in dependence on

composition and temperature at constant ambient pressure are presented in the phase diagram

in Figure 6. Further the metastable phase diagram close to the Fe70Pd30 composition is shown

and the thermoelastic transformation from a face centred cubic (fcc) to a face centred

tetragonal (fct) structure in this compositional region of interest is explained and compared

with recent theoretical investigations. To gain a fundamental understanding of phase

formation a detailed understanding of the equilibrium phase diagram of the binary Fe-Pd

system is required. This diagram depicts the equilibrium phases in dependence on

composition and temperature. When lowering the temperature below the Liquidus-line the γ-

phase forms, that is stable over the complete compositional range. This phase has a fcc

structure with a random (disordered) distribution of Fe as well as Pd atoms in the lattice. The

existence region of the γ-phase lies between the Liquidus-line (the transition from solid to

Page 26: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

14

liquid appears between 1577 K and 1828 K in dependence on composition) and reaches down

to temperatures of about 1173 K. Since this thesis focuses on the thermoelastic fcc to fct

transformation of samples near the Fe70Pd30 composition, the equilibrium as well the

metastable phase formation will be described. When considering an infinitesimal slow cooling

process the phase formation can be described as a sequence of thermal equilibrium states. At a

temperature of 1033 K the fcc Fe70Pd30 γ-phase decomposes into an α-Fe phase (bcc: body

centred cubic structure) with a Pd content < 3.5 at.% and a fcc γ-phase with decreased Fe

content according to the lever rule.54 Upon cooling along the γ-phase boundary the Pd content

increases up to a value of 46 at.% at 873 K.

Figure 6: Equilibrium phase diagram for the binary Fe-Pd system.55

When the temperature is lowered further, the Pd content in the bcc α-Fe phase decreases < 1

at.% (stabilized boundary solubility due to mixing entropies) and the fcc γ-phase stabilizes at

a Fe50Pd50 composition. This Fe50Pd50 phase has a L10 in layer ordered structure, that can be

described by a slightly distorted fcc structure. This slight distortion of the fcc structure

perpendicular to the layer planes results in a fct structure. A sequence of Fe and Pd layers

forms the L10 structure, where the lattice parameter in the layer plane is smaller than the value

for the out of plane direction. The phase formation in a bulk body is controlled by the

diffusion coefficients of Fe atoms. Since the diffusion coefficients and thus the diffusion

speed is limited, the formation of metastable phases can be achieved by rapid quenching to

Page 27: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

15

low temperatures. When a high-temperature equilibrium phase is quenched to low

temperatures there is still a significant driving force for decomposition and formation of new

phases. This is prohibited by a significant reduction of diffusion kinetics at low temperatures

that avoids decomposition and allows for the existence of metastable phases.

The interest on disordered Fe70Pd30 is based on the appearance of a reversible martensitic

transformation from an austenitic fcc (in the following termed as metastable/transforming fcc

phase) to a martensite fct phase. This fct phase occurs below room temperature and can be

achieved by quenching the sample from the γ-phase to temperatures < 573 K. Figure 7 a) to c)

shows three metastable phase diagrams describing the sequence of the different structural

phases as a function of Pd content and temperature. All diagrams show the presence of a

reversible martensitic transformation from a high-temperature fcc phase to a low-temperature

martensite fct phase. The transformation temperatures increase with increasing Fe content

until the fct structure transforms irreversibly into the bct structure.

Figure 7: Metastable phase diagrams for quenched alloys around the composition Fe70Pd30. Presented are the

structures as a functions of the Pd content and temperature. a) The full phase transformation ranging from a bcc,

over a bct and a fct to a fcc structure is reported by Sugiyama et al.56. The transformation from fcc to fct was

reported to be reversible, while the fct to bct transformation was found to be irreversible. b) Cui et al.57 reports

on an extension of the transforming region up to Pd contents of 29.3 at.% where samples still have a fct structure

and undergo a martensitic transformation. c) Matsui et al.58 describes the sequence from bcc (α), over fct (γ’) to

fcc (γ) as a function of composition and temperature. Further the variation of lattice parameters ratio c/a is

correlated to the different phases. This implies that the c/a ratio can be regarded as a parameter to describe this

structural transition through the different phases.

This dependency for other Fe-based alloys like Fe-Ni59 and Fe-Pt60 is often described in

literature by the valence-electron to atom ratio (e/a). In Figure 7 a) the structural sequence for

Fe-Pd foil samples with a thickness of about 40 µm is depicted. These samples were

fabricated by melting and subsequent homogenisation annealing at 1373 K for 60 h followed

by dicing and rolling into foils. To obtain the disordered fcc phase the foils were solution

Page 28: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

16

treated at 1173 K for 1 h and afterwards quenched in iced water. Due to the sample

preparation and treatment a polycrystalline structure with crystal lattice defects like

dislocations was observed. Further the existence range of the fct phase is increased to higher

Pd values. For Figure 7 b) single-crystal samples with a Fe70Pd30 composition were fabricated

using the Bridgman method from arc-melted buttons. This crystal was diced into several

pieces for investigation. Due to the fabrication method no hints for crystal defects were

observed. Here, the existence range for the fct phase is smaller than in a). The samples

presented in diagram c) were fabricated using a plasma jet furnace followed by a

homogenization annealing at 1423 K for 7 days encapsulated in a quartz glass tube. After

annealing the samples inside the glass tube were quenched in iced water to prevent

decomposition of the γ-phase. All diagrams show that the transformation from the fcc to fct

phase occurs around 300 K for the Fe70Pd30 composition. When the temperature is further

decreased the sample changes its structure irreversibly from a fct to a bct structure without

recovery of the fct phase upon subsequent heating. This irreversible transformation cannot be

used for the shape memory effect and thus needs to be avoided.61 The existence range of the

different structural modifications of the parental fcc phase differ slightly and thus also the

transformation temperatures in the presented diagrams. This is correlated to the different kind

of samples (single crystal, polycrystalline bulk and foils) and the processing routes. Especially

sample impurities and quenching rate affect significantly the developing microstructure as

reported by Matsui et al..58 This can also be determined by thermal hysteresis width that

differs for all three different sample types. The single-crystal sample from Cui et al.57 shows a

thermal hysteresis of about 5 K while this increases to about 10 K for the bulk samples

fabricated by Matsui et al. and reaches a maximum value of approximately 20 ± 10 K for the

foils. The thermal hysteresis is controlled among other factors by the crystal lattice misfit

between the austenite and the martensite phase. Therefore the single-crystal samples show the

smallest misfit between martensite and austenite indicated by a small amount of energy that

has to be given into the system to transform. In contrast the Fe-Pd foils need a significantly

higher amount of energy to undergo this transformation. This is related to the high amount of

structural defects, originated by the fabrication process, that constrain the martensitic

transformation. Stress effects originated by structural defects increase the transformation

temperatures, as mentioned in the previous chapter. This is observed for the bulk and foil

samples that exhibit higher transformation temperatures than the single-crystal samples. As

mentioned in Chapter 2.1 the stress state of a sample significantly affects the transformation

temperatures. Kato et al.22 report on an increase of temperatures for the fcc to fct

Page 29: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

17

transformation in dependence on stress by a factor of dσ/dT = 4.8 MPa/K. The amount of Fe

in a sample depends significantly on the kind/geometry of sample. It was found that thin film

samples generally allow to increase the content of Fe atoms in the fct phase. This is different

when compared to bulk samples that form Fe-rich precipitates at lower Fe contents. Sugimura

et al.61 reported on a Fe72.5Pd27.5 thin film sample in a single fct phase.

Figure 8: The three martensitic variants that can be formed upon transformation from the fcc austenite to the fct

martensite phase.57 The variants develop by contracting one of the base vectors (v1, v2 and v3). For the fct

structure the long axes are termed as a and the short as c-axis.

Starting from the fcc structure there are three different variants that can be formed upon

transformation to the fct phase. In Figure 8 these martensitic variants are shown. According to

this model the cubic austenite transforms into the tetragonal martensite by contracting the

lattice along one of the catersian axes and expanding the others correspondingly. In the

following, the short axis is defined as c-axis while the two longer axes are termed as a-axes.

In Fe70Pd30 single crystals the lattice parameters are determined to be afcc = 0.3756 nm for the

fcc austenite and afct = 0.3822 nm and cfct = 0.363 nm for the fct martensite structure.57 The

degree of tetragonality of the fct martensite phase can be calculated by calculating the c lattice

parameter over the a lattice parameter. From the above mentioned lattice parameters Cui et

al.57 determined c/a = 0.95 while Matsui et al.58 extrapolated a value of c/a = 0.914.

Considering Matsui’s value a complete alignment of martensitic variants by an external

magnetic field (ferromagnetic shape memory effect, see Chapter 2.3) can lead to a maximum

strain of 6%. Stern et al.62 calculated the maximum strain to be around 6.8% by using

electronic structure calculations. The Curie temperature TC for Fe70Pd30 is determined to be

Page 30: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

18

around 700 K and depends strongly on composition. Thus both the austenite as well as the

martensite are ferromagnetic with a saturation polarization of JS,fcc= 1.357 T and

1.508 < JS,fct < 1.885 T.57,62 The magnetic easy axis in the fcc austenite phase is aligned along

the [111] direction. The fct martensite phase has the magnetic easy axes aligned along the

[100] and [010] (a-axes) directions and the magnetic hard axis aligned to the c-axis of the

tetragonal unit cell.57

As shown in Figure 7 c) the sequence of different structures, developing in dependence on Pd

content and temperature is given by the c/a ratio of the fcc unit cell. This is expressed by the

Bain path formalism,63 which describes a continuous distortion of a unit cell from a fcc

austenite, over a fct and bct martensite to a bcc structure. According to Bain all these phases

can be defined as body centred tetragonal structures having different cbct- to abct-axis ratios.

Figure 9: Bain path distortion of a body-centred cubic (bcc) cell (“blue atoms”), with a c/a ratio 1, that can be

defined between two face-centred cells (“black atoms”). The bcc cell changes continuously its lattice parameters

c and a upon a tetragonal distortion. Starting from a bcc the lattice distorts to a body-centred tetragonal structure

with c/a = 1.02 upon elongation of the c-axis. When further distorted a face-centred tetragonal (fct) cell with

c/a = 1.33 develops, as defined by the black atoms. Finally the black atoms form a face- centred cubic (fcc) unit

cell with an c/a = 1.41 (adapted from Ref. 63).

Thus the structural sequence from a fcc to a bcc structure can be defined as a continuous

tetragonal distortion where the c/a ratio decreases from fcc to bcc. In Figure 9 the variation of

a unit cell upon tetragonal distortion along the Bain path is depicted. To understand the Bain

path formalism, there are two related systems based either on a face centred or a body-centred

lattice. The fcc austenite phase has a c/a ratio of cfcc/afcc=1 when using the face-centred

assumption (defined by “black atoms” in Figure 9). This can be also described by using the

Page 31: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

19

body centred assumption, where 41.12a/c bctbct == (“blue atoms” in Figure 9). Within this

body centred assumption cbct=cfcc, while the abct axis is aligned along the face diagonal of the

face-centred cell. From geometry fccfccbct c2/1a2/1a ⋅=⋅= can be derived for a body

centred cell within a face centred lattice. The degree of tetragonal distortion can be calculated

from the face- to the body- centred lattice by fccfccbctbct a/c2a/c ⋅= . Within this thesis the

body-centred assumption will be used in the following to describe the degree of tetragonal

distortion and related results. Along the Bain path the c-axis is elongated continuously while

the a-axis is compressed to keep a nearly constant unit cell volume. Both, the fct

(cbct/abct = 1.33) as well as the bct (cbct/abct = 1.02) structure only differ in the degree of

tetragonal distortion. Besides the bcc and the fcc phase only the fct and the bct structures are

stable at T < 200 K. For temperatures T > 200 K several intermediate states between the bct

and the fcc phase with 1.02 < c/a < 1.41 were observed. As described in Chapter 2.1, a

martensitic transformation is a first order transformation accompanied by a step-like change

of intrinsic properties like lattice parameter, magnetic saturation polarization and electrical

resistance. The martensitic transformation appearing in alloys around the Fe70Pd30

composition is of weak first order, since the c/a ratio changes continuously from a fcc to a fct

phase.9,64 At the beginning the c/a ratio increases rapidly when starting to decrease the

temperature to induce the martensitic transformation from the high-temperature fcc austenite

to the low-tempertaure fct martensite phase. This steep increase of the c/a ratio stagnates

when the structure is completely transformed into the fct phase but keeps on increasing

further when the temperature is decreased. Additionally the Fe70Pd30 based alloys exhibit a

small hysteresis in the range of 2-5 K being related to the rather weak first order

transformation.

This weak first order transformation from a fcc to a fct structure around the Fe70Pd30

composition can be interpreted in terms of a band-Jahn-Teller instability, which is originating

from a degeneracy in the electronic band structure at the Γ-point.65,66 This degeneracy gives

rise to an energy gain (∆E) of approximately 14 meV/atom when the structure is changed

from a high symmetric fcc to a lower symmetric fct structure. The energy difference ∆E

between all different structures (bcc to fcc) along the Bain path is rather small allowing to

easily change the structure at finite temperatures initiated by such an energy gain.

Calculations of the energy gain for the single elements Fe and Pd along the Bain path

(c/abct ratio) confirm a stabilization at the known crystal structures (bcc for Fe and fcc for Pd).

Page 32: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

20

Figure 10: Calculated energy landscape as a function of the c/abct ratio (from bcc to fcc) along the Bain path for

Fe (dark grey curve), Pd (grey curve) and Fe70Pd30 (red curve). The pure elements show a distinct energy

minimum at their equilibrium structures (bcc for Fe and fcc for Pd). For Fe70Pd30 a flattening of the energy

landscape (red curve) appears with a minimum around the fct structure.67

When ∆E is calculated for the Fe70Pd30 composition, the energy landscape flattens

significantly with an energy minimum around the fct structure as presented in Figure 10.67

This indicates that only a small amount of energy is needed to alter the structure from fcc to

fct. Figure 11 shows the total energy as a function of the tetragonal distortion of a lattice

along the Bain path. The upper curve is similar to Figure 10 and demonstrates the energy

landscape for Fe70.4Pd29.6 having an energy minimum around the fct structure. To investigate

the impact of relaxation upon tetragonal distortion on the total energy, a 108 atom supercell

was distorted at different c/abct ratios. The location of atoms was not fixed to predefined

position as in Figure 10. Instead a Conjugate-Gradient-Algorithm was used to calculate the

shifting of atomic locations with the lowest energy simulating the relaxation of the lattice

upon tetragonal distortion. Compared to the ideal positions, an energy gain ∆E ≥ 29 meV was

determined for all c/abct ratios upon relaxation of the lattice. The energy landscape of the

relaxed lattice is depicted by the lower curve in Figure 11. Now the energy minimum is

located at the bcc structure, defining a new structural ground state for this system. The energy

difference between bcc and fcc state is increased significantly to about 24 meV.65 This shows

that relaxation mechanisms have a significant impact on the structural ground state for

Fe70.4Pd29.6. Since all ab initio simulations were calculated at T = 0 K only estimations can be

given at finite temperatures. Especially thermal fluctuations and the impact of entropy for

finite temperatures can alter this energy profile leading to a different structural ground state.

Page 33: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

21

Figure 11: Upper curve: Total energy as a function of the c/abct ratio for fixed atomic positions. A rather flat

energy landscape appears with an energy minimum around the fcc structure. Lower curve: When allowing the

atoms to displace their position and to relax the lattice at different tetragonal distortions, a different picture

appears. Compared to fixed atomic positions the lower curve gains ≥ 29 meV in energy. Due to the movement of

atoms the bcc structure defines the new ground state with an energy difference of 24 meV between bcc and fcc

structure.65

For magnetic materials, either in bulk or in thin film geometry, the exchange integral and thus

the magnetic state and properties correspond directly to the ratio of interatomic distance a to

the radius r as known from the Bethe-Slater formalism.68 Therefore the lattice parameter can

be used to vary a material’s magnetic properties. In bulk material the lattice parameters of the

crystal structure can only slightly be varied. However, thin films can be grown epitaxially

where the film orientation is defined by the substrate, achieving a nearly single-crystal like

quality. Further the in-plane lattice parameter can be adapted by the growing film from the

substrate even if its equilibrium lattice parameter is different. When a thin film adapts the

lattice parameters of a substrate by straining itself, although the equilibrium parameter differs

significantly, strained coherent film growth occurs. Coherent film growth with strains in the

range of some percent is usually restricted to very thin layers (∼5 nm), since a large amount of

elastic energy is required to avoid relaxation of the strained lattice. For soft materials this

amount of elastic energy is reduced allowing to coherently grow strained films up to high

thicknesses.69 Therefore materials that show lattice instabilities are well suited regarding

Page 34: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

22

strained coherent growth. SMAs belong to this class of materials showing lattice instabilities

like a softening of the elastic constants near the martensitic transformation.

Figure 12: a) The shift of Fe70Pd30 (002) peak determined by XRD indicates the increase of the out-of-plane

lattice constant of the unit cell and thus the tetragonal distortion. Both the bcc and the fcc structure defining the

boundaries of the Bain path are depicted by dotted lines. b) The different stages of tetragonal distortion c/abct

along the Bain path between bcc (bottom: “red atoms”) and fcc (top: “dark blue atoms”). c) Variation of the

Fe70Pd30 unit cell as a function of substrate lattice spacing of the different buffer layers. The in-plane lattice

parameter abct (“open squares”) is identical to the substrate lattice spacing since it follows a straight line. The

c/abct values (“filled squares”) indicate a constant volume of the unit cell over the Bain path.67

A first attempt was reported by Godlevsky and Rabe70 who predicted the possibility to induce

a cubic to tetragonal distortion with c/a ratios from 0.95 to 1.25 in Ni2MnGa. Buschbeck et al.

adapted this idea to the Fe-Pd system and fabricated 50 nm thick, strained grown epitaxial

thin films at the Fe70Pd30 composition. These films were grown on single-crystal MgO

substrates with different buffer layers on top to adjust the in-plane lattice parameter. This

allowed to strain the Fe70Pd30 unit cell over the full range of the Bain path with

1.09 ≤ c/abct ≤ 1.39. Using XRD a shifting of the out-of-plane lattice parameter (shift of (002)

lattice peak indicated by black arrow) evidences the change of the c/abct-ratio on different

buffer layers as depicted in Figure 12 a). To clarify this, the tetragonal distortion of the unit

cell is sketched in Figure 12 b), with a bcc unit cell (bottom: red atoms) at c/abct = 1.09

towards a fcc unit cell (top: blue atoms) with decreasing in-plane lattice parameter. From pole

figure measurements the in plane lattice parameter was determined for all films on different

buffer layers to calculate the c/abct ratio value. In Figure 12 c) both the c/abct ratio and the in-

plane lattice parameter are presented as a function of the substrate lattice spacing (and thus of

the different buffer materials). The in-plane lattice parameter abct (open squares) is identical to

the substrate lattice spacing attesting the strained coherent growth of the film. The c/abct

Page 35: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

23

values (filled squares) as determined from in-plane lattice parameter measurements indicate a

constant volume of the unit cell over the full Bain path as marked by the dotted line. This

shows that the volume does not change upon tetragonal distortion. Magnetic investigations

revealed a strong variation of the magnetic properties upon tetragonal distortion of a Fe70Pd30

unit cell. Due to the structural variations the magnetocrystalline anisotropy constants K1 and

K3 change significantly. Also the Curie temperature TC shows an increase to 830 K with

decreasing c/abct. These results indicate that it is possible to adjust the crystallographic

structure (c/abct ratio) for a defined composition in the Fe-Pd system by strained epitaxial

growth. Especially the magnetic properties can be adjusted significantly in dependence on the

c/abct ratio allowing to tailor materials properties.

2.7 Ternary Fe-Pd-X systems

The Fe70Pd30 FSMA system exhibits several advantageous properties like the MFIS, high TC

and JS. Therefore this material is promising for sensor- and actuator-application. When

compared to Ni2MnGa, Fe-Pd is more ductile, has a higher magnetocrystalline anisotropy and

higher saturation polarization. For full implementation into technical devices there are several

properties that have to be tailored and further optimized. Especially an increase of

transformation temperatures and stability range of the fct martensite phase is required. A

promising route to enhance these properties is the alloying of third elements into Fe70Pd30.

This was reported in literature, where several third elements were added to Fe70Pd30 and the

change of properties was investigated. The effects of Co and Ni addition to Fe-Pd on the fcc-

fct transformation temperature and magnetic properties were investigated by Tsuchiya et al.71

who determined slightly increased values for saturation polarization compared to those for

Fe70Pd30. Co addition was reported to shift the fcc-fct transformation temperature slightly

higher, while Ni (isoelectronic to Pd) addition shifts it lower. These results suggest that the

relative stability between fcc and fct phases is affected by the electronic structure. Figure 13

shows the martensitic transformation temperature as a function of the e/a ratio. With

decreasing e/a ratio a shift of the martensitic transformation to higher temperatures is

observed. All ternary Fe-Pd-X with X = Co and Ni compositions can be correlated to a

polynomial function as indicated by a black line. Further investigation on Fe-Pd-Co and Fe-

Pd-Ni performed by Vokoun et al.72 and Sánchez-Alarcos et al.73, however, did not find

similar results. They reported a decrease of the fct-fcc transformation temperature upon

addition of Co into Fe70-XPd30CoX with X=2.02; 3; 4.1; 6.25 at.%. The non-reversible

Page 36: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

24

transformation from the fct to bct phase as described in Chapter 2.6 can be decreased to lower

temperatures upon addition of Co.

Figure 13: Martensitic transformation temperature as a function of the e/a ratio. A slight increase of

transformation temperature is observed for Fe68Pd29.9Co3.1.71

A decrease of the magnetostriction values in the fcc austenite phase was also identified due to

Co addition. Fabrication of Fe-Pd nanocrystalline particles with Co additions was reported by

Kovacs et al.74 to increase magnetic anisotropy and reduce formation temperature of the fcc

austenite phase. The addition of Ni into Fe-Pd results in a decrease of transformation

temperature accompanied by a lowered Curie temperature.71 Lin et al. reported an increase of

magnetostriction due to the addition of Ni into Fe70-XPd30NiX with X= 2; 4; 6 and 8 at.% when

compared to Fe70Pd30.75 Further they found, that a doping of Fe-Pd with Ni prevents the

decomposition of the fcc austenite phase when annealed at 673 K for 100 h and thus

stabilizing the parental phase against decomposition.76 An increase of anisotropy in Fe-Pd by

alloying with Pt was envisaged in Stern et al.77, but could not be verified experimentally.

Alloying Fe-Pd with Pt (isoelectronic to Pd) was found to decrease transformation

temperatures.78 Takeuchi et al. reported the combinatorial fabrication of Fe-Pd-Ga thin

films.79 Magnetic measurements were performed for this system and showed results, which

correlated well with the magnetic properties of the known sub-systems Fe-Pd and Fe-Ga.

Structural investigations of aged Fe66Pd30Rh4 were reported by Lin et al. using TEM, and

XRD measurements.80 They revealed a decrease of transformation temperature to T = 170 K

and the existence of a monoclinic intermediate structure between the fcc austenite and the fct

martensite structure. The alloying of Mn into Fe-Pd was performed by Alarcos et al.

motivated by the ferromagnetic-antiferromagnetic transition as described by the Bethe-Slater

Page 37: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

25

curve, which can enhance magnetic properties (increased magnetocrystalline anisotropy and

saturation magnetization).81 They found a significant increase of transformation temperatures

for Fe69.4-XPd30.6MnX (x = 0; 1; 2.5 and 5) compositions. All these reports are based on

experimental investigations giving only partial insight how a single element alters the

properties within the Fe-Pd system. For the Fe-Pd-Co and Fe-Pd-Ni system a defined

variation of transformation temperature was correlated to the e/a ratio without describing

significant trends that can be used to systematically control the system’s properties. A

systematic theoretical prediction to enhance the Fe-Pd alloy was given by Opahle on the basis

of DFT calculations.82 In addition to their important correlation of the formation of the fct

phase in the Fe-Pd system to the band-Jahn-Teller effect, the authors pointed out that

increasing the minority spin density of states at the Fermi level should destabilise the fcc

austenite. This is regarded as the origin for the formation of the fct martensite at higher

temperatures. Therefore they suggested to optimize the Fe content in order to systematically

shift the martensitic transformation temperature. This is also observed from experimental

results by various groups, where a shift of the martensitic transformation can be directly

linked to the Fe content in Fe-Pd-X with X = Co, Ni and Pt.

2.8 Routes for the development of novel Fe-Pd-X alloys

One of the aims of this thesis is to investigate how different fabrication methodologies and

sample designs affect the materials properties of the Fe-Pd system. The second aim is the

development of novel Fe-Pd-X FSMAs with improved properties. These properties are

defined by benchmark values in the following of this thesis.

As presented in the previous chapter, there is only sparse information on ternary Fe70Pd30

based FSMAs. There were less than 25 papers published within that field of research so far.

Thus, the subject of this thesis is a systematic approach to clarify how the properties of binary

Fe70Pd30 can be improved by the addition of third elements. Further a comprehensive

understanding will be provided in order to allow for a tailoring of properties regarding future

application. A fundamental prerequisite to enhance Fe70Pd30 is therefore to define which

criteria are important and which benchmark values have to be reached:

a) a high martensitic transition temperature Ms > 350 K

b) a reversible martensitic transformation ∆T < 5 K

Page 38: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

26

c) a high saturation polarization JS > 1 T

d) a high magnetocrystalline anisotropy at RT K1 > 100 kJ/m3

e) a high Curie temperature TC > 550 K (TC > MS)

f) highly movable twin boundaries (MFIS of several %)

g) a high blocking stress (σbl > 10 MPa)

A high martensitic start temperature (Ms) for Fe-Pd-X alloys is important to assure the

occurrence of the martensite phase that exhibits the MFIS, around operating temperature.

Many applications require a thermal stability of the implemented functional material in a

temperature regime up to 350 K, as for example required in the automotive sector. The

thermal hysteresis ∆T needs to be small in order to allow for a fast switching without

additional energy consumption (by varying the temperature) between austenite and martensite

phase. The saturation polarization JS contributes to the overall energy product of an alloy that

limits the maximum energy transduction from magnetic to mechanical energy output. This

value needs to be increased because it allows the use of low external magnetic fields to obtain

the maximum energy input at the anisotropy field HA = 2 K/JS and thus enhancing its

efficiency. As an intrinsic material property, the anisotropy constant K represents the

maximum energy density which can be supplied by an external magnetic field.83 The Curie

temperature TC defines the upper limit where the alloy is ferromagnetic. This value should not

interfere with the martensitic transformation and thus needs to be significantly higher than

MS. Further the ferromagnetic to paramagnetic transition has to be shifted to high values away

from operating temperature. This is related to continuous decrease of magnetization with

temperature that breaks down when the Curie point TC is reached. This lowers JS and the

maximum energy output significantly and thus leads to a decrease of efficiency around the

operating temperature regime. The maximum achievable strain by Fe-Pd-based FSMAs is

another important criterion to be fulfilled. Conventional SMAs exhibit high mechanical

strains but cannot be operated at high frequencies, since these are limited by the material’s

ability to loose heat. In contrast to this, high-frequency actuation is performed by materials

with rather low strains like piezoelectric materials. FSMA could combine these two important

criteria by allowing for high actuation frequencies (up to the kHz regime) and high strains at

the same time and therefore are of high interest for many actuation applications. An increase

of the maximum MFIS in Fe-Pd-based alloys from 3.5% to higher values, as known for

Ni2MnGa, is therefore required. The blocking stress σbl defines the force per area unit below

Page 39: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

27

which any twin boundary movement is blocked and presents the upper limit of mechanical

actuation load.

Regarding these benchmarks, a fundamental question arises, asking which elements should be

added into Fe-Pd to enhance the alloy and fulfil these prerequisites. The elements Ni and Co

are ferromagnetic around room temperature but do not enhance significantly the properties, as

reported in literature. A promising candidate to enhance intrinsic properties can be achieved

by altering the magnetic coupling and thus lattice parameters in terms of the amount of

tetragonal distortion (c/a ratio) in the fct phase. When adding antiferromagnetic elements into

Fe-Pd, a modification of the lattice by altering the exchange interaction can be achieved. As

mentioned before, the magnetic exchange interaction depends strongly on interatomic

distances. Magnetic inhomogeneities introduced in this way can alter the magnetism in

FSMAs and affect the lattice parameters. This influences the martensitic transformation and

thus can be used to significantly adjust the FSMA properties. Magnetic excitations can couple

to the system at finite temperatures, when spins of Fe and antiferromagnetic elements flip

against their preferred orientation at T = 0 K. Such effects were reported previously for

related systems, where magnetic inhomogeneities caused antiparallel spin ordering.84 A

promising candidate element to improve the properties of Fe70Pd30 is elemental Mn, which

can couple ferromagnetically as well as antiferromagnetically with Fe. Antiferromagnetic Mn

has a higher orbital moment when compared to Fe and could further increase the overall

orbital moment when added into Fe-Pd. Nevertheless, a general correlation of martensitic

transformation and composition within ternary Fe-Pd-X alloys was not identified up to now.

Several publications report a correlation of transformation temperatures with e/a ratio. This

was shown for Fe-Pd-Ni and Fe-Pd-Co alloys but did not reveal conclusive results for other

Fe-Pd-based alloys like Fe-Pd-Mn, Fe-Pd-Cu. Fe-Pd-Rh, Fe-Pd-Ga and Fe-Pd-Pt. To prove

that the e/a ratio is the decisive factor controlling the transformation temperatures, the

addition of Cu is promising. Cu is in a fcc structure under ambient conditions and allows to

increase the e/a ratio while keeping for example the Fe content constant and thus varying one

parameter without changing the other.

Regarding all these facts, this thesis focuses on investigating the impact of Mn and Cu as third

elements added into Fe-Pd in order to enhance intrinsic properties.

Page 40: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

28

2.9 Combinatorial materials science

Optimization of materials based on prior knowledge is a process that was performed since the

beginning of using tools and arms in mankind history. Due to technical and scientific

restrictions, the optimization of materials was a sequential and therefore tedious procedure.

Nowadays, rational planning and automation allow to accelerate novel developments in

materials science. This route focuses on detailed knowledge regarding relation of intrinsic

properties of materials to a set of their performance properties. Usually this knowledge was

obtained from experimental and simulation data. The first use of combinatorial techniques for

solid state materials can be ascribed to Boettcher et al.85 in 1955, while Kennedy et al.86

developed the first thin film composition spreads using co-evaporation in the late 1960’s. In

1970 Hanak introduced an integrated materials-development workflow that defines four key

aspects including a) a complete compositional mapping of a multicomponent system in one

experiment, b) simple rapid non-destructive compositional analysis, c) testing of properties by

a scanning device and d) computer data processing.87 This workflow can be considered as a

first attempt to establish a combinatorial approach for the development of new materials.

Xiang and Schultz reinitiated combinatorial routes in materials science in 1995.88 In the

combinatorial approach multiple samples with different compositions are synthesized and

investigated by high-throughput characterization methods. This was successfully performed to

discover new and to enhance known functional89, magnetic90, catalytic91 and optical92

materials, as well as polymers93. The advantage of combinatorial materials science can be

understood when calculating the number of possible Fe-Pd-X-Y systems. There might be a

possible number of 15 elements (alkaline earth metals, transition metals, lanthanides, non-

metals and metalloids) that could be promising to be added into the binary Fe70Pd30 system to

enhance the materials properties. This yields a number of 15 different Fe-Pd-X ternary

systems, each consisting of 5050 samples with different compositions when using a

compositional increment of 1 at.% per sample. Extending this to quaternary Fe-Pd-X-Y

systems the number of different systems increases to 105 with more than 500.000 samples per

system. This can be further extended when considering not only the composition of the

starting materials but also the different conditions for processing. For a simple catalytic

material the number of experimental runs is rapidly increased to several millions (see Figure

14).94

Page 41: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

29

Figure 14: Examples of diversity of materials compositions, process factors, and operation conditions applicable

for combinatorial screening. A) Dependence of the number of possible systems on the number of components.

Red: systems investigated up to now.95 B) Factors and their levels for one-step synthesis of diphenylcarbonate.96

C) Hyperspace of features of materials and measurements in chemical sensors.97 (Figure from Ref. 94)

All these considerations indicate the need for a highly automated approach to effectively find

new and enhance known materials by fabricating materials libraries within one single process.

These libraries have to be investigated by high-throughput methods in order to rapidly

characterize materials properties of interest. Correlating these properties a deeper

understanding for the structure-property-relationship in materials are aimed to be achieved in

a significantly reduced time frame.

In general, there are different types of materials libraries: In discrete libraries every sample is

locally separated and has an individual composition. In continuous libraries the samples are

not separated and a compositional gradient occurs over the whole substrate and appears even

over the sample. Diffusion couples can be regarded as a predecessor of continuous libraries in

bulk materials and were introduced by Zhanpeng98, Miller99 and Goldfarb100. Thin film

composition spreads were fabricated by co-deposition from different sources like

evaporation86 and sputter-deposition.99 Further multilayer deposition101 of wedge-type layers

was performed to induce a thickness and thus a compositional gradient across a substrate.

Page 42: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

30

After fabrication, automated high-throughput characterization is usually applied to gain

information about the materials library. Properties to analyse are composition of the samples

by energy- (EDX) and wavelength- dispersive X-ray analysis (WDX), Rutherford backscatter

diffraction (RBS), nuclear reaction analysis (NRA) and Auger-electron spectroscopy. The

structure can be determined by X-ray diffraction (XRD) and electron backscatter diffraction

(EBSD). Further magnetic properties are revealed by scanning superconducting quantum

interference device (SQUID) and magneto-optical Kerr effect (MOKE) measurements.

Mechanical properties can be investigated by using nanoindenter mappings. Next to these

common methods there are specialized techniques like optical methods for Hydrogen storage

materials102, photoluminescence of materials92, Seebeck microprobe103 and scanning-tip

microwave near-field microscope104. Further possibilities for high-throughput characterization

and processing of thin film materials libraries are devices fabricated by micro-electro-

mechanical system routes. In general these devices can be separated into active and passive

devices. Passive devices are shadow mask structures used during thin film deposition

processes and cantilever substrates90,105 that allow measuring a thin film’s stress states.

Further devices are micro-tensile106,107 and bulge test108,109 structures to determine mechanical

properties in thin film samples. Active devices have incorporated sensor or actuator structures

that allow for a processing of thin film samples deposited onto. Arrays of micro-hotplates110

can be used to in situ characterize a thin film sample under varying annealing parameters

(heating and cooling rates up to 104 K/s) way beyond what can be realized with a

conventional apparatus. Nanocaloric devices allow to determine transformation enthalpies of

nano-sized films in situ during annealing.111

After characterization, the materials properties have to be correlated in order to allow for an

understanding of structure-property relationship. A first attempt was reported by Hanak87,

who used computer processing in tabular, graphical and functional forms as part of the

“multi-sample-concept”. But his attempt lacked of resource capabilities, since computers were

not able to handle huge amounts of data in those days. Today data analysis is performed by

using dedicated software with various capabilities for cross-referencing and correlation

functions. Typical data correlation and mining software is developed by numerical computing

environments like MATLAB, ORIGIN and LabVIEW. The main application for such

software lies in the area of multi-property correlation, pattern recognition and clustering

analysis to identify areas of similarity. As an example, the XRDSuite package112 software

developed in the group of Prof. I. Takeuchi are here mentioned. This software allows

correlating spectral data with several integer data sets for many thousands of samples. Further

Page 43: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

31

cluster analysis and principal component analysis (PCA) are implemented within to facilitate

the laborious analysis of analyzing large sets of spectral data, by sorting patterns by their

similarity into discrete groups and subsequently deducing the representative basis X-ray

patterns.113

2.10 Thin film nucleation and growth

In a thin film deposition process atomic, molecular or ionic species are generated and

transported from a source to a substrate through a medium. In a sputter deposition process this

medium is the gas phase from which the atomic particles condensate on the substrate. When

the atomic species have a sufficiently small kinetic energy, they physiosorb onto the substrate

surface in the moment of condensation. Since the atomic species are not in equilibrium with

each other, they diffuse on the surface until they interact and form clusters. These clusters

decay directly or continue to grow until reaching a critical radius where they are

thermodynamically stable. These thermodynamically stable clusters are then defined as

nuclei, from which the thin film growth starts. The critical radius rcr is defined by:114

vcr G

2r

∆γ−= (8)

The specific surface free energy is γ and the volume free energy is denoted by ∆Gv. After

formation the film growth starts at these nucleation sites.

Figure 15: Schematic of film nucleation during thin film deposition. Depicted are atoms depositing out of the

plasma onto a substrate. Surface diffusion allows the atoms to conglomerate and to form nuclei. Blue arrows

indicate the energies involved in the deposition and conglomeration process. (Adapted from Ref. 115)

Young’s equation can be used to describe the different types of film growth in dependence on

surface energy γsv of a solid matter, surface tension energy γfv of a film nucleus droplet of the

Page 44: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

32

deposited material, interface energy γfs of this droplet and angle θ between solid matter and

droplet (Figure 15):116

fv

fssv)cos(γ

γ−γ=θ (9)

According to this equation, there is a restricted number of angles that can appear. In

dependence on the angle θ between solid matter and droplet of deposition material there are

different types of film growth. When θ > 0° and γsv < γfs + γfv the depositing atoms can diffuse

over the substrate and conglomerate to three-dimensional islands. With increasing number of

arriving atoms these islands grow and new islands are formed. In the initial state the substrate

is not fully covered. With increasing thickness the islands start to coalesce and a continuous

film is formed. This kind of growth is defined as Volmer-Weber growth117 as depicted in

Figure 16 a). When θ ≈ 0° several monolayers are formed before island growth occurs.

Figure 16: Different kinds of thin film growth. a) Volmer-Weber growth is defined by formation of islands that

grow in size (from 1 to 4) until they coalesce. b) Stranski-Krastanov growth forms monolayers (1 to 2) before

island growth starts (3). These islands increase in size with ongoing deposition (3 to 4). c) The formation of

monolayer after monolayer (1 to 4), each covering the full substrate surface is described by Frank-van-der-

Merwe growth. This figure is adapted from Ref. 116.

This growth mechanism is named after Stranski-Krastanov118 and is depicted in Figure 16 b).

When the angle between surface and film nucleus is θ = 0° the film grows monolayer per

monolayer on the substrate fully covering the surface before a further layer is growing (see

Figure 16 c)). This case is defined as Frank-van-der-Merwe growth119 where γsv ≥ γfs + γfv is

fulfilled. Next to the angle between film nucleus and solid matter, saturation of the vapour

phase affects the film growth further. When the vapour phase is saturated with atoms/atomic

clusters over a certain level, layer growth is observed, since the amount of Gibb’s free energy

to form a two-dimensional nucleus is lower than for a three-dimensional nucleus. Stranski-

Krastanov growth can occur when there is a higher misfit between forming film and substrate,

Page 45: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

33

although the interface energy would prefer a layer growth. In this case a first atomic layer is

growing which adapts to the substrate’s lattice parameter. When the film thickness increases,

elastic energy for adapting the lattice parameter is needed. If the adhesion energy for

conglomeration of the atoms is lower than the amount of elastic energy, three-dimensional

island growth occurs. Further the substrate temperature affects significantly the kind of film

growth and the film morphology. When the temperature is low the diffusion of atoms on the

substrate surface is restricted leading to an increase of small film nuclei. At higher

temperatures the atoms have enough energy to diffuse over long distances of the substrate

surface giving rise for less but larger film nuclei. According to Thornton’s structure zone

model120 a highly defect rich microstructure is formed when the substrate temperature Ts is

below a third of the source material’s melting temperature Tm. This model was developed for

polycrystalline thin film growth at high deposition rates. In this temperature range, shadowing

is the major effect leading to overgrowing of small by larger grains and thus to a high amount

of interfaces and pores. When the substrate temperature during deposition is increased to

0.3 < Ts/Tm < 0.5 surface diffusion of atoms is mainly occurring leading to a columnar grain

growth. At higher temperatures an increase of the grain size is observed.

In most cases thin film growth starts at different locations on the substrate that coalesce with

increasing amount of deposition materials to a closed film with a polycrystalline

microstructure. Thin films often show a preferred orientation/texture of the microstructure.

Many materials with a fcc (bcc) structure show a distinct (111) fibre-texture ((110) fibre-

texture). This is originated by the fact, that it is energetically favoured to have the lattice

planes with the highest number of atoms oriented to the film plane.121 Next to substrate

temperature, also the deposition rate, mutual inter-atomic reactions and the structure of the

substrate affect the film growth. When nucleation sites are forming on a surface having

similar lattice parameters, the coating atoms adapt to the structure of the surface. It is

energetically favoured for the nucleus to adapt to the periodicity of the underlying substrate

because the surface potential energy shows minima at the substrate’s atomic positions. When

the substrate’s lattice parameter is adapted by the film nucleus pseudo-morphed growth

occurs. This kind of growth is usually restricted to very thin films because there is often a

misfit between the substrate and the thin film lattice parameter. This misfit results in an

amount of elastic energy incorporated in the film structure and is often released at higher film

thicknesses by the formation of misfit dislocations.

Page 46: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

34

The presence of a defined relationship between crystal structure of substrate and thin film

material is defined as epitaxial growth. This relationship is defined by the lattice planes and

directions that are parallel in film and substrate:

(HKL) Substrate || (hkl)Film ; [UVW] Substrate || [uvw]Film (10)

In dependence on the difference in crystal structure and lattice parameters of substrate and

thin film material there can be a misfit. This misfit depends on the lattice parameter of

substrate dSubstrate and of the thin film dFilm and is defined by:

Substrate

SubstrateFilm

d

ddMF

−= (11)

The difference in crystal structures and the misfit can give rise to a film growth where the unit

cell is rotated at 45° to the substrate unit cell. This can occur when a fcc structure is forming

on top of bcc unit cells.122 Epitaxial growth can be achieved only in a small regime where the

substrate temperature is above a critical temperature and the amount of coating atoms is

sufficiently low. The atoms in a deposition process need a certain amount of energy to diffuse

over the surface and a certain amount of time to adapt to the substrate structure before further

atoms come in.

Page 47: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

35

3. Experimental methods

3.1 Fabrication and Processing

The methods used for the fabrication and the processing of all samples, especially the

fabrication of binary and ternary materials libraries performed in different sputter systems, are

explained in the following. A short description of the experimental planning prior to the

deposition of materials libraries is pointed out. Epitaxial growth of thin films is presented in

the next section and followed by the fabrication and processing of bulk samples.

3.1.1 Thin film materials libraries

The materials libraries were fabricated by physical vapour deposition (PVD) using an ultra-

high vacuum combinatorial sputter system (CMS 600/400LIN) from DCA, Finland and a CS

730 S cluster system from Von Ardenne. 4-inch Si (100) wafers (single side polished, prime

quality) with a 1.5 µm thermal SiO2, working as diffusion barrier, were used as substrates for

all materials libraries. After deposition a lift-off process was performed to structure the thin

film materials library into discrete 3 mm x 3 mm samples. This patterning allowed to

distinguish between the different samples and further to define every sample to a specific

location on the Si/SiO2 substrate. The sputter system consists of a load lock, a central

distribution chamber (CDC), a mask storage chamber and two sputter chambers. Within this

work, both sputter chambers were used to fabricate the materials libraries. The sputter

chamber K1 allows for a sequential deposition, while the second chamber K2 enables the

simultaneous deposition from elemental or alloy targets. In both chambers the substrate is

located on a rotatable and height-movable substrate holder with an integrated SiC heater (293

K ≤ T ≤ 1273 K). To pre-clean the substrate and to enhance adhesion of the thin films, a bias

voltage can be applied to the manipulator. The K1 sputter chamber is equipped with six

magnetron cathodes (three working in DC and three in RF mode) on a movable arm (Figure

17).

Page 48: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

36

Figure 17: Schematic of the sputter chamber (K1) for sequential deposition of wedge-type layers (drawn by D.

Grochla). The inset depicts three ignited sputter cathodes (the plasma colours indicate different elements to be

deposited). Two wedge-type layers were already deposited in an angle of 120° apart. The movable arm locates

the sputter cathode having green plasma above the Si/SiO2 substrate in order to deposit the next wedge-type

layer.

Target-substrate distance was set to 87.5 mm. A main shutter is placed directly below the

magnetron cathodes to control the deposition time during the process and to avoid

unintentional deposition. Four independently controlled shutters (4D shutters) are located

above the substrate and are placed 90° apart. These shutters can be used to create various

wedge-type layers and shield different parts of the substrate during deposition. To fabricate a

wedge-type layer a main shutter shields the substrate during deposition and retracts slowly

(moving speed between 1 to 5 mm/s) in order to vary the amount of material deposited across

the substrate. Different retraction lengths can be used to vary the area and thickness of the

wedge-type layer over the substrate. This allows controlling the film thickness of every wedge

layer down to zero thickness. The K2 sputter chamber has five 45° tilted magnetron cathodes

(four working in RF and one in DC mode) mounted in a circular geometry (similar to the

geometrical mounting in Figure 18). The focal point of each cathode is pointing to the middle

of the substrate having a target-substrate distance of 185 mm. All materials libraries were

prepared by sputtering from elemental targets (100 mm diameter, purity 99.99%) that were

pre-cleaned by sputtering prior to deposition in order to remove surface layers on the targets.

The typical base pressure before deposition in both chambers was lower than 4.8·10-5 Pa. As

deposition-gas Ar (6N) at a pressure of 0.67 Pa and under a flow of 20 sccm was used. All

Page 49: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

37

depositions in K1 and K2 were started by fabricating a homogeneous Fe layer (d ≈ 10 nm) to

enhance adhesion between substrate and thin films and were ended by depositing a

homogeneous Pd film (d ≈ 10 nm) to avoid oxidation in air.

Figure 18: Schematic of the confocal mounting of the sputter cathodes used for fabrication of materials libraries

(geometrical mounting similar to K2) by co-deposition (drawn by D. Grochla). The sputter cathodes are

confocally angled towards the substrate. The inset depicts three different cathodes during the co-deposition

process (the plasma colours are correlated to different elements to be deposited). The slopes of the elemental

wedges are fixed by the angle of the different cathodes. The angle between the elemental wedges is fixed due to

the position of the cathodes relative to the substrate.

The film thickness for all samples was set between 100 nm to 700 nm. Next to depositions

carried out in the DCA sputtering system, binary Fe-Pd materials libraries were fabricated

from a Fe70Pd30 alloy target (purity 99.95%) using a CS 730 S Von Ardenne cluster system.

This system has two sputtering chambers each equipped with four sputtering cathodes for

targets with 75 mm and 150 mm in diameter. The target to substrate distance during

deposition was around 50 mm. Further a load lock, a distribution chamber and a high-

frequency etching chamber is attached to the sputtering system. The typical base pressure

before deposition in both chambers was lower than 6·10-5 Pa. As deposition-gas Ar (6N) at a

pressure of 0.1 Pa and under a flow of 25 sccm was used. An aperture located below the

sputtering cathodes was used to protect the substrate from unintended deposition. Due to the

variation in plasma density below the target during sputter deposition, the composition of

deposited thin films differed from the nominal target composition across the substrate.

Page 50: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

38

Binary composition spreads were fabricated in both sputter chambers (K1 and K2). To

fabricate a binary composition spread the amount of each element has to be increased

continuously over the diameter of the substrate in a wedge-like fashion. Performing this for

two different elements having an angle of 180° apart from each other, a gradual change in

composition and thus a binary composition spread is fabricated. After deposition of the first

wedge the substrate was rotated 180° and a wedge from the second element was-deposited

opposing to the first wedge (Figure 19 a)). Both wedges have a thickness < 15 nm at the

thicker end to allow to form homogenous thin films by subsequent adequate annealing. This

procedure was repeated until the desired film thickness was achieved. Binary spreads by co-

deposition were fabricated in K2 where simultaneous deposition from two elemental targets

was performed. By changing the deposition-rates, the area covered in the binary composition

diagram, is varied. The overall thickness in this process is controlled by deposition time and

sputter rate. To fabricate binary thin film materials libraries with small variation in

composition, a Fe70Pd30 alloy target was used. Sputter deposition was performed at ambient

temperature (293 K) without heating the substrate, although the temperature increases by a

maximum value of ∆T = 10 K due to the sputtern deposition process. The deposition process

was carried out under static conditions without any relative movements between target and

substrate. An aperture was used to shield the substrate during ignition and was retracted when

the target reached working condition. The deposition rates were used to control the final

thickness of the materials library by deposition time.

Figure 19: a) Fabrication of a binary materials library spread by two opposing wedge-type thin film layers. Next

to the region, where both wedge-type layers are overlapping (green: A-B), there are additional pure elemental

regions located at the upper (light green: B) and lower part (carmine red: A) of the substrate. b) Alignment of the

three different wedge-type layers to fabricate a full ternary materials library. The colour indicates the different

deposition materials. All wedge-type layers are oriented to each other by an angle of 120°. The Si wafer

substrate is sketched by black lines and shows the areas of overlapping wedges. The triangular shaped area

depicted by A-B-C defines the region where the ternary materials library is located. The edge regions (A-B, A-C

and C-B) depict locations on the wafer where only binary compositions are deposited.

Page 51: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

39

Ternary composition spreads were fabricated by depositing three wedges of different

elements. In K1 the deposited wedge-type layers have an angle of 120° to each other. After

fabrication of a 10 nm thick homogenious Fe adhesion layer, the first wedge was-deposited by

retracting the main shutter over a defined distance (70 mm) across the substrate.

After this, the substrate was rotated 120° and the next wedge was deposited. Another rotation

of the substrate at 120° allowed to fabricate the last wedge. This procedure generates a full

ternary materials library including all binary compositions as depicted in Figure 19 b). To

fabricate thicker films this process was repeated until the final thickness was achieved. It

allows to cover the full area of all ternary compositions by one materials library. In K2

simultaneous deposition from three elemental targets was used to fabricate ternary Fe-Pd-X

materials libraries. Due to the geometric arrangement of the sputter cathodes in K2, the

elemental wedges cannot be aligned at 120°. Nevertheless, this method was used to fabricate

materials libraries covering the region of interest in ternary composition diagrams.

Figure 20: Ternary composition diagram showing the difference between calculated (blue triangles) and

experimentally determined (black squares) compositions of a materials library indicating a difference < 5 at.%.

All the presented techniques for preparation are very effective to fabricate materials libraries

consisting of a large number of single samples with different compositions. To control the

area within the ternary composition diagram, the composition in the centre of the substrate

was determined. This only allowed to determine the centre of the area in the composition

diagram, that is covered by the materials library. It did not give any information about the

variation of composition from neighbouring samples (step size of composition) and also did

not allow to determine the compositions at the border of the materials library. Thus only little

Page 52: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

40

information about the compositional range, covered by the fabricated materials library, was

known. In order to enhance process control a model was developed to calculate the

composition for each sample of the materials library in advance. Therefore the variation in

thickness for all three elemental thin film wedges was determined and the composition was

calculated by adding the different film thicknesses. Stripes of every element were deposited at

defined deposition parameters (retraction speed of main shutter, cathode power, Ar flow and

pressure) to determine the distribution of film thickness across the substrate. These film

thicknesses as a function of location on the substrate were then used to calculate the

composition for all samples by using the specific molar volume. All calculations were

performed using Origin software from OriginLab Corporation and self-written templates to

speed up and facilitate experimental planning. In Figure 20 the calculated and measured

compositions (energy-dispersive X-ray analysis) for all samples of a materials library are

presented. This shows that the difference in composition between calculation and

measurement is < 5 at.%, allowing to adequately target the compositional region covered by

the materials library.

After deposition, all materials libraries were annealed in a furnace (Schmetz IU 54 1F) under

N2 atmosphere at 80 kPa and then quenched to room temperature in N2 overpressure (40 kPa)

with a cooling rate of approximately 15 K/s to achieve the metastable transforming phase and

to avoid decomposition. After optimizing the annealing process (1073 K < annealing time <

1173 K; 15 min < dwell time < 70 min) to achieve transforming samples in a single fcc phase

state by using binary Fe-Pd samples, the ternary materials libraries were annealed at 1123 K

for 30 min.

3.1.2 Epitaxial thin films

In order to further investigate selected compositions showing enhanced intrinsic properties,

epitaxial growth of Fe-Pd(-X) thin films was performed in close collaboration with S. Fähler

and S. Weiss at the IFW Dresden. Epitaxially grown thin films are advantageous to determine

intrinsic materials properties like the Curie temperature (TC), anisotropic saturation

polarization, magnetocrystalline anisotropy and crystal structure. These properties are hardly

to investigate in polycrystalline thin films. The thin films were deposited in a Bestec sputter

system equipped with four 2-inch and one 4-inch sputter cathode. The base pressure prior

deposition was p < 5⋅10-7 Pa and the thin films were deposited from Fe, Pd and Cu (purity

99.99%) 2-inch and one 4-inch Fe69Pd31 alloy target (purity 99.99%). The deposition was

Page 53: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

41

carried out by using Ar (6N) at a pressure of 0.8 Pa and under a flow of 5.8 sccm. In

dependence on targeted composition the applied direct current (DC) power was set to 60 W

for Fe70Pd30, 120 W for Fe, 90 W for Pd and 50 W for Cu. The substrate to target distance

was approximately 90 mm. To investigate the variation of intrinsic properties as a function of

different c/a ratios (ratio of the long to the short axis), substrates with different lattice

parameters were used (discussed in detail in Chapter 4.2.3). Therefore epitaxial metallic

buffer layers (Cr (c/a=1.09), Au (c/a=1.12), Rh (c/a=1.33), Ir (c/a=1.39) and Cu (c/a=1.54))

serving as growth surface for the strained growth of Fe-Pd-X films were deposited at 573 K

onto MgO(100) single crystals of 10 mm x 10 mm x 0.15 mm in size. The nominal sequence

for all epitaxial grown thin films is: MgO / 50 nm buffer / 50-1000 nm Fe-Pd-X thin film.

These buffer layers were deposited by sputter deposition from 2-inch targets (Cr, Au and Cu)

as well as pulsed laser deposition (Rh and Ir). To promote epitaxial growth of buffer layers

having a fcc crystal structure (Au, Rh, Ir and Cu) a 5 nm thick Cr layer was-deposited onto

the (100) MgO substrate. Buffer layers of elements having a bcc crystal structure directly

grow epitaxially without the need for any intermediate layer to promote growth.

3.1.3 Bulk samples / Splats

Bulk samples of selected compositions, identified by thin film screening, were fabricated for

further investigation. The fabrication as well as the annealing for all splats was performed by

I. Claussen in the group of Prof. S. Mayr at the University of Göttingen. In order stabilize the

metastable Fe70Pd30 phase in the bulk samples, dedicated fabrication methods are needed. The

splat-quenching or ultra-rapid-quenching (URQ) technique123,124 allows to rapidly solidify

materials from the liquid phase at high cooling rates. Thus it is well suited to fabricate bulk

samples showing the metastable Fe70Pd30 phase after fabrication. The targeted composition

was weighted from elemental slugs of the different elements (purity 99.95%), allowing for a

compositional accuracy of approximately 0.01 at.%. Weighted slugs were then alloyed to

produce small ingots with a mass from 1 g to 2 g. These slugs were located on a water-cooled

Cu plate in an arc melting apparatus with a base pressure of 3⋅10-1 Pa. Prior to the fabrication

process the arc melter chamber was purged six times with Ar (purity 99.998%) to prevent the

samples from contamination and oxidation. The source material (slugs) was then melted

under an Ar atmosphere 6⋅105 Pa to fabricate the ingot. For homogenization the ingots were

remelted six times from both sides. Subsequently, the ingot was divided into samples of about

Page 54: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

42

0.18 g each, which were remelted by using the arc melter to form small spheres for splat-

quenching.

An URQ apparatus from Edmund Bühler GmbH was used to fabricate the bulk

samples/splats. The small spheres were located on a ceramic sample holder within the middle

of a high-frequency coil in this apparatus. The process chamber was pumped and purged by

Ar (purity 99.998%) six times to avoid contamination and a process pressure of 5⋅105 Pa Ar

was set. Then a high-frequency current was applied to the coil leading to an inductive melting

and a levitation of the liquefied sphere. When the material was completely melted the

electrical current was switched off and the liquefied sphere fell down.

Figure 21: Schematic of the splat-quenching process. a) A high-frequency coil inductively generates circling

currents leading to levitation and melting of a sphere. b) When the sphere is fully liquefied (yellow depicts a

glowing sphere) the current is switched off and the falling droplet is splatted between two Cu pistons creating a

thin foil sample (adapted from Bühler GmbH125).

On this way the sphere passed two photo sensors that activated two water-cooled Cu pistons

to splat the liquefied sphere (Figure 21). The resultant splats have diameters between 20 and

30 mm and thicknesses between 50 to 60 µm and are compositionally isotropic within the

resolution of the energy-dispersive X-ray analysis. Due to the high cooling rates in the range

of 106 K/s the microstructure of the splats shows a high defect-density and thus internal

stresses. To heal out these defects and to reduce the internal stresses, selected compositions

were annealed subsequently. Selected splats were encapsulated into quartz tubes under an Ar

(purity 99.998%) atmosphere of 6⋅105 Pa and annealed at temperatures between 873 and

1123 K for different dwell times ranging from 15 min up to 1 h. For rapid quenching the

quartz tubes were dropped into water after annealing. Slower cooling rates were performed by

Page 55: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

43

cooling the quartz tubes at ambient air (≈ 0.5 K/s) or keeping them in the tube furnace

(< 0.2 K/s) until the furnace reached ambient temperature (293 K).

3.2 Characterization

The methodologies to investigate the intrinsic and extrinsic properties of the samples are

presented in the following, which starts with a description of the energy-dispersive X-ray

analysis that was used to measure the composition of all samples. This is followed by Monte

Carlo simulations determining the penetration depth of the electron beam into a thin film

sample. Structural investigation methods are explained in the next subsection explaining the

different diffraction techniques and the importance of measurement parameters for the quality

of results. The high-throughput investigation technique to determine samples undergoing a

martensitic transformation is shown in the following subsection. Magnetic screening as well

as high resolution techniques are presented in the subsequent section. The last subsection

emphasizes on the determination of mechanical properties by nanoindentation.

3.2.1 Energy-dispersive X-ray analysis (EDX)

In order to investigate the chemical composition, all samples were measured by energy-

dispersive x-ray analysis (EDX). Samples were located inside a scanning-electron-microscope

(Leo 1430 VP or Jeol JSM 5800LV) equipped with an Oxford Instruments Inca EDX system

(polymer window, Si-Li detector with 30 mm2 area, Peltier-cooled). This system allows to

quantitatively determine elements from B to U. All samples were measured at an acceleration

voltage of 20 kV, 600 x magnification (integrating the measurement over an area of 600 µm

by 400 µm) and at a process time = 5 (giving a minimum energy resolution of 132 eV)

resulting in 20-30% dead time (Inca-EDX is a dead time corrected system) during

measurement. In order to provide the highest accuracy the measurement time was set to 60 s,

resulting in spectra with more than 250.000 counts. It has to be mentioned that longer

measurement times do not enhance the accuracy of composition measurement due to a

saturation of the signal-to-noise ratio in the EDX spectra (Bremsstrahlung and

background).126 Prior to every EDX-analysis a Co standard was measured at the same

parameters in order to determine the beam-current and thus correction coefficients for the

later EDX analysis. Fe70Pd30 and Fe60Pd30Mn10 alloy-standards (error < 0.05 at.%,

composition for both was quantified by inductively coupled plasma analysis) were measured

Page 56: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

44

before, allowing to calibrate the Inca-EDX system. All this led to a compositional error

< 0.2 at.% for all samples. This procedure was applied to all samples (polycrystalline thin

films, epitaxially grown thin films and bulk samples).

Figure 22: Trajectories of inelastically scattered electrons in a 0.5 µm thin film for different elements ( a) Fe, b)

Pd, c) Mn and d) Cu). The normalized X-ray yield originated from L and K shells in dependence on penetration

depth is depicted (different colours correspond to excited shell).

Monte Carlo simulations were performed by using CASINO software127,128 (simulated was

the inelastic scattering of 1000 electron trajectories until the energy of each trajectory was

below 50 eV due to energy loss by collision using a Mott by Interpolation129,130 physical

model) to ensure a sufficient penetration depth of the electron beam at an acceleration voltage

of 20 kV into the sample yielding compositional information through the whole thickness. In

Figure 22 a) to d) the distribution of inelastic scattered electrons in a 0.5 µm thick Fe, Pd, Mn

and film (bulk density values were chosen) are shown. Presented are cross-sections of 0.5 µm

thick thin films on a SiO2 diffusion barrier (1.5 µm) on top of a Si substrate and the

trajectories of electrons (blue) in dependence on depth and lateral dimension. On the left side

in each figure the normalized X-ray yield as a function of depth over the full film thickness is

presented. The different colours correspond to the excited X-rays of the different shells of the

different elements. Due to the elemental mass of Fe, Mn and Cu a high penetration depth of

Page 57: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

45

electrons can be observed in Figure 22 a) and b). Here the generation of characteristic X-rays

for elemental analysis occurs over the full thin film thickness including parts of the SiO2/Si

substrate. Heavier elements like Pd decrease this penetration depth significantly. For a Pd film

(Figure 22 b)) X-rays are mainly generated from a depth of about 0.5 µm, while a small

fraction of electrons fully penetrates the Pd film. These results confirm a sufficient

information depth for EDX analysis. Due to the fact that Fe-Pd-based alloys are investigated

in this work, the penetration depth of electrons and the generation of characteristic X-rays is

sufficiently high to quantitatively measure the composition of Fe-Pd-X (X = Mn and Cu)

films.

3.2.2 Structural analysis by X-ray diffraction

The crystallographic structure of thin film and bulk samples was investigated by X-ray

diffraction (XRD) in Bragg-Brentano-geometry using a Bruker-AXS system (GADDS area

detector, Cu-Kα radiation, 0.01° resolution, spot size 0.5 mm, integration time 300 s), a

PANalytical X’Pert PRO MPD (Pixel detector and Szintillation counter, Cu Kα and Co Kα

radiation, 5 mm mask, programmed divergence slit on both incident and diffracted beam path,

0.04 rad soller slit, step size 0.013° for phase analysis and 0.0065° for lattice parameter

determination). Pole-figures were measured in a four-circle set-up with Cu Kα radiation using

a Phillips X’Pert system. Calibration of the systems was performed on an Al2O3 standard

(Bruker) and a Si standard (PANalytical). Temperature-dependent X-ray diffraction, XRD(T),

was used to verify and analyze martensitic transformations and transition temperatures. For

the Bruker system XRD(T) measurements were conducted using an evacuated Be dome to

prevent ice formation on the samples (chamber pressure < 1 Pa). The temperature was varied

in the range from 253 K to 373 K (10 K steps) and measured using a NiCr-Ni thermocouple

(error < 0.5 K). For the PANalytical system XRD(T) was conducted by using an Anton Paar

TTK 450 (temperature range: 120 K to 723 K at a chamber pressure < 0.15 Pa). Further

structural investigation on selected samples was performed using synchrotron-based X-ray

microdiffraction at the 2-BM beam line at the Advanced Photon Source at Argonne National

Laboratory (photon energy 15 keV, image-plate detector MAR 345, beam size 15 µm x 15

µm refined by a set of 30 Be compound refractive lenses). To observe the martensitic

transformation the samples were measured between 300 K and 450 K for 20 s per spot to

obtain sufficient diffracted intensity. Prior to the measurement the system was calibrated

Page 58: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

46

using a LaB6 standard (National Institute of Standards and Technology). The Fit2d

software131 was used to extract the lattice parameters from integrated diffraction patterns.

In order to determine lattice parameters the d-spacing was determined by fitting of diffraction

peaks using a Cauchy function and then calculated by using Bragg´s law.132 The coherence

length of X-rays in the film was investigated by means of a modified Scherrer formula.133 The

full-width-at-half-maximum (FWHM) of a diffraction peak was determined by fitting with a

Lorentz function. Micro-stress analysis was performed using the FWHM of a Gaussian

function fitted to the diffraction peak according to Ref. 134. In order to determine the residual

stress in the thin film samples the sin2(ψ) method134,135 was used. The lattice parameter was

determined by a Pearson VII peak fitting136 and then plotted as a function of sin2(ψ). From the

slope of the d = F(sin2(ψ)) function the stress state and values were determined.137,138

Phase analysis and visualization of the diffraction data for the materials libraries was

conducted by using the MATLAB-based “XRDsuite” software package.139 To determine

samples with similar structural phases cluster analysis140 and non-negative matrix

factorization141 were used to reduce the complexity of the data. The identification of the

different phases was performed by using several databases like inorganic crystal structures

database (ICSD), international centre for diffraction data (ICDD), Pauling Files binaries and

Pearson's Crystal database. The “CaRIne” software package was used in order to calculate

XRD patterns of crystal structures predicted by density functional theory (DFT) simulations.

3.2.3 Microstructural analysis by Transmissionen-Electron-Microscopy

For further crystallographic investigations the most promising samples were selected to be

investigated by means of transmission electron microscopy (TEM). A Zeiss Libra 200

(CRISP) TEM was used for high resolution imaging and electron energy loss spectroscopy

analysis (EELS). Further a FEI Tecnai F20 G2 TEM, equipped with an EDX (EDAX) system,

a high-angle annular dark field (HAADF) detector and a scanning transmission electron

microscopy (STEM) unit was used. Cross-sectional TEM-lamellae were prepared by a 5 kV

Ga-Ion focused ion beam (Zeiss XB1540 Workstation (FIB)) and afterwards polished with

500 eV Ar-ions using a low voltage Ar-gun (PHI-AES970). Further sample preparation was

accomplished by cutting TEM lamellas out of the sample using an FEI Quanta 3D focused ion

beam system. Compositional depth profiles of selected film samples were determined by a

scanning Auger microscope (PHI AES970) having an error < 1 at.% for compositional

analysis.

Page 59: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

47

3.2.4 Temperature-dependent resistance measurements

Temperature-dependent resistance measurements were carried out for the rapid investigation

of thermoelastic transformations (hysteretic behaviour of resistance as a function of

temperature). Other established methods used to investigate structural transformations are

differential scanning calorimetry (DSC)142 and temperature-dependent X-ray diffraction

(XRD(T))143.

Figure 23: a) Overview of the high-throughput test stand for resistivity and magneto-optical Kerr-effect

(MOKE) investigations. b) 4 pin probe head for resistivity screening measuring on a 4-inch Si wafer located on

the hot/cold chuck. The close-up view depicts 4 pins touching a sample of a materials library. c) 4 pin screening

probe head for magneto-resistance measurement between the poles of an electromagnet. d) 5x4 pin probe head

for continuous resistance measurement.

The DSC method requires a tedious sample preparation and a significant amount of mass that

is usually only ensured for films with a thickness in the µm range. XRD(T) can only be

performed on samples with a maximum lateral size of 9 mm x 14 mm. Thus, both methods

are moderately suited for thin film samples and do not allow for a high-throughput

investigation of full materials libraries. The change of electrical resistance in a material is

correlated to its changes in crystal lattice, crystal phases (precipitates) and boundaries (grain

boundaries and lattice imperfections).144 Further this method is very well suited to investigate

thermoelastic transformations in thin films.145,146,147,148 By comparing temperature-dependent

resistance measurement (R(T)) with XRD and DSC measurements it was found, that

transformation temperatures determined by all three methodologies agree for Fe-Pd-based

Page 60: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

48

FSMAs.149 The resistance-screening of materials libraries was performed by using a Keithley

Multi Source Meter 2000 at a constant source current of 0.1 A. Therefore the materials library

was placed into a custom built fully automated high-throughput test-stand on a heatable and

coolable wafer chuck (Figure 23).150

A four-point probe head was used to determine the electrical resistance of each sample of the

materials library at ambient temperature. To gain an average value of the sample´s resistance

each sample was measured three times and the arithmetic middle-value was taken. To

determine the appearance of a martensitic transformation in the samples, R(T) measurements

from 170 K to 470 K (heating/cooling rate: 5 K/min) were performed. Generally this

measurement was arranged in a two-step procedure.

Figure 24: R(T) curves are presented for different samples where the heating is indicated by red and the cooling

cycle by blue coloured dots. a) The “linear” R(T) graph shows a metallic behaviour without hints for a

thermoelastic transformation. b) Graph of a screening measurement with a slight S-shape indicating the presence

of a thermoelastic transformation. c) Single mode R(T) measurement shows a hysteretic shape and indicating the

presence of a thermoelastic transformation. Further the transformation temperatures as well as the width of the

thermal hysteresis are depicted.

In the first step the screening for samples undergoing a martensitic transformation was

performed by using the four-point probe, which was automatically positioned to predefined

locations across the wafer in order to record the resistance of each point at a constant

temperature. After measuring the full materials library, the temperature was varied by a step

of 5 K, and the measurement sequence across the wafer was repeated. To identify the

occurrence of a martensitic transformation, individual R(T) curves were plotted for each

sample distinguishing between heating and cooling curve. Besides the well-known metal

behaviour (Figure 24 a)), where the resistance increases linearly with increasing temperature,

also hysteretic shaped R(T) curves were observed for various samples (Figure 24 b)). Selected

samples were further investigated by single-mode measurements using a 5 x 4 pin probe head,

where the sample´s resistance was measured continuously while the temperature was varied

(heating/cooling rate: 5 K/min). The resulting R(T) curves were used to determine the

Page 61: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

49

transformation temperatures. Transformation temperatures were determined by applying the

tangent method on the data gained by single-mode measurements (Figure 24 c)). In the

following the abbreviations for martensite-start-: Ms; martensite-finish-: Mf; austenite-start-:

As and austenite-finish-: Af - temperatures were used. The width of thermal hysteresis is

abbreviated by: ∆H.

3.2.5 Magnetic properties, screening and high-resolution measurements

For automated magnetic characterization of thin film materials libraries, a custom-designed

water-cooled electromagnet is integrated into the high-throughput test-stand (Figure 23),

supplying a maximum magnetic flux density of about 0.3 T between its pole shoes. The

magnet is attached to an adjustable frame, so the gap between the materials library and the

pole shoes is kept constant at approximately one millimeter, giving enough space to move the

materials library underneath the magnet. A magneto-optical Kerr-effect (MOKE) system was

used, consisting of a 5 mW multimode laser diode (λ = 670 nm), a polarizer, an analyzer and

a photodiode to measure the intensity of the reflected light. To measure the magnetic field

during the measurement, a Hall sensor positioned between the pole shoes was used. The

measurement signal of the photodiode is read out using a lock-in amplifier to detect small

intensity changes. A special probe head was designed to fit in the gap between the pole shoes,

thus enabling magnetoresistance measurements (resistance is measured while the magnetic

field is cycled). Both MOKE and magnetoresistance measurements can be performed as

temperature-dependent screening measurements across the materials library. This set-up was

used to identify ferromagnetic samples within the materials library. Although no quantitative

magnetization results can be achieved by this method, a qualitatively investigation of the

samples was possible by distinguishing samples that are hard and soft-magnetic.

Quantitative magnetization measurements were conducted using a Physical Property

Measurement System (PPMS, Quantum Design) equipped with a temperature-dependent

vibrating sample magnetometer (VSM) unit. For this measurement, the materials library was

cut into 4.5 mm x 4.5 mm pieces and selected samples were glued to a quartz sample holder.

Saturation polarization JS was extracted from in-plane hysteresis measurement, performed in a

field range from -0.2 T to 0.2 T (2.5 mT/s-1 sweep rate) in the martensite and the austenite

state. In order to determine the saturation polarization the saturated magnetization-signal (m)

was divided by the volume (V) of the sample (JS= µ0MS = m/V) previously determined by

multiplying lateral dimensions and film thickness. For epitaxially grown thin films JS was

Page 62: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

50

determined in-plane in <100> and <110> and out-of-plane in <001> direction in order to

determine the magnetocrystalline anisotropy constants. The Curie temperature, TC, was

determined from in-plane temperature-dependent magnetization measurements (stabilizing the

temperature every 10 K and re-calibrating the position every 20 K) in an applied magnetic

field (near the saturation polarization JS) of the sample) using Kuz’min’s fit.151,152

3.2.6 Mechanical properties investigated by nanoindentation

Investigation of the mechanical properties (Young´s modulus and hardness) was performed

by using a MTS Nanoindenter XP equipped with a Berkovich indenter. The measurements

were performed at ambient and elevated temperatures (353 K) using a heat stage. Single

samples were located on a stage inside the nanoindenter and each sample was measured up to

49 indents (array of 7 x 7, ∆x = ∆y = 50 µm) per composition for statistical purposes. To

eliminate large translations of the motorized stage and to avoid associated drift effects, a

serpentine positioning of the indents across the samples was conducted. Further the

penetration depth and loading rates (strain rate: 0.05 s-1) were chosen to allow for precise

measurements and reduce substrate influence.153 The continuous stiffness method (CSM) was

applied for all measurements and from the recorded load-displacement (P-h) curves, hardness

and Young’s modulus were determined. 154,155 All P-h curves were previously corrected for

residual zero-point errors and thermal drift. The values for Young’s modulus and hardness

were determined from an indentation depth ranging from 50 nm to 70 nm. The calibration of

the indenter area function (projected contact area versus contact depth) and of the ordinary

sample-size value of the machine compliance was performed by using a fused quartz sample

as reference material.156

Page 63: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

51

4. Results and Discussion

Within this chapter the results for binary Fe70Pd30 and ternary Fe-Pd-X samples in different

states are presented, as revealed by using the experimental methods described in Chapter 3.

This starts with section 4.1 where the results from binary Fe70Pd30 FSMAs are presented and

discussed. Section 4.1.1 emphasizes on the martensitic transformation in binary Fe70Pd30

samples having a polycrystalline microstructure. This is followed by investigating the

feasibility of the splat-quenching method to stabilize metastable phases and to develop

transforming Fe-Pd bulk samples. Finally the results for epitaxially grown Fe-Pd thin films,

which were investigated in close collaboration with various partners, are briefly described.

Chapter 4.2 presents the results from the different ternary Fe-Pd-X systems. In section 4.2.1

the Fe-Pd-Mn system is investigated in terms of phase stability and magnetic excitation due to

addition of an antiferromagnetic element. The last section within this chapter presents the

results revealed for the Fe-Pd-Cu system and shows how Cu can significantly improve the

materials properties.

4.1 Binary Fe-Pd Ferromagnetic Shape Memory Alloys

4.1.1 Polycrystalline Fe-Pd thin films

Several of the presented results were measured by H. Brunken as part of his diploma thesis,

which was supervised by the author of this thesis. The deposition of binary polycrystalline

FSMA Fe-Pd films was performed in a first attempt to investigate if thin films in the as-

deposited state undergo a martensitic transformation. Therefore these films have to be in the

transforming austenitic fcc phase after deposition on the Si/SiO2 substrate. Since the

transforming fcc austenite phase is metastable and forms only at temperatures above 1033 K,

rapid cooling is required to avoid decomposition when cooling a sample from T > 1033 K to

ambient temperature (300 K). Such high cooling/quenching rates can be obtained by physical

vapour deposition (PVD) techniques. The atoms and atomic clusters ejected from the target

escape with energies up to 10 to 50 eV, equivalent to a temperature between 104 and 105 K.157

When these highly energetic particles deposit on a substrate at ambient temperature (300 K),

high quenching rates in the range of 104 K/s are achieved. Such high quenching rates in a

Page 64: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

52

PVD process allow control of the structural state in thin film samples. These structural states

range from amorphous to nano-crystalline to fully crystalline structures that can be

thermodynamically stable as well as metastable.

Figure 25: State of matter as a function of energy and temperature.157

In Figure 25 the state of matter as a function of energy and temperature is presented. The

location of fixed atoms on lattice sites (deposited thin film) and independently moving

electrons and ions (plasma state) on both scales indicate the significant difference in energy as

well as temperature. The energy of sputtered atoms is about 100 times the energy of

evaporated atoms. This additional energy provides sputtered atoms with a higher surface

mobility on the substrate and thus can facilitate phase formation. To prove that as-deposited

thin films undergo a martensitic transformation, Fe-Pd binary films were deposited from a

Fe70Pd30 alloy target and elemental Fe and Pd targets mounted in a confocal geometry. Co-

deposition as well as deposition from an alloy target allow for homogeneous mixing on the

atomic scale over several millimetres in the as-deposited thin film. Since significant

compositional gradients are absent, long range diffusion is avoided. The absence of long

range diffusion avoids changes in composition at the nanoscale after annealing of the thin film

when compared to the moment of film nucleation. Thus, only crystallization is needed to form

the metastable fcc austenite phase that undergoes a martensitic transformation. In Figure 26

Page 65: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

53

the compositional variation across a 4 inch Si/SiO2 substrate of thin films deposited from an

alloy target is shown.

Figure 26: Colour-coded compositional variation of a) Fe and b) Pd content across a 4 inch Si/SiO2 substrate in

thin films, deposited from a Fe70Pd30 alloy target.

Due to small differences in magnetic flux density below the target, provided by permanent

magnets inside the magnetron sputtering cathode, the composition varies over the substrate.158

The Fe content ranges from approximately 69 to 73 at.% and the Pd content from 27 to 31

at.%. A binary Fe-Pd composition spread, fabricated by co-deposition is presented in Figure

27. The Fe (Pd) content varies over a broader range from 33 to 77 at.% (23 to 67 at.%) when

compared to the films deposited from an alloy target.

Figure 27: Binary Fe-Pd composition spread fabricated by co-deposition from elemental targets. a) The

variation in Fe content across a 4 inch Si/SiO2 substrate is depicted by colour coding. b) The compositional

variation of Fe (black circles) and Pd (red squares) is presented as a function of the Y position (starting from the

flat of the Si wafer and following the black arrow with dotted line).

Page 66: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

54

The crystal structure forming during the deposition process was examined by XRD for both,

the alloy and co-deposited materials libraries. Both the co-deposited as well as the thin films

deposited from an alloy target, did not reveal the austenitic fcc phase.

Figure 28: XRD linescan of a co-deposited binary Fe-Pd composition spread in the as-deposited state. The

colour coding indicates the diffraction peak height. Around the Fe70Pd30 composition undergoing a martensitic

transformation, a broad (111) peak of the stable Fe50Pd50 phase appears. Further elemental Fe precipitation can

be observed correlated to the (110) α-Fe peak. With decreasing Fe content the elemental Fe precipitation

disappears and the stable fcc Fe50Pd50 phase shifts to lower angles indicating a decrease in lattice parameter a due

to an increased amount of larger Pd atoms at the expense of smaller Fe atoms.

An XRD linescan along the gradient of a co-deposited composition spread in the as-deposited

state is depicted in Figure 28. Presented is the intensity of diffraction peaks by colour coding

as a function of the 2θ angle for different compositions from Fe71Pd29 to Fe24Pd76. Around the

Fe70Pd30 composition that undergoes a martensitic transformation, no clear indication of the

austenitic fcc phase is present. A small broad diffraction peak can be observed around 2θ =

40.5° originating from the (111) fcc Fe50Pd50 lattice plane. Further, there is a small peak at 2θ

= 44.5° that is correlated to elemental α-Fe. Around the Fe50Pd50 composition the stable Fe-

Pd fcc phase is formed and the elemental Fe precipitation disappears. Further it can be seen

that the (111) fcc diffraction peak shifts to lower angles with decreasing Fe content. This

indicates an increase of the (111) lattice plane distance and thus of the lattice parameter, a,due

to an increased amount of larger Pd atoms at the expense of smaller Fe atoms. The absence of

the fcc austenite phase in the as-deposited state can be correlated to the fact that the sputtered

Page 67: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

55

atomic clusters do not have sufficient energy for formation of the metastable fcc phase. In a

further experiment co-deposition of Fe-Pd films on a substrate heated to 573 K was

performed. It was investigated if the additional amount of energy, provided by thermal heat

from the Si/SiO2 substrate, would facilitate formation of the metastable fcc phase.

Figure 29: XRD diffraction patterns of the crystal structure formed for co-deposited thin films on a Si/SiO2

substrate heated to 573 K. Colour coding indicates peak intensities of the XRD patterns in dependence on the Fe

content.

The crystal structures formed were examined by XRD linescan measurements and are

depicted in dependence on the Fe content in Figure 29. The intensity of diffraction peaks is

shown by colour coding as a function of the 2θ angle at different compositions from Fe77Pd23

to Fe33Pd67. Around the Fe70Pd30 composition, a broad peak between 2θ = 41.2° (the (111)

peak originates from the stable Fe50Pd50 fcc phase) and 2θ = 41.7° (the (111) peak is

correlated to the fct martensite phase that transforms into the austenite phase upon heating)

appears. This proves that next to the stable Fe50Pd50 fcc phase also small amounts of the

Fe50Pd30 fcc austenite phase are present in the Fe70Pd30 film. Furthermore, there is a small

peak at 2θ = 44.5° that is correlated to elemental α-Fe precipitates. Around the Fe50Pd50

composition the stable Fe-Pd fcc phase is formed and the elemental α-Fe precipitation

disappears. At Fe contents < 62 at.% both the (200) stable Fe-Pd as well as the metastable

Fe70Pd30 fcc peak appear, hinting at a multiphase structure in the thin films. Additionally the

thin film samples show hints of the existence of the Fe-Pd bct phase at Fe contents < 60 at.%

Page 68: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

56

as indicated by the (002) bct Fe-Pd peak. The XRD results for the co-deposited binary

composition spread prove that heating the Si/SiO2 substrate during deposition allows to

partially achieve the desired fcc austenite phase within the thin films. Nevertheless the

substrate temperature of 573 K is still well below the existence value for the fcc austenite

phase of 1033 K as determined from the phase diagram. Thus, partial decomposition into the

stable Fe50Pd50 phase and elemental α-Fe precipitates occur within the thin films. Based on

the results, a subsequent heat treatment at T > 1033 K after deposition is mandatory to form

the fcc austenite phase and to achieve thin film samples undergoing a martensitic

transformation. Thus, Fe-Pd materials libraries fabricated from an alloy target and by co-

deposition from elemental targets were annealed after deposition at temperatures between

1073 K < T < 1173 K and for different dwell times of 20 min < t < 60 min followed by a

subsequent quenching of the samples at cooling rates > 15 K/s. Structural analysis of the

samples showed that annealing of thin film samples at 1123 K for 30 min, revealed a fcc

austenite single-phase without any precipitation. Annealing at lower temperatures led to an

increase of the dwell time needed to change the film structure from the stable fct phase to the

metastable fcc austenite phase, and to dissolve the α-Fe precipitates.

Figure 30: XRD linescan of a binary Fe-Pd co-deposited composition spread annealed at 1123 K for 30 min.

Due to the annealing treatment a single fct martensite phase, that should transform into the fcc austenite phase

upon heating, appears at the Fe70Pd30 composition. Due to the distinct diffraction peak having a small FWHM

width, the existence of additional precipitation phases can be ruled out.

Page 69: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

57

At higher annealing temperatures, the amorphous SiO2 diffusion barrier crystallized and

diffused into the thin film. Although the dwell time was reduced to form the fcc austenite

phase at T > 1123 K, the thin film sample developed additional Fe-Si phases. Concluding all

these findings, the optimum heat treatment parameters were identified as 1123 K for 30 min.

Thus, all binary Fe-Pd materials libraries were annealed at these parameters in the following.

In Figure 30 a XRD binary line scan of a co-deposited materials library annealed at 1123 K

for 30 min is presented. Around the Fe70Pd30 composition, only one significant peak with a

small FWHM appears at 2θ = 41.7°. This peak is correlated to the (111) lattice of the

metastable fct martensite that transforms into the fcc austenite phase upon heating. The

absence of the (200) and (002) fct martensite peaks at 2θ(200) = 47.3° and 2θ(002) = 50.8° can

be explained by the low signal to noise ratio that covers these weak peaks in the background

of the diffraction spectra. With increasing Fe content, the fct martensite phase changes to the

bcc phase as indicated by the developing (110) bcc peak. When the Fe content is lowered < 69

at.%, the fct phase changes to the fcc austenite phase indicated by the appearance of the (200)

fcc peak. A decrease in transformation temperatures below ambient temperature (293 K) and

with decreasing Fe content corresponds well to the metastable phase diagram presented in

Figure 7. With further decreasing Fe content, the (111) fcc Fe70Pd30 peak broadens and shifts

towards the position of the (111) peak of the stable Fe50Pd50 phase. Between 38 at.% < Fe <

45 at.% the fcc stable phase changes to the Fe-Pd bcc phase as indicated by a splitting of the

(200) fcc into the (110) and the (002) bcc peak. Below 38 at.% of Fe the Fe-Pd fcc phase

appears again.

After verifying that the thin film samples fabricated from an alloy target and by co-deposition

have a transforming fct/fcc single-phase structure without precipitates at ambient temperature,

it was investigated if these samples undergo a martensitic transformation. Therefore

automated temperature-dependent electrical resistance screening measurements were made on

these films, as described in section 3.2.4. Figure 31 shows the electrical resistance values as a

function of temperature for thin films deposited from an alloy target a) and by co-deposition

b). All thin films were annealed at 1123 K for 30 min followed by a rapid quenching to

ambient temperature. The R(T) measurements in both diagrams show an S-shaped curve

where the transformation temperatures are determined using the tangential method. Due to the

vanishing thermal hysteresis in some measurements, one value was defined for Mf and As as

well as Ms and Af. This vanishing can be correlated to the second-order-like character of the

martensitic transformation in Fe70Pd30. It is noted that the martensitic transformation shifts

towards higher temperatures with increasing Fe content. This behaviour follows the

Page 70: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

58

metastable phase diagram in Figure 7. Samples with a Fe content outside of 70 at.% < Fe < 73

at.% do not exhibit a martensitic transformation, as indicated by a linear R(T) relationship.

Figure 31: R(T) measurements on thin films deposited from an alloy target a) and by co-deposition b) that were

annealed at 1123 K for 30 min followed by quenching. The R(T) measurements in both diagrams show an S-

shaped curve where the transformation temperatures are determined using the tangential method. Due to the

vanishing thermal hysteresis Mf and As as well as Ms and Af were set equal. The martensitic transformation shifts

towards higher temperatures with increasing Fe content. Samples with a composition outside of the interval of

70 at.% < Fe < 73 at.% do not exhibit a martensitic transformation.

Since the R(T) measurement is an indirect method to determine the presence of a martensitic

transformation, additional XRD(T) measurements were performed to prove if these films

undergo a martensitic transformation.

Figure 32: XRD(T) measurements of a Fe70.8Pd29.2 thin film deposited from an alloy target and annealed at 1123

K for 30 min, followed by quenching. At 248 K the sample is in the fct martensitic phase indicated by the

presence of the (111) fct and the weak (200) fct peak. Upon heating a shift of the (200) fct peak towards higher

angles is observed. At 292 K the (200) fcc peak starts to grow in intensity and is fully developed at 333 K. This

structural behaviour was found to be reversible and thus evidences the occurrence of a martenstic transformation.

Page 71: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

59

In Figure 32 the XRD(T) of a Fe70.8Pd29.2 thin film between 248 K < T < 353 K is presented.

At 248 K the sample is in the fct martensite phase, showing a well-developed (111) fct peak at

2θ = 41.7° and a weak (200) fct peak at 2θ = 47.7°. Due to the signal-to-noise ratio, the (002)

fct peak at 2θ = 50.8° disappears in the background of the measurement. Further peaks are

originated by a Be dome that was covering the sample during measurement to prevent ice

formation due to humidity. When the temperature is increased, the (200) fct peak shifts

towards higher 2θ angles. At 292 K the (200) fcc peak at 2θ = 48.8° starts to grow in intensity

until reaching 333 K. This vanishing of the (200) fct peak and the development of the (200)

fcc peak upon heating was found to be reversible when repeated several times. Due to the

absence of any additional peaks and the reversibility of this behaviour, the sample is in a

single-phase state and exhibits a martensitic transformation. The transformation temperatures

determined from XRD(T) correlate well with the results from R(T) measurement. These

findings prove that it is not possible to achieve polycrystalline thin films exhibiting a

martensitic transformation in the as-deposited state fabricated from an alloy target or by co-

deposition. Thus, a heat treatment at 1123 K for 30 min followed by rapid quenching is

required to achieve samples in the transforming fct/fcc phase. Since the films have to be

annealed, also multilayer films can be used to fabricate transforming thin films near the

Fe70Pd30 composition that exhibit a martensitic transformation.

Figure 33: Binary composition spread fabricated using wedge-type multilayer thin films deposited from

elemental Fe and Pd targets. A larger compositional interval was covered by this method, since shutters were

used for the deposition. The Fe (Pd) content varies from 5 at.% to 85 at.% (15 at.% to 95 at.%).

Therefore wedge-type multilayer thin films were fabricated from elemental Fe and Pd targets

as described in Chapter 3.1.1. The thin films were fabricated as binary composition spread

and the variation in composition is depicted in Figure 33. The Fe (Pd) content varies between

5 at.% to 85 at.% (15 at.% to 95 at.%) with an almost linear increase over the wafer as shown

Page 72: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

60

in Figure 33 b). Using the multilayer fabrication method a larger compositional variation

along the binary composition spread can be achieved when compared to co-deposited films.

After annealing at 1123 K for 30 min and subsequent quenching, the multilayer thin films

show a similar structural sequence (bct-fct-fcc) with increasing Fe content like the co-

deposited binary composition spread in Figure 30.

Figure 34: a) R(T) screening measurements of films fabricated by multilayer deposition indicates the occurrence

of a martensitic transformation. With increasing Fe content a shift of the martensitic transformation towards

higher temperatures is observed. The Fe75Pd25 thin film does not show any hints of a structural transformation. b)

XRD(T) on the Fe71Pd29 thin film proves that this sample exhibits a martensitic transformation. The

transformation temperatures determined by R(T) and XRD(T) correlate well.

Thin films having a transforming fct/fcc structure were further investigated by R(T) and

XRD(T) to determine transformation temperatures. R(T) measurements presented in Figure

34 a) indicate the S-shaped slope in electrical resistance. The transformation temperatures

Page 73: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

61

were again determined by using the tangential method. With increasing Fe content, a shift of

the martensitic transformation to higher temperatures is observed. The Fe75Pd25 sample only

shows a linear increase of electrical resistance with temperature, indicating the absence of a

structural transformation. XRD(T) was used to further prove if samples that show an S-shaped

R(T) relationship undergo a martensitic transformation. XRD(T) measurements on the

Fe71Pd29 thin film are depicted in Figure 34 b). The structural change of this film is similar to

the thin film deposited from an alloy target. This sample is in the martensite phase at 293 K as

determined from the (111), the (200) and the (002) fct peaks. With increasing temperature,

both the (200) and the (002) fct peaks shift towards the (200) fcc peak position at 2θ = 48.8°.

When the film is around 309 K, a first weak (200) fcc diffraction peak appears at 2θ = 48.8°

and grows with increasing temperature. The transformation temperatures determined by R(T)

and XRD(T) measurements correlate well.

Figure 35: a) An annular dark field photograph of a Fe70Pd30 lamella fabricated by focussed ion beam (FIB)

milling and observed by a transmission electron microscope (TEM) at a magnification of x59.000 is shown. The

dark area on the left of this pictograph depicts the SiO2 on top of the Si/SiO2 substrate while the area on the right

in light grey colours presents the Fe70Pd30 film. A patterned structure on the surface of the lamella is observed

within an area marked by a red ellipse. This pattern is correlated to a martensitic twin structure. High-resolution

TEM was used to magnify the area in the ellipse. b) At high magnification a tweed like surface structure can be

observed that is related to a twinned martensitic structure within this sample (red frame corresponds to the red

ellipse). Selected-area electron diffraction in c) over the area depicted by a white ellipse in b) was used to verify

the martensite structure and to determine the lattice parameters. (FIB milling was performed by A. Sehrbrock

and TEM investigations were carried out by S. Irsen)

Page 74: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

62

To gain deeper insight into the microstructure of Fe-Pd thin films undergoing a martensitic

transformation as determined by R(T) and XRD(T) measurements, TEM was used. Further,

TEM investigations were conducted to clarify if the transforming Fe-Pd thin films are in a

single austenite/martensite phase without precipitates and if the elemental Fe and Pd

multilayer dissolve during the annealing process. Therefore a Fe70Pd30 thin film fabricated by

multilayer wedge-type deposition was milled using the focussed ion beam (FIB) technique to

produce a TEM lamella. In Figure 35 a) an annular dark field pictograph of a Fe70Pd30 lamella

fabricated by focussed ion beam (FIB) milling and observed by TEM is presented. The dark

area on the left in this pictograph shows the SiO2 parts of the Si/SiO2 substrate in the lamella.

The area in light grey colours located on the right of this pictograph presents the Fe70Pd30

film. It has to be noted that the microstructure does not show any indications of a multilayer

microstructure nor precipitation. This proves that the thin elemental Fe and Pd layers fully

dissolved during the annealing. When observing the microstructure in a) stripes can be

observed within the area marked by a red ellipse. These features can be interpreted as a

martensitic twin structure in this thin film lamella. High-resolution TEM magnifying the area

inside the ellipse in b) confirms that the tweed-like surface structure corresponds to a nano-

twinned martensitic structure within this sample (red frame corresponds to the red ellipse).

Selected-area electron diffraction in c) over the area depicted by a white ellipse in b) shows

the diffraction pattern of a twinned fct martensite structure observed from the (111) direction.

Both the nanostructure high-resolution pictograph as well as the diffraction pattern are

reported in literature for a twinned martensitic structure.159

Table 1: Fabrication and annealing parameters of all thin film samples investigated within this chapter. The film

thickness varies between 500 nm and 700 nm for films deposited from an alloy target and 450 nm to 700 nm for

the composition spreads.

Magnetic measurements were additionally performed on several samples fabricated by the

different sputtering methods. These measurements revealed that the Curie temperature,

Page 75: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

63

saturation polarization and magnetic coercivity are similar to literature values and no obvious

trend was determined in dependence on the fabrication technique.

The fabrication and processing parameters of all thin films investigated within this chapter are

summarized in Table 1. The first and second columns describe the compositional variation of

the samples and the sputter deposition technique. The following two columns present the

deposition parameters and the annealing conditions. The final column shows the film

thickness variation within the samples.

Figure 36: Martensite start temperature Ms as a function of the Fe and in dependence on the sputtering method.

Co-deposited thin films show significant higher transformation temperatures in comparison to films fabricated

from an alloy target or to wedge-type multilayers from elemental targets. The martensitic transformation is

lowest for thin films fabricated as wedge-type multilayers.

When comparing the transformation temperatures of all single-phase Fe-Pd thin films heat

treated at 1123 K for 30 min followed by quenching, significant differences can be observed.

In Figure 36 the martensite start temperature Ms for three different compositions is presented

as a function of the Fe content and in dependence on the fabrication method. All values for Ms

are considerably higher than reported in literature for bulk samples and single crystals.57 This

is related to stress induced by the difference in thermal expansion coefficient of thin film (in

this case Fe-Pd) and substrate (in this case Si).22,61 Such thin film stress shifts the martensitic

transformation to higher temperatures according to the Clausius-Clapeyron equation as

described in Chapter 2.1. As depicted in Figure 36, also the different methods (co-deposition,

deposition from a Fe70Pd30 alloy target and wedge-type multilayer deposition) do show

differences in Ms at similar Fe contents. The martensitic transformation is shifted to lowest

temperatures for thin films that were fabricated as wedge-type multilayer from elemental

Page 76: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

64

targets. The films deposited from a Fe70Pd30 alloy target show a slightly increased Ms while

the co-deposited films have the highest transformation temperatures.

Figure 37: a) Coherence length as a function of the Fe content and for the different fabrication methods (open

black symbols). The correlation coefficient R2, a measure for the peak fitting quality, is depicted for all methods

by red filled symbols. The highest coherence lengths between 70 nm and 80 nm are observed for thin films

fabricated as wedge-type multilayer. Co-deposited and from an alloy target fabricated films show a significantly

decreased coherence lengths between 20 nm and 30 nm. In b) the micro-stress inside the thin film is depicted as

a function of Fe content and fabrication technique (open black symbols). The correlation coefficient again

indicates a high fit quality (filled red symbols). It has to be noted, that the multilayer thin films show the lowest

values of micro-stress.

This can be interpreted in a way that the stress state in thin films is significantly related to the

fabrication methods. To clarify the impact of fabrication method on the thin film

microstructure, the coherence length and the micro-stress inside the thin films was

investigated by analysing the measured XRD patterns. The coherence length as a function of

the Fe content and for the different deposition methods is shown in Figure 37 a) by open

black symbols. The different symbols correspond to the different fabrication methods. Further

Page 77: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

65

the correlation coefficient R2 of the Lorentzian peak fitting is presented in red colour and by

filled symbols. All R2 are > 0.98, indicating a high peak fitting quality and thus a small error

for measuring the coherence length. Thin films deposited as multilayer exhibit the highest

values in coherence lengths ranging from 70 nm to 80 nm. Since the coherence length can be

interpreted as a measure for the grain size, it can be concluded that the films fabricated as

wedge-type multilayers have the largest grain size compared to films deposited by the other

techniques. The micro-stress state in these films is again presented as a function of Fe content

in Figure 37 b). The correlation coefficient R2 again ensures a high peak fitting quality and

thus a small error for the micro-stress measurements. The micro-stress inside the co-deposited

thin films is highest among all other fabrication methods, while it is lowest for the

multilayered thin films. Although the micro-stress state differs from the residual stress in thin

films, it can be generally considered as a measure for the residual stress distribution in thin

films.160 All these findings indicate that the microstructure in transforming thin films is highly

affected by the fabrication method, even after annealing. To understand this, the phase

formation during the annealing process has to be examined. Thin films fabricated by co-

deposition and from an alloy target already in the as-deposited state have a homogeneous

mixing of Fe and Pd atoms on the nanometre scale over several millimetres in lateral

dimension. Although nanometre sized grains of different phases already form during

deposition, there are no significant compositional gradients at the nanoscale in the thin film

samples. Thus, long range diffusion is avoided during annealing. This is different in thin films

fabricated as wedge-type multilayers. Due to the multilayer structure of pure elemental Fe and

Pd layers, there are high compositional gradients that accelerate the diffusion process when

the thin films are annealed. Within these elemental Fe and Pd layers, nanometre-sized grains

are forming. When this thin film is annealed Fe-Fe and Pd-Pd bonds are breaking to decrease

the compositional gradient according to Fick’s law161 and new Fe-Pd bonds are forming. The

energy of mixing by breaking elemental bonds and the formation of new heterogeneous bonds

is an exothermic process. This process is known to release energy in terms of heat especially

in multilayered thin films where a self-propagating exothermic reaction is often observed.162

This exothermic process induces an additional amount of thermal energy into the grain

growth during annealing and thus leads to a significantly increased grain size in thin films

fabricated as multilayers. Thin films with larger grain sizes do have less boundary interfaces

than films with small grains. Since the presence of boundary interfaces in thin films strongly

affects and can even increase the residual stress state201, it can be concluded that the co-

deposited thin films have the highest micro-stress and thus the highest transformation

Page 78: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

66

temperatures. In contrast to the multilayered thin films where the large grain size and the

reduced amount of interface boundaries provides a reduced micro-stress state.

4.1.2 Bulk / Splat Samples

All presented splat samples in this part were fabricated and annealed by I. Kock from the

Georg-August-Universität Göttingen. Binary Fe70Pd30 bulk samples were fabricated using the

splat-quenching technique as described in section 3.1.3. This method is well-suited to

maintain metastable structures within bulk samples and to avoid segregation into their

equilibrium phases. Especially for the Fe-Pd system, this method is therefore quite promising

to fabricate transforming Fe-Pd FSMAs.

Figure 38: Pictures of an as-splatted sample: a) Front view optical photograph. b) Cross-sectional view

indicating a highly columnar microstructure even in the as-splatted state. (Figure originally published in

Ref.163)

Due to the high quenching rate, the material crystallizes in the desired fcc austenite phase,

which can transform to the martensite fct phase upon cooling. When the liquid droplet is

rapidly solidified by the Cu pistons of the splat-quencher, the material flows from the centre

to the sides giving rise to a round shaped geometry. In general, the splats have a diameter of

20 mm and a thickness of around 60 µm as depicted in Figure 38 a). The droplet starts to

solidify at the coldest location which is the centre of the Cu pistons. This crystallization

process induces a stress gradient over the splat as indicated by a slight bending which is

present after fabrication. Because of the heat sink of the Cu pistons, the microstructure shows

a columnar growth orthogonal to the sample surface as observed by a light microscope in

Figure 38. Induced by the splat-quenching process, there is a large amount of stress and

defects incorporated in the sample, which can pin the martensitic phase boundaries.

Therefore, the samples were annealed at various temperatures (Tan= 873 K, 973 K, 1073 K,

1173 K and 1273 K) for 15 min, in order to heal out defects and improve the transformational

Page 79: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

67

behaviour. In Figure 39 the microstructure for three different samples is shown. To

investigate the grain shape and size, trenches were milled into the samples and lamellas were

fabricated using FIB in order to be investigated by SEM and TEM. Figure 39 a) shows an

overview on the trench milled into an as-splatted sample. Both the overview and the close-up

indicate a highly columnar grain shape with a width of about 180 nm. When the sample is

annealed at 1073 K for 15 min the columnar shape of the grains becomes more distinct and

the widths increase to about 440 nm. The inset shows a TEM picture of a lamella where

martensitic twins can be observed as fishbone-like structures on the polished grain surface.

The twin structure indicates the presence of the martensite phase and thus a possible

transforming sample.

Figure 39: a) SEM overview of a trench milled into an as-splatted sample using a FIB (FIB was performed by

H. Brunken). b) Close-up view of the as-splatted sample showing a columnar growth with a column width of

about 180 nm. c) and d) Upon annealing at 1073 K for 15 min the columnar grains can be distinguished better

due to the increased column widths of about 440 nm. The inset shows a TEM lamella of three grains with a

martensitic twin structure on the polished surface. In e) and f) TEM pictures of a splat annealed at 1273 K are

shown. The column width is increased further to about 640 nm. (TEM investigations were performed by D.

König and the figures were originally published in Ref. 163)

When the splat is annealed at higher temperatures of 1273 K for 15 min, the grain widths

increase further (Figure 39 e) and f)) to about 640 nm. In order to investigate if the splats

Page 80: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

68

undergo a martensitic transformation and to determine the transformation temperatures, R(T)

measurements were carried out.

Figure 40: a) Normalized R(T) measurements as a function of temperature for splats annealed at different

temperatures. The black dots in the inset define the transformation temperatures determined by using the

tangential method for a splat annealed at 1273 K. b) Temperature-dependent magnetization measurements of all

samples under a constant external magnetic flux density of 0.03 T. In both diagrams the curves are offset for

clarity. (Magnetization measurements were carried out by I. Claussen and both figures were originally published

in Ref. 163)

Figure 40 a) shows R(T) measurements of Fe70Pd30 splats for both the as-splatted and the

annealed state. The as-splatted as well as the samples annealed at 873 K and 973 K show only

a slight non-linearity in the R(T) curve. The hysteretic shape of the R(T) curves is barely

visible due to the low signal to noise ratio. Nevertheless, the transformation temperatures

were determined using the tangential method as depicted in the inset by light grey lines.

Splats annealed at higher temperatures indicate a distinct non-linearity in the R(T) curve. In

contrast to R(T) measurements of conventional SMAs164, these samples show only weak

changes in the R(T) curve with vanishing hysteresis widths. Therefore it is not possible to

distinguish between Ms and Af or Mf and As temperatures. Thus, only one temperature is

determined for the start (Af and Ms) and the end (As and Mf) of the martensitic transformation.

From this graph the transformation temperatures increase with increasing Tan.

To further prove the presence of a martensitic transformation, temperature-dependent

magnetization measurements M(T) at a constant magnetic flux density of 0.03 T were carried

out. In Figure 40 b) the magnetization curves as a function of temperature for the as-splatted

as well as the annealed samples are shown. At temperatures > 291 K the magnetization does

not vary significantly with temperature because the splats are in the austenitic state. When the

temperature is lowered the magnetization starts to decrease rapidly at the beginning of the

martensitic transition. Measurements down to 200 K (not shown here) indicate a continuous

Page 81: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

69

decrease of the magnetization. A hysteretic M(T) curve is observed for samples annealed at

873 K, 1073 K, 1173 K and 1273 K. The hysteresis extends to lower temperatures and does

not close even in the cycles that were measured down to 50 K. Since there is no distinct

change in the M(T) curves at low temperatures, it was not possible to determine values for Mf

and As. At a higher temperature regime, Ms and Af were determined by using the tangential

method. Splats annealed at 1173 K and 1273 K show a distinct decrease in magnetization and

a hysteretic M(T) curve progression. The sample annealed at 973 K does not indicate a

transformation, although the one annealed at 873 K clearly transforms. The decrease of

electrical resistance and magnetization with decreasing temperature, as observed for all

samples, can be correlated to a change in band structure and especially in the density of

electronic d-states near the Fermi energy as reported by Opahle et al..66 An alternative

explanation for this decrease of magnetization and resistance is the increase of magnetic

anisotropy energy in the martensite state with decreasing temperature.32 Under a low

magnetic flux density of 0.03 T, the individual magnetic moments rotate out of the direction

of the applied magnetic field and into the direction of the magnetic easy axis when the

magnetic anisotropy energy is sufficiently high. This leads to a demagnetization of the sample

and a change in magnetic domain structure, which influences the electronic scattering at

domain walls.165

To identify the phases appearing within the splats, XRD(T) measurements were performed.

Figure 41 presents the XRD diffraction patterns at different temperatures. At 298 K, the splat

annealed at 1073 K shows the (111) and the (200) fct martensite peaks of the transforming

Fe70Pd30 phase. When the temperature is increased, the (200) peak of the fct phase shifts to

higher 2θ angles and the (200) fcc austenite peak appears and grows in intensity. The

transformation is reversible as confirmed by cycling the temperature during XRD

measurement. The martensitic transformation is observed for all splats by the (200) fct peak

shift and the development of the (200) fcc peak. Even the sample annealed at 973 K shows a

slight shifting of the (200) fct and a development of the (200) fcc peak. Samples in the as-

splatted state and annealed at 873 K and 973 K show next to the (111) fct/fcc peak one broad

peak at the (200) fct and (200) fcc peak positions at low temperatures. With increasing

temperature this broadened peak shifts more and more to the (200) fcc peak position while the

austenitic state appears. As-splatted samples and those annealed at 873 K and 973 K show

additional phases due to a decomposition of the austenite phase into Fe50Pd50 and α-Fe. The

(110) peak of the α-Fe phase (splats annealed at 873 K and 973 K) and a (111) Fe50Pd50 peak

(as-splatted and annealed at 873 K and 973 K) occurs for various samples.

Page 82: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

70

Figure 41: XRD(T) diffraction patterns of a) as-splatted and for 15 min annealed samples at b) 873 K, c) 973 K,

d) 1073 K, e) 1173 K and f) 1273 K. The spectra are offset for clarity. (Figure originally published in Ref. 163)

The samples annealed at 873 K and 973 K show a weak (200) fcc peak. This is correlated to

the decreased amount of the Fe70Pd30 phase due to a decomposition of the transforming phase

when annealed at temperatures < 1073 K. Splats annealed at 1073 K and higher show a

distinct (200) fcc peak when they are heated over 279 K. It has to be noted that the position of

the fct peaks are slightly different in comparison to literature values.143 This can be explained

by the temperature dependence of the lattice parameters for Fe70Pd30 reported by Cui.57 At

248 K these splats are in a martensitic state but show a small amount of austenite remainders.

At temperatures above 293 K, the splats annealed at temperatures > 973 K transform into the

austenitic state without any martensite remainders. Splats annealed at ≤ 973 K are in an

intermediate state, where both the fct martensitic and the fcc austenitic structure occur even at

low temperatures (248 K). From the occurrence of the peaks during heating, the

Page 83: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

71

transformation temperatures (only Ms/Af and Mf/As), were determined with an error <5 K. In

Figure 43 the transformation temperatures determined by R(T), M(T) and XRD(T) are

presented as a function of the annealing temperature Tan. As described in the text previously,

it was not possible for several measurements to distinguish between Ms and Af or As and Mf

for several measurements. Therefore, Ms/Af and As/M f are combined to one value for R(T)

and XRD(T). As depicted in Figure 43 the As/M f curve shows a steeper increase than the

Af/Ms curve. A line as a guide to the eye was drawn through the experimental data

accentuating an Arrhenius-like increase of the Af/Ms temperatures with increasing Tan.

Figure 42: Transformation temperatures of the Fe70Pd30 splats determined by different measurement methods

and annealed for 15 min at different temperatures Tan. (Figure originally published in Ref. 163)

All three measurement methods indicate distinct changes in crystal structure, resistance and

magnetization of the splats upon phase transformation. The transformation temperatures

determined by different methods match within the error margins and rise with increasing Tan.

Only the sample annealed at 973 K shows significant deviations in all measurements, since

only a weak transformation is observed. The XRD diffraction patterns for this sample show

the presence of Fe50Pd50 and α-Fe precipitates while the EDX analysis of this sample

indicates slight inhomogeneities in the Pd content. This is assumed to be caused by some

irregularities during the splatting or annealing process.

In order to gain an understanding about the increase of transformation temperatures with

increasing Tan, a closer examination of parameters influencing the transformation behaviour is

needed. From literature it is known that the transformation temperatures are mainly

influenced by the Pd content57, the defect density166 and the stress in the sample.22 These

properties depend on compositional homogeneity, quenching rate, grain size and shape, and

Page 84: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

72

the presence of precipitates. Moreover, it is necessary to distinguish between factors affecting

the temperature T0 at which austenite and martensite phase are in thermodynamic equilibrium

and other factors shifting the actual transformation temperatures via the necessary

undercooling ∆T = (T0 - Ms) that acts as a driving force. In the present experiments, care was

taken to ensure adequate homogeneity of the Pd content in the splats. However, the possibility

of slight deviations in Pd content (< 0.3 at.%) cannot be ruled out by EDX. Since the

thermodynamic equilibrium temperature T0 strongly depends on Pd content57 (≈ 20 K/at.%),

the transformation temperatures can be shifted by several degrees, independent of any effects

originated by the annealing process. Furthermore, the defect density and distribution in the

samples annealed at different temperatures is not identical. Due to the fabrication process of

splat-quenching, there are differences between inner and outer parts of the splat. A high defect

density provides many nucleation sites for the martensite phase166, leading to high

transformation temperatures. These defects can heal out during the annealing process leading

to a decrease in Ms while Tan increases. On the other hand, defects can hinder the movements

of the phase boundaries through the material. When precipitates have formed during the

annealing process for samples annealed at Tan < 1073 K, the phase boundaries can be pinned.

This increases the necessary undercooling and the material transforms at lower

temperatures.167 Since the As and Mf curve increases much more distinctly with increasing Tan

than the progress of Ms and Af and there is hardly any hysteresis between heating and cooling,

the amount of the martensite phase is assumed to increase by movement of the phase

boundaries rather than by nucleation of new martensite nuclei.

The transformation temperatures can be further affected by the stress state inside the sample.

This is well known as stress-induced martensite.22 However, as the splats are not attached to a

rigid substrate, most of the stress incorporated during the splat-quenching process can relax

via bending or expansion of the sample. But there can also be contributions to the intrinsic

stress by lattice distortions at defects in rapidly solidified polycrystalline samples. The most

obvious property that should show distinct changes in the samples annealed at different Tan is

the grain size as depicted in Figure 39. The dependence of transformation temperatures on

grain size was discussed in several works. A strong increase of the transformation

temperatures with grain size was reported by Seki et al.168 and Kang et al.169. An increase of

grain size with Tan is expected according to R = R0 + α⋅t⋅exp(-Q/kBTan) giving rise to a shift of

transformation temperatures towards higher values.170 Here, the initial grain size is R0, the

activation energy for grain boundary movement is Q and α is defined as a proportionality

factor. The data in Figure 42 supports this model, when considering a short annealing time

Page 85: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

73

like 15 min. Deviations from the Arrhenius-like behaviour can be explained by variations in

the initial grain size due to the splat-quenching and slight deviations below the resolution

limit of EDX anaylsis, in the Pd content. As mentioned above, the martensite phase is

assumed to increase by growth rather than by nucleation. In a coarse-grained structure, there

are less grain boundaries pinning the phase boundaries, which leads to an increase of

transformation temperatures, especially of Mf, as the elastic energy stored during

transformation is smaller.171 XRD measurements at 313 K in the austenite phase were further

examined to clarify if the defect density and the intrinsic stress decreases upon annealing. The

(111) fcc peaks were fitted using a Gaussian function and the FWHM was used to calculate

the intrinsic stress as a function of Tan.

Figure 43: Intrinsic stress (black squares) as a function of annealing temperature Tan determined from XRD

measurements of splats in the austenite phase. The black line through the squares serves as guide to the eye. The

red circles represent the coefficient of correlation (R2) of the fit function to describe the Gaussian shape of the

(111) diffraction peak. An increase of intrinsic stress with Tan > 973 K is observed.

Figure 43 presents the change in intrinsic stress for the different splats. To calculate the

amount of stress, a Young’s modulus of 25 ± 5 GPa was used for the austenite phase.172 This

distribution indicates a maximum in intrinsic stress for the as-splatted sample. With

increasing Tan the intrinsic stress decreases due to out healing of crystal defects and a decrease

of interfaces like grain boundaries. For the splat annealed at 973 K the intrinsic stress shows a

minimum originated by the presence of equilibrium phases like Fe50Pd50 and α-Fe

precipitates. When Tan is increased to T > 973 K the amount of intrinsic stress increases again,

but is still below the value for the defect-rich structure of the as-splatted sample. Concluding

this, it is assumed that the martensitic transformation is shifted to higher temperatures with

Page 86: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

74

increasing Tan due to the increase in grain size and changes in stress state and defect density

during the annealing process.

Summarizing all the presented findings it can be stated, that splat samples represent the bulk

counterpart to thin films within the Fe-Pd system. The splat-quenching technique avoids

decomposition of the fcc phase and reveals transforming bulk samples. These samples are

well-suited to investigate the change of materials properties when up-scaling promising thin

film compositions in newly developed Fe-Pd-X systems.

4.1.3 Epitaxial Fe-Pd thin films

The application of FSMA alloys in technical devices requires highly crystalline materials in

order to allow for the MFIS. This is only given when restraints like grain boundaries and

lattice defects are absent because they hamper twin boundary motions. Further there are

several intrinsic properties that are difficult to measure in polycrystalline thin film and bulk

samples, like magnetocrystalline anisotropy constants, magnetic coercivity HC and remanence

BR. These properties are important to investigate when discovering new Fe70Pd30-based

materials. Single crystals are ideal samples, since they allow determination of all intrinsic

materials properties without affecting the measurement results by external properties like

microstructure and scaling effects. But single-crystal samples are difficult to fabricate and

suffer from expensive fabrication and processing routes. This becomes important when such

materials, which exhibit their actuating/sensing effects only in the single-crystal state, have to

be incorporated into miniaturized devices. Thus, the fabrication of single crystalline thin films

also known as epitaxial thin film growth, is promising. Thin films, fabricated by epitaxial

growth using different methods like pulsed laser deposition, e-beam evaporation and sputter

deposition, were investigated in this chapter. All samples were fabricated by collaboration

partners from the IFW Dresden (group of S. Fähler for pulsed laser deposited films), Georg-

August-University of Göttingen (group of Prof. S. Mayr for thin films fabricated by e-beam

evaporation) and Christian-Albrechts-University Kiel (group of Prof. E. Quandt for thick

freestanding films) and investigated at the Ruhr-Universität Bochum. All data on samples

fabricated by other groups, were investigated and/or analyzed by the author of this thesis.

Substrate attached epitaxial Fe-Pd thin films

In a first attempt, epitaxial Fe-Pd thin films were fabricated using pulsed laser deposition to

investigate the effect of post-annealing on the thin film properties. The aim was to improve

Page 87: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

75

the films’ microstructure by annealing of defects and relaxation of stress, in order approach

single-crystal-like behaviour. Investigations of phase stability, texture and phase

transformation behaviour are performed and correlated with magnetic properties. Fe-Pd films

with different Pd contents were deposited at room temperature on MgO (100) single-crystal

substrates using PLD from elemental targets by J. Buschbeck from IFW Dresden. Further

information about the fabrication can be found in Ref. 173. Fe-Pd films with 19 at.% < Pd <

37 at.% were annealed in the stability region of the fcc phase under vacuum at p = 10 -5 mbar

encapsulated in quartz tubes. After annealing at 1223 K for 10 min, the samples were cooled

by convection by taking the quartz tubes out of the furnace.

Figure 44: XRD patterns of epitaxial thin films with different Pd contents before and after annealing. Upon

annealing the diffraction peaks shift and the FWHM becomes smaller and more distinct, indicating a reduction in

defects and a structural transformation. The (110) bct diffraction peak was calculated for a c/abct - ratio = 1.06 as

observed in as-deposited films. Stars (*) denote diffraction peaks of Fe-rich bcc precipitates that occur in as-

deposited films. (XRD measurements on the as-deposited films were performed by J. Buschbeck and the figure

was originally published in Ref. 211)

Page 88: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

76

The samples cooled from 1223 K to 673 K within 2 min leading to a cooling rate of ≈ 4.6 K/s

By this procedure, a decomposition of the alloy is avoided. Due to the chemically inert

behaviour of MgO, interfacial reactions or intermixing are not observed even at higher

temperatures. In Figure 44 XRD results, measured in Bragg-Brentano geometry, are presented

for as-deposited as well as annealed thin films at different compositions. In dependence on

composition, the position of the diffraction peaks changes considerably. A comparison with

peak positions expected from bulk material suggests that at 33 at. % Pd (Figure 44 a)), the fct

martensite transformed to the fcc austenite upon annealing. A fct to fcc-transformation due to

heat treatment is also observed in films with 29 at.% Pd and 37 at.% Pd content. According to

the intense (200) fcc peak, heat treatment of the film with 28 at. % Pd results mainly in a

transformation from bct to fcc phase after annealing (Figure 44 b)). A slightly increased

intensity observed between the (200) fcc diffraction peak and the (110) bct peak indicates that

minor amounts of martensitic phases are present. High stress in as-deposited fct films causes

the formation of misoriented fractions by deformation twinning.173,201 This is accompanied by

the occurrence of the (111) peak. After annealing, the (111) diffraction peak vanishes. At the

same time, the coherence length of thin films with Pd ≥ 28 at.% increases from the as-

deposited state (3 to 8 nm) upon annealing (41 to 45 nm) as shown in Table 2. This can be

explained by grain growth and healing out of defects when annealed at high temperatures, and

the absence of martensitic twins due to the formation of the fcc austenite phase. Shifting of

peak positions in the diffractograms of the Fe75Pd25 film suggest that the bcc phase

transformed into a bct martensite after annealing (Figure 44 c)). This phase transformation is

also observed in Fe81Pd19 films and is accompanied by a rather surprising change in texture

from a (200) orientation to (110) orientation. The existence of the bct martensite phase was

checked by measuring the tetragonal distortion of the unit cell in the 4-circle diffractometer.

From the position of the 2θ diffraction angle of the (110) diffraction and the tilted (101) peak,

the c/abct-ratio = 1.09 was determined. Due to splitting up into variants, the formation of a

martensitic phase can counteract an increase in coherence length by grain growth. Formation

of the bct martensite thus explains why only minor changes in coherence length are observed

after annealing of films with Pd ≤ 25 at.%. Epitaxial, Fe-rich precipitates being present in as-

deposited films are denoted as stars (*) in Figure 44 and vanish after annealing. As expected

by the Fe-Pd phase diagram, this phase is dissolved in the matrix upon annealing at

temperature above 1173 K. Structural data of all thin films in the as-deposited and the

annealed state is summarized in Table 2.

Page 89: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

77

Table 2: Summarized structural data of Fe-Pd thin films at different Pd contents in the as-deposited and

annealed state, as determined by XRD. The normal lattice vector refers to the direction of the unit cell that is

closest to the substrate normal. (Table originally published in Ref. 211)

The change in texture of the thin films was investigated by pole figure measurements (Figure

45). As an example, pole figure measurements on the Fe67Pd33 film in the as-deposited and

the annealed state are presented for samples with ≥ 29 at.% Pd content. In the as-deposited

state, fourfold-symmetric intensities of the epitaxial orientation were measured at an angle of

ψ = 55°. Additional intensities were observed in the centre of the pole figure and around

ψ = 70°. These further texture components are caused by deformation twinning. After

annealing the intensities in the pole figure sharpened significantly around ψ = 55°. This

indicates the formation of a fcc structure having a sharp texture and a high coherence length,

as previously determined by Bragg-Brentano measurements. The epitaxial orientation

relationship for this thin film in the annealed state is MgO (001)[100] || Fe-Pd fcc (001)[100].

The crystal axes of the Fe-Pd fcc and the MgO unit cell are aligned in parallel. Besides the

epitaxial orientation, no additional texture components are observed. Misoriented fractions

being present in the as-deposited state vanish. When the Pd content is reduced, as presented

for the Fe72Pd28 film, the pole figure exhibits sharp intensities with fourfold-symmetry in the

as-deposited and the annealed state. Due to the low mismatch between MgO substrate and

Fe72Pd28 film, the bct phase grows in a single variant state.

Page 90: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

78

Figure 45: Pole-figures measured on thin films at different Pd contents in the as-deposited and annealed state. In

a) the Fe67Pd33 film has a fct structure in the as-deposited state. The intensities originate from nonepitaxially

oriented texture components that vanish upon annealing leading to a full epitaxial fcc oriented structure in b). c)

Upon annealing the Fe72Pd28 film, the intensity maxima in the pole figures shift along the ψ direction, indicating

a phase transformation from the bct to the fcc phase. e) The Fe75Pd25 film exhibits a bcc structure in the as-

deposited state that transforms to a bct martensite state by annealing (f)). This bct martensite phase has a twinned

microstructure as depicted by a pronounced splitting of the intensities. The strong MgO (200) substrate intensity

located in the centre of the bcc (110) and the bct (101) pole figure is covered by a white spot for a better

understanding of the crystal orientation in the thin films. (Pole figure measurements were performed by J.

Buschbeck and the figure was originally published in Ref. 211)

No additional splitting of the intensities was observed. After annealing, the intensities remain

sharp but shift significantly from ψ =47° to ψ =55°. Together with the results from Bragg-

Brentano measurements, this evidence reveals the transformation from epitaxial bct phase to

epitaxial fcc phase. The epitaxial orientation relationship of this fcc phase to the substrate is

identical to the fcc phase in films with higher Pd content (as mentioned for the Fe67Pd33 film).

Page 91: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

79

When the Pd content is reduced further, the intensities of the Fe75Pd25 film are sharp in the as-

deposited state. Upon annealing, the diffraction peaks in the pole figure measurement broaden

and split in intensity between ψ ≈ 53° and 60°. This behaviour can be explained by the

formation of a bct martensite phase with a twinned microstructure. In agreement with Bragg-

Brentano XRD, the pole figure measurement of the (110) bct diffraction peak verifies the

alignment of (110) bct lattice planes in parallel to the substrate plane. A single intensity is

observed in the centre of this pole figure (see Figure 46).

Figure 46: Illustration of the formation of the bct martensite phase with (110)bct planes orienting in parallel to

the substrate upon annealing. a) Starting from two fcc unit cells sitting on top of each other, the Bain

transformation results in formation of a bct unit cell with an in-plane aligned c-axis. b) As expected for this

alignment of the (110)bct planes a central diffraction peak is observed in the corresponding pole figure

measurement. (Figure originally published in Ref. 211)

The formation of this special alignment of the bct martensite unit cells can be understood in

terms of a phase transformation induced by annealing of the sample. Annealing was

performed in the stability region of the disordered fcc phase. According to the phase diagram,

the bcc phase in the as-deposited films transforms to the fcc phase at high temperatures.

Considering the annealing temperature of 1223 K, it is likely that this transformation is not

martensitic but derives from diffusion processes. The epitaxial orientation relationship of this

fcc phase is expected to be similar to the fcc phase at higher Pd contents: MgO (001)[100] ||

Fe-Pd fcc (001)[001]. During cooling, the high temperature fcc austenite phase transforms

into the bct martensite phase. Due to symmetry-breaking, three different alignments of the bct

unit cell on the MgO substrate are possible:

Page 92: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

80

One with c-axis perpendicular to the film plane:

MgO (001)[100] || Fe-Pd bct (001)[110]

Two orientations with their c-axis oriented into the film plane:

MgO (001)[100] || Fe-Pd bct (110)[001]

MgO (001)[100] || Fe-Pd bct (110)[001]

Variant 1) is formed via the usual way of illustrating the Bain transformation with two

neighbouring fcc unit cells sitting beside each other on the substrate.67 In this case the

substrate plane is parallel to the a-b-plane of the forming bct unit cell. A (002) bct diffraction

peak belonging to variant 1), however, cannot be identified in XRD (Figure 44). In addition

this orientation should give intensities around ψ =45° in the (101) pole figure measurement

(Figure 45 f)). However, only a weak intensity is observed. Thus, there is no indication that

significant amounts of variant 1) are present in the film. Formation of variants 2) and 3) is

schematically illustrated in Figure 46. In this case, two fcc unit cells sitting on top of each

other have to be considered. According to the Bain transformation, a bct variant forms having

its (110) plane aligned in parallel to the film plane (Figure 46 a)). Formation of these variants

also explains the high (110) intensities in Figure 44. Since the (110) plane of these bct

martensite variants lies in the film plane, their c-axis does as well. Due to fourfold-symmetry

of the substrate, a second variant must form with its c-axis rotated by 90° in the film plane.

Thus, the Bain transformation explains the unusual texture change from (200) orientation in

the as-deposited state to (110) orientation in the bct martensite upon annealing. However,

since the Bain transformation is only a simplified model that does not consider a habit plane,

it does not explain the tilt and rotation of the crystal axes involved in the martensitic

transformation as depicted in Figure 45. The observed orientation of the bct martensite

forming after annealing is different to the growth texture of the bct phase in the as-deposited

state of Fe72Pd28 samples, where the cbct-axis was aligned perpendicular to the film plane in a

single variant state. Pole figures of defect rich, deformation twinned, fct films in the as-

deposited state (≥ 29 at.% Pd), exhibit considerable improvement in texture after annealing,

due to recrystallization (Figure 45 b)). The driving force for recrystallization is the reduction

of defects (grain boundaries and other interfaces) and stress (elastic energy) within the

microstructure. During the recrystallization process, defect-poor regions in the microstructure

grow at the expense of defect-rich regions. In the as-deposited films, only regions close to the

substrate interface (≈ 20 nm thickness) have an epitaxial orientation to the substrate.201 During

annealing, misoriented texture components vanish while epitaxy is maintained.

Page 93: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

81

Figure 47: Schematic illustration of texture improvement by recrystallization. The as-deposited fct-film exhibits

a heterogeneous microstructure consisting of two layers: a) epitaxial orientation at the interface to the substrate

and b) a misoriented, deformation-twinned layer on top. During annealing, the defect-rich, misoriented layer is

overgrown by the epitaxial layer. In result, a fully epitaxial, single-crystal-like film is obtained. (Figure originally

published in Ref. 211)

This suggests that the epitaxial region at the substrate interface grows at the expense of the

defect-rich, deformation-twinned part of the film. This process is schematically illustrated in

Figure 47, showing the vanishing of misoriented areas in the thin film’s structure upon

annealing. According to the metastable phase diagram, films that exhibit a fcc structure at

room temperature should transform into martensitic phases at lower temperatures. XRD(T)

was performed on an annealed Fe72Pd28 thin film sample within an interval from 275 K to

175 K in steps of 25 K during cooling (Figure 48). Besides peaks originated by the substrate

and sample holder, only the (200) fcc diffraction peak is observed until 200 K. When the

sample is cooled below 200 K, the intensity of the (200) fcc peak decreases and a new

diffraction intensity appears at lower angles. This peak belongs to the (200) fct plane of the

martensite phase. This indicates the presence of a fcc-fct transformation upon cooling of the

sample. When the temperature is decreased further, the (200) fct peak shifts towards smaller

angles, indicating a decrease of the unit cell’s c/a-ratio. The length of the a-axes are

determined from the centre of the (200) fct peak. It is assumed that the volume of the unit cell

is constant during martensitic transformation (Vfct = Vfcc = 0.05336 nm3). Accordingly, the c/a

ratio of the fct martensite changes from c/a = 0.95 at 200 K to c/a = 0.94 at 125 K. Besides the

shift, a broadening of the (200) fct peak is observed with decreasing temperature. The

tetragonal distortion of the fct martensite strongly depends on the undercooling from the

martensite start temperature.64 It is known, that martensitic transformations can be constrained

by the interface to the substrate.174

Page 94: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

82

Figure 48: Phase transformation from fcc to fct determined for an annealed Fe72Pd28 film by XRD(T)

measurements. Diffraction peaks from the sample holder are indicated by *. (Figure originally published in Ref.

211)

Considering this, there are two possible explanations for the peak broadening of the (200) fct

peak: 1) An inhomogeneous tetragonal distortion forms in the film during cooling. The peak

width suggests that c/a ratios ranging from approximately 0.95 to 0.93 coexist in the film at

125 K. A possible reason for this could be the hindering of the martensitic transformation at

the substrate interface. This corresponds to the fact that the fcc austenite diffraction peak does

not fully vanish until 125 K, revealing that a significant fraction of residual austenite remains

in the film. 2) A temperature-dependent reduction of the variant width due to decreasing c/a

ratio. Splitting up into smaller variants allows for compensation of elastic energy that is

caused by the tetragonal distortion. Similar reduction in the variant size has been observed

before, as a consequence of a decreasing c/a ratio in fct single crystals.64 Further XRD(T)

measurements on a Fe67Pd33 thin film sample did not reveal a phase transformation. This is

known in literature, and is correlated to the fact that the increased Pd content in this sample

shifts the martensitic transformation to lower temperatures and thus stabilizes the fcc phase

down to 125 K (see Figure 7).

Magnetic properties were investigated by hysteresis measurements at room temperature by

using a PPMS. In Figure 49 the hysteresis curves measured on Fe75Pd25, Fe72Pd28 and

Fe67Pd33 thin films are presented as examples.

Page 95: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

83

Figure 49: Structure-dependent shape of magnetic hysteresis curves measured from the annealed films at room

temperature. (Figure originally published in Ref. 211)

The shape of the hysteresis curves can be interpreted as a magnetic fingerprint of the phases

present in the Fe-Pd films. While the fcc films exhibit a low HC and low saturation field, the

bct film exhibits a significantly higher coercivity and is harder to saturate. Magnetically softer

behaviour of the fcc phase is expected by its high crystal symmetry. In contrast, the shape of

the hysteresis curve measured from the bct phase was rather unexpected. As-deposited bct

martensite films can be saturated easily and exhibit a lower coercivity when compared to

films with a fct structure at Pd contents ≥ 29 at.%. Although the phases are similar (both

structures differ slightly in tetragonal distortion c/a ratio), this can be explained by the

significant difference in texture. While the bct phase in as-deposited films grows in a single

variant, untwinned state with c-axis out-of-plane, the bct phase forming after annealing has its

c-axis oriented in the plane and is in a twinned state. Thus, there are two possible

contributions that can explain the different shape of the hysteresis curves: 1) Pinning of

magnetic domain walls at twin boundaries and 2) the alignment of both hard and easy

magnetization axes along the measurement direction in annealed films. The different shape of

the hysteresis curves enable tracking of martensitic transformations. In contrast to XRD,

magnetic measurements can be performed down to lower temperatures. Hysteresis

measurements at 400, 200, and 50 K are depicted in Figure 50. Both the Fe72Pd28 and Fe71Pd29

compositions exhibit indications of a magnetic transition from the magnetization

measurements. From temperature-dependent hysteresis measurements, both, coercivity HC

and saturation field HS were extracted and are summarized in Figure 51. HS was taken from

the applied field where J reaches 90% of the polarization at 1 T.

Page 96: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

84

Figure 50: Hysteresis curves of annealed thin films at different temperatures. The shape of the hysteresis loop is

independent of temperature for Fe67Pd33 (fcc at RT) and Fe75Pd25 (bct at RT). For Fe72Pd28 and Fe71Pd29 the

observed temperature-dependent change in shape of the hysteresis loop indicates a fcc to bct transformation.

(Figure originally published in Ref. 211)

Coercivities and saturation fields of the Fe75Pd25 and Fe67Pd33 films having a bct and fcc

structure at room temperature, exhibit an almost linear, temperature-dependent behaviour.

Thus, there is no evidence for a phase transformation in these samples. The slight, continuous

increase in HC and HS observed with decreasing temperature for the Fe75Pd25 sample is

expected for a normal anisotropic ferromagnetic material. In contrast to this, both films with

intermediate Pd content (Fe72Pd28, Fe71Pd29) exhibit a strong increase in coercivity and

saturation field starting below 250 K. Both films are in the fcc phase at room temperature.

From XRD it is known that the Fe72Pd28 film transforms to the fct martensite phase during

cooling. As expected by a tetragonal distortion, the hysteresis measurements reveal a

transformation from fcc to phases with increased magnetocrystalline anisotropy. In agreement

with the metastable phase diagram, the transformation is shifted to lower temperatures at

higher Pd content. The magnetic behaviour indicates that the martensitic transformation

proceeds further below the temperature of 125 K than was accessible in XRD. At 50 K, the

magnetic properties of both films are comparable to those of the more Fe-rich sample

exhibiting the bct phase. Formation of bct martensite is expected from the metastable phase

diagram and supported by XRD(T). The gradual decrease in tetragonal distortion that was

observed could finally result in the formation of the bct martensite at lower temperatures.64

Page 97: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

85

Figure 51: Saturation field HS and coercivity field HC extracted from the hysteresis loop measurements in Figure

50. Films with highest and lowest Pd content only show minor changes in the magnetic properties. In Fe72Pd28

and Fe71Pd29 thin films indications of the phase transformation from fcc to bct are observed. (Figure originally

published in Ref. 211)

From literature it is known that pulsed laser deposited Fe-Pd films can exhibit a

heterogeneous microstructure in the as-deposited state. This microstructure is usually

originated by stress-induced deformation twinning.201 Recrystallization during postannealing

enables significant reduction in defects. As illustrated in Figure 52, annealing also has

considerable impact on the phase stability. Stress-induced fct martensite is present in as-

deposited films with ≥ 29 at. % Pd173. After annealing, the stress within the thin films is

released and the samples transform into the fcc austenite state. At low Pd content the bct

martensite phase is observed. Compared to bulk material, the bct phase is stabilized at lower

Pd contents. Interface and elastic energies stored in as-deposited fct films provide a large

driving force for grain growth. Since stress and defect density at the substrate interface are

highest201, the recrystallization starts from there. During annealing, the epitaxial fraction of

the as-deposited film overgrows the entire misoriented, deformation twinned part of the film

and deformation twins vanish. Single-crystal-like misorientation-free texture is obtained in

the films. Formation of the fcc phase shows that by recrystallization significant stress

reduction is achieved. Defect and stress reduction in annealed films result in a significantly

increased coherence length. Due to epitaxial (100) growth, both, the fcc and the fct phase can

be clearly identified by sharp, well separated (200) fcc and (200) fct peaks. The peak shift

enables easy following of the transformation path from fcc-fct. Comparable to experiments on

Page 98: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

86

Ni-Mn-Ga174, the presented results indicate that the rigid substrate interface stabilizes the

austenite in Fe-Pd.

Figure 52: Comparative illustration of the phase stability ranges of Fe-Pd at room temperature in bulk material,

as-deposited films and annealed films. In bulk material the existence-range of fct is very narrow. Due to stress in

as-deposited films the epitaxial fct-phase is stabilized up to high Pd content. After annealing this stress is relaxed

and the fcc phase is observed. At low Pd contents either bcc or bct phases are observed. White areas have not

been investigated. (*) Indicates a compositional region where XRD measurements indicate a fcc phase with

minor amounts of martensite. (Figure originally published in Ref. 211)

It is observed that both bct martensite and fct martensite in the annealed films preferably form

variants with their c-axis aligned in the plane, corresponding to a two-variant state. In a

martensitic transformation, variants are formed to compensate the lattice distortion. In an

unconstrained state, equal fractions of all three possible variants would be expected. For the

present case, the different thermal expansion coefficients of film and substrate may result in a

stressed state. Since the thermal expansion coefficient of common metals exceeds those of

oxides, cooling from the annealing temperature results in tensile stress (this argument neglects

the invar anomaly around room temperature). Tensile stress may result in a variant selection,

but they should rather favour alignment of the unit cells with their a-axes in the plane and c-

axis out-of-plane. This, however, contradicts to the present observations. Since thin films are

investigated, it can be also suspected that the observed preference of particular variants is an

effect of the substrate, constraining the martensitic transformation. Indeed, recent model

calculations indicate that during a cubic to tetragonal transformation, the substrate constraint

may result in an alternating alignment of variants having parallel c-and a-axes in the film

plane while an a-axis is aligned perpendicular to the film plane.175 To enable the formation of

an invariant plane, this transformation requires the presence of two habit planes. Between the

habit planes and the interface to the substrate some residual austenite remains. This was

calculated for the transformation from cubic austenite to tetragonal martensite in Ni-Mn-Ga

and a similar scenario can be expected in Fe-Pd. The temperature dependence of the magnetic

coercivity and saturation field suggests that magnetocrystalline anisotropy increases during

Page 99: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

87

martensitic transformation, from fcc to fct and bct structure. In general, an increase in

anisotropy can be expected due to the symmetry breaking of the formerly cubic unit cell.

However, strained epitaxial film growth showed that the magnetocrystalline anisotropy is

reduced if the tetragonal distortions of the unit cell exceeds those of the fct phase.67 To

explain this discrepancy, the microstructure of the bct martensite has to be considered. The

bct martensite exhibits a twinned microstructure. Hence a large number of pinning sites are

present, which commonly results in an increased coercivity. When considering that stronger

tetragonal distortion favours splitting up of the martensite into smaller variants, pinning can

explain the increasing coercivity with decreasing c/a-ratio. Moreover, decreasing the c/a-ratio

increases the twinning angle and the tilt of the crystal axes of the martensite unit cell with

respect to the former austenite unit cell.174 Due to the splitting of the martensite into variants,

a broad orientation distribution of the easy magnetization axes forms in the martensite,

explaining the roundish shape of the magnetic hysteresis measured for the bct martensite.

Concluding the presented results it is shown that annealing is suitable to achieve significant

reduction in defects and homogenization of microstructure in pulsed laser deposited epitaxial

Fe-Pd films. Single-crystal-like fcc films are obtained. Due to recrystallization, deformation

twins that were present in the as-deposited state completely vanish. Because of the epitaxial

(100) orientation, transformation from fcc austenite to fct martensite was investigated in detail

in XRD experiments. Residual austenite is observed even at temperatures of 50 K, below the

start of the martensitic transformation along with a complex formation of martensitic variants.

Both observations show that the substrate constrains the transformation from fcc to fct phase.

Different to bulk material, magnetic measurements indicate a continuous transformation from

fct to bct martensite at low temperatures. According to our results annealing of pulsed laser

deposited thin films is a promising route to obtain single-crystal-like epitaxial Fe-Pd films.

Freestanding Fe70Pd30 epitaxial thin films

When a film is grown epitaxially, the substrate underneath is constraining twin boundary

movement in the thin film lattice. Therefore it is important to release epitaxially-grown films

from the underlying substrate to allow twin boundary movement and therefore the MFIS.

Epitaxial growth of Fe70Pd30 thin films was performed by T. Edler using electron-beam

evaporation on MgO (100) substrates heated to T = 963 K. The structural relationship for the

thin film on the substrate is defined by MgO(001)[100] || Fe70Pd30fcc(001)[100] as depicted in

Figure 53.

Page 100: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

88

Figure 53: Epitaxial relationship of the Fe70Pd30 thin film deposited on a (100) oriented MgO substrate.

Key parameters that have to be adjusted in this process are composition and the quality of the

crystal lattice in terms of defects and stress state within the thin film after deposition. Further

information about fabrication and processing can be found in Ref. 159. In order to release

these thin films from the substrate, two routes to etch the substrate were developed and

performed by Edler et al..176 The approach used to release the Fe70Pd30 thin films from the

substrate uses a saturated sodium bicarbonate solution that dissolves the MgO substrate at a

low rate of ≈ 80 nm/h. This procedure allows releasing epitaxially-grown Fe70Pd30 thin films

perfectly from the underlying MgO substrate without affecting composition, structure or

surface. XRD performed after fabrication as well as after releasing the film from the MgO

substrate revealed single crystalline thin films without a high crystal lattice defect density.

Figure 54 a) shows XRD spectra measured in Bragg-Brentano geometry for the substrate-

attached Fe70Pd30 thin film. Next to the substrate peaks of the (200) and (400) MgO planes,

there are additional peaks from Cu Kβ and W Lα radiation (blue text) diffracted on the MgO

planes, since the Cu X-ray radiation source is not monochromatic. A further peak comes from

the stage of the X-ray diffraction system. The Fe70Pd30 thin film shows only the presence of

the (200) fcc peak indicating a highly textured growth of the film on the underlying MgO

substrate. After releasing the thin film, the MgO peaks disappear, leaving only the (200) fcc

Fe70Pd30 peak. In Figure 54 b) a pole figure measurement of the (111) fcc planes of the

Fe70Pd30 thin film after MgO etching is presented to further prove if epitaxial growth is

present.

Page 101: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

89

Figure 54: a) Comparison of θ/2θ XRD measurements before (red line) and after (green line) MgO etching

confirming the structural integrity of thin films after etching. The blue annotations indicate additional diffraction

peaks from Cu Kβ and W Lα radiation diffracted on the MgO planes and from the XRD stage. b) Pole figure of

the (111) fcc Fe70Pd30 planes proving the high degree of crystallinity after film release. (Pole figure measurement

was performed by T. Edler and originally published in Ref. 177)

Due to the four-fold symmetry and the significantly small FWHM of the (111) peaks, a high

degree of crystallinity and no structural changes after MgO etching is confirmed. In order to

investigate if the samples undergo a martensitic transformation, XRD(T) measurements were

performed on substrate-attached as well as on freestanding films in the as-deposited state.

These measurements show a variation of the fcc lattice parameter with temperature but do not

indicate a martensitic transformation. Since the Invar effect was observed to occur in the Fe-

Pd system, Figure 55 presents the lattice parameter of the fcc unit cell as a function of

temperature for both the substrate-attached and the freestanding Fe70Pd30 thin films in the as-

deposited state. For T > 300 K the latter remains constant, while the former increases with

temperature, although at a slightly reduced slope. This is considered to be a substrate effect,

originated by thermal expansion of the MgO substrate while the Fe70Pd30 thin film is attached

on top. Since calculations predict a very soft material and therefore a low resistance to

straining67, the thin film is expanded by the thermal expansion of the underlying MgO. This is

further corroborated by calculating the thermal expansion coefficient from the change in

lattice constant. The thermal expansion coefficient determined from the slope of the curve for

T > 300 K is αMgO+Fe70Pd30 = 12⋅10-6 K-1 which is in good agreement with the literature value

for MgO (αMgO = 14⋅10-6 K-1).

Page 102: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

90

Figure 55: Lattice parameters as a function of temperature for the substrate-attached and the freestanding

samples in the as-deposited state. Dashed lines serve as guide for the eye. The Invar effect is observed as

stagnation of thermal expansion of the lattice parameter for T > 300 K in the freestanding sample. (Figure

originally published in Ref. 177)

While the substrate constraints are considered to hinder any structural changes of the lattice,

defect pinning was identified to constrain any martensitic transformation in as-released thin

films. Therefore, an annealing treatment with subsequent quenching was performed to heal

out crystallographic defects and thus to reduce pinning sides. The freestanding thin film

sample was sealed in a quartz tube under an Ar atmosphere of 600 mbar and annealed for 30

min at 1173 K. Subsequent quenching in water kept the sample in the metastable fcc phase

and avoided decomposition. Using XRD(T), a fractional thermoelastic fcc to bcc

transformation of the freestanding sample upon cooling was identified. Figure 56 presents

XRD(T) patterns measured at different temperatures. Above 278 K a single (200) fcc phase

occurs, consistent with the pole figure measurement. Additional peaks at 44.5° and 50.6°

originate from the sample holder and at 47.8° from a supporting Si plate. The step-like feature

at 47° is caused by a Ni filter that was used to weaken the Cu Kα2 and Kβ radiation. Below

278 K the (110) bcc peak starts to grow and increases in intensity upon cooling down to 128

K. At 228 K, small amounts of the fct phase occur, as indicated by the presence of the (200)

fct peak in the diffraction pattern. With decreasing temperatures down to 128 K, this peak

also increases in intensity. By heating the sample again up to 400 K, both the (110) bcc at 278

K and the (200) fct peak at 253 K decrease reversibly and disappear. At low temperatures

only small fractions transform from the fcc to the fct phase as inferred from the (200) fct peak

intensity. Thus a fractional martensitic transformation is considered to occur.

Page 103: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

91

Figure 56: XRD(T) results of an annealed freestanding Fe70Pd30 sample. A reversible fractional fcc to bcc

transformation upon cooling was observed by the appearance of the (110) bcc peak. By further cooling down to

253 K, the (200) fct peak occurs, indicating a fractional martensitic fcc to fct transformation.

Since the (200) peak of the fcc phase occurs at all temperatures, it is reasonable to assume that

parts of the freestanding film transform simultaneously and reversibly from bcc to fcc and

from fct to fcc. The slight shift observed for all peaks with changing temperature is attributed

to the thermal expansion of the sample. It would have been interesting to measure the

structural changes at even lower temperatures, but this was not possible, since the Anton Parr

TTK 450 was not able to lower the temperature < 120 K. Instead, magnetic measurements

were performed to investigate the magnetic properties down to 25 K in order gain hints for

further structural changes below 120 K. These measurements were preformed by T. Edler

using a Quantum Design superconducting quantum interference device (SQUID)

magnetometer. The thin films were investigated upon cycling of the temperature several times

between 25 K and 400 K under a constant magnetic flux density of 0.01 T as presented in

Figure 57. The results from SQUID measurements confirm the findings of the XRD(T). The

temperature-dependent magnetization M(T) follows the Curie-Weiss behaviour for the as-

deposited freestanding thin film, while the annealed sample shows a more complex scenario.

First the magnetization increases during cooling (from 1 to 2), but changes its slope around

258 K as determeind by the tangential method. This coincides well with a fractional structural

change from fcc to fct around 253 K as determined by XRD(T). During the martensitic

transformation, the magnetic anisotropy increases significantly as reported by Kakeshita and

Fukuda.178 This anisotropy counteracts the external magnetic field, and thus leads to a

Page 104: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

92

deviation from a Curie-Weiss behaviour towards lower values. When the sample is heated

again (from 2 to 3), the magnetization follows the cooling curve and again changes its slope

around 258 K, where the martensitic transformation occurs.

Figure 57: M(T) of as-deposited (blue squares) and annealed (red circles) freestanding samples measured in a

constant external magnetic flux density of 0.01 T. (Measurements were performed by T. Edler and the figure was

originally published in Ref. 177)

When the sample is heated above room temperature for the first time, quenched-in defects

(such as vacancies) can heal out. This reduces the pinning sites in the sample, and thus leads

to an increase in magnetization when compared to the initial cooling curve. As the

temperature is decreased again from 400 K to low temperatures (from 3 to 4), the slope

follows the curve of the as-deposited film. As observed for the initial cycle, the magnetization

shows a defined kink at 253 K as determined by the tangential method. When the temperature

is increased again, the curve fully follows the cooling curve. The Curie temperatures for the

freestanding film in the as-deposited and annealed states were examined according to

Kuz’min’s model.151 The parameters β = 1/3 and p = 5/2 were kept constant while the shape

parameters s, the Curie temperatures and the magnetizations at 0 K were fitted. The Curie

temperatures found for both the as-deposited (TC = 643 K) and annealed (TC = 625 K) films

agree well with bulk samples reported in the literature.52 The two fractional, thermoelastic

transformations from fcc to fct and fcc to bcc structure, as observed by XRD(T) in Figure 56,

can be understood using the Bain-path formalism as mentioned in Chapter 2.6 Figure 9. Using

the (200) fcc peak, a lattice parameter of c = 0.3788 nm for the fcc unit cell is observed at 400

K. The (110)-planes of the bcc cell correspond to the (200)-planes of the fcc cells and are thus

Page 105: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

93

parallel to the substrate surface. At 128 K a lattice parameter of a = 0.2950 nm for the bcc cell

and of c = 0.3844 nm for the fct cell is observed. The fcc to bcc transformation in bulk Fe-Pd

is generally considered to be non-thermoelastic and irreversible.179 However, the

transformation behaviour of a Fe-Pd sample not only depends on the chemical composition,

but is also affected by its microstructure and other constraints. For example, the

transformation temperature of polycrystalline bulk Fe69.9Pd30.1 samples with a grain size of

about 2.6 µm is roughly 40 K lower than a sample having a grain size of 6.1 µm. A single-

crystal-like value of 292 K was found for a grain size of 266 µm.168 The microstructure after a

martensitic transformation in bulk material is determined by boundary conditions from the

matrix in which the martensite forms. To dissipate the total elastic energy, a twinning of the

martensite structure occurs. This elastic and twin-boundary energy counteracts the tendency

to form the martensitic phase, and leads to a hysteresis in the martensitic transformation.171 In

this sense, open surfaces can be considered as an ultimate way of relieving all constraints - in

particular if they are accompanied by a strongly reduced dimension normal to the surface. A

freestanding film can easily accommodate stress load by bending, expanding or stress

relaxation at the surfaces. Once a new structure has formed over the whole film thickness, all

transformation stresses (apart from the interface to the untransformed portion of the film) are

considered to be relaxed. Thus, the martensitic transformation does not stop in the fct phase,

but continues along the Bain path to the final bcc structure. Nevertheless, the boundary of this

transformation will be accompanied by large lattice strains, as it moves through the film. This

boundary can be pinned at defects, such as remaining dislocations, which originate from film

growth180 and have not completely healed out during annealing; this explains the remaining

fcc structure. A stress-induced transformation into the fct structure can be envisaged at this

interface.22 The reverse bcc transformation is related to the fact that the fcc phase retains a

significant volume fraction in the film. When the temperature is increased, these areas can act

as nucleation sites for the fcc phase and may greatly facilitate this transformation. Such

behaviour is not observed in bulk samples, as this would cause a large straining of the

surrounding matrix of the transforming area. However, in the freestanding film, this can,

again, be easily accommodated at the film surfaces. In conclusion, a fractional transformation

from the fcc structure at 400 K to a bcc structure at 278 K and a fct structure at 253 K,

respectively, was observed in freestanding epitaxially grown films. Even at temperatures as

low as 128 K the fcc, bcc and fct structures coexist. These reversible transformations can be

interpreted with the Bain-path formalism: due to the flat energy landscape between the

different crystal structures in materials showing a Bain path behaviour181, an adequate driving

Page 106: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

94

force can induce such a fcc to bcc transformation, while free surfaces can provide a relaxation

source for the strains of the matrix.

Substrate attached Fe70Pd30 thick films

Ultra-thin films in the nanometre range are very challenging to be implemented into

microsystems, since their small thickness makes handling complicated. Further the MFIS can

be destabilized by this thickness due to a decrease in blocking stress making significantly

thicker films more desirable. Fabrication of active Fe-Pd films in the micrometre range would

provide a route towards application for this class of materials. Although epitaxial growth of

thin films is usually restricted to nanometre thicknesses, this approach was extended to obtain

freestanding Fe70Pd30 films of micrometer thickness, thus fulfilling both key requirements for

the integration into microsystems as well as prerequisites for FSMA films, that is, martensitic,

ferromagnetic at RT, freestanding, and single-crystalline-like.

Figure 58: Schematic of the orientation relationship of MgO substrate, Cr adhesion layer, Au buffer and

Fe70Pd30 film.

These films were fabricated by C. Bechtold from Christian-Albrechts-Universität Kiel and

investigated in close collaboration with the Ruhr-Universität Bochum. Fe70Pd30 films with a

thickness of 1.2 µm were fabricated by sputter deposition at very low deposition rates in the

range of 0.024 nm/s on a Au-buffer layer (d = 50 nm), deposited on a Cr adhesion layer (d = 5

nm) on (001) oriented MgO, polished single-crystal substrates. For further information on the

fabrication process refer to Ref. 182. Due to the coherent epitaxial growth on a Au buffer

layer deposited onto a MgO substrate, the Fe70Pd30 films are deposited in a bct martensitic

Page 107: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

95

structure. The structural relationship between the substrate, the Au buffer layer and the

Fe70Pd30 film is:

MgO(001)[100] || Au(001)[100] and Au(001)[100] || Fe70Pd30fcc(001)[110]

as depicted in Figure 58. XRD measurements in Bragg-Brentano geometry reveal the (200)

Au diffraction peak of the buffer layer as well as the Fe70Pd30 (002) reflection. Assuming a

constant volume of the Fe70Pd30 unit cell compared to the cubic austenite, the lattice

parameters were calculated to be a = 0.287 nm and c = 0.321 nm (±0.001 nm), which

constitutes a c/a = 1.12. XRD(T) indicate no significant change in the crystal structure in a

temperature range between 150 and 375 K as presented in Figure 59 a). A peak shift of the

(200) Au and the (200) Fe70Pd30 diffraction peak dependent on temperature is observed. This

is correlated to the thermal expansion with temperature in both layers. Pole figure

measurements of the (101) diffraction peak at 2θ = 42.15° reveal a four-fold symmetry (see

Figure 59 b)). The maximum intensity is obtained at an average of 47.278° at ϕ = 45°. The

peak in ϕ direction is sharp with a small FWHM, indicating a well-oriented growth of the bct

unit cell rotated by 45° compared to the edges of the MgO cell. The increased FWHM in the

ψ direction indicates a relaxation of the lattice. Furthermore, no hints of relaxation such as

twinning in the film structure is found. All these results indicate the high quality of the single-

crystal structure in the Fe70Pd30 film even at such a high film thickness of 1.2 µm.

Figure 59: a) XRD(T) measured in Bragg-Brentano geometry reveals the (200) growth of the Au buffer layer

and the (002) growth of Fe70Pd30. A martensitic transformation of the bct Fe70Pd30 phase is not observed. b) The

four-fold symmetry of the (101) pole figure indicates epitaxial growth without misorientations. (Pole figure

measurement was performed by C. Bechtold and the figure was originally published in Ref. 182)

Page 108: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

96

To gain a deeper insight into the microstructure of the film, TEM investigations were carried

out by B. Erkatal from Christian-Albrechts-University Kiel. A lamella sample was prepared

by FIB milling. Cross-sectional TEM and electron diffraction investigations revealed an

entirely epitaxial heterostructure. In Figure 60 a) a 2 µm wide cross-section bright-field TEM

image of the Fe70Pd30 film is presented. No signs of columnar growth or small-angle grain

boundaries are observed in the film’s cross-section, proving the rather perfect single-

crystalline structure. The bending contours observed in TEM are caused by the FIB milling

due to differences in lamella thickness. Compositional investigations performed by using

EDX in scanning mode (STEM) verified the Fe70Pd30 composition. Selected area electron

diffraction (SAED) patterns of the Fe70Pd30 film and MgO are shown in Figure 60 b) and c).

Based on the slight bending of the sample, only marginal readjustments of the zone axis

orientation were needed to observe the same SAED pattern over the complete area of the

cross section. A SAED pattern taken along the [100] zone axis of the film confirmed the

presence of the bct structure in the Fe70Pd30 film. The high-resolution TEM (HRTEM) image

of the rough interface between the Fe70Pd30 film and the Au buffer layer is depicted in Figure

60 d) and indicates an orientation relationship as described in Figure 58.

Figure 60: a) Cross-sectional bright field TEM image of Fe70Pd30 (thickness ≈ 1.2 µm) on a Au buffer layer

(thickness 50 nm) and MgO(100). b) Electron diffraction pattern of Fe70Pd30 (location: white circle in (a)). c)

MgO diffraction pattern revealing identical lattice distances of MgO(020) and Fe70Pd30 (110). d) HRTEM image

of the interface between the thick Fe70Pd30 film and Au (location: white box in (a)), showing the epitaxial growth

of the Fe70Pd30 film. (TEM measurements were performed by B. Erkatal and the figure was originally published

in Ref. 182)

Page 109: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

97

The Young’s modulus of the epitaxially grown Fe70Pd30 film was determined by

nanoindentation under the guidance of J. Pfetzing-Micklich (Figure 61 a)). Averaging over 25

measurements gave an average value of 140 (±10) GPa. To release the Fe70Pd30 film from the

substrate, wet chemical etching of the sacrificial Au buffer layer was performed using an

aqueous solution of potassium iodide and iodine. Owing to shadowing effects during the

sputtering of Fe70Pd30, access to the Au buffer layer was blocked at the vertices. Therefore the

5 mm x 5 mm MgO substrates were broken into 1 mm x 5 mm strips. The small contact

surface between the buffer layer and the etchant results in long etching times of up to 36 h.

KI/I 2 is highly selective and does not significantly attack the Fe70Pd30 film during etching as

determined by the absence of etching pits using optical microscopy. XRD measurements on

the released, freestanding films show only a slight lattice relaxation compared to the films

restricted by the rigid substrate. The out-of-plane lattice parameter changed from 0.321 nm to

0.318 nm (±0.001 nm).

Figure 61: a) Nanoindentation depth profile of the Fe70Pd30 film deposited on a single crystalline MgO substrate.

(nanoindentation was performed under the guidance of J. Pfetzing-Micklich) b) Tensile tests show a reversible

deformation of 1% (εmax ≈ 1.5%, εirr ≈ 0.4%). (Tensile testing was performed by C. Bechtold and the figure was

originally published in Ref. 182).)

Mechanical properties of the released films were further investigated with a uniaxial tensile

testing machine. Since the film thickness is very small for tensile testing, the fixation within

the Al2O3 clamping jaws is critical. Careful clamping of the film ensured no slip was observed

up to strains of 1.5% (Figure 61 b)). A maximum strain of 1.5% was obtained at a stress level

of 150 MPa. In total, a plastic deformation of less than 0.4% is observed. The slopes of

unloading and loading branches coincide well, proving an elastic behaviour in these regions.

From the linear slope of the first unloading, the Young’s modulus was estimated to be ≈ 15

Page 110: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

98

GPa. This is contrary to the nanoindentation measurements and indicates an unusual

elastically soft behaviour. The difference in mechanical properties determined by tensile

testing and nanoindentation can be explained by differences in mechanical loading and

deformation of the material. While in tensile testing uniaxial mechanical loading is applied

along [110] bct, a multiaxial stress state is present in the material under the tip of the

nanoindenter. The difference between the two measurements reflects the high mechanical

anisotropy that is characteristic for SMA.183 Nanoindentation measures an average of the

Young’s modulus in epitaxial films while tensile testing probes a particular direction.

Moreover, significant plastic deformation occurs in particular in front of the Berkovich tip.

By this deformation, the tetragonal lattice that has been forced by the coherent epitaxial

growth will get lost. The plastically deformed material at the nanoindenter tip is thus expected

to have significantly different mechanical properties compared to the rest of the material. All

these results prove that the constraints of an epitaxial interface allow coherent epitaxial

growth of Fe70Pd30 films up to large thicknesses as required for microsystems. Epitaxial

growth to such high thicknesses and without relaxation of the crystal structure implemented

by the seed layer is related to the mechanically soft behaviour originating from the

structural/martensitic instability. This is confirmed by stress-strain measurements of a

freestanding, 1.2 mm thick film. The observed Young’s modulus of 15 GPa is more than one

order of magnitude smaller than values known for common metals. Neither releasing the film

from the substrate nor heating it up to 375 K and straining resulted in a reverse transformation

to austenite. Since a stabilizing substrate is absent, alternative reasons for the absence of a

reverse transformation have to be considered. During a forward transformation, the tetragonal

martensite has a lower free energy than the cubic austenite below Ms. In order to minimize the

elastic energy at the habit plane, differently aligned unit cells form that are connected by twin

boundaries. Since one lattice constant of the martensite is shorter and one longer compared to

the austenite, this is geometrically possible. The reverse transformation commonly nucleates

at some residual austenite or at twin boundaries, where length compensation is possible. The

present film, however, was grown artificially, directly in the martensitic state. Considering a

diffusionless transformation from a single variant martensite to cubic austenite, the austenite

has no degree of freedom to compensate the tetragonal deformation. Hence the absence of the

reverse transformation (within the parameter ranges examined) can be attributed to the

different origin of the martensite stabilized in the film. The development of such artificial

structures as presented, are promising for micro-sensors probes. In these materials, length

changes can result in a significant change of magnetic properties and therefore can be

Page 111: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

99

detected even without contact. Since a tetragonal distortion of Fe70Pd30 films already resulted

in strong changes of the anisotropic magnetic properties184, the further orthorhombic

distortion during uniaxial strain is expected to give even higher effects. Compared to

superplastic stress-strain curves of FSMA within the martensitic state, the present film reveals

a reversible behaviour, not requiring an external restoring system, and almost linear behaviour

is observed over a range of > 1% strain, which is beneficial for unambiguous data evaluation.

The key advantage, however, is that a single variant state is kept and the magnetic properties

are not averaged over different variant orientations.

Page 112: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

100

4.2 Ternary Fe-Pd-X Ferromagnetic Shape Memory Alloys

4.2.1 The Fe-Pd-Mn System

4.2.1.1 Polycrystalline Fe-Pd-Mn thin films

The addition of Mn into Fe-Pd is quite promising as explained in Chapter 2.8, since it

preferentially couples antiferromagnetically to the surrounding Fe atoms and thus allows

utilization of the interdependence of the structural stability of the lattice and magnetism.

Investigation of Fe-Pd-Mn FSMAs bulk samples revealed that small Mn additions favourably

suppress the fct-bct transition that is known to be irreversible in bulk samples (see Chapter

2.6). Further, it was also reported that the addition of Mn also shifts the martensitic

transformation to higher temperatures.73

To clarify the impact of Mn on the structural and magnetic properties when added into Fe-Pd,

several Fe-Pd-Mn materials libraries were fabricated by co-deposition on Si/SiO2 substrates.

After deposition these materials libraries were subsequently annealed at 1123 K for 30 min

under N2 atmosphere followed by quenching. The high-throughput mapping techniques

mentioned in Chapter 3.2 were used to identify a compositional region of interest. Favourable

materials properties for new FSMAs were defined in Chapter 2.8 with respect to

transformation temperatures and magnetic properties. Thus, promising new FSMA thin film

samples have to undergo a martensitic transformation at an increased temperature when

compared to the binary Fe70Pd30 alloy. Further, these samples have to be in a ferromagnetic

state at T ≥ 550 K and in a single martensite phase. An example of the the identification of

this region of interest is shown for one materials library, and the findings are then summarized

for all three libraries fabricated, processed and investigated under identical conditions within

this thesis.

High-throughput characterization for identifying novel FSMAs

In a first step resistance mappings were performed across the materials libraries at 300 K. This

allowed identification of areas with similar values in electrical resistance. It has to be noted, that the

electrical resistance correlates with the conducting medium. This can lead to significantly different

values in electrical resistance for samples with similar materials properties when there is a difference

in film thickness. The film thickness of all samples in the presented materials library varies between

350 nm and 500 nm and therefore only changes its value in a range ≤ 30%. This variation in film

Page 113: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

101

thickness is negligible, since the resistance mapping is used to get a coarse overview of boundaries

connecting different phases. Different phases as well as single- and multiple phase regions exhibit

different amounts of boundary interfaces within the samples and thus shift electrical resistance inthese

samples to higher values.

Figure 62: a) Electrical resistance mapping at 300 K of a Fe-Pd-Mn materials library showing 3082 colour-

coded resistance values. b) Partial section of a ternary composition diagram with colour-coded resistance values.

Lines are added for clarity to distinguish between the different regions. (Figure originally published in Ref. 185)

For this mapping 3082 points were measured at a step size of 1.5 mm in x- and y-directions. Each

point was measured three times for better statistics. In Figure 62 the variation in electrical resistance

(colour-coded) is presented as a function of the location on the substrate in a) and as a function of

composition in b). For clarity only the region of interest is displayed colour-coded, while the areas

with significantly higher and lower resistance are displayed in grey. During measurement, contact

problems between the sample and the pins can occur, leading to measurement values in the range of

MΩ: these points also have been excluded (shown in grey). For displaying the resistance data (3082

values) in a ternary composition diagram (301 compositions measured), the results of the EDX

measurements were interpolated. This is valid because the measurement points are close enough to

each other such that the concentration gradients between them are almost linear. By interpolating the

data, each resistance value pairs up with a composition and can be displayed in the ternary

composition diagram using colour-code. The results indicate the presence of different regions

with significantly different values in electrical resistance. Although the electrical resistance

changes only from 0.15 Ohm to 0.28 Ohm, distinct areas with similar values can be identified.

In the next step, R(T) measurements were performed to identify thin film samples that

undergo a martensitic transformation. This was conducted using the mapping mode in a

temperature range of 273 K to 423 K with temperature steps of 5 K.

Page 114: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

102

Figure 63: a) Visualization of a Fe-Pd-Mn materials library showing the categorized results of R(T)

measurements. In b) R(T) results are summarized in the ternary composition diagram, where the region

containing transforming samples is encircled by a black line. Sample showing no transformation

(Fe51.1Pd42.8Mn6.1) are shown in c) while sample exhibiting a non-linearity and thus a martensitic transformation

(Fe68.5Pd26.8Mn4.7) are exemplarily presented for a specific composition in d). (Figure originally published in Ref.

185)

The results are presented in Figure 63: R(T) curves of samples that do not undergo a

martensitic transformation have a quasi-linear R(T) relationship (see Figure 63 c)), while

those exhibiting a martensitic transformation show an S-shaped non-linearity as presented in

Figure 63 d). As described, R(T) measurement curves give detailed information about the

transformation temperatures and temperature-hysteresis of transforming samples. The

analysis is time-consuming, since generally this has to be done manually. However, to get a

quick overview of the transforming sections of the materials library, the curves were

categorized to a scale ranging from 0 (non-transforming) to 1 (transforming). To do so the

smoothed second derivative of each curve was taken as an indication if a transformation

occurs. From the derived curve, the difference between maximum and minimum was taken

Page 115: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

103

and collected for each curve before normalizing all values. The thin film samples in the area

depicted by a black line in Figure 63 b) were identified to undergo a martensitic

transformation. To prove if the samples which show a martensitic transformation are

ferromagnetic, MOKE measurements were carried out. Using the MOKE setup described in

Chapter 3.2, 301 compositions from the Fe-Pd-Mn materials library were measured at 300 K

(measurement grid: 4.5 mm). The magnetic flux of the electromagnet was cycled between -

300 mT and +300 mT.

Figure 64: Colour-coded visualization of the Fe-Pd-Mn materials library showing 301 categorized MOKE

measurement points. b) The same results shown in a section of the ternary composition diagram. The

ferromagnetic region is encircled by a black line. The following graphs depict different types of MOKE

measurement curves. A non-ferromagnetic sample (Fe63.2Pd19.7Mn17.1) is shown in c). Measurement curves for

ferromagnetic samples with detectable (Fe72.4Pd15.8Mn11.8) and well-defined magnetic hysteresis

(Fe82.1Pd12.3Mn5.6) are shown in d) and e). (Figure originally published in Ref. 185)

While the non-ferromagnetic points show no hysteresis curve (Figure 64 c)), the magnetic

sections of the materials library display hysteresis curves with step-like changes in MOKE

intensity (Figure 64 d) and e)). To compare the resulting data, the measurement curves were

automatically categorized defining a scale ranging from 0 (no hysteresis) to 1 (well-

pronounced hysteresis). This was done by taking again the second derivative of the original

curve to analyze the change in intensity of the MOKE signal. Since the different measurement

Page 116: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

104

curves exhibit different noise levels, the resulting curve was divided by a factor proportional

to this noise level. The difference between maximum and minimum values of this data was

collected for each curve and finally normalized. These values correspond well with the quality

of the curves. The screening of the magnetic properties of the Fe-Pd-Mn materials library

using MOKE revealed ferromagnetic and non-ferromagnetic regions in the composition

diagram as visualized in Figure 64. On the Fe-rich side of the phase diagram (Fe > 80 at.%)

all samples are ferromagnetic, while in the Fe-poor region (Fe < 55 at.%) and the region with

Mn content > 12 at.% no ferromagnetic behaviour is observed. The remaining area shows

both ferromagnetic and non-ferromagnetic sections depending on the Pd- and Mn content.

The R(T) measurements revealed a region in the phase diagram where samples undergo a

martensitic transformation. In the room temperature resistance map of the materials library,

the same section can be is identified as well because it exhibits a higher resistance compared

to the surrounding points. In particular, those points are of interest which reveal both,

magnetic and phase-transformation behaviour. There is not a complete overlap of these zones,

but rather a small section includes samples of all regions as shown in Figure 65.

Figure 65: Superposition of all measurement results in the Fe-Pd-Mn materials library defining a region of

interest. All thin film samples within this region are ferromagnetic at 300 K and undergo a martensitic

transformation. (Figure originally published in Ref. 185)

This area lies in the range of (60 - 66 at.%) Fe, (6 - 8 at.%) Mn and (28 - 32 at.%) Pd. This

suggests that for Fe-Pd-Mn FSMAs Mn has to be added to the Fe70Pd30 system at the expense

of Fe (keeping the Pd content approximately constant) in order to achieve a transforming

phase, similar to Fe70Pd30.

The presence of a martensitic transformation determined by the R(T) measurement was

finally proven by using XRD(T). This is presented in Figure 66 where a Fe68Pd29Mn3 sample

was investigated between 300 K and 398 K. At 300 K the sample is in the fct martensite

phase indicated by the presence of the (111) and the absence of the (200) fcc diffraction peak.

Page 117: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

105

Figure 66: XRD(T) measurement to verify that thin films showing an S-shaped non-linearity in the R(T)

screening undergo a martensitic transformation.

The (200) and the (002) fct peaks cannot be detected, since the signal to noise ratio is very

low in this measurement, hiding both diffractions in the background of the spectra. With

increasing temperature the (200) fcc peak of the austenite phase appears at 333 K and

develops in intensity fully at 363 K. This growing and disappearing of the (200) fcc peak was

found to be reversible, indicating that thin films with an S-shaped R(T) curve undergo a

martensitic transformation.

Structural investigation of all Fe-Pd-Mn materials libraries was conducted by XRD

measurements. Figure 67 presents samples that are in a single martensite/austenite phase

(depicted by filled red and blue circles) within the compositional region of interest. This

region is located within multi-phase areas. Besides the phases already known from the binary

Fe-Pd alloy (Fe50Pd50 with α-Fe precipitates and bcc Fe-Pd) new Mn-rich phases appear. At

Fe contents < 57 at.% a fcc structure, similar to the Fe50Pd50 phase and elemental Pd and Mn

precipitates (open green circles) appear. By increasing the amount of Fe > 67 at.%, the Fe-Pd

fcc/fct phases decompose and precipitates form. As described before, FSMAs have to be in a

single-phase state to show the MFIS, since defects such as grain boundaries hamper twin

boundary motion.

Page 118: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

106

Figure 67: Diagram of the metastable phases occurring in the Fe-Pd-Mn thin film system. Besides multi-phase

regions, a small single-phase region is present (filled red and blue circles correspond to samples in a single

martensite/austenite phase at 300 K). (Figure originally published in Ref. 190)

Therefore all further investigations concentrate on Fe-Pd-Mn thin films being in the fcc/fct

single-phase state as depicted in Figure 67. The metastable phase diagram for the Fe-Pd-Mn

system indicates that the stability of the phases is preserved by trend, where the amount of Mn

increases at the expense of Fe for all single-phase samples. This is not unexpected since it is

known that Fe and Mn form several intermetallic compounds.186 Furthermore, an increase of

the lattice spacing was found in all fct/fcc Fe-Pd-Mn single phases when compared to binary

Fe70Pd30. This is related to the larger covalent atomic radius of Mn and the increased

spontaneous volume magnetostriction in Fe-Mn alloys.218 In addition, it is known from

literature that the crystal structure as well the intrinsic properties in Fe-Mn alloys vary

significantly dependent on the composition and annealing procedure.218 This is in good

agreement with previously reported results from literature.73 Although the solubility of Mn in

Fe-Pd was reported to be < 2.5 at.% it was found that Mn contents up to 10 at.% are soluble if

the Fe content is reduced accordingly.73

Page 119: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

107

In Figure 68 the martensite start temperature Ms is correlated to the different crystal phases

determined within the Fe-Pd-Mn materials libraries. The compositional range in which single-

phase Fe-Pd-Mn thin films undergo a martensitic transformation follows the line in the phase

diagram, where the Mn content is increased at the expense of Fe content. The values

determined for Ms are the highest transformation temperatures reported for any binary and

ternary Fe-Pd-X FSMAs so far.

Figure 68: Phase diagram of single-phase Fe-Pd-Mn thin films with Ms transformation temperature depicted by

colour-code.

Magnetic properties of Fe-Pd-Mn thin films

Further changes in the magnetic properties are observed within the Fe-Pd-Mn single-phase

region when compared to Fe70Pd30. This is caused by the different magnetic alignments of Fe

and Mn in the Fe-Pd matrix. In Figure 69 the analyzed data revealed by magnetic

measurements for two thin films is presented as an example. Figure 69 a) shows the magnetic

hysteresis curves for a Fe66.1Pd30.1Mn3.8 and a Fe60Pd31Mn9 thin film in the martensitic (50 K)

and the austenitic (400 K) state. It can be seen, that both samples do not show a significant

difference in HC but do in JS. Further, the martensite phase has a significantly higher

saturation polarization JS than the austenite phase as explained previously for the binary

Fe70Pd30 alloy. In Figure 69 b) the normalized magnetization J/J50 K as a function of

temperature between 50 K and 400 K and under a constant magnetic flux density of µ0H =

0.05 T is depicted for both samples.

Page 120: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

108

Figure 69: a) Magnetic hysteresis measurement for the martensite (50 K) as well as for the austenite (400 K)

state of a Fe66.1Pd30.1Mn3.8 (open circles) and a Fe60Pd31Cu9 (filled squares) thin film. b) Normalized

magnetization J/J50 K as a function of temperature and under a constant magnetic flux density of µ0H = 0.05 T is

presented. The data points are fitted using the Kuz’min equation for the Fe66.1Pd30.1Mn3.8 (open circles and red

curve) and the Fe60Pd31Cu9 (filled squares and blue curve) thin film.

Using the Kuz’min relation a fit (red: Fe66.1Pd30.1Mn3.8; blue: Fe60Pd31Mn9) was applied

through the data points and extrapolated to the temperature, where J/J50 K approaches zero.

Both the JS and TC values revealed for single-phase samples are presented in Figure 70. Here,

the saturation polarization as a function of Fe content (colour coding indicates the Mn

content) for both the martensitic (squares, determined at 50 K) and the austenitic structure

(triangles, determined at 400 K) is shown in a). The saturation polarization decreases with

increasing Mn content from JS = 1.56 T for Fe66.7Pd28.8Mn4.5 and reaches a minimum at

JS = 1.22 T for Fe61.6Pd30Mn8.4 in the martensitic fct phase (determined at T = 50 K). At Mn

contents > 8.4 at.% JS increases again slightly to JS = 1.29 T for the Fe60Pd31Mn9 thin film.

This behaviour originates from the addition of antiferromagnetic Mn at the expense of

ferromagnetic Fe. The antiferromagnetic interactions are considered to be strong and thus an

increase of TC with decreasing Fe content is observed. This is shown in Figure 70 b) where TC

and the shape fitting parameter s as a function of Fe content are presented.

Page 121: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

109

Figure 70: a) JS as a function of Fe content for single-phase samples in both the martensite (50 K) and austenite

phase (400 K). Highest values for JS are identified for the highest Fe content. b) TC and shape parameter s as a

function of Fe content, show a significant decrease with increasing Fe content. The Mn content is indicated by

colour-coding.

This behaviour is not unusual, since the corresponding trend in the binary alloy exhibits an

increase of TC with decreasing Fe content until it reaches its maximum value for the Fe50Pd50

phase.187 Nevertheless, it can be concluded that the addition of Mn generally lowers TC when

compared to the binary Fe70Pd30 alloy. The addition of 9 at.% Mn at the expense of Fe content

lowers TC (TC for Fe70Pd30 = 700 K to Fe60Pd31Mn9 = 499 K) by about 28 %.

Transmission electron microscopy

TEM was performed to investigate the existence of a single ternary phase in the sample

Fe66.1Pd30.1Mn3.8. The prepared lamella consist of the following layers: SiO2 substrate layer,

the homogeneous Fe-Pd-Mn layer and a Pt-protection layer deposited prior to FIB-

Page 122: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

110

preparation. Figure 71 a) shows the HRTEM image of a selected region of the

Fe66.1Pd30.1Mn3.8 layer. A red coloured square indicates a region where a diffraction pattern

was calculated using a fast Fourier transformation. The difference in colour across the

overview photograph in Figure 71 a) is related to the orientation contrast of differently

orientated grains in the lamella. The diffraction pattern taken from the region is depicted by a

red square in Figure 71 b) and reveals the occurrence of lattice planes with different distances.

The different distances are ascribed to the following lattice planes: d(002) fct = 0.189 nm,

d(111) fct = 0.22 nm and d(111) fcc = 0.223 nm.

Figure 71: a) High resolution transmission electron microscope (HRTEM) overview of a Fe66.1Pd30.1Mn3.8 film.

The inset b) shows a fast Fourier transformation conducted over a regions depicted by a red square. The

diffraction patterns in this region revealed a mixture of the martensite fct and the austenite fcc phases. (FIB

preparation was performed by C. Zamponi and TEM by B. Erkatal from the group of Prof. L. Kienle at the

Christian-Albrechts-University of Kiel)

The determined lattice spacings confirm that the presented lamella of a Fe66.1Pd30.1Mn3.8 thin

film has a microstructure consisting of both the martensite as well as the austenite phase.

Since this measurement was performed at ambient temperature (300 K) the martensite phase

would be expected judging from the increased transformation temperatures compared to

Fe70Pd30. Nevertheless, the sample where the lamella was extracted is a thin film. This kind of

Page 123: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

111

sample usually exhibits high residual stresses due to the difference in thermal expansion

between thin film and substrate. This can lead to a significant increase in transformation

temperatures due to stress-induced martensite as mentioned previously. After extraction of a

lamella out of the surrounding thin film, the residual stress can partially relax through a

bending of the lamella, allowing it to transform back into the austenite phase as shown here.

Further a local heating of the observed region by the electron beam cannot be excluded. This

can lead to increased observation temperatures than the ambient value of 300 K and thus can

induce the martensitic transformation in the lamella. The determined lattice distances are

significantly higher compared to the distances reported for the binary Fe70Pd30 alloy. This is

in good accordance with the synchrotron measurements that revealed an increase of the lattice

parameters upon addition of Mn at the expense of Fe, which are presented in the following

section. Compositional analysis by EDX further revealed no decomposition or separation of

Mn and Fe-Pd. Thus it can be summarized that Mn is homogeneously dissolved in the Fe-Pd

matrix and no composition or phase segregation was detected within the analyzed regions.

Stress effects in Fe-Pd-Mn thin films

To further clarify how the addition of Mn varies the lattice parameters within the Fe-Pd-Mn

thin films, temperature-dependent synchrotron radiation measurements were performed.

Further residual stress measurements using the sin2(ψ) method were used to determine the

crystal structure and the stress state of Fe-Pd-Cu thin films in the martensite (300 K) as well

as in the austenite (393 K) state. In Figure 72 the area detector diffractograms of a

Fe66.1Pd30.1Mn3.8 thin film at two different temperatures are presented as examples. The area

diffractograms depict the colour-coded intensity (blue for low and white for high intensity) in

terms of X-ray counts as a function of d-spacing and ψ angle. At an angle of ψ = 90°, the ψ-

vector points to the thin film surface at an angle of 90° and thus is parallel to the thin film

normal vector. This is indicated by a dashed white line, that divides the upper half (90°-160°)

to one half-sphere and the lower part (20°-90°) to the other half-sphere above the sample.

Instead of intensity rings known for polycrystalline samples, intensity spots appear. These

spots originate from texturing and correspond to the specific d-spacings of different phases. In

Figure 72 a) the crystal structure of a Fe66.1Pd30.1Mn3.8 thin film in the martensite phase at

300 K is presented. The martensite phase can be identified by the presence of the (220), (002),

(200) and (111) fct diffraction peaks. Upon heating to 393 K the thin film transforms into the

austenite phase by developing a strong (200) fcc peak.

Page 124: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

112

Figure 72: Area detector diffractograms of a Fe66.1Pd30.1Mn3.8 thin film at different temperatures. a) At 300 K the

thin film is in the martensite phase, proved by the presence of the (200) and (002) fct peaks. When heating the

sample to 393 K in b), the sample transforms into the austenite phase and the (200) fcc peak appears. The red

stars denote diffraction intensities originated from the Si/SiO2 substrate. (Synchrotron measurements were

performed by H. Brunken)

In order to verify the lattice spacings determined by TEM, the lattice spacings in the

martensite (fct (111) and fct (200) diffraction peaks at 300 K) and austenite (fcc (111)

diffraction peak at 393 K) state were determined. The variation of lattice spacings in

Page 125: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

113

dependence on the Fe content (Mn content is colour-coded) is depicted in Figure 73 for a

Fe66.1Pd30.1Mn3.8 thin film. The values for both the (111) fct martensite (open circles) and

(111) fcc austenite (open squares) lattice spacings correspond to the left scaling (black

arrows). Values describing the variation of the (200) fct martensite (open triangles) lattice

spacing correspond to the right scale (dark blue). From this data it can be easily seen that the

lattice parameters increase upon addition of Mn, which originates from the significantly

higher covalent atomic radius compared to Fe (rMn = 0.135 nm < rFe = 0.126 nm).188

Figure 73: Shown is the lattice spacing revealed from the different diffraction peaks in the martensite (circles:

fct (111) and triangles: (200) diffraction peaks at 300 K) and the austenite (squares: fcc (111) diffraction peak at

393 K) state as a function of the Fe content for a Fe66.1Pd30.1Mn3.8 thin film. The values for (111) peaks

correspond to the scaling on the left (black). Values determined from the fct (200) correspond to the right scale

(dark blue).

It has to be mentioned that the lattice spacing of the fcc (111) diffraction peak was measured

at a higher temperature (393 K) compared to TEM performed at ambient temperature (300 K).

Since the lattice parameters especially in Fe-Pd-based systems depend on temperature, slight

differences have to be expected. Further, stress effects can alter the lattice parameters,

especially when considering that the TEM lamella can partially relax while the thin film was

substrate attached during synchrotron radiation measurements. Thus it can be concluded, that

values determined from synchrotron measurements correlate well with values determined

from TEM diffraction on a Fe66.1Pd30.1Mn3.8 thin film within the error margins.

In order to determine the residual stress state of this thin film sample in the austenite phase,

the lattice parameter afcc of the thin film was determined from the (111) fcc peak at different

ψ-angles. The variation of lattice parameter with tilt angle afcc = sin2(ψ) is shown in Figure

Page 126: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

114

74 a). From the slope of this relationship the stress state and the strain in this thin film can be

determined. The data points are fitted by using a linear function and the slope of this function

is used to determine the residual strain. The positive slope indicates a tensile residual strain

that appears for thin films showing Invar-related anomalies with smaller thermal expansion

coefficient than the substrate they were deposited onto. In Figure 74 b) the FWHM and c) the

intensity of the (111) fcc diffraction peak as a function of sin2(ψ) is shown to ensure no

texture related errors. Since the sin2(ψ) method only reveals the strain of the sample, the

Young’s modulus needs to be determined further for every thin film.

Figure 74: Lattice parameter as a function the ψ-angle for a Fe66.1Pd30.1Mn3.8 thin film. a) A linear function was

fitted through the data points to determine the slope. The positive slope indicates a tensile strain state in this

sample. For every lattice spacing afcc in a) that corresponds to a diffraction peak, the FWHM b) and the intensity

c) for this peak was determined.

This was performed by nanoindentation measurements of the Fe-Pd-Mn thin films. The

measurements were conducted at ambient and elevated temperatures (353 K) to determine the

Young’s modulus in the austenite phase. Examples of the nanoindentation measurements are

depicted in Figure 75 a) to c) for a Fe66Pd30Mn4 thin film sample. In Figure 75 a) the load on

the sample as a function of the displacement starting from the thin film surface is shown. In

Figure 75 b) the data evaluated to measure the Young’s modulus are presented.

Page 127: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

115

Figure 75: Examples of measurement curves for a) load, b) Young’s modulus and c) hardness as a function of

displacement of the Berkovich tip into the thin film surface.

The area defined by hatched stripes depicts the indentation depth that was used for Young’s

modulus measurement and describes a displacement ranging from 50 nm to 70 nm. The

values for thin film hardness are illustrated in Figure 75 c). Here, the hardness values again

were calculated from indentation depth between 50 nm to 70 nm. The choosen indentation

depth follows the rule that up to an indentation depth of one-tenth of the overall film

thickness, the influence of the substrate is negligible.

Figure 76: Summary of nanoindentation measurements at 353 K for Fe-Pd-Mn single-phase thin films.

Presented are the a) Young’s modulus and the b) thin films’ hardness as a function of Fe content with colour-

coded Mn content. With decreasing Fe content (and increasing Mn content) a lowering of Young’s modulus and

hardness are observed. Further data from binary Fe70Pd30 reference thin films that were fabricated and processed

under identical conditions are included in both diagrams.

Nanoindentaion measurements are summarized for single-phase Fe-Pd-Mn thin films in

Figure 76. In Figure 76 a) the Young’s modulus values as a function of Fe content (Mn is

colour-coded) are illustrated. From the distribution of the data, it can be stated that the

Page 128: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

116

Young’s modulus increases with increasing Mn content. The values for the binary Fe70Pd30

reference thin films that were fabricated and processed under identical conditions, are marked

by a hatched area around 140 GPa. Since the values are significantly higher than the value for

the underlying Si/SiO2 substrate, a significant influence of the Si substrate can be excluded.

Hardness values revealed from nanoindentation measurements are shown in Figure 76 b) and

follow a similar trend as observed for the variation Young’s modulus with Mn content.

The determined values for the Young’s modulus were used to calculate the amount of residual

stress in the Fe-Pd-Mn single-phase thin films. In Figure 77 the residual stress as a function of

the Mn content is depicted for samples with a Pd content between 28.7 at.% and 30.5 at.%.

The colour-code indicates the martensite start temperature Ms. A significant increase in

residual tensile stress is observed with increasing Mn. The values for low Mn content > 7 at.%

exhibit high tensile stress values, while Ms is in the range of 356 K to 364 K. Upon decrease

of Mn content, the amount of tensile stress lowers significantly while Ms increases. At a Mn

content of 6.7 at.% the amount of tensile stress is lowered to 658 MPa although the

martensitic transformation is shifted to the highest value of Ms = 379 K.

Figure 77: Presented is the amount of residual stress as a function of Mn content. To separate the impact of Pd

on thermal expansion behaviour and thus the stress state, only thin film samples with a small variation in Pd

(28.7 at.% < Pd < 30.5 at.%) content are compared. Upon addition of Mn a significant increase of tensile stress is

observed. The change of Ms temperature (colour-coded) does not follow a well-defined trend and has average

values for samples with highest tensile stress amount.

Thin films with a 4.5 at.% < Mn < 6 at.% exhibit Ms temperatures > 350 K. Nevertheless, the

lowest value for Ms = 336 K is found for a Fe66.1Pd30.1Mn3.8 thin film with a stress of

Page 129: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

117

339 MPa. From this data it can be concluded that no obvious correlation between tensile

stress amount and transformation temperatures exists.

Variation of martensitic transformation in the Fe-Pd-Mn system

In order to identify the dominant parameter that can be used to describe the shift of

transformation temperatures, the martensite start temperature Ms of all samples in a single-

phase state is plotted as a function of e/a ratio and the Fe content. The amount of Mn in each

sample is indicated by the colour-code of the symbols. The data is presented in a in a three-

dimensional diagram. For clarity, only the projections of data points into the three base planes

are depicted. Ms (coloured triangles) and the e/a ratio (coloured circles) as a function of Fe

content do not show a simple analytic trend.

Figure 78: Martensite start temperatures, Ms as a function of e/a ratio and Fe content presented in a three-

dimensional graph with colour-coded Mn content. Due to the addition of Mn, Fe content and e/a ratio are two

independent variables. When looking at the martensite start temperature Ms as a function of the e/a ratio, a linear

dependency becomes visible. In the other two projections no clear dependency is present. (Figure originally

published in Ref. 190)

However, an approximately linear behaviour is observed when Ms is depicted as a function of

the e/a ratio. It should be kept in mind that the e/a ratio is independent of the Fe content when

one or more additional elements are added into Fe-Pd alloys. At e/a = 8.51 (Fe64.3Pd29Mn6.7)

the highest transformation temperature in the system is Ms = 379 K and reaches down to a

value of 335 K at e/a = 8.56 (Fe66.1Pd30.1Mn3.8). The description of Ms in terms of the e/a ratio

is common in literature, especially for Fe-based systems. Although this relation was mostly

Page 130: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

118

applied to binary alloys such as Fe-Pd, Fe-Pt and Fe-Ni,189 where the Fe content and the e/a

ratio are related to each other, this dependence holds also for some ternary Fe-Pd-X (X = Ni

and Co) alloys including the Fe-Pd-Mn system.71 After identification of the decisive factor

that describes the change of the martensitic transformation, the origin of the increased

transformation temperatures has to be clarified.

Figure 79: Energy gain due to displacements from the ideal lattice positions compared for different ternary

compositions Fe70.4Pd25X4.6, where X represents the third component. Further information can be found in Ref.

65. The choice of antiferromagnetic elements such as X = Cr and Mn lead to considerable energy gain compared

to binary Fe70.4Pd29.6. (Figure adapted from Ref. 190)

A first hint comes from ab initio calculations performed by M. E. Gruner presented in Figure

79. Here, a significant energy gain (approximately 27 meV/atom) due to displacements from

the ideal lattice positions is found after addition of 5 at.% of Mn. This is significantly smaller

if the lattice sites are occupied by (ferromagnetic) Fe and underlines that antiferromagnetic

Mn causes local structural distortions (even at low temperatures) which significantly impact

the system when compared to other ferromagnetic contributions like Fe, Co, and Ni.191

However, the influence of the magnetoelastic effects at finite temperature are not easy to

predict, since - on the one hand - a magnetically excited Mn atom might still show a parallel

alignment to the average magnetization, while - on the other hand - the thermal stability of the

ferromagnetic phase itself is compromised by large amounts of Mn. A shift of martensitic

transformation due to thin film stress effects is not considered, since Ms does not distinctly

scale with the stress amount.

By summarizing all the shown findings an enhancement of properties due to the addition of

Mn into Fe-Pd occurs. The Fe-Pd-Mn system exhibits the highest transformation temperatures

Page 131: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

119

up to 379 K measured for Fe-Pd-based FSMAs so far. The single-phase region in the

compositional diagram is extended, allowing to alloy up to 9 at.% of Mn in Fe-Pd without

segregation of the fct/fcc phase. Considering the magnetic properties, neither TC nor JS are

drastically lowered. From theoretical predictions an energy gain due to displacements from

the ideal lattice positions upon addition of Mn was reported, explaining the increased amount

of energy needed to induce the martensitic transformation. Finally the e/a ratio was found to

be the dominating parameter, describing the change of the martensitic transformation.

4.2.2 The Fe-Pd-Cu System

4.2.3.1 Polycrystalline Fe-Pd-Cu thin films

In order to investigate the materials properties without significantly affecting the magnetic

properties Cu was added into Fe-Pd as described in Chapter 2.8. The change of the e/a ratio at

constant Fe content, the variation of the unit cell volume and the impact of Cu on the

stabilization of the metastable and disordered Fe70Pd30 phase against decomposition was

investigated. Due to the higher number of valence electrons of Cu in comparison to Fe and

Pd, the e/a ratio is supposed to increase. Nevertheless if the Fe content is increased at the

expense of Pd, also e/a ratios below 8.6 can be achieved. This allows for a variation of the Fe

content while keeping the e/a ratio constant and thus enables the identification of the decisive

factor that characterizes the martensitic anomaly. The colvalent atomic radius of Cu lies

between those of Fe and Pd and thus decreases the unit cell volume, if the Cu content

increases at the expense of Pd. Finally, Cu with its full d-shell is known to be an austenite

stabilizer eventually allowing stabilization of the high temperature fcc austenite phase.

In order to gain a deeper understanding of basic magnetic and magnetoelastic properties, first-

principles calculations in the framework of density functional theory (DFT) were carried out.

The theory results by M. E. Gruner (University of Duisburg-Essen) were then compared to

experimental results. The calculations were carried out with the Korringa-Kohn-Rostoker

Green-function method (KKR) in connection with the atomic sphere approximation (ASA)

and the coherent potential approximation (CPA) for the description of configurational

disorder, as implemented within the fully relativistic Spin Polarized Relativistic Korringa-

Kohn-Rostoker (SPR-KKR) package from Ludwig-Maximilians-Universität Munich.192,193

The calculations employed the generalized gradient approximations (GGA) for the description

of the exchange correlation potential in the formulation of Perdew, Burke and Ernzerhoff.194

Separate calculations were carried out for all compositions exhibiting martensitic

Page 132: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

120

transformations as determined in the experiments. For each composition, the equilibrium

lattice constants for the austenite ferromagnetic and magnetically disordered configuration

were determined; the latter was simulated by mixing Fe-species of different spin orientations

by means of CPA, in the spirit of the disordered local moment approach.195,196 Furthermore,

spin and orbital magnetic moments were obtained assuming a ferromagnetic spin

configuration as the ground state. In the angular momentum expansion, orbitals were included

up to lmax = f states; 1639 irreducible k-points were used for the Brillouin zone integration. In

a separate calculation, magnetic exchange-parameters were determined for use within a

Heisenberg-model.197 From this, the Curie temperatures within the mean-field approximation

were obtained, which allows a comparison of compositional trends between theory and

experiment.

Structural behaviour of Fe-Pd-Cu thin films

The measured compositions of the ternary Fe-Pd-Cu materials libraries are depicted in a

partial composition diagram in Figure 80 a) (600 samples). Open grey circles represent

compositions measured by EDX, while black lines indicate the calculated e/a ratios of the

respective samples. Filled circles represent all samples showing a thermoelastic non-linearity

in the R(T) measurement, with the martensite start temperature being indicated by colour-

coding. The composition area investigated, is centered around Fe70Pd30 (from Fe40 to Fe94)

with variation of Cu from 1 to 18 at.%. In order to identify compositions showing phase

transformations, all compositions were investigated by R(T) screening measurements.

Figure 80: a) Partial Fe-Pd-Cu composition diagram showing the compositional area covered by two materials

libraries. Constant e/a ratios are defined by black lines. Grey circles denote the fabricated samples of the

materials libraries. Transforming samples are indicated by filled circles, with colour-coding indicating the

martensite start temperature Ms; b) example R(T) curve of the transforming sample Fe71.8Pd26.6Cu1.6. The

determination of transformation temperatures by the tangential method is indicated. (Figure originally published

in Ref. 199)

Page 133: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

121

Figure 80 b) shows an example R(T) curve from a Fe71.8Pd26.6Cu1.6 thin film, which indicates

a phase transformation due to its non-linear S-shaped R(T) curve, from which transformation

temperatures were determined by the tangential method.144 In order to assign the observed

non-linearities in the R(T) curves to a martensitic transformation, XRD(T) was performed.

Figure 81 shows an example of XRD(T) data for the sample Fe71.8Pd26.6Cu1.6. At 253 K, (111)

and (200) fct reflections are observed, as expected for a polycrystalline martensitic structure

comparable to binary Fe70Pd30.61 No additional peaks are observed, which indicates the

absence of large precipitates. With increasing temperature, the (111) fct peak does not change;

instead the (200) fct peak decreases while the (200) fcc peak starts to grow, starting at

approximately 293 K. With further increasing temperature, the (200) fcc peak continues to

increase in intensity up to 343 K. For temperatures above 343 K, no further growth of the

(200) fcc peak is observed. Thus, martensitic transformation temperatures deduced from the

R(T) measurement are confirmed by the structural analysis and the transformation

temperatures determined by these methods showed no significant differences.

Figure 81: XRD(T) of the sample Fe71.8Pd26.6Cu1.6 measured between 253 K and 373 K. (Figure originally

published in Ref. 199)

In order to determine the structure of the samples, room temperature XRD mapping of the Fe-

Pd-Cu samples was performed. Figure 82 a) shows the ternary Fe-Pd-Cu composition

diagram, indicating all investigated compositions. Transforming compositions are colour-

coded according to their martensite start temperature, Ms, while the symbols represent the

respective phases. The squares in Figure 82 a) define single-phase compositions showing a

non-linearity in the R(T) measurement. Diamonds designate multi-phase samples where the

Page 134: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

122

transforming Fe70Pd30 phase dominates but additional precipitate phases occur and are thus

not considered any further. The dark grey triangles represent samples having a Fe50Pd50

structure with additional Fe-Cu and Pd-Cu precipitates. For Fe-concentrations above 72.5

at.% only a bcc structure and precipitate phases (α-Fe, Fe-Cu and Pd-Cu) are present (grey

circles). Open squares indicate samples having a Fe70Pd30 austenite phase with additional Fe-

Cu and Pd-Cu precipitates present at larger Cu contents, and not showing an S-shaped R(T)

measurement curve.

Figure 82: a) Ternary diagram of the Fe-Pd-Cu system showing the distribution of phases (symbols) and the

transforming compositions (Ms colour-coded). The arrow indicates a constant Fe content. Diffractograms along

this composition are summarized in b). For Cu > 6 at.% the transforming fct/fcc structure decomposes into a bcc

structure with Fe8Cu2 precipitates. (Figure originally published in Ref. 199)

Page 135: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

123

All single-phase compositions are located in a zone with Fe content between 69 and 72 at.%

and having a Cu content from 1 to 5 at.%. The distribution of transforming single-phase

samples in the ternary compositional diagram suggests an increase of Cu at the expense of Pd.

Samples with Cu contents <5 at.% transform up to a maximum Fe content of 72 at.%. For

binary films an expansion of the transforming region from Fe70Pd30 to Fe72Pd28 was

previously observed.143 However, in the ternary system, a considerably larger maximum

Ms = 359 K for compositions around Fe71.8Pd26.6Cu1.6 was observed. In order to understand

how Cu alloying into Fe-Pd affects the structure, a “binary cut” was defined through the

ternary composition diagram (arrow in Figure 82 a)). Along this line, Cu and Pd contents

vary, while the Fe content is constant at 70 at. %. Figure 82 b) illustrates the structural change

of these samples, in the form of a colour-coded (red = high intensity, blue = low intensity) top

view of the diffraction patterns. A martensitic structure with (111) and (200) peaks is

observed at room temperature up to 6 at.% Cu. No additional peaks originating from

precipitate phases are observed. For Cu contents > 6 at.%, a (110) bcc phase occurs in

addition to the transforming fct/fcc phase. At 8 at.% Cu, the (110) Fe8Cu2 peak appears. Due

to an increase of Cu at the expense of Pd content, a slight shift of the (111) fct/fcc peak is

observed over the whole compositional range as shown in Figure 82 b).

While austenitic Fe70Pd30 has a lattice constant of a = 0.3756 nm57, alloying with small

amounts of Cu should change this. The lattice constant a of the austenite fcc phase was

determined from the (200) fcc peaks of selected Fe-Pd-Cu samples using XRD at 373 K.

Figure 83: Calculated and experimentally determined lattice constants for fixed Fe contents (70 to 72 at.%) in

dependence on Cu. Inset: Lattice constants a of the fcc cubic austenite phase (measured at 373 K) for single-

phase transforming Fe-Pd-Cu samples within the composition diagram. (Lattice parameters calculated by M. E.

Gruner and originally published in Ref. 199)

Page 136: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

124

Results are shown in Figure 83 for experimental (black) and calculated (red) lattice

parameters for selected Fe-Pd-Cu samples with a constant Fe content (Fe70: squares; Fe71:

triangles; Fe72: circles). The inset shows the locations of the experimental samples in the

ternary composition diagram. Since the covalent atomic radius for Cu (0.128 nm) lies

between that of Fe (0.126 nm) and Pd (0.137 nm), a decrease of lattice constants with addition

of small amounts of Cu at the expense of Pd should occur.188 As expected, the lowest lattice

constant is observed for samples with the smallest Pd and highest Cu content. However, the

lattice parameters of the Fe-Pd system are known to deviate from Vegard´s law198 and at

elevated temperatures the Invar-behaviour of Fe-Pd leads to a significant reduction of the

thermal expansion coefficient in a broad range of temperature. In fact, the calculated values

for a (determined for the ferromagnetic ground state) are slightly higher than the lattice

constants measured experimentally. Since calculations represent the state at 0 K and

measurements were performed at 373 K, this is opposite to what one would expect from

normal thermal expansion and thus another indication for the presence of Invar anomalies in

these films. This observation is of importance for the interpretation of the influence of

substrate-induced stress on the martensitic transformation and will be discussed in the

following.

Transmission electron microscopy

TEM was used to confirm the existence of a single ternary phase in the sample

Fe69.5Pd26.7Cu3.8. The prepared lamella (not shown) consist of the following layers: SiO2

substrate layer, a thin FeOx intermediate layer, the homogeneous Fe-Pd-Cu layer and a Pt-

protection layer deposited during FIB-preparation. The thin oxidized Fe layer at the film-

substrate interface is related to diffusion effects during annealing and has a thickness of

approximately 40 nm. The grain structure of the layer, revealed by diffraction contrast,

verified a homogenous chemical composition in the Fe-Pd-Cu layer. Figure 84 a) shows the

HRTEM image of a selected region of the Fe-Pd-Cu layer. The martensitic twin structure can

be seen in the lattice image as well as on the Fast Fourier Transform (inset b) in Figure 84).

Page 137: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

125

Figure 84: High resolution transmission electron microscope (HRTEM) image showing twinned lattice planes

of a Fe69.5Pd26.7Cu3.8 film. The inset shows the fast Fourier transform revealing the twinned microstructure. (FIB

preparation was performed by A. Sehrbrock and TEM by S. Irsen at caesar Bonn and the figure was originally

published in Ref. 199)

Summarizing these results, Cu is homogeneously dissolved in the Fe-Pd matrix. In the

investigated area, no composition or phase segregation was detected. The results were

confirmed by Auger-electron spectroscopy depth profiles.

Magnetic properties of Fe-Pd-Cu thin films

Since determination of magnetic anisotropy is hampered by the polycrystalline microstructure

of the films, the experimental investigations concentrate on the other key properties, the

saturation polarization, JS and the Curie temperature TC. In Figure 85 magnetic measurements

for a Fe71Pd27.7Cu1.3 (open circles) and a Fe70.2Pd25.5Cu4.3 (filled squares) thin film are

presented as examples. Figure 85 a) shows the magnetic hysteresis curve in the martensitic

(blue symbols; at 243 K) and the austenitic (red symbols; at 393 K) state of both samples. As

explained in the previous chapter, the martensite phase shows a significantly higher saturation

polarization JS than the austenite phase.

In Figure 85 b) the normalized magnetization J/J50 K as a function of temperature between

315 K and 500 K and under a constant magnetic flux density µ0H = 0.05 T is depicted for

both samples.

Page 138: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

126

Figure 85: a) Magnetic hysteresis measurement for the martensite (243 K) as well as for the austenite (393 K)

state of a Fe71Pd27.7Cu1.3 (open circles) and a Fe70.2Pd25.5Cu4.3 (filled squares) thin film. b) Normalized

magnetization J/J50 K as a function of temperature and under a constant magnetic flux density of µ0H = 0.05 T is

presented. The data points are fitted using Kuz’min’s equation for the Fe71Pd27.7Cu1.3 (open circles and red curve)

and the Fe70.2Pd25.5Cu4.3 (filled squares and blue curve) thin film.

Again using the Kuz’min relation, a fit function (red: Fe71Pd27.7Cu1.3; blue: Fe70.2Pd25.5Cu4.3)

was applied through the data points and extrapolated to the temperature, where J/J50 K

approximates zero. The determined values for both JS and TC are shown in Figure 86. Figure

86 a) presents the saturation polarization as a function of Cu content (colour coding indicates

the Fe content) for both the martensitic (squares, determined at 243 K) and the austenitic

(triangles, determined at 393 K) structure.

Figure 86: a) Saturation polarization for both the martensite fct and austenite fcc phase determined at 243 K and

393 K respectively, in comparison to the calculated ground state total (spin + orbital) magnetic moments, for

selected compositions along the path marked in Figure 82 a). The straight line through the experimental data

illustrates the deviations from a linear behaviour. b) Normalized values of Curie temperature TC in dependence

on Cu content determined by calculation and experiments. (Values for JS and TC were calculated by M. E.

Gruner and originally published in Ref. 199)

Page 139: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

127

Error bars indicate a maximum variance < 5% of JS. Generally, the saturation polarization for

Fe-Pd-Cu samples is about 10% lower than for Fe70Pd30 (1.5 T for martensite and 1.35 T for

austenite phase). The higher JS of the martensite phase compared to the austenite reflects the

expected decrease of spontaneous magnetization with increased thermal fluctuations. For both

martensite and austenite, JS has a general tendency to decrease with increasing Cu content,

which is in agreement with the composition dependence of the ground state magnetic

moments obtained from ab initio calculations (performed by M. E. Gruner). However, for Cu

contents from 5 to 6 at.% the experimental JS values increase abruptly to a maximum of 1.1 T

for the austenite and 1.35 T for the martensite phase. Although a direct evidence of large

precipitates was not observed with XRD, the rather unexpected increase of JS indicates that

the formation of small precipitates below the sensitivity of XRD already starts at 5 at.% Cu.

Below 5 at.% the homogeneous solution of Cu, which has a filled d-shell and thus shows only

a negligible induced spin-polarization on the order of 0.08 µB, decreases magnetization. At Cu

contents between 5 and 6 at.%, decomposition of the Fe-Pd-Cu sample into a Fe70Pd30 phase

with a high JS and Cu-rich precipitate phases results in the overall increase of magnetization.

When the Cu content increases to values > 6 at.% the volume fraction of non-magnetic

precipitates increases and thus leads to a decrease in JS. This agrees well with observations for

the Fe-Pd-Cu splats with a Cu content > 5 at.% as shown in the following. Figure 86 b) shows

the experimentally determined and calculated Curie temperatures, TC, as a function of the Cu

content, which are in both cases presented relative to TC of binary Fe70Pd30. Experimentally,

TC was extrapolated from a fit of the temperature dependence of saturation curves to

Kuz’min’s parameterization.151,152 Because accurate results cannot be achieved by a

simultaneous fit of the shape parameters s and p and the critical exponent of the

magnetization β, while determining the spontaneous magnetization M0 and TC by

extrapolating the magnetization curve, the values were restricted to s = 1, p = 5/2 and β = 1/3

as determined by Buschbeck et al..67 The choice of a large value s=1 (also valid for other 3d

transition metals) is consistent with the largely decreased spin-wave stiffness of Fe-Pd alloys

reported in the vicinity of the fcc-fct transition.200 It should also be noted, that the calculated

TC values are systematically too large on an absolute scale, since M. E. Gruner used the

mean-field approximation to the Heisenberg-model, which furthermore neglects Invar-typical

longitudinal spin fluctuations. Nevertheless, both approaches show the same trend - a slight

decrease as a function of composition for Cu-concentrations < 4 at.%. This is again expected

from the dilution of the magnetic sites by the Cu-atoms, but it obviously does not lead to a

severe degradation of the magnetic properties. Between 4 and 5 at.% Cu, a kink to lower

Page 140: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

128

TC/TC(Fe70Pd30) ratios is observed in the experimental data - similar to the anomaly in the

magnetization presented in Figure 86 b). This kink is again attributed to the decomposition of

the single ternary phase into a multiphase structure for Cu > 5 at.%. The highest TC of 560 K

for a single phase is observed at a composition with 72 at.% Fe. This value is slightly lower

compared to 600 K for Fe70Pd30.123 Samples having a Cu content of 5 at.% exhibit a smaller

TC - but following the previous arguments this is very likely attributed to the formation of a

complex phase mixture. In addition, we expect uncertainties originating from slight changes

of the composition due to the limited accuracy of the EDX measurements and the formation

of interfacial layers.

Decisive for a large MFIS is a strong coupling between the orientation of magnetic moments

and the tetragonal distortion of the martensitic twins, which usually arises from spin-orbit

coupling. This can be quantified in terms of the magnetocrystalline anisotropy energy, which

is connected to the change of orbital moments upon variation of the magnetization direction.

These quantities are difficult to measure in a high-throughput approach and therefore further

experimental characterization of orbital magnetism of the Fe-Pd-Cu system must be left for

future work. Fully relativistic first-principles calculations as employed in this investigation,

however, provide a straight forward approach to grasp the principal trends of related

quantities, such as orbital moments in dependence on composition. Figure 87 provides the

variation of the total orbital moment per atom for all compositions of interest. The obtained

variations are small and at the limits of the methodological resolution. Nevertheless, a few

discernable trends are discussed below.

Figure 87: Total orbital moments µorb of Fe-Pd-Cu as a function of composition obtained from first-principles

calculations. Only moderate variations of µorb are encountered within the martensitically transforming

concentration range. (Calculated by M. E. Gruner and originally published in Ref. 199)

Page 141: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

129

The total orbital moment of the binary reference composition Fe70Pd30 is with 0.054 µB/atom

among the largest value for the considered concentration range and thus it could be concluded

that addition of Cu does not significantly affect the magnetocrystalline anisotropy of the

material. Slightly larger values are only obtained for compositions with an Fe content of about

72 at.% and low Cu content. The Fe-species provide the largest orbital moment of about 0.065

µB/atom with an overall variation on the order of 1%. Thus, changes of its fractional

contribution according to composition will provide the most relevant trend. The Pd-atoms

exhibit an orbital moment of about 0.025 µB/atom (with a variation of 5% over all

compositions), which is proportional to the magnitude of the Pd-spin moment. The latter, in

turn, is induced by the surrounding Fe moments. This raises the expectation that a large Fe

content should have a beneficial influence on the magnetocrystalline anisotropy, since it is

commonly traced back to the hybridization of the 3d and 4d electrons in this alloy. Cu

exhibits only a small induced moment, which is connected with a rather tiny orbital

contribution of the order of 0.01 µB/atom.

Stress effects in Fe-Pd-Cu thin films

In comparison to bulk samples, thin films exhibit considerably higher transformation

temperatures, regardless of their composition. This can partly be attributed to the high stress

state of the films due to the different thermal expansion coefficients of thin film and substrate.

It is known that Fe-Pd FSMA thin films on SiO2 and MgO substrates are subject to high

tensile stresses.201,202 These lead to stress-induced martensitic transformations at

systematically increased transformation temperatures. Kato et al.22 estimated a proportionality

factor of 4.8 MPa/K for the Fe-Pd system. From other experiments on Fe70Pd30 thin films

fabricated and processed in a similar way as in the present study, tensile stresses in the range

of 0.2 GPa to 0.3 MPa after annealing were observed.61 These values for tensile stress are

sensitive to the composition and decrease systematically with increasing Fe content.203 This

behaviour was attributed to the Invar effect, which is defiend as the anomalous reduction of

thermal expansion, and is also present in Fe-Pd alloys in this composition range. The Invar

effect originates from a thermodynamic repopulation of electronic configurations, which are

characterized by different local magnetic moments and equilibrium volumes. This effect

becomes increasingly prominent for Fe-rich compositions204 and may even lead to negative

thermal expansion coefficients. The Invar properties of the ternary alloy are - apart from the

changing Fe content - also influenced by Cu addition. As measurements of thermal expansion

coefficients are difficult in film/substrate composites, the spontaneous volume

Page 142: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

130

magnetostriction ωs0 was calculated by M. E. Gruner to give a qualitative indication. The

quantity ωs0 is defined as the relative change (VFM-VPM)/VPM between the volume VFM of the

ferromagnetic phase and the volume VPM of the paramagnetic phase, both extrapolated to T =

0. Alternatively, ωs0 can be interpreted as the relative volume expansion induced by the onset

of ferromagnetism below TC, which is measured in comparison to a (hypothetical) reference

alloy which remains paramagnetic at all temperatures.

Figure 88 shows the variation of ωs0 as a function of the composition, demonstrating an

increase of up to 10% with Fe concentration as the dominating trend in the transforming

composition range. As a secondary trend, at constant Fe content, an increase of ωs0 with

increasing Cu concentration is observed, which is opposite in direction with respect to the

corresponding valence electron concentration e/a. This clearly demonstrates that the moment-

volume interaction in the ternary system cannot be consistently interpreted as a function of e/a

alone. It further can be concluded that increasing Fe as well as Cu contents, will enhance the

Invar-typical anomalies and thus further reduce thermal expansion and elastic constants below

TC. Thus, the addition of Cu can be expected to further decrease the tensile stress compared to

a binary film with the same Fe content.

Figure 88: Calculated spontaneous volume magnetostriction ωs0 of Fe-Pd-Cu as a function of composition. A

clear increase of ωs0 of about 10% is encountered for increasing Fe concentrations. (Calculated by M. E. Gruner

and originally published in Ref. 199)

The increase of Invar-related anomalies and the decrease of thin film stress due to the addition

of Cu was further investigated by synchrotron-based measurements. Residual stress

measurements using the sin2(ψ) method were used to determine the stress state of Fe-Pd-Cu

Page 143: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

131

thin films in the martensite (300 K) as well as in the austenite (393 K) state. Figure 89 shows

area detector diffractograms at two different temperatures of a Fe71Pd26Cu3 thin film.

Figure 89: Area detector diffractograms of a Fe71Pd26Cu3 thin film at different temperatures. a) At 300 K the

thin film is in the martensite phase proved by the presence of the (200) and (002) fct peaks. When heating the

sample to 393 K in b), the sample transforms into the austenite phase and the (200) fcc peak appears. The red

stars denote diffraction intensities originated from the Si/SiO2 substrate. (Synchrotron measurements were

performed by H. Brunken)

Page 144: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

132

The area diffractograms depict the colour-coded intensity (blue for low and white for high

intensity) in terms of X-ray counts as a function of d-spacing and ψ angle. At an angle of ψ =

90° the ψ-vector points to the thin film surface at an angle of 90° and thus is parallel to the

thin film normal vector. This is indicated by a dashed white line, that divides the upper half

(90°-160°) to one half-sphere and the lower part (20°-90°) to the other half-sphere above the

sample. Instead of intensity rings known for polycrystalline samples intensity spots appear.

These spots originate from texturing and correspond to the specific d-spacings of different

phases. In Figure 89 a) the synchrotron diffraction pattern of a Fe71Pd26Cu3 thin film in the

martensite phase at 300 K is presented. The martensite phase can be identified by the presence

of the (220), (002), (200) and (111) fct diffraction peaks. Upon heating to 393 K, the thin film

transforms into the austenite phase and develops a (200) fcc peak. In order to determine the

residual stress state of this sample in the austenite phase, the lattice parameter afcc of the thin

film was determined from the (111) fcc peak at different ψ-angles.

Figure 90: Lattice parameter as a function the ψ-angle for a Fe71Pd26Cu3 thin film. a) A linear function was

fitted to the data. The positive slope indicates a tensile strain state in this sample. For every lattice spacing afcc in

a) that corresponds to a diffraction peak, the FWHM b) and the intensity c) for this peak were determined.

An example of the variation of lattice parameter with tilt angle afcc = sin2(ψ) is shown in

Figure 90 a). From the slope of this curve the stress state and the strain in this thin film can be

Page 145: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

133

determined. The data points are fitted by a linear function and the slope of this fit function is

used to determine the residual strain. The positive slope indicates a tensile residual strain that

appears for thin films showing Invar-related anomalies having smaller thermal expansion

coefficient than their substrate. In Figure 90 b) the FWHM and c) the intensity of the (111)

fcc diffraction peak at different ψ angles and thus as a function of sin2(ψ) is shown to ensure

no texture related errors. Since the sin2(ψ) method only reveals the strain of the sample, the

Young’s modulus needs to be determined separately for each thin film. This was performed

by nanoindentation measurements of the Fe-Pd-Cu thin films. The measurements were

conducted at ambient and elevated temperatures (353 K) to determine the Young’s modulus

in the austenite phase. A summary of nanoindentation measurements is depicted in Figure 91.

For these measurements the Young’s modulus was determined for an indentation depth

ranging from 50 nm to 75 nm. This indentation depth follows the rule that up to an

indentation depth of one-tenth of the overall film thickness, the influence of the substrate is

negligible.154 Presented are the a) Young’s modulus and b) the thin film hardness as a

function of Pd content with colour-coded Cu content at 353 K. Figure 91 b) shows hardnesses

values of the thin films revealed from nanoindentation measurements in dependence on Pd

content. Since the error bars in both measurements are quite large, no significant trend can be

observed. As mentioned previously, the amount of Cu increases at the expense of the Pd

content in the Fe-Pd-Cu system. Since Pd and Cu have similar Young’s moduli (EPd = 121

GPa and ECu = 117 GPa) and number of valence electrons (eCu = 11 and ePd = 10) it can be

expected that the Young’s modulus of the Fe-Pd-Cu single-phase does not vary significantly

with changing Cu content. The same behaviour can be expected for the variation of hardness

with composition. All values for the Young’s modulus are higher than the values determined

for a binary Fe70Pd30 reference thin films that was fabricated and processed under identical

conditions. The values determined for the Young’s modulus in Figure 91 a) were used to

calculate the residual stress in the Fe-Pd-Cu single-phase thin films. Figure 92 depicts the

residual stress as a function of the Cu content for samples with a constant Fe content of 71

at.%. The colour-coding indicates the martensite start temperature Ms. A significant decrease

in residual tensile stress is observed with increasing amount of Cu. The values for low Cu

content in the range of 1 at.% exhibit high tensile stress values, while Ms is in the range of

343 K to 353 K. Upon further increase of Cu content, the amount of tensile stress lowers

significantly while Ms increases.

Page 146: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

134

Figure 91: Summary of nanoindentation measurements at 353 K for Fe-Pd-Cu single-phase thin films. Presented

are the a) Young’s modulus and the b) thin films hardness as a function of Pd content with colour-coded Cu

content. No distinct trend can be observed with changing composition. Further binary Fe70Pd30 reference thin

films that were fabricated and processed under identical conditions are included in both diagrams.

At a Cu content of 1.6 at.% the amount of tensile stress is lowered to 673 MPa. Thin films

with a Cu content > 1.6 at.% exhibit Ms temperatures > 350 K. The data point with a Cu

content of 3.9 at.% has a tensile stress amount of 667 MPa and shows a significant decrease in

Ms. This decrease can be attributed to the fact that this sample has a Cu content that is close to

the limit (Cu > 4 at.%) where the Fe-Pd-Cu single-phase starts to decompose by forming Cu-

rich precipitates. It cannot be ruled out that this thin film already shows a tendency on the

nanoscale to decompose. However, no hints regarding this were observed.

Page 147: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

135

Figure 92: Residual stress as a function of Cu content. To separate the impact of Fe on thermal expansion

behaviour and thus the stress state, only thin film samples with a constant Fe content of 71 at.% are shown. Upon

addition of Cu a significant decrease of tensile stress is observed. The Ms temperature (colour-coded) does not

change significantly and has highest values for the samples with lowest tensile stress.

Considering these findings, it can be concluded that the addition of Cu decreases the thin film

residual stress state. The Ms temperature does not drop for the thin films with smallest

amounts of tensile stress, indicating that the increase in transformation temperatures upon

addition of Cu in Fe-Pd is not controlled by the formation of stress-induced martensite. The

presented tensile stress values are high, when compared to literature values for binary Fe-Pd

thin films.61 Nevertheless, the considerable increase of volume magnetostriction revealed

from ab initio calculations and the decrease of thermal expansion with increasing Cu content

is a good explanation for these high tensile stress values. Further investigations of the thermal

expansion coefficient of these samples would be interesting. This is not possible with these

samples, since the substrate constrains the thermal expansion expansion of the film.

4.2.3.2 Fe-Pd-Cu Splats

Substrate constraints are not present in bulk-like samples. Therefore it is desirable to cross-

check the findings for corresponding systems. In this part Fe-Pd-Cu splats were fabricated,

processed and investigated. All splats were fabricated and annealed by I. Kock from the

Georg-August-Universität Göttingen. The samples were produced by the splat-quenching

technique to obtain thin foils in a metastable phase and to avoid demixing during cooling. It is

known that a high cooling rate is beneficial for the formation of the fct phase.52 The splats

Page 148: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

136

were annealed to improve their microstructure and transformation behaviour. The influence of

quenching and annealing on binary Fe70Pd30 splats was described in Chapter 4.1.2. Figure 93

presents room temperature XRD data of different splats with the nominal compositions of

Fe70Pd30, Fe68.4Pd29.3Cu2.3 and Fe70Pd23.8Cu6.2. All splats were annealed at the same

temperature and time (1123 K, 1 h) as used for the thin film materials libraries. In order to

investigate the stability of the metastable phase in the Fe-Pd-Cu splats, cooling rates were

varied and compared with the results from Fe70Pd30 splats. For rapid cooling (> 100 K/s), the

splats, which were sealed in quartz tubes under inert gas, were water-quenched. For slow

cooling, the splats were cooled in air (< 0.5 K/s), while for very slow cooling the splats rested

in the furnace after annealing until room temperature was reached (cooling rate < 0.2 K/s,

temperature after 12 h ~300 K). The binary Fe70Pd30 splats show the transforming Fe70Pd30

phase for both the quenched and air cooled samples, indicated by the (111) and (200) fcc

peaks. No additional precipitate phases are observed. The oven-cooled Fe70Pd30 splat shows

fractions of the Fe50Pd50 and α-Fe phase besides the transforming Fe70Pd30 phase. The

Fe68.4Pd29.3Cu2.3 splats, that were water-quenched and air-cooled, show a single-phase

structure in Figure 93.

Figure 93: XRD patterns of Fe70Pd30, Fe68.4Pd29.3Cu2.3 and Fe70Pd23.8Cu6.2 splats processed with different cooling rates after annealing (Quenched in water, air-cooled and oven-cooled). (Splats were fabricated and annealed by I. Kock and originally published in Ref. 199)

Two additional peaks from Fe50Pd50 and Fe8Cu2 phases occur only for the very slowly cooled

splats, with the highest fraction occurring in the Fe70Pd30 sample. Comparing the (111) fcc

and (111) Fe50Pd50 peaks for the oven-cooled Fe70Pd30 to the Fe68.4Pd29.3Cu2.3 splat, a

Page 149: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

137

significant decrease in the (111) fcc peak intensity (and corresponding increase in (111)

Fe50Pd50 peak intensity) is observed. Thus a high fractional decomposition from the

transforming Fe70Pd30 phase into the Fe50Pd50 equilibrium phase with precipitates occurs for

the binary Fe70Pd30 splat upon very slow cooling. By addition of Cu, this decomposition can

be partially suppressed. In contrast, the water-quenched Fe70Pd23.8Cu6.2 splat does not show

the single transforming phase: instead the (111) fcc/fct peak is shifted to higher angles and

shows a broad peak at 42.9° which is correlated to the (110) bcc phase. For the oven-cooled

sample, a further decomposition into precipitate phases (Fe50Pd50 and Fe8Cu2) occurs. Thus

the decomposition of the high-temperature and metastable phase to equilibrium phases is

suppressed by the addition of small amounts of Cu into binary Fe-Pd FSMA even for slow

cooling rates. In contrast, Fe70Pd23.8Cu6.2 splats do not show a single ternary phase due to

decomposition upon cooling.

Figure 94 a) shows the change of lattice constants as a function of temperature during the

martensitic transformation for a water-quenched Fe68.4Pd29.3Cu2.3 splat. Lattice constants were

determined from XRD(T) measurements by peak fitting. The error for these results is smaller

than the size of the symbols presented in Figure 94 a). From 128 K (cfct = 0.360 ± 0.001 nm,

afct = 0.382± 0.001 nm) to 233 K, only the low temperature tetragonal martensite (fct) phase

occurs. Within this range, an increase of c and a decrease of a is observed as the structure

comes close to the transformation temperature. At 230 K the transition from the martensite to

austenite phase starts, along with the appearance of (200), (220) and (311) fcc peaks (not

shown here). With further increasing temperature, the (200), (002), (202) and (311) martensite

peaks decrease and finally disappear at 238 K. When the temperature increases further a

growing of the (200), (220) and (311) fcc austenite peaks is observed, which finishes at 278

K. At 300 K the lattice constant for the austenite phase was determined to be afcc = 0.374 ±

0.001 nm. In Figure 94 b) the linear thermal expansion coefficient as a function of Cu content

for splat samples with different compositions is presented. The colour-code indicates the Fe

content of the splats. The linear thermal expansion coefficient for every Fe-Pd-Cu splat was

calculated from XRD(T) measurements. The variation of lattice parameters afcc of the

austenite phase was determined as a function of temperature between 220 K and 400 K and

fitted by using a linear function.

Page 150: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

138

Figure 94: a) Change of lattice constants by the transformation from the low temperature (fct) to the high

temperature phase (fcc), determined by XRD(T) for a water-quenched Fe68.4Pd29.3Cu2.3 splat. b) Linear thermal

expansion coefficients as a function of Cu content for splat samples with different compositions. Colour-coding

indicates the amount of Fe content in the different splats. The addition of Cu produces a significant decrease of

the linear thermal expansion coefficient. (Figure a) originally published in Ref. 199)

The error for the linear thermal expansion coefficient is smaller than the symbol size in Figure

94 b). Starting from a Fe70Pd30 reference splat that was fabricated and processed identically

and is in good accordance with literature values, the thermal expansion coefficient decreases

upon addition of Cu for a Fe68.6Pd29.8Cu1.6 splat. To separate the impact of Fe on the variation

of the thermal expansion the Pd content is kept constant at 29.8 at.% for the following Fe-Pd-

Cu splats. Although the Fe content is decreasing from 68.6 at.% to 67.9 at.% the thermal

expansion coefficient further decreases to a value of 1.4 x 10-6 K-1, which is well below the

value for the Fe70Pd30 splat. These findings further corroborate the previously presented

findings for Fe-Pd-Cu thin films.

Page 151: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

139

At 128 K the Fe68.4Pd29.3Cu2.3 splat has a tetragonality of c/a = 0.942, which is higher than

reported for a Fe68.8Pd31.2 single crystal.37 Due to the change of the lattice constants for

Fe68.4Pd29.3Cu2.3, a shape change consisting of an expansion of (a128K - a300K)/a300K = 2.14%

along the a-axis and a contraction of (c128K - a300K)/a300K = 3.74% can be estimated. Thus a

maximum strain caused by conversion of the variants upon cooling from 300 K to 128 K of

approximately 5.88% can estimated. This is slightly lower than the value of about 6% known

for the Fe68.8Pd31.2 single crystal. Since these investigations were performed on polycrystalline

samples, where grain boundaries hinder magnetic field induced strain and the influence of

micro-stress on the transformational behaviour was not investigated, these values just give an

upper estimate for MFIS in Fe68.4Pd29.3Cu2.3.

Variation of martensitic transformation in the Fe-Pd-Cu system

Fe-Pd-Cu thin film materials libraries were fabricated covering a range of 40 < Fe < 95 at.%,

5 < Pd < 55 at.% and 1 < Cu < 18 at.% and exhibiting transformation temperatures of up to

359 K. TEM confirmed the presence of a single ternary phase and showed the occurrence of

nano-twins. XRD investigation of Fe68.4Pd29.3Cu2.3 splats, processed with different cooling

rates after annealing, showed an enhanced stability against decomposition of the high-

temperature metastable phase in comparison to Fe70Pd30. Ab initio calculations and

experiments reveal the same trends for TC and JS, which are not significantly affected by the

dilution of the magnetic sites in the ternary alloy. Slightly decreased orbital moments and an

enhanced volume magnetostriction are predicted from ab initio calculations for Cu-enriched

compositions.

The presented findings deliver a comprehensive overview of the compositional dependence of

the martensitic transformation temperatures. The complete analysis, however, must consider

that a systematic influence by substrate induced-stresses may be present. On the other hand,

the fabrication method used for bulk-like splats is not feasible for producing the large amount

of data necessary to reveal the relevant trends. Therefore, it is worthwhile to combine the

results of both approaches with previous results from literature for the binary Fe-Pd system.

This has been done in Figure 95 a), which summarizes the martensite start temperature, Ms, as

a function of the composition in a three-dimensional diagram. The composition is defined in

terms of the valence electron concentration e/a (see Figure 80 for the whole materials library)

and Fe content. In the introduction, these quantities were identified as the two most influential

parameters determining the physics of the martensitic transformation in the ternary alloy.

Again, the full three-dimensional representation (Figure 95 b)) is difficult to interpret. Thus,

Page 152: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

140

the discussion is restricted to projections of the data onto each of the three base planes, which

cover the essential physics.

Figure 95: Martensite start temperatures, Ms, as a function of e/a ratio and Fe content presented in a 3D graph

with colour-coded Cu content. Binary systems are shown by black symbols (squares and triangles). For clarity,

only projections of the 3-dimensional data onto three different planes are depicted in a); the spatial distribution

of the data is given in b). (Circles refer to the Ms obtained from thin film experiments in this study, diamonds

refer to values obtained for splats. For comparison, literature data for binary alloys57,143 are included (black open

squares and black filled triangles). Black lines are guides to the eyes and visualize general compositional trends

for the three kinds of data. Red dotted lines mark the border lines beyond which Cu-rich ternary thin films

decompose into different phases. Outside of this volume, only data points corresponding to single-phase samples

are presented. (Figure originally published in Ref. 199)

These are shown exclusively in Figure 95 a). The bulk samples (69 at.% < Fe < 71 at.% and

8.575 < e/a < 8.613), published by Cui et al.57 show a linear distribution within all three

planes, depicted by a black dashed line. For the Fe-Pd-Cu thin film martensite start

temperatures, a clear linear dependence on both the Fe content and the e/a ratio is observed.

The slope for these ternary thin films is considerably lower than for the binary reference data.

As discussed before, this can be traced back to the tensile stresses, which are largest for the

samples with the smallest Fe content and decrease significantly with increasing Fe content.

Tensile stresses can also be partially responsible for the considerable increase of the

transformation temperatures with respect to bulk material. Splats as bulk references show

much lower transformation temperatures in comparison to Fe-Pd-Cu thin films. For

Page 153: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

141

Fe68.4Pd29.3Cu2.3 MS is 334 K for the thin film and 238 K for the splat, while for the binary

composition, MS of splat and bulk systems coincide. If the observed difference of Ms between

film and splat is solely attributed to the stress state of the film, we obtain a value in the range

of 460 MPa using the factor of 4.8 MPa/K determined by Kato et al..22 A maximum Ms of 359

K was observed for a sample with an Fe content of 71.8 at.% and an e/a ratio of 8.58. This is

one of the highest value obtained in the Fe-Pd system so far and significantly higher than the

values obtained for corresponding binary systems,143 demonstrating that the addition of Cu

can substantially improve the relevant FSMA properties. This trend cannot be ascribed to a

stress-related variation of the martensitic transformation temperature, since, as shown before,

the Invar-related reduction of thermal expansion, which partially compensates tensile stresses,

is in fact enhanced in the Cu-doped alloys. The clear linear dependence between Ms and the

compositional parameters vanishes if the Fe content reaches a threshold value (indicated in

Figure 95 by red dotted lines), below which no transforming single component samples were

obtained. Above the critical concentration, the stability region is depicted by a triangle

(marked by the broken black lines on the bottom plane of Figure 95). In comparison to results

concerning binary bulk samples57 and thin films143 the compositional range in which single-

phase transforming samples can be found is significantly extended to lower Fe contents. From

the discussion of the binary alloys and ternary thin films alone, it cannot be decided whether

the Fe content or the valence electron concentration represents the dominating variable which

needs to be optimized in order to improve the FSMA behaviour in a ternary system. However,

a conclusive indication comes from the splat experiments. Here, the Ms value of the ternary

splats coincides with the extrapolation of the data from the work of Cui et al.,57 if plotted

versus the Fe content, while it deviates significantly for larger e/a values from the respective

projection of the bulk values. This interpretation is also supported by the thin film data of

Inoue et al..143 These, however, do not show a well-defined linear behaviour. This may be

related to the variation of the fabrication parameters and the changing film thicknesses of the

samples. Nevertheless, in the range between 70.5 and 71.5 at % Fe, the Ms values of the

binary thin films (film thickness 200 nm) fall almost on top of Fe-Pd-Cu data if plotted

against the Fe content, but deviates for the e/a projection. The films with low Fe content agree

better with the bulk curve because of the higher film thickness (2 and 4 µm).

Based on these observations, it can be concluded, that the addition of Cu is a very promising

route to develop improved Fe-Pd-based FMSAs. Cu extends the stability of the

thermoelastically transforming phase to Fe concentrations up to 72 at.% and therefore is

Page 154: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

142

expected to increase the martensitic transformation temperatures to values which are suitable

for technological applications.

4.2.3.3 Epitaxial Fe-Pd-Cu thin films

This chapter presents the fabrication of epitaxial Fe70Pd30-XCuX (X = 3, 7 at.%) thin film

samples and investigates the influence of Cu alloying on the structural and magnetic

properties. The motivation for this was to grasp the idea of coherent epitaxially grown Fe-Pd

thin films on varying buffer layers from Buschbeck et al.67 and S. Kauffmann-Weiss et al.206

and to apply this to new Fe-Pd-based FSMAs found by combinatorial materials science. One

aim was to gain a deeper insight and to further investigate the new Fe-Pd-Cu system with

regard to properties, which are hardly to measure in polycrystalline thin films. Most

promising Fe-Pd-Cu compositions were chosen by the author and fabricated by using

coherent epitaxial growth.

All epitaxial thin film were fabricated and partially investigated at the IFW Dresden in close

collaboration with S. Kauffmann-Weiss in the group of S. Fähler during a scientific stay of

the author. Further all presented findings were developed in close collaboration with

S. Kaufmann-Weiss. The calculated results shown in this chapter were carried out by M. E.

Gruner in the group of Prof. P. Entel from the University of Duisburg-Essen. Findings from

calculated results were further developed in close collaboration with M. E. Gruner.

Single crystals are well suited to determine materials properties such as the

magnetocrystalline anisotropy, which can be affected by finite size, texture and stress effects.

These are difficult to measure in polycrystalline samples. Nevertheless single crystals are

difficult to fabricate and often expensive as large amounts of the material are needed.

Epitaxial films represent the thin-film counterpart to bulk single crystals and allow

determination of the anisotropic magnetic properties. Furthermore, coherent epitaxial growth

on cubic buffer layers with different lattice parameters can be used to adjust the tetragonal

distortion of the martensite as explained in Chapter 2.6. This approach allows the c/abct ratio

to be controlled almost along the entire Bain transformation path between bcc (c/abct = 1) and

fcc (c/abct = 1.41). Through this kind of epitaxial film growth, artificial single-variant states

are realized, allowing K to be measured along the different crystallographic directions (K1 and

K3).67 Coherent epitaxial growth causes the lattice constants of the buffer to alter the in-plane

lattice parameters of Fe70Pd30-XCuX films and thus c/abct ratio. The c/abct was controlled by

depositing Fe70Pd30-XCuX films on Cr (1.07), Au (1.09), Ir (1.31), Rh (1.34) and Cu (1.57)

buffer layers that were previously deposited on MgO substrates. The intended structural

Page 155: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

143

orientation relation between substrate, buffer and Fe70Pd30-XCuX thin film is schematically

depicted in Figure 96 for each different buffer material.

Figure 96: Structural orientation relation between MgO substrate, buffer materials and Fe70Pd30-XCuX thin film.

Next to a) Cr as buffer material, all following buffers b) Au, c) Ir, d) Rh and e) Cu have a Cr adhesion layer

below. With varying buffer material the c/abct ratio increases as illustrated by a tetragonal distortion of the

Fe70Pd30-XCuX unit cell.

In order to analyse the crystal structure of Fe70Pd30-XCuX films grown on different buffer

layers, XRD scans in Bragg-Brentano geometry were performed (Figure 97). This method is

only suitable to reveal the out-of-plane lattice parameter. With changing buffer material from

Au to Cu, the position of the (002)fcc diffraction peak of the fcc buffers (blue pentagons)

increases to higher 2θ angles. By using a bcc buffer, such as Cr, the peak position of the (002)

diffraction peak occurs at considerably higher 2θ angles.

This reflects the shift of the XRD peak with the lattice parameter for several buffer layers

with different structures. When varying the buffer material, the position of the (002)

diffraction peak of Fe70Pd27Cu3 (red circles) shifts from 2θ = 70.5° (c/abct = 1.07) to a lower

Page 156: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

144

value at 2θ = 60.6° (c/abct = 1.31). For a film on the Cu buffer (c/abct = 1.57), the (002) peak

shifts to 2θ = 52.6°.

Figure 97: Bragg-Brentano scans of Fe70Pd27Cu3 films on different buffer layers. Red circles mark (002)

diffraction peaks of Fe-Pd-Cu films and blue pentagons indicate (002) diffraction peaks of the cubic buffers.

With increasing c/abct ratio, a shift of (002) Fe-Pd-Cu peaks to lower angles is observed. Red dotted lines indicate

the boundaries of the Bain transformation path. (Figure originally published in Ref. 205)

These XRD measurements show the presence of the (002) diffraction peak only in the

Fe70Pd27Cu3 thin films, which indicates a highly textured microstructure. The absence of any

further diffraction peaks suggests a coherent epitaxial growth of the Fe70Pd27Cu3 films. Since

the unit cell volume remains almost constant, this results in an increase of the out-of-plane

lattice parameter due to the shrinking of the in-plane lattice parameter. The in-plane lattice

parameters were measured by conducting 2θ scans of the (101)bct lattice planes under tilted

conditions at a tilt angle ψ. The (101)bct diffraction peaks were then used together with the

measurements in Bragg-Brentano geometry to calculate the in-plane lattice parameter of the

films. The results revealed identical in-plane lattice parameters for the Fe70Pd27Cu3 films and

the underlying buffer material (Figure 98 a)). This indicates a coherent growth in all thin films

deposited on the different buffers and reveals the following relation: dbuffer = abct.

The c/abct ratio for each thin film sample was then calculated by using results from Bragg-

Brentano and tilted XRD measurements. This is shown in Figure 98 b), depicting the

dependency of the tetragonal deformation (c/abct ratio) for all buffer materials. This

deformation behaviour is similar to investigations for thin binary Fe70Pd30 films which are

added for comparison.67,206

Page 157: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

145

Figure 98: a) In-plane lattice parameters abct for Fe70Pd30 (rectangle) and Fe70Pd30-XCuX thin films with X = 3, 7

at.% (triangle) are fixed by the substrate’s lattice spacing d of different buffers (as marked at the top). b) The

change in tetragonal deformation (c/abct) with buffer material is smaller for the Fe70Pd30-XCuX system (triangle)

than for the binary Fe70Pd30 (rectangle67,206). The curves illustrate the expected change in deformation at constant

volume of the unit cell. (Figure originally published in Ref. 205)

In contrast to Fe-Pd films, the c/abct ratio of Fe70Pd27Cu3 is found to be slightly smaller,

suggesting that the addition of smaller atoms, such as Cu, reduces the volume of the Fe70Pd30

unit cell. This is in good accordance with results on lattice parameters determined for

polycrystalline Fe-Pd-Cu thin films and presented in Figure 83 in Chapter 4.2.3.1. The

presented thin-film samples with varying c/abct ratio allow investigation of the intrinsic

properties as a function of the tetragonal deformation. The variation of the c/abct ratio is used

in the following, next to the composition, as a parameter to control the magnetic properties.

For a detailed understanding of the anisotropic magnetic properties, the quality of coherent

epitaxial growth and the orientation relationship of thin film unit cells and the underlying

buffer material have to be determined.

The structural quality of the thin film samples was confirmed by pole figure measurements on

the (101)bct diffraction plane. Due to the 4-fold surface symmetry of the (001)-oriented MgO

substrate, only one quadrant of the pole figure measurement for Fe70Pd27Cu3 films on

different buffers is depicted in Figure 99 a).

Page 158: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

146

Figure 99: a) (101) bct pole figures of Fe70Pd27Cu3 films on different buffer layers. With increasing c/abct ratio a

shift of the (101) pole towards higher ψ angles is observed. b) Drawing illustrating the geometric relation

between the tetragonal distortion (c/abct ratio) and the tilt angle ψ within a bct unit cell. (Pole figure

measurements were conducted together with S. Kaufmann-Weiss and the figure was originally published in Ref.

205)

The MgO [100] edges are oriented parallel to the edges of the figure. This reveals an

orientation relationship of the bct unit cell of the film and MgO (001) substrate as planned in

Figure 96 and evidences the successful coherent epitaxial growth of Fe-Pd-Cu thin films:

Fe70Pd30-XCuX(001)[110] || fccbuffer(001)[100] || Cr(001)[110] || MgO(001)[100] or

Fe70Pd30-XCuX(001)[110] || bccbuffer(001)[110] || MgO(001)[100].

The corresponding pole figures in Figure 99, illustrated by using custom-made software207,

indicate a high quality epitaxial growth of the films since well-defined and sharp peaks were

observed. The unit cells are rotated by 45° with respect to each other - a shift in ϕ direction of

about 45° - to reduce misfits between different layer materials. These pole figure

measurements can be further used to determine the c/abct ratio, since the tilt angle ψ of the

(101)bct plane is connected to the tetragonal distortion by tan ψ = c/abct (Figure 99 b)). This

relation predicts a tilt angle of the (101) pole at ψ ≈ 45° for a bcc structure (c/abct = 1) and at

ψ ≈ 54° for a fcc structure (c/abct = 1.41). When the buffer material is varied, the ψ angle shift

Page 159: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

147

of the poles reflects the structural changes from bcc towards fcc structure. For a film on Cr

buffer the (101) pole is at a tilt angle ψ ≈ 47°. This confirms an almost cubic structure with

c/abct = 1.07. However, an Ir buffer layer shifts ψ to ≈ 52°, giving a c/abct ratio of 1.31 which

corresponds to a fct structure. The pole-figure measurement is therefore a second independent

measurement confirming the c/abct ratios obtained from θ-2θ scans with accuracy in the range

of 1%. A further increased tetragonal distortion is visible for the film on a Cu buffer

(c/abct = 1.57). In Bragg-Brentano scans (Figure 97) the (002) diffraction peak of the

Fe70Pd27Cu3 thin film is observed beyond the limits of the Bain path. This tetragonal

distortion agrees well with c/abct = 1.57 derived from the position of the pole in the (101)bct

pole figure at ψ ≈ 57°. The intensity in this measurement further exhibits a substantial

broadening and splitting in the ϕ direction. This observation agrees well with a report on

highly strained binary Fe70Pd30 films on Cu buffer layers.206 The broadening can be ascribed

to adaptive nanotwinning, which reduces the huge elastic energy induced by the coherent

epitaxial strain through the formation of twin boundaries. The rotation of the tetragonal

variants that form (101)fct twin boundaries leads to the small observed additional intensities.

The same set of XRD measurements was performed for a series of Fe70Pd23Cu7 thin films

with increased Cu content. These measurements are not shown here, since they are practically

identical to the ones obtained for lower Cu content. Concluding these results, no Cu-rich

precipitates occur even at 7 at.% Cu in epitaxially grown films. This is in contradiction to

polycrystalline annealed Fe-Pd-Cu thin films which decompose at Cu contents above 4 at.%.

This behaviour can be attributed to the difference in fabrication process routes. Epitaxial films

are deposited at room temperature without any annealing treatment, avoiding precipitation

compared to annealed films.

The lattice parameters determined for Fe70Pd23Cu7 are equal to those of Fe70Pd27Cu3. No

change in tetragonal deformation with increased Cu content is observed. Accordingly the

symbols in Figure 98 depict both the Fe70Pd27Cu3 and the Fe70Pd22Cu7 composition.

Previously presented experiments on polycrystalline annealed films have shown a decrease of

lattice constant with increasing Cu content, since the atomic radius for Cu (0.128 nm) is

smaller than for Pd (0.137 nm). However, it is known that alloys such as Fe-Pd, which exhibit

the Invar effect, deviate from the rule of mixture of Vegard’s law. This has to be considered

for the Fe-Pd and the second binary Cu-Pd systems.208

Epitaxial Fe70Pd30-XCuX thin films show improved growth behaviour when compared to

binary Fe70Pd30 films. In order to avoid unfavourable relaxation mechanisms, such as (111)fcc

deformation twinning,201 very low deposition rates of 0.024 nm/s were required for epitaxial

Page 160: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

148

growth of binary Fe70Pd30 films.182 The present Fe-Pd-Cu films, were grown under a

deposition rate one order of magnitude higher (0.3 nm/s). Nevertheless, the poles in the (101)

pole figure and the diffraction peaks in the Bragg-Brentano measurement reveal a small full

width at half maximum. By using a modified Scherrer equation134, the determined coherence

length is in the range of the overall thickness of the Fe-Pd-Cu films, confirming a high quality

epitaxial growth. This enhancement of crystal growth can be understood in terms of the

structural variation of the total energy.

Total energy in epitaxially grown Fe-Pd-Cu films

Calculations of the total energy were performed by M. E. Gruner from the University of

Duisburg-Essen. These simulations were obtained from first principles within a supercell

description involving a full relaxation of the atomic position which minimizes the interatomic

forces for each tetragonal stage. The disordered arrangement was realized using a 500 atom

supercell and a random distribution of the 340 Fe atoms and 160 Pd atoms, or respectively,

135 Pd and 25 Cu atoms. In order to avoid statistical uncertainties, which hamper a direct

comparison, 25 randomly chosen Pd atoms were exchanged by Cu to model a comparable

ternary distribution, while the remaining elements were left untouched. In addition,

composition-dependent and finite-temperature magnetic properties using the Korringa-Kohn-

Rostoker (KKR) approach as implemented in the Munich SPR-KKR code (version 5.4) were

used.192 Further computational details can be found in Ref. 65 and 190. Here, the disordered

nature of the Fe-Pd alloy was modelled in terms of averaging the electronic scattering

properties within the coherent potential approximation (CPA). The CPA enables the economic

use of small cells but does not allow for structural relaxations. The magnetic exchange

parameters were determined for use within a Heisenberg model following the approach by

Liechstenstein et al.197, starting from the ferromagnetic state. The corresponding Curie

temperatures were estimated within the mean-field approximation. According to Figure 100

a), only a small variation in elastic energy E is involved if the c/abct ratio is varied along the

entire Bain transformation path. In addition to the global minimum at bcc we observed a

second local minimum beyond the Bain path (c/abct > 1.41). Such a local minimum can be

attributed to the formation of a nanotwinned microstructure in the simulation cell and is

further confirmed by pole figure measurements depicted previously. This minimum is

observed for the binary Fe68Pd32 solid solution, but also for ternary Fe68Pd27Cu5, where the

feature is comparatively shallow. Replacing 16% of Pd atoms by Cu increases the overall

valence electron number e/a by 0.05, which tends to stabilize the fcc austenite.209

Page 161: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

149

Figure 100: a) Variation of the total energy per atom as a function of the tetragonal distortion c/abct obtained

from ab initio calculations of a Fe68Pd32 (black squares) and Fe68Pd27Cu5 supercell (red triangles). The

calculations were carried out at constant volume of 13.1⋅10-3 nm3/atom and include a full geometric optimization

of the atomic positions. Apart from the bcc ground state minimum at c/abct = 1 (which defines the energy

reference for each composition), in both cases a second local minimum around c/abct = 1.5 (beyond fcc) is

obtained, which corresponds to the appearance of a finely twinned adaptive superstructure in the 500 atom

simulation cell. b) The variation of the average ground state magnetic moment (normalized per atom) for both

configurations as a function of c/abct. (Calculations performed by M. E. Gruner and originally published in Ref.

205)

This is reflected in calculations from M. E. Gruner by the noticeably decreased energy

difference between the bcc ground state at c/abct = 1 and the fcc austenite at c/abct = 1.41.

Interestingly, the energy profile does not change significantly at low deformations for

c/abct < 1.3. Changes appear mainly in the vicinity of the local fcc maximum, which is now

embedded in a flatter energy landscape. A flat profile in the vicinity of the martensitic phase,

however, can be considered beneficial for the formation of the metastable fct phase, which is

required for the MFIS. The decrease in the fcc-bcc energy difference implies a decreased

temperature for the onset of the martensitic transformation, but this can be compensated by

the simultaneous replacement of Pd by Fe, as shown for polycrystalline Fe-Pd-Cu thin films.

The simultaneous substitution of Pd by Cu and Fe therefore opens a way to specifically

design the profile of the binding surface around 41.12a/c bctbct == which determines the

stability range of the fct phase. The total difference in elastic energy between the bcc and fcc

Page 162: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

150

state, ∆E, is reduced from 15 to 12 meV atom-1 for Fe68Pd32 and Fe68Pd27Cu5, respectively.

This reduces the driving force for the transformation and all associated relaxation processes

and is thus favourable for strained epitaxial growth.

Magnetic remanence, coercivity and saturation field

The hard magnetic axis in a tetragonal lattice of a bulk Fe70Pd30 single-crystal is aligned along

the c-axis while the magnetic easy directions lie within the basal plane.57,210 Within the easy

plane, not all directions are equivalent and a slight anisotropy is observed, which favours both

[110]bct easy axes.

Figure 101: Determination of magnetocrystalline anisotropy within the basal plane by measuring hysteresis

curves of thin films with different tetragonal deformation ratios a) to e) at 300 K. Solid lines indicate

measurements along the [110]bct direction, broken lines along [100]bct direction. f) Comparison of hysteresis

loops measured in out-of-plane orientation along [001]bct at 300 K. The inset presents a magnification around

µ0H = 0 T. From these measurements, the magnetic constants JR, JS and HC to calculate the magnetocrystalline

anisotropy were determined. (Figure originally published in Ref. 205)

Page 163: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

151

To investigate structure-dependent magnetic properties of epitaxial Fe70Pd30-XCuX films,

magnetic hysteresis curves were measured at 300 K in in-plane orientation for thin film

samples having different tetragonal deformation (Figure 101 a) to e)). The magnetization of

the thin film samples differs significantly along different crystallographic in-plane directions

([100]bct and [110]bct) when compared to binary Fe70Pd30 thin films. The most evident

differences occur for c/abct = 1.31, which is located close to the middle of the Bain

transformation path. For this c/abct ratio both crystallographic directions have different values

for remanent polarisation JR and saturation fields HS, while the values for magnetic coercivity

HC do not change significantly. Furthermore, the shape of the magnetization curves and thus

the magnetic characteristic changes with the c/abct ratio. The magnetic hysteresis measured in

the [100]bct direction shows a step-like switching behaviour for the thin film with a low

tetragonal deformation (c/abct = 1.07). In the [110]bct direction a similar switching process

occurs, but additionally a nearly linear increased magnetization at higher fields is observed.

This behaviour originates from a rotation of the magnetization in the thin film during

measurement, indicating that the [110]bct direction is the harder axis within the basal plane,

while JR is reduced in this direction. By increasing the c/abct ratio to 1.31 and higher, an

opposite behaviour occurs. This is correlated to a change in sign of the magnetic anisotropy

within the basal plane.

A summary of the extracted values for magnetic coercivity HC and saturation field HS along

the [110]bct is depicted in Figure 102 a) and b). The trend of µ0HC and µ0HS determined from

in-plane measurements of films with different tetragonal distortions is similar to epitaxial

Fe70Pd30 films (black squares in Figure 102) and epitaxial Fe1-XPdX films deposited on

MgO(100) substrates as reported in literature.67,211 Films having a c/abct ratio close to the bcc

structure exhibit a low HC and HS due to the high crystal symmetry. Thin films with a bct

structure exhibit a slightly increased HC and slightly reduced HS values. Thin films having a

fct structure exhibit a different loop shape with the highest values for HC and HS due to the

large lattice deformation. When the tetragonal distortion comes close to the fcc phase, HC and

HS are reduced because of the high crystal symmetry of the fcc structure. All fabricated

Fe70Pd30-XCuX thin films follow this behaviour. Both values HC and HS approximate to zero

when the tetragonal distortion comes close to both cubic structures bcc (c/abct = 1) and fcc

(c/abct = 1.41). In contrast both values show a maximum value for tetragonal structures. The

HC value decreases and shows only slight differences between both in-plane curves for

Fe70Pd30-XCuX and Fe70Pd30 thin films, distorted beyond the Bain path (c/abct = 1.57, Figure

101 e)).

Page 164: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

152

Figure 102: Presented is the in-plane a) coercivity field µ0HC || [100]bct and b) saturation field µ0HS || [100]bct in

dependence on c/abct ratio. c) The out-of-plane saturation field µ0HS || [001]bct is extracted from magnetic

hysteresis measurements in Figure 101. Black rectangles represent the results for Fe70Pd30 films, red circles

Fe70Pd27Cu3 films and blue triangles Fe70Pd23Cu7 films. All lines are visual guides for the eye. (Figure originally

published in Ref. 205)

This behaviour will be discussed together with the magnetic anisotropy constants K1 and K3

in the following.

To obtain the magnetic anisotropy constant K1, out-of-plane magnetic hysteresis curves along

the [001]bct direction were measured (Figure 101 f)). Along the in-plane directions, all thin

films are magnetically saturated at magnetic flux densities of 0.1 T. In contrast to this, flux

densities > 1 T are required to saturate the same thin films in out-of-plane hysteresis

measurements. The shape anisotropy dominates the magnetization behaviour for an ideal thin

film (demagnetization factor N = 1). When neglecting magnetocrystalline anisotropy,

magnetic saturation is expected at µ0HS = JS. The difference between µ0HS and JS will be used

later to calculate the magnetic anisotropy constant K1. When increasing the tetragonal

deformation to c/abct < 1.41, the HS value decreases (Figure 102 a)).

For all measurements a small hysteresis is observed (see inset in Figure 101 f)). This

switching process is attributed to a slight angular misalignment of the thin film sample normal

with respect to the field during measurement. For a pinning-controlled magnetic coercivity

mechanism, HC exhibits a Kondorsky-like strong increase when the direction of the magnetic

field approaches the hard direction.212 At the same time the remanence-to-saturation ratio

decreases. As a detailed analysis requires a more accurate control of the tilt angle, the HC

Page 165: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

153

value is not used for the out-of-plane direction here. Values for HC and HS along the [110]bct

and [001]bct directions from the in-plane and out-of-plane hysteresis curves were extracted for

Fe70Pd27Cu3 and Fe70Pd23Cu7 films, following same trends (see blue triangles in Figure 102).

Variation of Curie temperature and spontaneous polarization

Previous experiments have indicated that the degree of tetragonal distortion of the lattice

significantly affects the Curie temperature.67 For all epitaxial Fe70Pd30-XCuX thin films,

temperature-dependent magnetization curves were measured in-plane along the [110] bct

direction. An applied magnetic flux density in the range of ≈ 1 T in this direction is sufficient

to saturate the thin film sample. To avoid the destructive thermal decomposition of the

metastable alloy,213 the temperature-dependent magnetization curves were only measured up

to 400 K, which is below the values for TC. As in the previous sections, the TC values of the

thin films were determined from an extrapolation of the magnetization curves using

Kuz’min’s empirical fit.151,152 The results are presented in Figure 103 a).

Figure 103: Variation of a) experimental and b) calculated Curie temperature TC in dependence on c/abct and

composition. With increasing c/abct < 1.41, TC decreases. For c/abct = 1.57, similar values as at c/abct = 1.11 are

observed. The theoretical Curie temperatures are obtained using the mean-field approximation to the Heisenberg

model and corrected by a factor 0.75. The dashed lines in b) represent results for the adaptive nanotwinning

concept. These data agree well with the experimental values. (Figure originally published in Ref. 205)

Black rectangles represent the values for Fe70Pd30, red circles Fe70Pd27Cu3 and blue triangles

Fe70Pd23Cu7. Both the Fe70Pd30 as well as the Fe70Pd30-XCuX systems show a similar variation

of TC with c/abct. The values for TC decrease monotonically as the c/abct ratio increases up to

Page 166: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

154

1.41. The value of TC = 652 K at c/abct = 1.39 approximates the value of 600 K reported for

the bulk fcc phase (c/abct = 1.41).52 The change in TC by addition of Cu is within the range of

experimental error. Thus, compared to other FSMAs, such as Ni-Mn-Ga (TC < 370 K), the

Curie temperature TC in Fe-Pd-based systems is significantly higher.214

In order to determine the compositional and structural trends with regard to TC and the ground

state JS independently, ab initio calculations were performed by M. E. Gruner. The

calculations corroborate the experimental findings. Values for TC are obtained from a classical

Heisenberg model which is parameterized with first-principles magnetic exchange constants.

In the present case, the mean-field approximation can already give a reasonable estimate of

the structural and compositional trends governing the magnetic transformation. Nevertheless,

it has to be considered that this approach systematically overestimates critical temperatures

(typically by 20-30%) due to neglecting spin fluctuations. Furthermore, the induced nature of

the Pd moments, which vary in magnitude according to the field of the surrounding localized

Fe moments, is also not taken into account.215 However, a comparison with a numerically

exact Monte Carlo treatment of the Heisenberg model216 demonstrates that for a heuristic

prediction these shortcomings can be compensated sufficiently by correcting TC

systematically with an empirical factor of 0.75. The quantitative values for TC that are

calculated along the Bain path between bcc and fcc (Figure 103 b)) agree with the values from

experiments. However, no significant variation on TC in dependence on the Cu content (3

at.% and 7 at.%) is observed.

For tetragonal distortions beyond the Bain path at c/abct > 1.41, the experimentally determined

TC value increases again. The TC value of 822 K at c/abct = 1.56 is similar to values for a

structure with c/abct = 1.11. In contrast, the calculations exhibit only a small increase in TC

(solid line in Figure 103 b)). These assume a single variant with c/abct = c/abox > 1.41

according to the substrate constraint. The agreement is much better if an energetically more

favourable nanotwinned fct configuration is assumed, where the individual twins with

c/abct < 1.41 form a microstructure which suits the substrate constraints. From a simple

geometric consideration an approximated relation between the epitaxial constraint due to the

substrate c/abox and the tetragonality of the twins c/atwin has been derived:206

2

twinbox a

c21

a

c−

+= (12)

By using this relation to extrapolate TC for c/abct > 1.41 beyond fcc, a reasonable agreement

with experimental values is obtained for the thin films deposited on a Cu buffer layer (dashed

line in Figure 103 b)).

Page 167: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

155

Another important intrinsic parameter for the FSMAs is the high spontaneous magnetic

polarisation JS. Figure 104 a) shows how JS depends on the tetragonal distortion for different

Cu contents. These values were extracted from the deformation-dependent changes in the

magnetic hysteresis curves in Figure 101 f). According to the structural variation of TC, a

minimum of JS = 1.19 T, taken at a constant temperature of 300 K, at fcc (c/abct = 1.41) is

observed. Near to the bcc (c/abct = 1.09) and at huge strains, JS reaches values of 1.76 T. When

compared to binary films, both Fe70Pd27Cu3 and Fe70Pd23Cu7 films reach only 80% of the JS

of Fe70Pd30. This is consistent with previously shown findings for polycrystalline Fe-Pd-Cu

thin films in Chapter 4.2.3.2. Nevertheless, this value for JS is still significantly higher than

the value JS = 0.76 T obtained for the most prominent FSMA Ni-Mn-Ga.217 While the main

trends are identical, where JS decreases with increasing Cu content having a minimum value

at the fcc structure, the experimental values obtained at 300 K close to TC exhibit a

significantly stronger variation with both composition and c/abct ratio than the results from the

ab initio calculations at T = 0 K (Figure 104 b)).

Figure 104: a) Spontaneous polarization JS determined for different c/abct ratios and compositions at 300 K. JS

for Fe70Pd30-XCuX is approximately 80% of the value for Fe70Pd30. Lines are guides for the eye. b) Dependence of

calculated ground state total (spin + orbital) magnetic moments on c/abct and composition at 0 K. The dashed

lines represent results for the adaptive nanotwinning concept. (Figure originally published in Ref. 205)

The ground state magnetic moments vary only by a few percent while the changes found

experimentally are almost one order of magnitude larger. The larger reduction for the fcc

phase around 300 K is expected, since the fcc phase exhibits the lowest TC. This might also be

Page 168: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

156

correlated to longitudinal spin fluctuations, which can occur in Invar materials such as fcc

Fe70Pd30 218,219 and are known to reduce the values for TC in the austenite phase further.

Magnetocrystalline anisotropy energy

As mentioned before, one of the key intrinsic magnetic properties of FSMAs is their

magnetocrystalline anisotropy energy (MAE). For the MFIS, the MAE represents the

maximum energy input possible by applying an external magnetic field. Hence the MAE

limits the energy available to move twin boundaries or conduct external work. In general the

MAE of a tetragonal lattice can be described by the following equation:220

MAE = K1 sin2(α) + K2 sin4(α) + K3 sin4(α)cos(4β) (13)

Here Ki are the anisotropy constants, α is the angle between the magnetization direction and

the c-axis and β is the angle between the field and the a-axes within the basal plane of a

tetragonal lattice. In most compounds without rare-earth elements, higher-order terms (K2)

can be neglected. For the Fe70Pd30 system, this was reported by Cui et al..57 Thus, it is

sufficient to consider here only K1, which describes the work required to magnetize the

sample along the hard magnetization c-axis, and K3, which characterizes the 4-fold anisotropy

within the basal plane. In the fabricated Fe70Pd30-XCuX thin films, the c/abct ratio is controlled

by strained epitaxial film growth. This allows determination of Ki for all distortions along the

Bain path at one temperature (300 K), which is not possible in bulk samples. Due to the

martensitic transformation and the temperature dependency of the c/aratio, it is not possible to

separate the influence of T and c/a in the bulk. From previously shown hysteresis

measurements on Fe70Pd30-XCuX thin films (Figure 101), the [001]bct direction was determined

to be the hard magnetization axis while the [100]bct and [110]bct directions form the easy

plane.

From the present thin film experiments K1 can be determined from the measurements of HS

along the hard [001]bct direction. The shape anisotropy was considered by using a

demagnetization factor N = 1 of an ideal infinite film, and the anisotropy field HA can be

calculated by µ0HA = µ0HS - NJS. The anisotropy constant is then converted by:

2

JHK SA

1⋅−= (14)

In Figure 105 a) K1 is plotted as a function of the c/abct ratio and the Cu content. The

tetragonal deformation of a Fe-Pd unit cell results in the formation of an easy plane,

Page 169: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

157

corresponding to K1 < 0. The maximum effect is observed around c/abct = 1.34, corresponding

to the fct structure, which exhibits the MFIS in bulk. For binary fct Fe70Pd30 the value was

found to be K1 = -1.6 x 105 J/m3. This agrees well with literature values reported for fct single

crystals (open rectangles in Figure 105 a)) and DFT calculations (stars in Figure 105 a)).67,42

In contrast, Fe70Pd30-XCuX films exhibit an increased absolute value of magnetocrystalline

anisotropy of K1 ≈ -2.4 x 105 J/m3. The absolute values for K1 at 300 K of Fe70Pd27Cu3 and

Fe70Pd23Cu7 even exceed the magnetocrystalline anisotropy constants reported for any other

FSMAs like the Ni-Mn-Ga system: K1 = 1.65 x 105 J/m3 for a 10 M single-variant single-

crystal and K1 = 1.7 x 105 J/m3 for the 14 M structure.214 A larger MAE of similar magnitude

is only observed for Ni-Mn-Ga at significantly lower temperatures217,221 according to the

considerable variation of the MAE with temperature in uniaxial magnets222. When changing

the tetragonal deformation close to highly symmetric cubic structures (bcc and fcc), K1 is

reduced for all compositions. At c/abct > 1.41, the magnetocrystalline anisotropy increases

again, but does not reach the values of the fct structure.

Figure 105: Magnetocrystalline anisotropy constants a) K1 and b) K3 as a function of c/abct for various Cu

contents (solid symbols, T = 300 K). Also shown are the values for a bulk Fe68.8Pd31.2 single-crystal (open

rectangle42) and calculations for disordered Fe70Pd30 (stars, T = 0 K67). The errors for K1 are shown by error bars

and for K3 they are in the range of the symbol size. All lines are guides for the eye. (Figure originally published

in Ref. 205)

This is again consistent with the presence of adaptive nanotwins in the film.206 For a very

small martensitic variant width it is no longer possible to form a complete 90° domain wall at

twin boundaries since the exchange energy favours a parallel alignment of magnetization.

Page 170: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

158

According to the random anisotropy model of Herzer, the critical length for this is the

magnetic exchange length lexch.223 This exchange length lexch is in the order of 18-85 nm (and

depends on the composition) for fct Fe-Pd67, which far exceeds the width of adaptive

nanotwins206. Due to the 4-fold symmetry of the basal plane a deviation from an idealized

easy plane behaviour can be observed in films (Figure 101) and bulk samples.57,210 K3 is a

measure of the anisotropy within this basal plane and defines the work that is necessary to

magnetize along the respective directions of the bct unit cell:

2K3 = W[100]bct - W[110]bct (15)

The magnetocrystalline anisotropy constant K3 can be extracted from the area enclosed by the

magnetic hysteresis curves (Figure 101) measured along both the [110]bct and [100]bct

directions. These values of K3 are two orders of magnitude smaller than for K1 (Figure 105

b)) and change sign within the Bain path. For c/abct values close to bcc, positive values are

observed. Films with c/abct ratios close to fcc have negative K3 values. In between these

values no significant differences were observed for Fe70Pd30 and Fe70Pd30-XCuX films. As for

|K1|, |K3| exhibits a maximum of 1.8 x 105 J/m3 for the fct structure (c/abct = 1.34).

The presented findings and analysis of the structure and magnetism in ternary Fe-Pd-Cu

epitaxial thin films suggests that the addition of small amounts of Cu is able to significantly

enhance the functional properties of the Fe70Pd30 FSMA. The combination of thin film

experiments and ab initio simulations yield an insight into the “frozen stages” of the

martensitic transformation process within the limits of the Bain path, which is enforced by the

epitaxial relation to selected buffer materials. This is in particular favourable for a detailed

analysis of the anisotropic magnetic properties since the measurements are affected neither by

a magnetically-induced reorientation nor by the continuous variation of the tetragonal

distortion with temperature, which both occur in bulk samples. One main result is that Cu

flattens the energy landscape, which suppresses common relaxation mechanisms and thus

allows for a much better film quality in combination with faster growth. More importantly, an

increase of the magnetocrystalline anisotropy constant K1 by ≈ 40% is obtained, which is a

substantial improvement of a key property necessary for the MFIS. The values obtained for

the fct structure even exceed those reported for the prototype Ni-Mn-Ga system, which makes

the Fe-Pd-Cu system of significant interest for microsystems with a high energy density. A

minor drawback is the reduction of the spontaneous polarization JS by about 15-20% due to

the addition of Cu.

Page 171: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

159

XRD measurements indicate that the use of room temperature deposition techniques prevents

decomposition, which must otherwise be expected for more than 5 at.% Cu. However, upon

increasing the Cu content from 3 to 7 at.% no variation in the trend of magnetic properties

was observed. This might in turn be taken as an indication for the presence of structural or

compositional inhomogeneities at the nanoscale (as for example a tendency towards a slight

L10 short-range order with antiphase boundaries). Recent Mössbauer experiments on Fe-Pd-

Cu splats with <1 at.% Cu suggest some chemical short-range order.224 Moreover, first-

principles calculations of Fe-rich Fe-Pd predict a certain preference for forming a layered type

of order (L10 or Z1),225,226 while the cubic Fe3Pt-type L12 order appears to be ruled out for

energetic reasons65. In the martensitic state, ageing behaviour resulting in symmetry-

conforming short-range order at this length scale does not necessarily inhibit shape memory

applications227 and could tentatively induce a kind of beneficial two-way behaviour. This

aspect, however, cannot be probed for the films tested with the presented methods and

requires further investigation.

Page 172: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

160

5. Conclusion and Outlook

In the presented work, combinatorial fabrication and high-throughput characterization

methods were employed to develop new Fe-Pd(-X) ferromagnetic shape memory alloys

(FSMAs) with improved properties. Broad regions of the full binary and ternary Fe-Pd(-X)

compositional diagram were fabricated as thin film materials libraries by wedge-type and

confocal thin film deposition and were investigated by high-throughput characterization:

namely, energy dispersive X-ray analysis (EDX) for composition, temperature-dependent

resistance (R(T)) and magneto-optical Kerr effect (MOKE) measurements for characterization

of phase transformation and magnetic properties. Further X-ray diffraction (XRD) and

synchrotron measurements for structural, and nanoindentation at elevated temperatures for

mechanical, properties were investigated. Promising thin films were further investigated by

non-high-throughput methods such as vibrating sample magnetometry (VSM), temperature-

dependent X-ray diffraction (XRD(T)) and transmission electron miocroscopy (TEM) for

detailed analysis.

Polycrystalline Fe-Pd thin film materials libraries revealed transforming composition regions

in dependence of sputter deposition method. Lowest transformation temperatures were found

for films, fabricated by wedge-type deposition. This was correlated to the exothermic energy

of mixing during annealing and the decrease of internal stress due to a reduction of interfaces

in the microstructure.

The splat quenching method was identified to be well-suited for fabricating single-phase

Fe70Pd30 bulk samples that exhibit a martensitic transformation. Annealing at different

temperatures resulted in a shift of transformation temperatures and was correlated to the

columnar grain growth and the difference in grain sizes.

Epitaxial Fe-Pd thin films attached to a substrate revealed different structural phases in

dependence on composition without undergoing a martensitic transformation in the as-

deposited state. After annealing a reversible transformation was identified in a Fe72Pd28 thin

film.

Annealed freestanding Fe70Pd30 thin films showed an unusual reversible martensitic

transformation from the bcc over fct to the austenite fcc phase. This sequence of phases was

theoretically expected from the Bain path formalism but was never observed before for Fe-Pd.

Epitaxial Fe70Pd30 films were fabricated and investigated with regard to demonstrate the

feasibility of fabricating freestanding 1.2 µm thick films that can be implemented as

sensing/actuating devices in microsystems.

Page 173: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

161

The Fe-Pd-Mn system was developed as a new FSMA showing improved properties. A

broadening of the transforming single-phase region was identified in the composition

diagram, allowing to alloy up to 9 at.% of Mn at the expense of Fe content. Transformation

temperatures up to 379 K were measured in Fe-Pd-Mn thin films, exceeding all values known

for Fe-Pd-based FSMAs from literature so far. Magnetic properties such as Curie temperature

TC and saturation polarization JS in this system were found to be only slightly lowered

compared to the Fe-Pd system. Stress measurements allowed to exclude the formation of

stress-induced martensite as the cause for the increased transformation temperatures. From

theoretical predictions an energy gain due to displacements from the ideal lattice positions

upon addition of Mn was reported. This explained the increased amount of energy in terms of

temperature that is needed to induce the martensitic transformation. Finally the e/a ratio was

found to be the dominating parameter, describing the change of martensitic transformation

temperatures.

The Fe-Pd-Cu system indicated a broadening of the transforming single-phase region, where

up to 4 at.% of Cu can be added at the expense of the Pd contents. Increased transformation

temperatures were determined for this system, while TC and JS are not significantly lowered.

In accordance to ab initio results, an increase of spontaneous volume magnetostriction and

thus a decrease of thermal expansion was postulated. This was confirmed by synchrotron

experiments revealing a decrease of residual thin film stress and a decrease of thermal

expansion coefficient in Fe-Pd-Cu splats with increasing Cu content. Finally, the Fe content

was found to be the decisive factor controlling the shift of martensitic transformation. The

addition of Cu also enabled to increase the Fe content to values above the limit known for the

binary Fe-Pd system, without the formation of precipitates. Epitaxial Fe70Pd30-XCuX (X = 3

and 7 at.%) thin films deposited on different buffer layers, varying the in-plane lattice

parameter, revealed a significant change of magnetic properties in dependence of the c/a-ratio

in the unit cell. Finally, the highest value for the magnetocrystalline anisotropy constant

K1 = 2.4 x 105 J/m3 at 300 K was identified for c/abct = 1.34. This value increases by about

40% due to the addition of Cu and is higher than any other value reported for FSMAs so far.

The discovered results in the new Fe-Pd-Mn and Fe-Pd-Cu thin film systems exceed the

benchmark values defined in Chapter 2.8. Both systems showed martensitic start temperatures

with Ms > 350 K without increasing the thermal hysteresis to values ∆T > 5 K. The saturation

polarization was lowered in both systems, but never exceeded JS < 1 T in the martensite

phase. The Curie temperature was found to vary between 460 K < TC < 550 K and thus is

smaller than the benchmark value. The magnetocrystalline anisotropy constant K1 can be

Page 174: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

162

considered as the most important value for the MFIS. Here, the benchmark value of

K1 = 1.0 x 105 J/m3 was significantly exceeded for a Fe70Pd27Cu3 thin film. All other

benchmark values were not proven and will be subject for future work.

Furthermore it has to be mentioned, that no distinct rules were identified for enhancing Fe-Pd-

based FSMAs upon addition of further elements from the present findings. It was found, that

the shift of martensitic transformation is controlled by different dominating parameters in

both ternary systems. While the transformation temperatures vary with Fe content in the Fe-

Pd-Cu system, the Fe-Pd-Mn system is controlled by the e/a ratio. This indicates that the

impact of the third element on the martensitic transformation, structure and magnetic

parameters is tremendous and alters the materials properties significantly.

All presented findings indicate the advantage of combinatorial fabrication and

characterization methodologies to develop new FSMAs with enhanced properties. The impact

of third elements on the composition-structure-property relationship can be determined up to a

certain extent in polycrystalline thin film materials libraries. However, key properties for

FSMAs like the magnetocrystalline anisotropy cannot precisely be measured in

polycrystalline thin film samples. Therefore the combination of different sample designs such

as materials libraries, epitaxial films and bulk samples is well-suited for an accelerated

understanding, followed by a tailored development of new materials with improved

properties. This approach can be further extended to other materials that undergo reversible

structural or magnetic transformations and are of high interest for industrial application, such

as Fe- and Co-based magnetocaloric materials.

Page 175: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

163

6. References

1 Janocha H 1999 Adaptronics and Smart Structures: Basics, Materials, Design and Applications (Berlin:

Springer) 1st ed 2 Otsuka K and Ren W 2005 Physical metallurgy of Ti-Ni-based shape memory alloys Prog. Mater Sci. 50 pp 511-678 3 Ullakko K, Huang J K, Kanter C, Kokorin V V and O’Handley R C 1996 Large magnetic field induced strains in Ni2MnGa single crystals Appl. Phys. Lett. 69 p 1966 4 James R D and Wuttig M 1998 Magnetostriction of martensite Philos. Mag. A 77 p 1273 5 Gehanno V, Marty A, Gilles B and Samson Y 1997 Magnetic domains in epitaxial ordered FePd (001) thin films with perpendicular magnetic anisotropy Phys. Rev. B 55 18 pp 12552-12555 6 Chen M and Nikles D E 2002 Synthesis of spherical FePd and CoPt nanoparticles J. Appl. Phys. 91 8477 7 Duplessis R R, Stern R A and MacLaren J M 2004 Phase stability criteria for cubic and orthorhombic Fe3Pt and Fe3Pd J. Appl. Phys. 95 6589 8 Jessen K 1962 Über die Invar-Eigenschaften der Eisen - Palladium-Legierungen Ann. Phys. 464 5-6 pp 313-315 9 Matsui M and Adachi K 1983 Magnetostriction of Fe-Pd Invar J. Magn. Magn. Mater. 31-34 pp 115-116 10 Okazaki T, Tanaka M, Ogasawara N, Furuya Y, Saito C and Imaizumi N 2009 Development of Magnetic-Field-Driven Micro-Gas Valve Mater. Trans. 50 3 pp 461-466 11 Chopra H D and Wuttig M 1995 Precursor Shear Elastic Pseudo-Isotropy in Near Second Order Martensitic Phase Transitions J. Phys. IV 05 C8 pp 157-160 12 Muto S, Oshima R and Fujita F E 1990 Elastic softening and elastic strain energy consideration in the f.c.c.-f.c.t. transformation of Fe-Pd alloys Acta Metall. Mater. 38 4 pp 685-694 13 Otsuka K. and Wayman C M 1999 Shape memory materials (Cambridge Univ Press) 1st ed 14 Martens A 1887 Stahl und Eisen 7 pp 235-242 15 Haasen P 1994 Physikalische Metallkunde (Springer) 3rd ed 16 Wayman C and Duerig T 1990 Introduction to Martensite and Shape Memory, in Engineering Aspects of

Shape Memory Alloys (Elsevier Butterworth-Heinemann Oxford) ed T Duerig, K Melton, D Stöckel and C Wayman

17 Hornbogen E 1985 The effect of variables on martensitic transformation temperatures Acta Metall. 33 pp 595-601

18 Bowles J S and Mackenzie J K 1954 The crystallography of martensite transformations Acta Mater. 2 pp 129-234

19 Lieberman D S, Wechsler M S and Read T A 1955 Cubic to orthorhombic diffusionless phase change-experimental and theoretical studies of AuCd J. Appl. Phys. 26 4 pp 473-484

20 Nagasawa A 1976 Nature of Martensitic phase transition in β phase alloys J. Phys. Soc. Jpn. 40 4 pp 1021-1026

21 Wollants P, Bonte M and de Roos J R 1979 A thermodynamic analysis of the stress-induced martensitic transformation in a single crystal Z. Metallkd. 70 113

22 Kato H, Liang Y and Taya M 2002 Stress-induced FCC/FCT phase transformation in Fe-Pd alloy Scripta Mater. 46 471

23 Kästner J, Petry W, Shapiro S M, Zheludev A, Neuhaus J, Roessel T, Wassermann E and Bach H 1999 Influence of atomic order on TA1[110] phonon softening [4] and displacive phase transition in Fe72Pt28 Invar alloys EPJB 10 641

24 Oikawa K, Saito R, Anzai K, Ishikawa H, Sutou Y, Omori T, Yoshikawa A, Chernenko V A, Besseghini S, Gambardella A, Kainuma R and Ishida R 2009 Elastic and Superelastic Properties of NiFeCoGa Fibers Grown by Micro-Pulling-Down Method Mater. Trans. 50 4 pp 934-937

25 Miyazaki S, Fu Y Q and Huang W M 2009 Thin Film Shape Memory Alloys: Fundamentals and Device Applications (Cambridge: University Press) 1st ed

26 Du Trémolet De Lacheisserie E 1993 Magnetostriction, Theory and Applications of Magnetoelasticity (Boca Raton, Ann Arbor, London, Tokyo: CRC Press) 1st ed

27 Krevet B and Kohl M 2010 Thermodynamic modelling of ferromagnetic shape memory actuators Mater. Sci. Forum 635 p 175 28 Omori T, Watanabe K, Umetsu R Y, Kainuma R and Ishida K 2009 Martensitic transformation and magnetic

field-induced strain in Fe-Mn-Ga shape memory alloy Appl. Phys. Lett. 95 8 082508 29 Chmielus M, Witherspoon C, Wimpory R C, Paulke A, Hilger A, Zhang X, Dunand D C and Müllner P 2010

Magnetic-field-induced recovery strain in polycrystalline Ni-Mn-Ga foam J. Appl. Phys. 108 123526 30 Quandt E and Wuttig M 2004 Comparison of joule and twin induced magnetostriction Bremen Actuator Conf.

Proc. pp 359-362

Page 176: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

164

31 Bozorth R M 1993 Ferromagnetism (Piscataway, New Jersey: IEEE Press) 1st ed 32 Kakeshita T, Fukuda T and Takeuchi T 2006 Magneto-mechanical evaluation for twinning plane movement

driven by magnetic field in ferromagnetic shape memory alloys Mater. Sci. Eng. A 438-440 12-17 33 O’Handley R C, Murray S J, Marioni M A, Nembach H and Allen S M 2000 Phenomenology of giant

magnetic-field-induced strain in ferromagnetic shape-memory materials J. Appl. Phys. 87 pp 4712-4717 34 Heczko O, Scheerbaum N and Gutfleisch O 2009 Magnetic Shape Memory Phenomena In Nanoscale

Magnetic Materials and Applications (Heidelberg: Springer) ed J P Liu, E Fullerton, O Gutfleisch, D J Sellmyer pp 399-439

35 Heczko O 2005 Magnetic shape memory effect and magnetization reversal J. Magn. Magn. Mater. 290-291 787-794

36 Ullakko K, Huang K, Kokorin V V and O’Handley R C 1997 Magnetically controlled shape memory effect in Ni2MnGa intermetallics Scripta Mater. 36 pp 1133-1138

37 Kakeshita T and Fukuda T 2002 Giant magnetostriction in Fe3Pt and FePd ferromagnetic shape-memory alloys Mater. Sci. Forum 394-395 p 531

38 Lavrov A N, Komiya S and Ando Y 2002 Antiferromagnets: Magnetic shape-memory effects in a crystal Nat. Mater. 418 p 385

39 Rhyne J J, Foner S, McNiff E J and Doclo R 1968 Rare earth metal single crystals. I. high-field properties of Dy, Er, Ho, Tb, and Gd J. Appl. Phys. 39 p 892

40 Sozinov A, Likhachev A A, Lanska N and Ullakko K 2002 Giant magnetic-field-induced strain in NiMnGa seven-layered martensitic phase Appl. Phys. Lett. 80 p 1746

41 Chmielus M, Zhang X X, Witherspoon C, Dunand D C, Müllner P 2009 Giant magnetic-field-induced strains in polycrystalline Ni-Mn-Ga foams Nat. Mater. 8 p 863

42 Kakeshita T, Fukuda T and Takeuchi T. 2006 Magneto-mechanical evaluation for twinning plane movement driven by magnetic field in ferromagnetic shape memory alloys Mater. Sci. Eng. A 438-440 pp 12-17

43 Hausch G 1973 Magnetovolume effects in invar alloys: Pressure dependence of the curie temperature Phys. Status Solidi A 16 2 371-376

44 Cullity B D and Graham C D 2008 Introduction to Magnetic Materials (Piscataway, New Jersey: IEEE Press) 2nd ed

45 Guillaume C E 1897 Les Comptes Rendus de l'Académie des sciences 125 pp 235-238 46 Kaufman L, Clougherty E V and Weiss R J 1963 The lattice stability of metals-III. Iron Acta Metall. 11 p 323 47 Wassermann E F 1989 New Developments on the Invar-Effect. Phys. Scr. T. 25 pp 209-219 48 Weiss R J 1963 The Origin of the ‘Invar’ Effect Proc. Phys. Soc. 82 p 281 49 Kondorsky E I and Sedov V L 1960 Effect of the Polarity of the III-V Intermetallic Compounds on Etching J.

Appl. Phys. 31 331 50 Kakehashi Y 1989 Magnetovolume effect and finite-temperature theory of magnetism in transition metals and

alloys Physica B 161 pp 143-152 51 Schlosser W 1971 Occurrence of Invar Anomalies in Binary Alloys of Fe with Ni, Pd, Pt, and Mn J. Appl.

Phys. 42 1700 52 Matsui M, Shimizu T, Yamada H and Adachi K 1980 Magnetic Properties and thermal expansion of Fe-Pd

Invar Alloys J. Magn. Magn. Mater. 15-18 pp 1201-1202 53 Nakayama T, Kikuchi M and Fukamichi K 1980 Young's modulus and the ∆E effect of Fe-Pd Invar alloys J.

Phys. F. Metal. Phys. 10 715 54 Haasen P 1994 Physikalische Metallkunde (Berlin: Springer) 3rd ed 55 Burton B P 1990 Fe-Pd (Iron-Palladium), Binary Alloy Phase Diagrams (ASM Alloy Phase Diagrams Center

2nd ed) ed T B Massalski vol 2 pp 1749-1751 56 Sugiyama M, Oshima R and Fujita F E 1984 Martensitic transformation in the iron-palladium alloy system

Mater. Trans. 25 9 pp 585-592 57 Cui J, Shield T W and James R D 2004 Phase transformation and magnetic anisotropy of an iron-palladium

ferromagnetic shape-memory alloy Acta Mater. 52 1 pp 35-47 58 Matsui M, Yamada H and Adachi K 1980 A New Low Temperature Phase (fct) of Fe-Pd Invar J. Phys. Soc.

Jpn. 48 6 pp 2161-2162 59 Matsui M and Chikazumi S 1978 Analysis of Anomalous Thermal Expansion Coefficient of Fe-Ni Invar

Alloys J. Phys. Soc. Jpn. 45 pp 458-465 60 Sumiyama K, Shiga M, Morioka M and Nakamura Y 1979 Characteristic magnetovolume effects in Invar type

Fe-Pt alloys J. Phys. F 9 p 1665-1677 61 Sugimura Y, Cohen-Karni I, McCluskey P and Vlassak J J 2005 Stress Evolution of Sputter-Deposited Fe-Pd

Shape-Memory Thin Films J. Mater. Res. 20 9 pp 2279-2287 62 Stern R A, Willoughby S D, MacLaren J M, Cui J, Pan Q and James R D 2003 Fe3 Pd ferromagnetic shape

memory alloys J. Appl. Phys. 93 p 10

Page 177: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

165

63 Bain E C and and Dunkirk N Y 1924 The nature of martensite Trans. Am. Inst. Min. Metall. Eng. 70 p 25 64 Oshima R 1981 Successive Martensitic Transformations in Fe-Pd Alloys Scr. Metall. 15 829-833 65 Gruner M E and Entel P 2011 Impact of local lattice distortions on the structural stability of Fe-Pd magnetic

shape-memory alloys Phys. Rev. B 83 214415 66 Opahle I, Koepernik K, Nitzsche U and Richter M 2009 Jahn-Teller-like origin of the tetragonal distortion in

disordered Fe-Pd magnetic shape memory alloys Appl. Phys. Lett. 94 7 072508 67 Buschbeck J, Opahle I, Richter M, Rössler U K, Klaer P, Kallmayer M, Elmers H J, Jakob G, Schultz L and

Fähler S 2009 Full tunability of strain along the fcc-bcc bain path in epitaxial films and consequences for magnetic properties Phys. Rev. Lett. 103 21 216101

68 Slater J C 1930 Atomic shielding constants Phys. Rev. 35 pp 509-529 69 van der Merwe J H 1963 Crystal Interfaces. Part I. Semi-Infinite Crystals J. Appl. Phys. 34 123 70 Godlevsky V V and Rabe K M 2001 Soft tetragonal distortions in ferromagnetic Ni2MnGa and related

materials from first principles Phys. Rev. B 63 134407 71 Tsuchiya K, Nojiri T, Ohtsuka H and Umemoto M 2003 Effect of Co and Ni on martensitic transformation and

magnetic properties in Fe-Pd ferromagnetic shape memory alloys: Structural and functional control of materials through solid-solid phase transformations in high magnetic field Mater. Trans. JIM 44 12 pp 2499-2502

72 Vokoun D, Wang Y W, Goryczka T and Hu C T 2005 Magnetostrictive and shape memory properties of Fe-Pd alloys with Co and Pt additions Smart Mater. Struct. 14 p 261

73 Sánchez-Alarcos V, Recarte V, Pérez-Landazábal J I, Gómez-Polo C, Chernenko V A and González M A 2008 Reversible and irreversible martensitic transformations in Fe-Pd and Fe-Pd-Co alloys Eur. Phys. J. Special Topics 158 p 107

74 Kovacs A, Sato K and Hirotsu Y 2007 Strong perpendicular magnetic anisotropy of Fe-Pd nanocrystalline particles enhanced by Co addition J. Appl. Phys. 101 033910

75 Lin Y C, Lin C F, Yang J B and Lee H T 2011 Microstructures and magnetostriction of two-phase Fe66-Pd30-Ni4 high-temperature ferromagnetic shape memory alloys J. Appl. Phys. 109 07A912

76 Lin Y C and Lin C F 2012 Effects of Ni addition on the magnetostriction and microstructures of Fe70-xPd30Nix high-temperature ferromagnetic shape memory alloys J. Appl. Phys. 111 07A902

77 Stern R A, Willoughby S D, Ramirez A, MacLaren J M, Cui J, Pan Q and James R D 2002 Electronic and structural properties of FePd-Pt ferromagnetic shape memory alloys J. Appl. Phys. 91 10 pp 7818-7820

78 Wada T, Tagawa T, Taya M 2003 Martensitic transformation in Pd-rich Fe-Pd-Pt alloy Scripta Mater. 48 pp 207-211

79 Takeuchi I and Wuttig M 2007 High-throughput screening of magnetic properties of quenched metallic-alloy thin-film composition spreads J. Appl. Surf. 254 pp 734-737

80 Lin Y, Lee H T, Jen S U and Chen Y T 2007 Magnetic structure of an Fe-Pd-Rh alloy J. Appl. Phys. 101 09N514

81 Sánchez-Alarcos V, Recarte V, Perez-Landazabal J I, Gonzalez M A, Rodriguez-Velamazan J A 2009 Effect of Mn addition on the structural and magnetic properties of Fe-Pd ferromagnetic shape memory alloys Acta Mater. 57 pp 4224-4232

82 Opahle I, Koepernik K, Nitzsche U and Richter M 2009 Jahn-Teller-like origin of the tetragonal distortion in disordered Fe-Pd magnetic shape memory alloys Appl. Phys. Lett. 94 072508

83 Heczko O 2005 Magnetic shape memory effect and magnetization reversal J. Magn. Magn. Mater. 290 p 787 84 Liot F and Abrikosov I A 2009 Local magnetovolume effects in Fe65Ni35 alloys Phys. Rev. B 79 014202 85 Boettcher A, Haase G and Thun R 1955 Strukturuntersuchung von Mehrstoffsystemen Durch Kinematische Elektronenbeugung Z. Metallkde. 46 5 pp 386-400 86 Kennedy K, Stefansky T, Davy G, Zackay V F and Parker E R 1965 Rapid Method for Determining Ternary-Alloy Phase Diagrams J. Appl. Phys. 36 3808 87 Hanak J J J 1970 The “multiple-sample concept” in materials research: Synthesis, compositional analysis and

testing of entire multicomponent systems Mater. Sci. 5 pp 964-971 88 Xiang X D, Sun X, Briceno G, Lou Y, Wang K A, Chang H, Wallace-Freedman W G, Chen S W and Schultz

P G 1995 Combinatorial Approach to Materials Discovery Science 268 pp 1738-1740 89 Takeuchi I, Lauterbach J and Fasolka M J 2005 Combinatorial materials synthesis Mater. Today 8 pp 18-26 90 Takeuchi I, Famodu O O, Read J C, Aronova M A, Chang K S, Craciunescu C, Lofland S E, Wuttig M,

Wellstood F C, Knauss L and Orozco A 2003 Identification of novel compositions of ferromagnetic shape-memory alloys using composition spreads Nat. Mater. 2 pp 180-184

91 Tuinstra H E and Cummins C H 2000 Combinatorial Materials and Catalyst Research Adv. Mater. 12 pp 1819-1822

92 Wang J, Yoo Y, Gao C, Takeuchi I, Sun X, Chang H, Xiang X D and Schultz P G 1998 Identification of a Blue Photoluminescent Composite Material from a Combinatorial Library Science 279 pp 1712-1714

Page 178: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

166

93 Hoogenboom R, Meier M A R and Schubert U S 2003 Combinatorial Methods, Automated Synthesis and

High-Throughput Screening in Polymer Research: Past and Present Macromol. Rapid. Comm. 24 15-32 94 Potyrailo R, Rajan K, Stoewe K, Takeuchi I, Chisholm B and Lam H 2011 Combinatorial and High-

Throughput Screening of Materials Libraries: Review of State of the Art ACS Comb. Sci. 13, 579-633 95 Jansen M 2002 A concept for synthesis planning in solid-state chemistry Angew. Chem. 41 pp 3746-3766 96 Cawse J 2001 Experimental Strategies for Combinatorial and High-Throughput Materials Development Acc.

Chem. Res. 34 pp 213-221 97 Göpel W 1998 Chemical imaging: I. Concepts and visions for electronic and bioelectronic noses Sens.

Actuators B 52 pp 125-142 98 Zhanpeng J A 1981 A study of the range of stability of σ phase in some ternar systems Scand. J. Metall. 10 pp

279-287 99 Miller N C and Shirn G A 1967 Co-sputtered Au-SiO2 cermet films Appl. Phys. Lett. 10 pp 86-88 100 Goldfarb I, Zolotoyabko E, Berner A and Shechtman D 1994 Novel sample preparation technique for the

study of multicomponent phase diagrams Mater. Lett. 21 pp 149-154 101 Dahn J R, Trussler S, Hatchard T D, Bonakdarpour A, Mueller-Neuhaus J R, Hewitt K C and Fleischauer M

2002 Economical Sputtering System To Produce Large-Size Composition-Spread Libraries Having Linear and Orthogonal Stoichiometry Variations Chem. Mater. 14 pp 3519-3523

102 Gremaud R, Broedersz C P, Borsa D M, Borgschulte A, Mauron P, Schreuders H, Rector J H, Dam B and Griessen R 2007 Hydrogenography: An Optical Combinatorial Method To Find New Light-Weight Hydrogen-Storage Materials Adv. Mater. 19 pp 2813-2817

103 Wu K H, Hung C I, Ziolkowski P, Platzek D, Karpinski G, Stiewe C and Mueller E 2009 Improvement of spatial resolution for local Seebeck coefficient measurements by deconvolution algorithm Rev. Sci. Instrum. 80 pp 105104-105108

104 Chang H, Gao C, Takeuchi I, Yoo Y, Wang J, Schultz P G, Xiang X D, Sharma R P, Downes M and Venkatesan T 1998 Combinatorial synthesis and high throughput evaluation of ferroelectric/dielectric thin-film libraries for microwave applications Appl. Phys. Lett. 72 pp 2185-2187

105 Woo N C, Ng B G and van Dover R B 2007 High-throughput combinatorial study of local stress in thin film composition spreads Rev. Sci. Instrum. 78 072208 5

106 Sharpe W N, Yuan B, Vaidyanathan R and Edwards R L 1996 New test structures and techniques for measurement of mechanical properties of MEMS materials. in Microlithography and metrology In Micromachining II, ed M T Postek and C Friedrich pp 78-91

107 Hemker K and Sharpe W N 2007 Microscale Characterization of Mechanical Properties Annu. Rev. Mater. Res. 37 pp 93-126

108 Gally B J, Abnet C C, Brown S and Chui C 2000 Wafer scale testing of MEMS structural films In Thin films-stresses and mechanical properties VIII, ed R Vinci, O Kraft, N Moody, P Besser and E Shaffer pp 195-200

109 Gaspar J, Schmidt M, Held J and Paul O 2007 Reliability of MEMS Materials: Mechanical Characterization of Thin-Films Using the Wafer Scale Bulge Test and Improved Microtensile Techniques Material Research Society Symposium (Boston USA) ed D A LaVan, M G DaSilva, S M Spearing and S Vengallatore vol 1052 DD01 supplement: 02 pp 15-20

110 Hamann S, Ehmann M, Thienhaus S, Savan A and Ludwig A 2008 Micro-hotplates for high-throughput thin film processing and in situ phase transformation characterization Sens. Actuators A 147 pp 576-582

111 McCluskey P J, Zhao Z W, Kfi O and Vlassak J J 2011 Precipitation and thermal fatigue in Ni-Ti-Zr shape memory alloy thin films by combinatorial nanocalorimetry Acta Mater. 59 13 pp 5116-5124

112 Takeuchi I, Long C J, Famodu O O, Murakami M, Hattrick-Simpers J, Rubloff G W, Stukowski M and Rajan K 2005 Data management and visualization of x-ray diffraction spectra from thin film ternary composition spreads Rev. Sci. Instrum. 76 pp 62223-62228

113 Long C J, Hattrick-Simpers J, Murakami M, Srivastava R C, Takeuchi I, Karen V L and Li X 2007 Rapid structural mapping of ternary metallic alloy systems using the combinatorial approach and cluster analysis Rev. Sci. Instrum. 78 072217

114 Lüth H 1993 Solid Surfaces, Interfaces and Thin Films (Berlin: Springer) 5th ed pp 88-131 115 Ohring M 2001 Material Science of Thin Films (Waltham, Massachusetts: Academic Press) 2nd ed 116 Bauer E 1958 Phänomenologische Theorie der Kristallabscheidung an Oberflächen Z. Kristallogr. 110 372 p

1820 117 Volmer M and Weber A 1926 Keimbildung in übersättigten Gebilden Z. Phys. Chem. 119 277 118 Stranski J N and Krastanov L 1938 Zur Theorie der orientierten Ausscheidung von Ionenkristallen

aufeinander Ber. Akad. Wiss. Wien 146 797 119 Frank F C and Van der Merwe J H 1949 One-Dimensional Dislocations. I. Static Theory Proc. R. Soc.

London A 198 205 120 Thornton J A 1977 High rate thick film growth Annu. Rev. Mater. Sci. 7 1 pp 239-260

Page 179: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

167

121 Walton D, Rhodin T N and Rollins R W 1963 Nucleation of silver on sodium chloride J. Chem. Phys. 38

2698 122 Chang C A 1992 Metal alloy-films containing body-centered cubic and face-centered cubic metals - Relation

between structures and lattice spacings J. Appl. Phys. 72 p 4463 123 Jones H 1973 Splat cooling and metastable phases Rep. Prog. Phys. 36 11 p 1425 124 Duwez P, Willens R H and Klement W J 1960 Continuous Series of Metastable Solid Solutions in Silver-

Copper Alloys J. Appl. Phys. 31 1136 125 Edmund Bühler GmbH: Rascherstarrungstechnologie, Splat Cooling, April 2010 http://www.edmund-

buehler.de/i-rascherstarrungstechnologie.pml 126 Oxford Instruments Inca Manual 2009 127 Casino software available at www.gel.usherb.ca/casino 128 Drouin D, Couture A R, Joly D, Tastet X, Aimez V and Gauvin R. 2007 A Fast and Easy-to-use Modeling

Tool for Scanning Electron Microscopy and Microanalysis Users Scanning 29 pp 92-101 129 Czyzewski Z, MacCallum D O, Romig A and Joy D C 1990 Calculations of Mott scattering cross section

J. Appl. Phys. 68 7 pp 3066-3072 130 Drouin D, Hovington P and Gauvin R 1997 A New Monte Carlo Code in C Language for Electron Beam

Interaction - Part II: Tabulated Values of the Mott Cross Section Scanning 19 pp 20-28 131 http://www.esrf.eu/computing/scientific/FIT2D/ 132 Cullity B D 1956 Elements of X-Ray Diffraction (Massachusetts: Addison-Wesley Publishing Company, Inc.)

1st ed 133 Warren B E X-Ray Diffraction (Boston: Addison-Wesley Publishing Co.) 1st ed pp 251-254 134 Spieß L, Schwarzer R, Behnken H and Teichert G 2005 Moderne Röntgenbeugung (Wiesbaden: B. G.

Teubner) 1st ed pp 263-284 135 Birkholz M 2006 Thin Film Analysis by X-Ray Scattering (Weinheim: Wiley-VCH) 1st ed 136 Prevéy P S 1986 The use of pearson VII distribution functions in X-ray diffraction residual stress

measurement Adv. X-Ray Anal. 29 pp 103-111 137 Noyan I C and Cohen J B 1987 Residual Stress Measurement by Diffraction and Interpretation ed B Ilschner

and N J Grant (Berlin: Springer) 1st ed 138 Cheng W and Finnie I 2007 Residual Stress Measurement and the Slitting Method (Berlin: Springer) 1st ed 139 Takeuchi I, Long C J, Famodu O O, Murakami M, Hattrick-Simpers J, Rubloff G W, Stukowski M and Rajan

K 2005 Data management and visualization of x-ray diffraction spectra from thin film ternary composition spreads Rev. Sci. Instrum. 76 62223-62228

140 Long C J, Hattrick-Simpers J, Murakami M, Srivastava R C, Takeuchi I, Karen V L and Li X 2007 Rapid structural mapping of ternary metallic alloy systems using the combinatorial approach and cluster analysis Rev. Sci. Instrum. 78 072217

141 Long C J, Bunker D, Li X, Karen V L and Takeuchi I 2009 Rapid identification of structural phases in combinatorial thin-film libraries using x-rays diffraction and none-negative matrix factorization Rev. Sci. Instrum. 80 103902-103906

142 Cho H, Kim H Y and Miyazaki S 2005 Fabrication and characterization of Ti-Ni shape memory thin film using Ti/Ni multilayer technique Sci. Tech. Adv. Mater. 6 pp 678-683

143 Inoue S, Inoue K, Fujita S and Koterazawa K 2003 Fe-Pd Ferromagnetic Shape Memory Alloy Thin Films Made By Dual Source DC Magnetron Sputtering Mater. Trans. 44 2 298

144 Lo Y C and Wu S K 1993 A study of B2-B19-B19´ two-stage martensitic transformation in a Ti50Ni40Cu10 alloy, Acta Metall. Mater. 41 3 pp 747-749

145 Faulkner M G, Amalraj J J and Bhattacharyya A 2000 Experimental determination of thermal and electrical properties of Ni-Ti shape memory wires Smart Mater. Struct. 9 pp 632-639

146 Uchil J 2002 Shape memory alloys - characterization techniques Pramana Indian Academy of Sciences 58 1131-1139

147 Wu S K, Lin H C and Lin T Y 2006 Electrical resistivity of Ti-Ni binary and Ti-Ni-X (X = Fe, Cu) ternary shape memory alloys Mater. Sci. Eng. A 438-440 pp 536-539 148 Wan D and Komvopoulos K 2005 Thickness Effect on Thermally Induced Phase Transformations in

Sputtered Titanium-nickel Shape-memory Films J. Mater. Res. 20 1606 149 Mohanchandra K P, Ho K K and Carman G P 2004 Electrical Characterization of NiTi Film on Silicon

Substrate J. Intell. Mater. Syst. Struct. 15 387 150 Thienhaus S, Zamponi C, Rumpf H, Hattrick-Simpers J, Takeuchi I and Ludwig A 2006 High-throughput

characterization of shape memory thin films using automated temperature-dependent resistivity measurements Mater. Res. Soc. Symp. Proc. 894 0894-LL06-06.1

151 Kuz’min M D 2005 Shape of Temperature Dependence of Spontaneous Magnetization of Ferromagnets: Quantitative Analysis Phys. Rev. Lett. 94 107204

Page 180: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

168

152 Kuz’min M D, Richter M and Yaresko A N 2006 Factors determining the shape of the temperature

dependence of the spontaneous magnetization of a ferromagnet Phys. Rev. B 73 100401R 153 Pharr G M, Strader J H and Oliver W C 2009 Critical issues in making small-depth mechanical property

measurements by nanoindetation with continuous stiffness measurement J. Mater. Res. 24 653 154 Oliver W C and Pharr G M 1992 An improved technique for determining hardness and elastic modulus using

load and displacement sensing indentation experiments J. Mater. Res. 7 1564 155 Oliver W C and Pharr G M 2004 Measurement of hardness and elastic modulus by instrumented indentation:

Advances in understanding and refinements to methodology J. Mater. Res. 19 3 20 156 Pfetzing J 2010 Nanoindentation von NiTi-Formgedächtnislegierungen PhD thesis (Institute for Materials,

Ruhr-Universität Bochum, Germany) 157 Sree Harsha K S 2006 Principles of Physical Vapor Deposition of Thin Films Elsevier Science & Technology 158 Kaufmann S, Buschbeck J, Heczko, Schultz L and Fähler S 2008 Epitaxial growth of Ni-Mn-Ga:

Consequences of Magnetron configuration on martensitic behaviour Proc. ICOMAT (Santa Fe) 159 Kühnemund L, Edler T, Kock I, Seibt M and Mayr S G 2009 Epitaxial growth and stress relaxation of vapor-deposited Fe-Pd magnetic shape memory alloy films New J. Phys. 11 113054 160 De Keijser T H, Langford J I, Mittemejer E J and Vogels A B P 1982 Use of the Voigt function in a single-line method for the analysis of X-ray diffraction line broadening J. Appl. Cryst. 15 pp 308-314 161 Fick A 1855 Ueber diffusion Ann. Phys. 94 pp 59-86 162 Gavens A J, Van Heerden D, Mann A B, Reiss M E and Weihs T P 2000 Effect of intermixing on self-

propagating exothermic reactions in Al/Ni nanolaminate foils J. Appl. Phys. 87 3 pp 1255-1263 163 Kock I, Hamann S, Brunken H, Edeler T, Ludwig A and Mayr S G 2010 Development and characterization of Fe70Pd30 ferromagnetic shape memory splats Intermetallics 18 pp 877-882 164 Zarnetta R, Savan A, Thienhaus S and Ludwig A 2007 Combinatorial study of phase transformation

characteristics of a Ti-Ni-Pd shape memory thin film composition spread in view of microactuator applications Appl. Surf. Sci. 254 743 8

165 Viret M, Samson Y, Warin P, Marty A, Ott F and Søndergård E 2000 Anisotropy of domain wall resistance Phys. Rev. Lett. 85 3962

166 Guenin G and Gobin P F 1982 A localized soft mode model for the nucleation of thermoelastic martensitic transformations: Application to the β → 9r transformation Metall. Mater. Trans. 13 1127 34

167 Dondl P and Bhattacharya K 2007 The effect of precipitates on the evolution of a martensitic phase boundary Proc. Appl. Math. Mech. 7 1151207

168 Seki K, Kura H, Sato T, Taniyama T 2008 Size dependence of martensite transformation temperature in ferromagnetic shape memory alloy FePd J. Appl. Phys. 103 063910

169 Kang S, Lee Y, Lim Y, Nam J, Nam T and Kim Y 2008 Relationship between grain size and martensitic transformation start temperature in a Ti-30Ni-20Cu alloy ribbon Scripta Mater. 59 1186 9

170 Levine R D 2009 Molecular Reaction Dynamics (Cambridge: University Press) 1st ed 171 Wollants P, Roos J, Delaely L 1993 Thermally- and Stress-Induced Thermoelastic Martensitic

Transformations in the Reference Frame of Equilibrium Thermodynamics Progr. Mater. Sci. 37 227 88 172 Kock I 2010 Phasenbildung, Phasenübergang und mechanische Eigenschaften des Funktionsmaterials Eisen-

Palladium PhD thesis (I. Physical Institut, Georg-August-Universität Göttingen) 173 Buschbeck J, Lindemann I, Schultz L and Fähler S 2007 Growth, structure, and texture of epitaxial Fe100-xPdx

films deposited on MgO(100) at room temperature: An x-ray diffraction study Phys. Rev. B 76 205421 174 Buschbeck J, Niemann R, Heczko O, Thomas M, Schultz L and Fähler S 2009 In situ studies of the

martensitic transformation in epitaxial Ni-Mn-Ga films Acta Mater. 57 2516 175 Onisan A T, Bogdanov A N and Rössler U 2010 Domain models for ferromagnetic shape-memory materials

Acta Mater. 58 13 4378-4386 176 Edler T and Mayr S G 2010 Film lift-off from MgO: Freestanding single crystalline Fe-Pd films suitable for

magnetic shape memory actuation - and beyond Adv. Mater. 22 4969 177 Edler T, Hamann S, Ludwig A and Mayer S G 2011 Reversible fcc ↔ bcc transition in freestanding epitaxially grown Fe-Pd ferromagnetic shape memory films Scripta Mater. 64 pp 89-92 178 Kakeshita T and Fukuda T 2006 Magnetic field-control of microstructure and function of materials exhibiting

solid-solid phase transformation Sci. Technol. Adv. Mater. 7 350 179 Oshima R and Sugiyama M 1982 Martensite transformations in iron-palladium alloys J. Phys. Colloq. 43 C4 180 Kühnemund L, Edler T, Kock I, Seibt M and Mayr S G 2009 Epitaxial growth and stress relaxation of vapor-

deposited Fe-Pd magnetic shape memory films New J. Phys. 11 113054 181 Gruner M E, Adeagbo W A, Zayak A T, Hucht A and Entel P 2010 Lattice dynamics and structural stability

of ordered Fe3Ni, Fe3Pd and Fe3Pt alloys using density functional theory Phys. Rev. B 81 064109

Page 181: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

169

182 Bechtold C, Buschbeck J, Lotnyk A, Erkartal B, Hamann S, Zamponi C, Schulz L, Ludwig A, Kienle L,

Fähler S and Quandt E 2010 Artificial single variant martensite in freestanding Fe70Pd30 films obtained by coherent epitaxial growth Adv. Mater. 22 1-4

183 Wagner M F X and Windl W 2008 Lattice stability, elastic constants and macroscopic moduli of NiTi martensites from first principles Acta Mater. 56 6232

184 Buschbeck J, Opahle I, Fähler S, Schultz L and Richter M 2008 Magnetic properties of Fe-Pd magnetic shape memory alloys: Density functional calculations and epitaxial films Phys. Rev. B 77 174421

185 Thienhaus S, Hamann S and Ludwig A 2011 Modular high-throughput test-stand for the versatile screening of thin film materials libraries Sci. Technol. Adv. Mat. 12 054206 186 Acet M, Schneider T, Gehrmann B and Wassermann E F 1995 The Magnetic aspects of the martensitic transformations in Fe-Mn Alloys J. Phys. IV Colloque C8 p 379 187 Klimars S, Hesse J and Huck B 1985 Mössbauer spectroscopy on iron rich Fe-Pd alloys including fcc invar J. Magn. Magn. Mater. 51 183 188 Periodic table 2006 (Illinois, Skokie: Sargent-Welch Scientific Company) 189 Acet M, Wassermann E F, Andersen K, Murani A and Schärpf O 1997 The anomalous temperature dependence of the paramagnetic response of Fe-rich fcc Fe-Ni Europhys. Lett. 40 93 190 Kauffmann-Weiss S, Hamann S, Gruner M E, Buschbeck J, Ludwig A, Schultz L and Fähler S 2012 Understanding the Magnetic Shape Memory System Fe-Pd-X by Thin Film Experiments and First Principles Calculations Adv. Eng. Mater. 14 8 pp 724-749 191 Gruner M E, Hamann S, Brunken H, Ludwig A and Entel P 2012 Compositional trends and magnetic excitations in binary and ternary Fe-Pd(-X) magnetic shape memory alloys J. Alloys Compd. <doi:10.1016/j.jallcom.2012.02.033> 192 Ebert H 2009 Lecture notes in physics: Electronic structure and physical properties of solids 535 191 (Berlin:

Springer Verlag) ed H Dreysse 193 Ebert H <http://olymp.cup.uni-muenchen.de/ak/ebert/SPRKKR> 194 Perdew J P, Burke K and Ernzerhoff M 1996 Generalized gradient approximation made simple Phys. Rev.

Lett. 77 3865 195 Gyorffy B L, Pindor A J, Staunton J, Stocks G M and Winter H J 1985 A first-principles theory of

ferromagnetic phase transitions in metals J. Phys. F: Met. Phys. 15 1337 196 Khmelevkyi S, Turek I and Mohn P 2003 Large negative magnetic contribution to the thermal expansion in

iron-platinum alloys: quantitative theory of the Invar effect Phys. Rev. Lett. 91 037201 197 Liechstenstein A I, Katsnelson M I, Antropov V P and Gubanov V A 1987 Local spin density functional

approach to the theory of exchange interactions in ferromagnetic metals and alloys J. Magn. Magn. Mater. 15-18 1201

198 Okamoto H 1993 Phase Diagrams of Binary Iron Alloys (Materials Park/Ohio: ASM International) 199 Hamann S, Gruner M E, Irsen S, Buschbeck J, Bechtold C, Kock I, Mayr S G, Savan A, Thienhaus S, Quandt E, Fähler S, Entel P and Ludwig A 2010 The ferromagnetic shape memory system Fe-Pd-Cu Acta Mater. 58 pp 5949-5961 200 Kuang J P, Kontani M, Matsui M and Adachi K 1988 Electronic, phonon and magnon specific heats of FePd

alloys Physica B 149 209 201 Edler T, Buschbeck J, Mickel C, Fähler S and Mayr S G 2008 Mechanisms of stress generation and relaxation

during pulsed laser deposition of epitaxial Fe-Pd magnetic shape memory alloy films on MgO New J. Phys. 10 063007

202 Kock I, Edler T and Mayr S G 2008 Growth behavior and intrinsic properties of vapor-deposited iron palladium thin films J. Appl. Phys. 103 046108

203 Kim H J, Han J H, Kaiser R, Oh K H and Vlassak J J 2008 High-throughput analysis of thin-film stresses using arrays of micromachined cantilever beams Rev. Sci. Instrum. 79 045112

204 Khmelevskyi S and Mohn P 2004 Magnetostriction in Fe-based alloys and the origin of the Invar anomaly Phys. Rev. B 69 140404R

205 Kauffmann-Weiss S, Hamann S, Gruner M E, Schultz L, Ludwig A and Fähler S 2012 Enhancing magnetocrystalline anisotropy of the Fe70Pd30 magnetic shape memory alloy by adding Cu Acta Mater. 60 pp 6920-6930 206 Kauffmann-Weiss S, Gruner M E, Backen A, Schultz L, Entel P and Fähler S 2011 Magnetic Nanostructures by Adaptive Twinning in Strained Epitaxial Films Phys. Rev. Lett. 107 206105 207 Kauffmann A 2011 <CorrVert http://sco.ifw-dresden.de> 208 v Steinwehr H E 1967 Ursachen der Abweichungen von der Vegardschen Regel Z. Kristallogr. 1967 125 360 209 Nishiyama Z 1978 Martensitic transformation (New York: Academic Press) 210 Kakeshita T and Fukuda T 2006 Energy evaluation for twinning plane movement under magnetic field in ferromagnetic shape memory alloys Int. J. Appl. Electromagn. Mech. 23 45

Page 182: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

170

211 Buschbeck J, Hamann S, Ludwig A, Holzapfel B, Schultz L and Fähler S 2010 Correlation of Phase Transformations and Magnetic Properties in Annealed Epitaxial Fe-Pd Magnetic Shape Memory Alloy Films J. Appl. Phys. 107 113919 212 Kondorsky E J 1940 A mechanism of magnetic hysteresis in heterogeneous alloys Phys (USSR)2 161 213 Buschbeck J, Heczko O, Ludwig A, Fähler S and Schultz L 2008 Magnetic Properties of epitaxial Fe-Pd films measured at elevated temperatures J. Appl. Phys. 103 07 B334 214 Straka L and Heczko O 2003 Magnetic anisotropy in Ni-Mn-Ga martensites J. Appl. Phys. 93 8636 215 Polesya S, Mankovsky S, Sipr O, Meindl W, Strunk C and Ebert H 2010 Finite-temperature magnetism of FexPd1−x and CoxPt1−x alloys Phys. Rev. B 82 214409 216 Gruner M E, Entel P, Minar J, Polesya S, Mankovsky S and Ebert H J 2012 Electronic and magnetic trends in martensitically transforming Fe-Pd alloys Magn. Magn. Mater. <http://dx.doi.org/10.1016/j.jmmm.2012.02.081> 217 Tickle R and James R D 1999 Magnetic and magnetomechanical properties of Ni2MnGa J. Magn. Magn. Mater. 195 627 218 Pepperhoff W and Acet M 2001 Engineering materials (Berlin: Springer Verlag) vol VIII 219 Wassermann E F 1990 Ferromagnetic materials (Amsterdam: Elsevier) ed K H J Buschow and E P Wohlfarth vol 5 p 237 220 Buschow K H J and de Boer F R 2004 Physics of magnetism and magnetic materials (New York: Kluwer) vol 97 221 Klaer P, Eichhorn T, Jakob G and Elmers H J 2011 Microscopic origin of magnetic anisotropy in martensitic Ni2MnGa Phys. Rev. B 83 214419 222 Callen E R and Callen H B 1960 Anisotropic magnetization J. Phys. Chem. Solids 16 310 223 Herzer G J 1992 Nanokristalline soft magnetic materials Magn. Magn. Mater. 112 258 224 Claussen I, Brand R, Hahn H and Mayr S G 2012 Relaxation scenarios in Fe-Pd and Fe-Pd-Cu ferromagnetic shape memory splats Scripta Mater. 66 163 225 Barabash S V, Chepulskii R V, Blum V and Zunger A 2009 First-principles determination of low-temperature order and ground states of Fe-Ni, Fe-Pd, and Fe-Pt Phys. Rev. B 80 220201 226 Chepulskii R V, Barabash S V and Zunger A 2012 Ab initio theory of phase stability and structural selectivity in Fe-Pd alloys Phys. Rev. B 85 144201 227 Ren X and Otsuka K 1997 Origin of the Rubber-like Behavior in Metal Alloys Nature 389 579

Page 183: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

171

Publications

Peer reviewed publications on ferromagnetic shape memory alloys

1) Kauffmann-Weiss S, Hamann S, Gruner M E, Schultz L, Ludwig A and Fähler S

2012 Enhancing magnetocrystalline anisotropy of the Fe70Pd30 magnetic shape

memory alloy by adding Cu Acta Mater. 60 pp 6920-6930

2) Kauffmann-Weiss S, Hamann S, Gruner M E, Buschbeck J, Ludwig A, Schultz L and

Fähler S 2012 Understanding the Magnetic Shape Memory System Fe-Pd-X by Thin

Film Experiments and First Principles Calculations Adv. Eng. Mater. 14 8 pp 724-749

3) Gruner M G, Hamann S, Brunken H, Ludwig A and Entel P 2012 Compositional

trends and magnetic excitations in binary and ternary Fe-Pd(-X) magnetic shape

memory alloys J. Alloys Compd. (accepted)

4) Thienhaus S, Hamann S and Ludwig A 2011, Modular high-throughput test-stand for

the versatile screening of thin film materials libraries, Sci. Technol. Adv. Mat. 12

054206

5) Lai Y W, Hamann S, Ehmann M and Ludwig A 2011 High-throughput

characterization of stresses in thin film materials libraries using Si cantilever array

wafers and digital holographic microscopy Rev. Sci. Instrum. 82 063903

6) Edler T, Hamann S, Ludwig A and Mayr S G 2011 Reversible fcc ↔ bcc transition in

freestanding epitaxially grown Fe-Pd ferromagnetic shape memory films Scripta

Mater. 64 1 pp 89-92

7) Hamann S, Gruner M E, Irsen S, Buschbeck J, Bechtold C, Kock I, Mayr S G, Savan

A, Thienhaus S, Quandt E, Fähler S, Entel P and Ludwig A 2010 The ferromagnetic

shape memory system Fe-Pd-Cu Acta Mater. 58 pp 5949-5961

8) Bechtold C, Buschbeck J, Lotnyk A, Erkartal B, Hamann S, Zamponi C, Schulz L,

Ludwig A, Kienle L, Fähler S and Quandt E 2010 Artificial single variant martensite

in freestanding Fe70Pd30 films obtained by coherent epitaxial growth Adv. Mater. 22

24 p 2668

Page 184: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

172

9) Buschbeck J, Hamann S, Ludwig A, Holzapfel B, Schultz L and Fähler S 2010

Correlation of Phase Transformations and Magnetic Properties in Annealed Epitaxial

Fe-Pd Magnetic Shape Memory Alloy Films J. Appl. Phys. 107 113919

10) Kock I, Hamann S, Brunken H, Edler T, Mayr S G and Ludwig A 2010 Development

and characterization of Fe70Pd30 ferromagnetic shape memory splats Intermetallics 18

pp 877-882

11) Ludwig A, Zarnetta R, Hamann S, Savan A and Thienhaus S 2008 Development of

multifunctional thin films using high-throughput experimentation methods

Int. J. Mater. Res. 99 10 pp 1144-1149

Proceeding papers on ferromagnetic shape memory alloys

1) Hamann S, Gruner M E, Thienhaus S, Savan A and Ludwig A 2011 Combinatorial

Development of Fe-Pd-X thin film systems with improved intrinsic properties 3rd

International Conference on Ferromagnetic Shape Memory Alloys Proceeding,

(Dresden Germany) ed S Fähler 18.-22.07.2011 pp 140-141

2) Bechtold C, Buschbeck J, Lotnyk A, Erkartal B, Hamann S, Zamponi C, Schulz L,

Ludwig A, Kienle L, Fähler S and Quandt E 2010 Development Towards MSM

Active FePd Thick Films Actuator 2010 Conference Proceedings (Bremen Germany)

ed H Borgmann pp 299-302

3) Hamann S, Savan A, Thienhaus S and Ludwig A 2008 Combinatorial Development

of Fe-Pd-Mn Ferromagnetic Shape Memory Thin Films Actuator 2008 Conference

Proceedings (Bremen Germany) ed H Borgmann pp 271-274

Other peer reviewed publications

1) Hamann S, Brunken H, Salomon S, Meyer R, Savan A and Ludwig A 2013 Synthesis

of Au micro-wires by selective oxidation of Au-W thin film composition spreads

Sci. Technol. Adv. Mat. 14 1 p 015003

2) Grochla D, Siegel A, Hamann S, Buenconsejo P J S, Kieschnick M, Brunken H,

König D and Ludwig A 2013 Time and space resolved high-throughput

Page 185: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

173

characterization of stresses during sputtering and thermal processing of Al-Cr-N thin

films J. Phys. D: Appl. Phys. 46 8 084011

3) Meyer R, Hamann S, Ehmann M, Thienhaus S, Jaeger S, Thiede T, Devi A, Fischer R

A and Ludwig A 2012 Micro-gradient-heaters as tools for high-throughput

experimentation ACS Comb. Sci. 14 10 pp 527-578

4) König D, Buenconsejo P J S, Grochla D, Hamann S, Pfetzing-Miklich J and Ludwig

A 2012 Thickness-dependence of the B2-B19 martensitic transformation in nanoscale

shape memory alloy thin films: zero hysteresis in 75 nm thick Ti51Ni38Cu11

Acta Mater. 60 1 pp 306-313

5) Zimmer C M, Schubert J, Hamann S, Kunze U and Doll T 2011 Nanoscale

photoelectron ionisation detector based on lanthanum hexaboride Phys. Status Solidi A

- Applications and Materials Science 208 6 pp 1241-1245

6) Meyer R, Hamann S, Ehmann M, König D,Thienhaus S, Savan A and Ludwig A

2011 Small Scale Deposition of Thin Films and Nanoparticles by Micro-Evaporation

Sources J. Microelectromech. S. 19 pp 1264-1269

7) Lai Y W, Koukourakis N, Gerhardt N C, Hofmann M R, Meyer R, Hamann S,

Ehmann M, Hackl K, Darakis E and Ludwig A 2010 Integrity of micro-hotplates

during high-temperature operation monitored by digital holographic microscopy,

J. Microelectromech. S. 19 5 pp 1175-1179

8) Hamann S, Ehmann M, Thienhaus S, Savan A and Ludwig A 2008 Micro-hotplates

as processing and characterization platforms for combinatorial materials science and

high-throughput experimentation Sensor. Actuat. A-Phys. 147 pp 576 - 582

Page 186: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

174

Acknowledgments

The presented results, findings and interpretations were performed at the Centre for Advanced

European Studies and Research (caesar), Bonn in the beginning and finished at the Institute

for Materials, Faculty of Mechanical Engineering at the Ruhr-University of Bochum from

February 2007 to December 2012.

I would like to express my gratitude to my doctoral adviser Prof. Dr.-Ing. Alfred Ludwig, for

giving me the opportunity to perform my PhD thesis and for his trust and encouragement

during this time. Further I have to thank Prof. Dr.-Ing. Alfred Ludwig especially for allowing

me to have insights into many other fields of scientific work next to the topic of my PhD.

Additionally I have to thank many current and former members of his group, who helped me

and my scientific research in many ways and provided a pleasant working atmosphere.

Especially gratefulness goes to the following colleagues: Sigurd (for technical support in

many ways), Alan (for helping in many ways but especially sputtering), Michael Ehmann (for

MHPs and GHs, we got quite a few papers out), Hayo (for company and more starting from

the 1st semester), Dennis (for fruitful discussions and everything else), Robert (for his critical

considerations and being my cigarette break buddy), Alexander (for technical help and many

more things), Dario (for helping and cheering me up) and Pio (for inspiring discussions and

ideas). Further thanks go to all students that supported my scientific work.

Thanks also to Prof. Dr. Gunther Eggeler and many of his group members (Janine, Jan, Klaus

and Christoph) for the cooperative atmosphere and scientific support.

Additional gratitude goes to Prof. Dr.-Ing. F. Peters for supervising the process of my Ph. D.

defense.

Further, I would like to acknowledge many colleagues from other research institutions for the

nice and productive cooperation: Markus Gruner for very fruitful discussions bringing

understanding into the dark of experimental findings; Sebastian Fähler for inspiring

discussions and many advices; Jörg Buschbeck and Sandra Kauffmann-Weiss for this great

collaboration with the IFW in Dresden; Anja Backen and Robert Niemann for several fruitful

discussions; Christoph Bechtold and Eckhard Quandt for the nice company at caesar; Burak

and Lorenz for deep insight by TEM; Iris Claussen, Tobias Edler and Stefan Mayr for very

Page 187: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

175

fruitful collaboration. Further many thanks to all others that I might have forgotten and

especially to all persons that participated in publishing this huge amount of papers.

Finally, I have to thank Christina for her support (especially during hellweek) and everthing

she did and my parents Geli and Gerd as well as my brother Dennis and his girlfriend Lina. I

additionally want to thank all my other relatives and friends for their moral support by

cheering me up in moments when I was down.

Page 188: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

176

CURRICULUM VITAE

PERSONAL INFORMATION

Surname Hamann

Name Sven

Date of Birth 30 August 1979

Place of Birth Essen (Ruhr), Germany

Nationality German

WORK EXPERIENCE

Oct 2006 – present Research associate and lecturer

with Prof. Dr.-Ing. A. Ludwig, Chair for MEMS materials - Institute for Materials - Faculty of Mechanical Engineering, Ruhr-Universität Bochum, Germany

Project “Combinatorial development of Fe-Pd based ferromagnetic shape memory alloys“, part of the Deutsche Forschungsgemeinschaft (DFG)-priority programme SPP 1239

Mar – Apr 2011 Guest researcher

Leibniz Institute for Solid State and Materials Research Dresden (IFW Dresden), Germany

Epitaxial growth of Fe-Pd based ferromagnetic shape memory alloys.

Jan – May 2005 Student research assistant

Chair of Automatic Control and Systems Theory - Faculty of Mechanical Engineering, Ruhr-Universität Bochum, Germany

Programming of the waste-gallery system in the Airbus A380 (industrial project for AEROTEC)

Mar – Sept 2006 Diploma thesis “Die Charakterisierung und Anwendung von Mikro-Heizplatten im Bereich der kombinatorischen Materialforschung von dünnen Schichten“ (Characterization and application of micro hotplates in the field of combinatorial material science of thin films)

Prof. Dr.-Ing. A. Ludwig, Combinatorial Material Science (CMS), centre of advanced european studies and research (caesar), Bonn, Germany

Page 189: Development of Fe-Pd-based ferromagnetic shape memory alloys ...

177

EDUCATION

Oct 2003 – Sept 2006 Main Studies “Mechanical Engineering” (focus field Micro-Engineering) and diploma degree “Dipl.-Ing. Mechanical Engineering”

Ruhr-Universität Bochum, Germany

Feb – Aug 2003 Semester abroad

Universidad Politecnica De Valencia, Spain

Oct 2000 – Jan 2003 Basic studies “Mechanical Engineering” and intermediate diploma

University of Duisburg-Essen, Germany

1999 – 2005 Training as reserve officer / reservist lieutenant, including course of studies “Leadership”

Deutsche Bundeswehr (German Army)

1991 – 1999 Secondary School, German university entrance qualification

Helmholtz-Gymnasium, Essen, Germany

ACHIEVEMENTS

23 – 28 AUG 2010 Invited lecture at Joint European Magnetic Symposia (JEMS): “Structural, magnetic and phase transformation properties of Fe-Pd-X thin films”, Krakow, Poland

MAR 2010 “Best presentation” during the evaluation of the last funding period of the Deutsche Forschungsgemeinschaft (DFG)-priority programme SPP 1239

FEB – AUG 2003 ERASMUS-scholarship for an academic semester abroad

2002 – 2006 Scholarship holder of the “Prof. Dr. Koepchen Academic Foundation” (RWE AG, Essen, Germany)

NATIVE LANGUAGE German

LANGUAGES English (fluent), Spanish (advanced knowledge), French (basic knowledge)