larson acta mat review paper 04

Upload: nat55nat

Post on 10-Apr-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/8/2019 Larson Acta Mat Review Paper 04

    1/16

    Overview No. 138

    Information storage materials: nanoscale characterisation bythree-dimensional atom probe analysis

    D.J. Larson a,*, A.K. Petford-Long b, Y.Q Ma b, A. Cerezo b

    a Seagate Technology LLC, Recording Heads, One Disc Drive, NRE304, Bloomington, MN 55435, USAb Department of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, UK

    Received 15 December 2003; received in revised form 11 March 2004; accepted 13 March 2004

    Available online 22 April 2004

    Abstract

    The development of nanoscale magnetic materials for applications in information storage systems relies heavily on the ability to

    engineer the properties of the layered structures from which such materials are fabricated. These properties are strongly dependent

    on the nature of the interfaces between the individual nanoscale magnetic layers, so knowledge of the interface chemistry is crucial.

    In this paper, we discuss the application of three-dimensional atom probe analysis to the characterisation of layered magnetic

    materials, including details of specimen preparation techniques required for this type of analysis. Recent results are presented on the

    characterisation of interfaces in Co/Cu or CoFe/Cu multilayers, which form part of the read sensor in magnetic recording heads, and

    Co/Pd multilayers, which are being considered for use as perpendicular recording media.

    2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

    Keywords: Atom probe; Layered structures; Magnetic thin films

    1. Introduction

    In recent years there has been increasing interest in

    thin-film layered structures because of the novel prop-

    erties that they exhibit. Multilayer film (MLF) structures

    are formed by alternate deposition of two or more dif-

    ferent elements or compounds, usually with layers that

    are 15 nm in thickness. Of particular interest are layered

    structures formed from one or more transition-metal

    magnetic materials because they have applications in the

    field of data recording. Examples include high density

    longitudinal or perpendicular magnetic recording media,

    and sensors based on the giant magnetoresistance

    (GMR) phenomenon such as those used for read heads

    [1,2]. For a recent review of modern magnetic materials

    in data storage see [3].

    Layered structures based on Co, Ni, Fe, Cu and

    their alloys form sub-structures of devices such as the

    spin-valve [4], which exhibit GMR [2]. Changes in

    microstructure (e.g. interfacial roughness, chemical

    intermixing, etc.) are believed to influence the magni-

    tude of the GMR effect and there has been much

    research devoted to the dependence of the GMR ratio

    on the specific nature of the interfaces in layered

    structures. However, inadequate structural character-

    ization has led to contradictory results arising from

    different research groups even for the same materials

    system (see, for example, [57]). This arises in part

    because of the difficulty in distinguishing between

    chemical mixing and roughness for most of the tech-

    niques used to characterise the films. Interface

    roughness may be defined as a spatial variation of

    the interface dividing two layers and intermixing as

    the chemical mixing of the elements that comprise the

    two layers, at or near the interface. These two mi-

    crostructural features are likely to have very different

    effects on the properties, therefore establishing the

    correlation between structure and properties requires

    the ability to distinguish between roughness and

    intermixing.

    When characterising layered structures, the very

    small layer thicknesses makes it extremely difficult to

    * Corresponding author. Tel.: +1-952-402-7140; fax: +1-952-402-

    7734.

    E-mail address: [email protected] (D.J. Larson).

    1359-6454/$30.00 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

    doi:10.1016/j.actamat.2004.03.015

    Acta Materialia 52 (2004) 28472862

    www.actamat-journals.com

    http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/
  • 8/8/2019 Larson Acta Mat Review Paper 04

    2/16

    determine the nature of each individualinterface unless a

    technique is used that allows independent interface

    analysis with high spatial resolution. Low-angle X-ray

    diffraction (X-ray reflectivity) has been widely used to

    study layered thin film materials (see for example [8]).

    However, the X-ray reflectivity profiles that are obtained

    are summed data that include contributions from all theinterfaces in the structure. It can therefore prove ex-

    tremely difficult to obtain a unique fit to the data for

    multilayer films if the interfaces between different layers

    do not contain the same degree of roughness and/or

    interdiffusion through the multilayer stack. Neutron

    reflectivity has also been applied to the study of inter-

    faces in epitaxial superlattice structures and to poly-

    crystalline sputtered films but this technique is not as

    good at determining structural information as X-ray

    reflectivity, and is therefore more usually applied using

    polarised neutrons to determine magnetic structure (see

    [9] for a recent review).

    One of the most commonly used techniques that can

    be used to analyse each interface independently is

    transmission electron microscopy (TEM) of cross-sec-

    tion samples, which can be used to study both the mi-

    crostructure and compositional profile across the layers

    [10]. However, a limitation of TEM is that it is a pro-

    jection technique and integration through the specimen

    thickness can make it difficult to distinguish interfacial

    mixing and fine-scale roughness except under very spe-

    cific conditions. An example is given by Hytch of sep-

    arating these two effects in Co/Cu multilayer films using

    detailed image analysis of through-focal series of Fresnel

    fringe images [11].The technique of three-dimensional atom probe

    (3DAP) analysis [1215] is unique in having the capa-

    bility to characterise internal interfaces, grain bound-

    aries and precipitates with sub-nanometre resolution in

    all three dimensions. With this technique it is therefore

    possible to measure the extent of interdiffusion or in-

    terface segregation at the atomic scale, and separate

    these effects from nanometre-scale topological features.

    In a 3DAP, pulsed field evaporation is used to re-

    move individual atoms from the surface of a specimen,

    which is in the form of a needle with end radius $50 nm.A position-sensitive detector is used for measurement of

    both flight time (to identify atomic species) and impact

    position (which enables the original position of each

    atom on the specimen surface to be recorded). The po-

    sition of an atom in the depth direction is calculated

    from the sequence of ion detection. The 3DAP has a

    depth resolution of a single atomic layer and sub-

    nanometre lateral resolution. An intrinsic requirement

    of the 3DAP microscopy technique is the production of

    a very high electric field (2050 V/nm) at the surface of

    the specimen in order to achieve field evaporation. The

    field is produced by applying a high voltage (typically

    $10 kV) to a needle-shaped specimen with an apex ra-

    dius of curvature of 10100 nm. For most materials,

    these sharp needles can be produced by electrochemi-

    cally polishing fine wires, whiskers or blanks cut from

    bulk material [1216]. However, preparing a specimen

    from certain sample geometries can be quite compli-

    cated. Examples include certain multi-phase alloys,

    semiconductors, ceramics and, in particular, thin filmmaterials. Over the years, various methods have been

    applied to 3DAP specimen preparation in order to

    overcome these problems [13,14].

    This paper gives an overview of the results of recent

    3DAP studies, both from our group and from other

    groups, of magnetic layered film structures (e.g. Ni/Cu,

    Co/Cu, CoFe/Cu, Co/Pd, etc.), concentrating on mate-

    rials with potential applications in information storage

    systems. These materials science studies have been made

    possible by recent progress in the application of high

    resolution focused ion beam systems to the preparation

    of 3DAP specimens from complex layered systems, and

    this is also reviewed briefly.

    2. Development of 3DAP specimens from thin film

    samples

    In order to investigate thin layered structures in the

    atom probe, these features must be present in the apex

    region of the specimen. The preparation of 3DAP

    specimens from thin film structures is difficult, but is a

    crucial step in obtaining 3DAP data that provide the

    highest possible spatial resolution across the feature of

    interest, in this case the interfaces in the layered struc-ture. The specimen preparation process is difficult be-

    cause the total sample thickness (region of interest

    containing the thin layers) is often only of the order of

    20100 nm. There are three main methods that have

    been used to prepare 3DAP specimens from thin film

    structures and these are outlined below.

    A number of papers have been produced detailing the

    results of 3DAP studies carried out on thin film mate-

    rials in which the films of interest were deposited directly

    onto pre-evaporated needle-shaped specimens [1730].

    Fig. 1(a) shows a schematic of the geometry of a speci-

    men that has been prepared in this way, but note that

    the curvature of the tip relative to the layer thickness is

    not necessarily to scale (the tip radius is of the order of

    50 nm and the layer thickness is typically a few nano-

    metres). The small volume from which the 3DAP data

    are taken (of the order of 20 nm in diameter) means that

    within the volume analysed the substrate approximates

    to a flat surface, and the structure of the layers is often

    assumed to be reasonably similar to that of the same

    layers deposited on a planar substrate. However the

    microstructure is unlikely to be exactly the same, in part

    because the substrate material (in this case a metallic tip)

    does not correspond well to the substrates used in ap-

    2848 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862

  • 8/8/2019 Larson Acta Mat Review Paper 04

    3/16

    plications (most commonly an amorphous underlayer

    followed by an oriented seed layer). This fact, combined

    with the curvature of the end of the tip, will mean that

    the stress configuration in the film is likely to be different

    from that in a standard layered film. In addition, in

    order to make a direct correlation of the structural and

    composition results obtained by 3DAP with magnetic or

    transport measurements, 3DAP analysis must be per-

    formed on layered films deposited on flat substrates

    because it is not possible to make standard magnetic and

    transport measurements on the films deposited on tips.

    It is also often desirable to investigate the microstructure

    of films that have been deposited in standard industrial

    deposition systems, rather than in special deposition

    chambers designed specifically to enable the study of

    films deposited onto pre-evaporated needles, as has also

    been pointed out by Lang and Schmitz [30]. In spite of

    these detriments, this technique has been widely usedand does have certain advantages.

    A preparation method that addresses some of the

    above issues has been reported by Hasegawa et al. for

    specimens containing a single layer alloy film [31]. In this

    method the film to be analysed is deposited on a sub-

    strate, such as a Si wafer, covered in photoresist. The

    film is then lithographically patterned to form small

    samples with a shape similar to that shown in Fig. 1(b).

    Following patterning the photoresist is dissolved, leav-

    ing the samples that are comprised of the layer(s) of

    interest. An individual sample must then be picked up

    and attached to the end of a sharpened metallic tip using

    conducting epoxy. Final shaping of the post section of

    the sample, which will form the specimen to be analysed,

    is carried out using a pulsed micropolishing technique.

    This method results in a specimen with the film(s) in its

    plane and has been used to investigate Co- and Fe-based

    magnetic films [3234].

    The Hasegawa technique does result in the fabrica-

    tion of a specimen from a layered film deposited on a flat

    substrate using a standard deposition system, but it

    requires the film material(s) to be deposited onto a

    non-standard seed layer and to be amenable to electro-

    chemical polishing. Problems arise when the different

    layers in the film require very different polishing solu-tions, as is the case for Cu and Co, two commonly used

    materials in information storage applications. In such

    cases, another method of specimen sharpening, such as

    ion beam milling, must be used.

    In principle, specimen preparation using ion milling

    [3538] has several advantages over electrochemical

    techniques [12,39]. These advantages include the elimi-

    nation or reduction of contamination and the reduction

    of preferential etching problems in multi-phase materi-

    als such as layered films. The application of ion milling

    to prepare 3DAP specimens from thin film(s) or near-

    surface regions has not been widely investigated. Liddle

    et al. [40], Larson et al. [41] and Kvist et al. [42] used a

    broad ion beam to mill without the capability to image

    the specimen during preparation. In-situ imaging and

    ion milling of atom probe specimens using a gallium

    liquid metal ion source was first reported by Waugh

    et al. [43] using an instrument with a focused ion spot

    size of$50 nm. Further work on ion milling of atomprobe specimens was performed by Alexander et al. [44]

    using a Gatan model 645 precision ion mill with a

    minimum spot size of approximately 1500 nm. Although

    the above investigations show the feasibility of concur-

    rent ion milling and secondary electron imaging, the full

    Fig. 1. (a) Schematic diagram of the method of thin film deposition

    onto curved sample surface, (b) schematic diagram of the modified

    lithographic specimen preparation method described by Hasegawa

    et al. [31] and Larson et al. [47] and (c) schematic diagram of the

    specimen geometry that results from silicon post method used by the

    current authors [50].

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2849

  • 8/8/2019 Larson Acta Mat Review Paper 04

    4/16

    capability of the new generation of commercial high

    resolution (

  • 8/8/2019 Larson Acta Mat Review Paper 04

    5/16

    There are, of course, disadvantages associated with

    ion milling methods. For example, milling may intro-

    duce implantation and defects into the near-surface

    region of the specimen but these can usually be re-

    moved by field evaporation before atom probe analysis

    proceeds. The FIB techniques described above lead to

    a region of the specimen that has been partially dam-aged by the Ga ion beam [48]. An example of the

    problems associated with the use of a focused ion beam

    for specimen fabrication using the modified Hasegawa

    technique is shown in Fig. 5. No layers were visible in

    the field ion image of a specimen containing a (Co2 nm/

    Cu2 nm)100 multilayer film milled using the FIB at

    10 keV [45]. Subsequent atom probe analysis showed

    the gallium content vs analysis depth to range from

    $6.5 at.% Ga initially to $0.5% Ga at a depth of$15nm, as seen in Fig. 5(a). For comparison, the results of

    analysis of a specimen fabricated using 30 keV gallium

    ions are also shown. The gallium implantation level

    was still 1020% even up to a depth of $20 nm intothe specimen. Fig. 5(b) shows atom maps of the Ga

    and Cu taken from the 3DAP analysis of the specimen

    milled at 10 kV and whose Ga content is shown in

    Fig. 5(a). The layered structure appears after $20 nmof field evaporation, as shown by 3DAP (note that

    Fig. 5(b) does not include the first 10 nm of the profile

    shown in Fig. 5(a)).

    The FIB-damaged region can be removed and mon-

    itored during 3DAP analysis. Alternatively, the dam-

    aged region can be removed by DC field evaporation

    while field ion imaging, or by low-energy ion sputtering

    using field emission in the presence of Ne gas [53]. Note

    that only analyses that are relatively free of Ga im-

    plantation (less than $12% Ga) should be consideredas representative of the original thin film microstructure.

    3. Studies of layered thin films

    Over the last decade or so, the three-dimensional

    atom probe technique has been applied to various lay-

    ered systems of interest for information storage mate-

    rials. These include Ni/Cu [21], Co/Cu [27,46], Ni/Al

    [29], CoFe/Cu [5456], NiFe/Cu [57], CoFe/NiFe [58],

    Fe/Cr [59], Co/Pd [60], as well as various magnetic ma-

    terial-based single film structures. The results of some of

    these studies are discussed in more detail below.

    3.1. Co/Cu and CoFe/Cu multilayer films

    The modified Hasegawa technique was used by

    the current authors to prepare specimens from an

    Fe/(Cu2 nm/Co2 nm)100 multilayer film. An Fe seed layer

    was used in these samples to impart a strong (1 1 1)

    Fig. 3. Silicon post (a) prior to pattern placement at low ion current (inset at high ion current), (b) the evolution of the end of the post as milling

    proceeds using the$1 lm inner radius pattern, (c) the post after milling has reduced the radius of the unmilled region of the specimen to$0.4 lm and(d) the post after milling at low current has reduced the radius of the unmilled region of the specimen to less than $0.1 lm.

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2851

  • 8/8/2019 Larson Acta Mat Review Paper 04

    6/16

    crystallographic texture to the MLF. The as-deposited

    films showed a magnetoresistance ratio of$5% at roomtemperature, and the magnetic data suggested that a

    significant fraction of the films were coupled ferromag-

    netically (contrary to the expectation of antiferromag-

    netic coupling for this copper interlayer thickness). The

    coercivity was $60 Oe, which is relatively high for filmsof this type. A field-ion image of the multilayer,

    Fig. 6(a), shows the brightly-imaging cobalt layers: in

    some regions the layers are relatively straight and in

    other regions the layers are non-parallel, with a wavi-

    ness amplitude of$2 nm and period of$20 nm [45].

    The specimen preparation method used in this workresults in layers that run parallel to the long axis of the

    specimen and the layers are thus viewed in cross-section

    in the field ion image. The waviness of the layers was

    also evident in the 3DAP data obtained from the same

    film, as seen in Figs. 6(b) (map of Cu atoms only) and

    Fig. 6(c), which shows the Co composition profile. The

    fact that the layers are wavy, as shown in Fig. 6, suggests

    that the ferromagnetic coupling is a result of Neel or-

    ange peel type magnetostatic coupling between adja-

    cent cobalt layers. In addition, in some areas adjacent

    cobalt layers appear to be in contact, which will further

    increase the ferromagnetic coupling. The relatively high

    coercivity is expected to be a result of the poor layer

    planarity, which leads to a high number of domain wall

    pinning sites.

    Co/Cu layered structures have also been studied by

    3DAP using the method of film deposition onto a pre-

    formed W needle [27]. In this work, a layered structure

    of NiFe25nm/Cu20nm/Co10nm was ion beam sputtered

    onto a W needle and analysed in the as-deposited and

    annealed (0.5 h at 350 C) states. Schleiwies at al. [27]

    observed an interfacial solid solution zone approxi-

    mately 1.5 nm in width at the interfaces in these samples

    (as-deposited), as well as segregation of Fe through the

    Fig. 4. (a) Side view of the specimen after the $0.4 lm inner radiuspattern stage of milling. The metal film(s) are visible as a bright region

    on the end of the specimen. (b) Low and (c) high magnification images

    of the final specimen shape. The radius is less than 100 nm and the

    region of interest for 3DAP analysis is positioned at the apex of the

    needle.

    Fig. 5. (a) Composition profile in Co/Cu multilayer film specimen

    showing the degree of Ga incorporation caused by FIB milling. (b)

    3DAP atom maps for Ga and for Co from final $20 nm depth ofspecimen milled at 10 kV showing onset of appearance of the Co/Cu

    layers following the Ga implantation damage. (Reproduced from [47]

    with permission from Institute of Physics Publishing.)

    2852 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862

  • 8/8/2019 Larson Acta Mat Review Paper 04

    7/16

    Cu layer to the Co/Cu interface after annealing,

    although no physical mechanism for this effect was

    proposed.

    Fig. 7(a) shows atom maps of Ni, Co and Cu from a

    3DAP dataset of a repeated (Ni20Fe5 nm/Co10Fe4 nm/

    Cu3 nm/Co10Fe4 nm) multilayer film fabricated from a

    sample deposited onto Si posts and prepared by the

    annular FIB milling method as described above [50]. As

    can be seen in the figure, for samples prepared in this

    geometry, the interfaces run across the apex of the

    specimen (normal to the axis) and thus the 3DAP

    technique can achieve atomic resolution across the in-

    terfaces, enabling accurate measurement of their

    roughness and intermixing. This is clearly seen in

    Fig. 7(b), which shows the (1 1 1) atomic planes across a

    NiFe/CoFe bilayer, indicating the high degree of crys-

    tallographic texture in the sample. In addition, micro-

    structural results from specimens prepared in this

    geometry can be compared directly with devices grown

    for magnetic and transport measurements.

    Fig. 8(a) shows an atom map indicating the atomic

    positions of the Co (blue), Fe (yellow) and Cu (red)

    atoms in a CoFe/Cu/CoFe trilayer (a subsection of the

    multilayer structure described in the previous para-graph). The atom map shows qualitatively that there is

    more intermixing at the CoFe-on-Cu interface than

    there is at the Cu-on-CoFe interface, resulting in the

    CoFe-on-Cu interface being wider than the Cu-on-CoFe

    interface. However the strength of the 3DAP technique

    lies in the fact that in addition to being able to show

    such effects in a qualitative manner, the extent of the

    interdiffusion can be measured quantitatively by taking

    composition profiles through small sections of the data

    perpendicular to the local interface plane. Fig. 8(b)

    shows a composition profile across the same trilayer

    region. The direction of growth is from left to right in

    the profile. Measurement of the interface widths from

    the composition profile (using 1090% of the Cu con-

    centration) gave values of 1.08 (0.18) nm for the CoFe-on-Cu interface and 0.4 (0.14) nm for the Cu-on-CoFeinterface [61].

    3.2. Comparison of 3DAP data with modelling

    Molecular dynamics (MD) simulation of the growth

    of the same materials system [61], Fig. 9, supports the

    3DAP results presented above. As can be seen, the

    match between the 3DAP data and the simulation is

    extremely good, with the MD simulation giving a widthfor the CoFe-on-Cu interface of$1.44 nm and for theCu-on-CoFe interface of $0.33 nm. Further 3DAPquantitative analyses of this type gave values of 0.82

    (0.10) and 0.47 (0.15) nm for the CoFe-on-Cu andCu-on-CoFe interfaces, respectively [54,58]. The differ-

    ence in interface width is believed to be the result of the

    lower surface free energy of Cu compared to that of Co

    or Fe [62]. The molecular dynamics modelling suggests a

    difference in the exchange probability during deposition

    of Cu with Co and Fe at the two interfaces, with the

    exchange probability at the CoFe-on-Cu interface being

    higher, resulting in more mixing at the interface. Of

    course, this will result from a combination of driving

    forces (due to the lower surface energy for Cu) and the

    lower activation barrier for exchange (due to the lower

    melting temperature of Cu), and it is difficult to distin-

    guish the two effects in the model. Note that the model

    only takes into account rearrangements resulting from

    atomic impacts, and does not cover the much longer

    time scale over which thermally activated surface and

    bulk diffusion may occur and contribute to intermixing.

    However, since the agreement between model and ex-

    periment is so good, such effects are not likely to con-

    tribute greatly to the observed intermixing, at least at

    Fig. 6. (a) Field ion microscope image showing Co (bright) and Cu

    (dark) layers, (b) 3DAP atom map of the Co/Cu multilayer showing

    degree of curvature of the Co and Cu layers (Cu atoms only for clarity)

    and (c) selected-region Co composition profile across the layers.

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2853

  • 8/8/2019 Larson Acta Mat Review Paper 04

    8/16

    the deposition rates (%0.1 nms1) and substrate tem-peratures (300330 K) used here. The good agreement

    between the simulated and experimental data shows that

    3DAP analysis is really the first type of experiment that

    can be used to check this model at the correct length

    scale. A favourable comparison between model and

    experiment also served to validate the MD simulation

    parameters used for these materials and has encouraged

    both research groups to proceed with further compari-

    sons, as described below.

    The data in Figs. 8 and 9 show a layered film in which

    the top surface of the CoFe layer is relatively flat.

    However growth of multilayer films with a strong [1 1 1]

    crystallographic texture often results in the top surface

    of the layers being rough, or wavy, with a period of a

    few nanometres. In many multilayer systems this peri-

    odicity is transferred through to subsequently deposited

    layers, resulting in correlated roughness that usually

    becomes worse as more layers are deposited [8]. An

    example of this is shown in Fig. 10. The roughness at the

    top of the NiFe layer has been transferred up to the next

    CoFe layer, as can be seen from the isoconcentration

    surface (surface of constant concentration of one or

    more elements) shown in Fig. 10(a), plotted at a com-

    position of$45 at% Co. A very interesting feature is the

    fact that the upper surface of the Cu layer is not con-formal with the upper surface of the underlying CoFe

    the deposition of the Cu has acted to smooth out the

    correlated roughness [54]. This smoothing effect had

    been suggested by Eckl et al. [7] on the basis of in-situ

    resistivity measurements but was not confirmed by a

    microstructural study.

    The question then arises as to how the Cu grows in

    order to smooth out the roughness. Eckl et al. pro-

    posed that the initial Cu layer was deposited confor-

    mally on the underlying Co, followed by nucleation of

    islands of Cu in the troughs at the grain boundaries.

    While the 3DAP data in Fig. 10 show the Cu

    smoothing effect, the actual growth process cannot be

    understood just from the 3DAP data as the technique

    does not allow us to follow the process as a function

    of time. However, this can be achieved using the same

    MD simulation technique as discussed above [61].

    Fig. 11 shows a series of images obtained during a

    MD simulation of the growth of a Cu layer on a

    rough CoFe layer [63]. Fig. 11(a) shows that the initial

    monolayers of Cu are indeed deposited conformally

    on the CoFe surface, as proposed by Eckl et al. The

    MD simulations indicate that as the Cu layer gets

    thicker, facets with (11 1) crystallographic planes

    Fig. 7. (a) 3DAP atom maps of Ni, Co and Cu atoms in the multilayer (volume is $20 nm$20 nm$35 nm), (b) selected-region compositionprofile across two of the layers shown in (a).

    2854 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862

  • 8/8/2019 Larson Acta Mat Review Paper 04

    9/16

    Fig. 8. (a) 3DAP atom map of Co (blue), Fe (yellow) and Cu (red)

    atoms in a CoFe/Cu/CoFe trilayer and (b) a composition profile across

    the same region (growth direction is left to right in the profile). (Re-

    produced from [61] with permission from Elsevier Publishing.)

    Fig. 9. (a) Molecular dynamics simulation showing an atom map of Co

    (blue), Fe (yellow) and Cu (red) atoms in a a CoFe/Cu/CoFe trilayer

    and (b) a composition profile across the same region (MD-simulated

    growth direction is left to right in the profile). (Reproduced from [61]

    with permission from Elsevier Publishing.)

    Fig. 10. (a) Isoconcentration surface at $45% Co for a NiFe/CoFe/Cu/CoFe multilayer, and (b) and (c) 3DAP atom maps of Co (blue), Fe

    (yellow) and Cu (red) atoms in the same region. (b) Shows the first few

    atoms of the Cu layer and (c) shows the structure up to the point at

    which deposition of the Cu ends.

    Fig. 11. Molecular dynamics simulation [63] of the growth of a thin Cu

    layer onto a rough CoFe layer. (a) The first few monolayers of Cu

    grown conformally on the CoFe and (b) and (c) further Cu growth

    results in a smooth film surface.

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2855

  • 8/8/2019 Larson Acta Mat Review Paper 04

    10/16

    begin to form at the peaks and troughs of the un-

    dulations. The (1 1 1) surfaces are low-energy surfaces

    and to maximise their area, Cu atoms lying on the

    sides of the troughs diffuse across the growing surface

    and fill in the troughs as seen in Figs. 11(b) and (c).

    The final result is that the upper surface of the Cu

    layer is much flatter than the CoFe surface on whichit grew. Both this MD simulation and the 3DAP data

    (Fig. 10(b)) show that deposition of Cu onto a rough

    CoFe layer results in relatively little intermixing with

    the CoFe. When the next CoFe layer is deposited, the

    enhanced interdiffusion at the CoFe-on-Cu interface

    serves to remove any remaining correlated roughness

    at the upper surface of the Cu layer, as shown in

    Fig. 10(c). The combination of the 3DAP data and

    the MD simulation can thus be used to provide a full

    picture of the way in which the growth of the Cu

    layer smoothes out the CoFe roughness and at the

    same time yields an intermixed upper Cu surface.

    3.3. Effect of oxygen during growth on thin film micro-

    structure

    One factor that can significantly affect the GMR

    properties of thin films structures is the presence of

    impurities, such as oxygen, in the sputtering chamber

    during deposition. Different proposals have been put

    forward to explain this effect, such as a smoothing of

    the interfaces during growth via a reduction of surface

    free energy [64], a reduction in pinhole density across

    the Cu spacer layers [65] and a reduction in grain size

    and interface roughness caused by partial oxidation ofthe interfaces [66]. However, to date, relatively little

    nanoscale characterisation has been carried out to

    support these suggestions. There are therefore several

    issues to be clarified with respect to oxygen-doped

    multilayer growth such as the amount and position

    of residual oxygen trapped in the layered structures

    and the exact effects of oxygen on the nature of the

    interfaces.

    The 3DAP has been used to investigate the thin

    film microstructure resulting from the addition of an

    oxygen surfactant during growth of the Cu layers in

    {Si//Ni alloy5 nm/Co10Fe3 nm/(Cu1:8 nm/Co10Fe3 nm)5/

    cap50nm} films [55]. In this structure, the Ni-based seed

    layer leads to a h1 1 1i crystallographic orientation inthe films, and this can be seen in the field ion image of

    the multilayer in Fig. 12 which shows strong crystal-

    lographic orientation along the specimen axis, together

    with with the bright (CoFe) and dark (Cu) contrast

    regions which indicate the layers in the structure. The

    sputter gas pressure during deposition of the Cu and

    CoFe layers was of the order of several mTorr, and

    deposition of the Cu layers was carried out either in Ar

    or in an Ar + oxygen gas mixture with the total pres-

    sure being kept constant.

    Magnetoresistance values measured on films depos-

    ited on unpatterned samples (using a standard four-

    point probe measurement) were found to be $1% and$7% for the Ar and Ar + oxygen Cu deposition condi-tions, respectively, as shown in Fig. 13. Fig. 14 shows

    composition profiles across the CoFe/Cu layers for de-

    positions (a) without and (b) with oxygen present for

    deposition of the Cu layers [55]. The most obvious dif-

    ference is the degree of intermixing in the Cu layers,

    which is considerably reduced for the Ar + oxygen de-

    position, resulting in a higher Cu content at the centre of

    the layers. As expected from the surface free energy

    argument discussed above, the CoFe-on-Cu interface

    shows more intermixing than the Cu-on-CoFe interface,

    and the effect of the oxygen has been to reduce the width

    of both interfaces (see Table 1). See [55] for further de-

    tails of the interfacial width measurements in these

    structures.

    Fig. 12. Field ion image of four repeats of the CoFe/Cu bilayer

    showing the expected [1 1 1] crystallographic texture.

    0

    1

    2

    3

    4

    5

    6

    7

    -1500 -1000 -500 0 500 1000 1500

    H (Oe)

    dR/R(%)

    O-doped as-grownO-doped,annealedNo-O, as-grownNo-O,annealed

    Fig. 13. GMR curves for as-deposited and annealed CoFe/Cu multi-

    layer films for which the Cu layers were deposited either in Ar or in

    Ar + oxygen.

    2856 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862

  • 8/8/2019 Larson Acta Mat Review Paper 04

    11/16

    As can be seen from Table 1, the Ar + oxygen depo-

    sition has resulted in a decrease in the width of the

    CoFe-on-Cu interface by a factor of approximately two

    relative to the deposition without oxygen. This clearlyshows that the presence of oxygen suppresses mixing

    during growth. Egelhoff et al. [67] have suggested that

    this occurs because the oxygen favours the formation of

    CoO (or FeO) bonds rather than CuO bonds (the

    heats of formation of the oxides are: CoO)58 kcal/mol and CuO)38 kcal/mol [68]). The lowest energystate thus has Co at the surface rather than Cu, leading

    to less interfacial mixing. This argument would tend to

    suggest that the width of the Cu-on-CoFe interface

    might increase in the presence of oxygen. However the

    higher cohesive energy of Co compared to that of Cu

    may restrict the atomic exchange [61] at the Cu-on-CoFe

    interface, so that no increase in intermixing is observed.

    Indeed the intermixing is found to be reduced slightly,

    which suggests that the presence of oxygen on the

    growing surface reduces the number of atomic ex-

    changes at the surface. This will also contribute to

    the reduction in interface width at the CoFe-on-Cu

    interface.

    Another interesting observation is that there is anincrease of only approximately a factor of 2 in the re-

    sidual oxygen level (average of $260 appm oxygen forthe no-oxygen condition compared to an average of

    $530 appm oxygen) for the Ar + oxygen condition [55].The oxygen present in the films grown without added

    oxygen is most likely a result of the dissociation of water

    vapour in the deposition chamber and of oxygen in-

    corporated into the sputter deposition targets. The rea-

    son that the incorporated oxygen content increases only

    by a factor of two, despite a large difference in the ox-

    ygen pressure in the chamber, may be the very low

    solubility of O in Cu and Co [69]. Further evidence for

    the low incorporation of oxygen for the Ar + oxygen

    deposition can be seen in the elemental map shown in

    Fig. 15. The oxygen atoms are not positioned prefer-

    entially at the layer interfaces but are distributed rela-

    tively randomly through the layers, and there is no

    evidence for the formation of an oxide phase in these

    samples, as has been suggested in the literature [66].

    However, there is some evidence for preferential reten-

    tion of oxygen near grain boundaries, as previously

    suggested [64].

    A further effect of the presence of oxygen during

    sputtering is a decrease in the conformal roughness of

    the layers, as observed previously by Miura et al. [66],which would also be expected to increase the GMR

    ratio. An estimate of the conformal interfacial rough-

    ness was made from TEM cross-section images by

    measuring the maximum peak-to-peak amplitude of the

    spatial variations in each layer. The mean amplitude

    values (averaged across all five bilayers) are 1.89 (0.08)nm and 1.62 (0.08) nm for the Ar and Ar + oxygendeposition conditions, respectively [55]. The presence of

    oxygen is believed to reduce long-range surface diffusion

    [70] and thus decrease grain boundary grooving this

    would drive the system away from equilibrium and thus

    lead to decreased grooving, as observed in the TEM

    data.

    Fig. 14. Composition profiles across CoFe/Cu multilayers for deposi-

    tions of the Cu layers in (a) Ar and (b) Ar + oxygen. The growth di-

    rection is from right to left in both profiles. (Reproduced from [55]

    with permission from The American Physical Society.)

    Table 1

    Interface widths for CoFe/Cu MLFs for which the Cu layers were deposited either in Ar or in an Ar + oxygen mixture, for the as-deposited films and

    for films annealed at 300 C for 30 min

    Cu deposition condition Interface widths (nm) Ave. Cumax Conc. (%)

    CoFe-on-Cu Cu-on-CoFe

    Ar as-grown 1.21 (0.05) 0.49 (0.03) 65.113.3A r + O2 as-grown 0.55 (0.03) 0.31 (0.02) 77.111.72Ar (300 C/0.5 h) 1.16 (0.09) 0.61 (0.02) 72.17.6A r + O2 (300 C/0.5 h) 3.01 (0.12) 2.38 (0.12) 55.67.4

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2857

  • 8/8/2019 Larson Acta Mat Review Paper 04

    12/16

    3.4. Annealing studies

    Diffusivity in Ni/Cu layered films (as-deposited and

    annealed) has been studied by Keilonat et al. [21]. Ac-

    tivation energies for interdiffusion, as a function of Cuconcentration, were measured and compared well with

    previous literature data. Interfacial mixing in NiFe/Cu

    multilayers was reported using 3DAP by Larson et al.

    [57]. Highly intermixed interfacial regions at the NiFe-

    on-Cu and Cu-on-NiFe interfaces were reported al-

    though the geometry of the samples was such that the

    interfaces were parallel to the specimen axis (Fig. 1(b))

    so that the accuracy of the measurements can only be

    considered to be $1 nm.The interaction of an Al film deposited onto Ni (using

    ion beam sputtering and electron beam evaporation) has

    been studied by Jeske et al. [29] also using the method of

    film deposition onto a preformed needle of Ni or W. It

    should be noted that although Ni/Al layers are not of

    interest per se in magnetic information storage systems,

    the formation of the aluminium oxide tunnel barriers in

    spin tunnel junction structures [3] is usually carried out

    by depositing metallic Al on the underlying ferromag-

    netic layer (such as a Ni, Fe or Co alloy) and then ox-

    idising. Interdiffusion of the Al and the underlying layer

    can therefore occur, which can lead to a decrease in the

    spin polarisation in the ferromagnetic layer. In the as-

    deposited state, Jeske et al. identified an initial solid

    solution zone about 1 nm in width in both types of

    samples. After annealing, an off-stoichiometry B2 NiAl

    phase was observed to form in samples formed by

    electron beam deposition of an Al layer onto a Ni nee-

    dle, in which no grain boundaries were observed.However, in samples where a Ni/Al bilayer was sputter

    deposited onto a W needle, NiAl3 regions were observed

    to form at grain boundaries in the Al layer. The authors

    ascribed this to a difference in the flux of Ni into and out

    of the thin inter-reaction layer at the Al/Ni interfaces.

    In films containing grain boundaries, rapid diffusion

    of Ni along the grain boundaries to the NiAl3 precipi-

    tates leads to depletion of nickel in the interfacial region

    and thus inhibits the formation of the intermetallic

    phase [29].

    Annealing of the Cu2 nm/Co2 nm multilayer films de-

    scribed above has been studied by the current authors.

    Samples that were annealed for 1h at 360 C showed an

    increase in grain size but no apparent change in the in-

    terlayer mixing between Cu and Co [71]. Studies of the

    CoFe/Cu films deposited in the presence of oxygen,

    following annealing, tend to confirm that the presence of

    oxygen drives the system further from equilibrium. This

    can be seen from the changes in the GMR ratio for the

    two types of film, as shown in Fig. 13. For the film de-

    posited in Ar, the decrease in GMR ratio following

    annealing at 300 C for 30 min was from 1% to 0.5%,

    whereas for the film deposited in Ar + oxygen, the de-

    crease was from 7% to 0.2%. The changes to the trans-

    Fig. 15. Atom map from 3DAP data for the as-deposited CoFe/Cu

    MLF deposited using Ar + oxygen. The oxygen atoms (large, spheres)

    and Cu atoms (small, dark spheres) are shown. The presence of a

    columnar grain boundary is indicated. (Reproduced from [55] with

    permission from The American Physical Society.)

    0

    10

    2030

    40

    50

    60

    70

    80

    90

    100

    0 1 2 3 4 5 6 7 8

    Distance (nm)

    Con

    centration(at%)

    %Co %Cu

    0

    10

    20

    30

    40

    50

    60

    70

    0 5 10 15 20 25 30 35

    Distance (nm)

    Concentration(at%)

    Co% Cu%

    (a)

    (b)

    Fig. 16. Composition profiles (showing only Co and Cu) across CoFe/

    Cu multilayer films following annealing at 300 C for 30 min (a) for

    deposition of Cu layers in Ar, and (b) for deposition of Cu layers in

    Ar + oxygen.

    2858 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862

  • 8/8/2019 Larson Acta Mat Review Paper 04

    13/16

    port properties can be accounted for by changes to the

    microstructure that occur during annealing. Fig. 16

    shows composition profiles taken from 3DAP analyses

    across annealed CoFe/Cu multilayer films with the Cu

    layers deposited (a) in Ar and (b) in Ar + oxygen. As can

    be seen from Table 1, the increase in interfacial mixing is

    relatively small for the film deposited in Ar (the Culayers are still well defined), but for the film deposited in

    the presence of oxygen, the CoFe-on-Cu interface width

    increases by a factor of 5, and the Cu-on-CoFe in-terface by a factor of 6.8. This is observed even thoughthe distribution of the oxygen surfactant does not ap-

    pear to change after annealing, as can be seen from a

    comparison of the atom maps shown in Fig. 15 (as-de-

    posited) and Fig. 17 (following annealing at 300 C for

    30 min). It is likely that in the layers that are grown with

    oxygen, and thus are further from equilibrium, the re-

    duced conformal roughness provides a driving force for

    intermixing at the CuCoFe interfaces to occur during

    annealing. On the other hand intermixing in the layers

    grown without oxygen appears limited to that which

    occurs by atomic exchange during sputter deposition

    [61]. However it is not clear from our results whether the

    intermixing has occurred by long-range changes in the

    interface position in order to establish the equilibriumshape, or whether the non-equilibrium configuration

    raises the free energy of the system sufficiently to make

    the CoFe and Cu layers miscible [56].

    3.5. Media materials

    New information storage systems require not only the

    development of new read/write heads but also of new

    storage media. The media used currently in hard disk

    drives consist of CoCr alloys (to which additional al-

    loying elements such as Ta are frequently added) in

    which the information is stored with the magnetisation

    vectors in the plane of the film (longitudinal media). The

    high storage density is achieved in the media by having a

    polycrystalline film in which the grains are small and are

    ideally magnetically decoupled from each other. Mag-

    netisation studies of the films suggested that this occurs

    because of compositional inhomogeneities in the CoCr

    alloy layer [72] and this was confirmed by TEM studies

    [73]. However the exact way in which the inhomogenei-

    ties were distributed in three-dimensions had not been

    well understood. The grain size in this type of material is

    typically of the order of a few nanometres and this is

    therefore an ideal topic to be addressed by 3DAP.

    A small number of atom probe field ion microscopystudies had been made of CoCr alloy thin films, both by

    deposition of the films on curved surfaces [22] and using

    the Hasegawa method [32] and more recently studies

    have been made using 3DAP. Hono et al. [74] used the

    Hasegawa method to prepare thin films of a Co

    22at%Cr alloy that had been deposited on a heated

    substrate, as would be the case for the thin film media

    layer on a hard disk. Their results showed that the films

    consisted of a ferromagnetic phase and a paramagnetic

    Cr-rich phase, both with a lamellar-type structure, which

    coexist within each grain of the film. The lamellar shape

    of the Co-rich ferromagnetic phase results in the particles

    having a high shape anisotropy which increases their

    thermal stability, and the presence of the paramagnetic

    phase acts to isolate the ferromagnetic lamellae mag-

    netically from each other. Further work on a similar

    system was carried out by Nishikima et al. [75] on a Co

    12Cr2Ta alloy thin film deposited at elevated temper-

    ature. The addition of the Ta to the alloy resulted in

    segregation of the Cr to the grain boundaries, leading to

    Co-rich ferromagnetic grains, as shown in the Co con-

    centration map in Fig. 18(a), enveloped by a Cr-rich non-

    magnetic phase, seen in Fig. 18(b), but still with some

    non-uniformity in the Cr composition within the grains.

    Fig. 17. Atom map from 3DAP data for the CoFe/Cu MLF deposited

    using Ar+ oxygen, following annealing at 300 C for 30 min. The

    oxygen atoms (large, pale spheres) and Cu atoms (small, dark spheres)

    are shown.

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2859

  • 8/8/2019 Larson Acta Mat Review Paper 04

    14/16

    The authors proposed that the enhanced segregation

    occurred because the Ta atoms are larger than both Co

    and Cr and thus reduced the migration energy of the Cr

    to the boundaries. The Ta was found to be uniformly

    distributed through the film, as seen in Fig. 18(c), and did

    not segregate to the grain boundaries with the Cr.

    One recent innovation has been the use of perpen-

    dicular magnetic storage in which the bits of informa-

    tion are written into a magnetic layer with the

    magnetization vectors pointing perpendicular to the

    plane of the magnetic film [76]. This allows for a con-

    siderably higher information density than can be

    achieved with conventional longitudinal storage media

    [77]. A number of materials are currently being consid-

    ered for perpendicular storage media, one of the most

    popular being a Co/Pt or Co/Pd multilayer film in which

    the Co layers are very thin (typically

  • 8/8/2019 Larson Acta Mat Review Paper 04

    15/16

    composition profile taken from 3DAP data for the Co/

    Pd multilayer film. Four Pd layers and three Co layers

    are visible plus part of the NiFe cap (which shows atomic

    planes). As can be seen, the Co and Pd layers are not

    clearly differentiated; the Co concentration varies from

    $20% to $40% across the Co layers. The binary phasediagram for CoPd [69] shows complete solubility be-

    tween the two elements and, in addition, the deposition

    was carried out in a relatively low Ar pressure, leading to

    high adatom energy and increased intermixing. Thus it is

    not entirely surprising that a high degree of intermixing

    was observed in the Co/Pd samples. However, such a

    high degree of intermixing was unexpected from a con-

    sideration of the magnetic data, which show a degree of

    perpendicular anisotropy greater than would be expected

    without inclusion of an interfacial contribution arising

    from the Co/Pd interfaces.

    4. Summary and conclusions

    The aim of this paper has been to show that 3DAP

    analysis is a technique that can be very effectively ap-

    plied to the understanding of the correlation between

    microstructure, composition and physical properties in

    thin films with applications in information storage. The

    technique is particularly appropriate to films in which

    the composition is inhomogeneous on the nanometre

    scale, for example because the film consists of thin lay-

    ers, or because of chemical segregation at grainboundaries, and for which the properties can vary

    greatly even with atomic scale differences in chemical

    distribution. The validity of the technique has been

    demonstrated by comparison with molecular dynamics

    models of the growth mode in CoFe/Cu layered films,

    which allows us to place reliance on the experimental

    data that are obtained using 3DAP.

    However, as for all techniques, there are outstanding

    issues with using 3DAP for the analysis of thin films, of

    which one is specimen preparation and another is the size

    of the analysis volume. In order to make valid correla-

    tions of the composition profile and microstructure with

    the magnetic and transport properties, the authors be-

    lieve that it is crucial to analyse films that have been de-

    posited on planar substrates. As has been illustrated

    above, this is now an achievable goal, and the results that

    have been obtained have been invaluable in enabling

    phenomena such as the effect of surfactants on GMR in

    CoFe/Cu multilayers to be elucidated. The analysis vol-

    ume in 3DAP remains limited, although novel instrument

    designs are currently helping to address this issue too. In

    any case, careful comparison with other techniques such

    as TEM can ensure that the data that are obtained using

    3DAP are typical of the structure as a whole.

    In conclusion, we believe that atomic-scale nano-characterisation using 3DAP, possibly combined with

    molecular dynamics modeling, provides a unique tool

    that can be applied both to the design of nanostructured

    magnetic materials for information storage applications

    and to the further understanding of the fundamental

    physical properties on which their operation relies.

    Acknowledgements

    The authors thank Drs. T.P. Nolan, P.H. Clifton,

    S.P. Bozeman, H.L. Brown, E.W. Singleton, R. Grove,

    R. Hipwell (Seagate Technology LLC), M. Abraham, O.Dimond, R.M. Langford and G.D.W. Smith FRS

    (University of Oxford) for their various contributions to

    this research. The authors especially thank Dr. X.W.

    Zhou (University of Virginia) for his provision of the

    MD model results on a rough substrate. We are also

    grateful to the EPSRC for funding.

    References

    [1] Baibich MN, Broto JM, Fert A, Nguyen Van Dau F, Petroff F,

    Eitenne P, et al. Phys Rev Lett 1988;61:2472.

    Fig. 20. 3DAP data from Co/Pd multilayer film deposited on Si at

    room temperature showing (a) the positions of Co atoms (large dark

    spheres), Pd atoms (large light spheres) and the Ni capping layer (small

    dark dots) and (b) the corresponding composition profile.

    D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862 2861

  • 8/8/2019 Larson Acta Mat Review Paper 04

    16/16

    [2] Parkin SSP. Annu Rev Mater Sci 1995;25:357.

    [3] Comstock RL. J Mater Sci Mater Elec 2002;13(9):509.

    [4] Dieny B, Speriosu VS, Metin S, Parkin SSP, Gurney BA,

    Baumgart P, et al. J Appl Phys 1991;69:4774.

    [5] Kudrnovsky J, Drchal V, Turek I, Blaas C, Weinberger P, Bruno

    P. J Met 2000;52(7):29.

    [6] Kief MT, Egelhoff WF. Phys Rev B 1993;47(16):10785.

    [7] Eckl T, Reiss G, Bruckl H, Hoffmann H. J Appl Phys

    1994;75(1):362.

    [8] Payne AP, Clemons BM. Phys Rev B 1993;47(4):2289.

    [9] Zabel H, Theis-Brohl K. J Phys Condens Mater 2003;15:S505.

    [10] Portier X, Petford-Long AK. J Phys D: Appl Phys 1999;32:R89.

    [11] Hytch MJ, Walls MG, Chevalier J-P. Ultramicroscopy

    2000;83:217.

    [12] Miller MK, Smith GDW. Atom probe microanalysis: principles

    and applications to materials problems. Pittsburgh: Materials

    Research Society; 1989.

    [13] Miller MK, Cerezo A, Hetherington MG, Smith GDW. Atom

    probe field ion microscopy. Oxford: Oxford University Press; 1996.

    [14] Miller MK. Atom probe tomography. New York: Kluwar

    Academic/Plunum Publishers; 2000.

    [15] Cerezo A, Larson DJ, Smith GDW. MRS Bull 2001;26(2):102.

    [16] Melmed AJ. J Vac Sci Technol B 1991;9:601.

    [17] Cerezo A, Hetherington MG, Petford-Long AK. J Phys 1989;50-

    C8:349.

    [18] Petford-Long AK, Cerezo A, Hyde JM. Ultramicroscopy

    1992;47:367.

    [19] Melmed AJ. J Phys 1989;50-C8:547.

    [20] Camus PP, Shull RD, Melmed A. J Phys 1989;50-C8:343.

    [21] Keilonat C, Camus E, Wanderka N, Wollenberger H. Appl Phys

    Lett 1994;65(16):2007.

    [22] Pundt A, Michaelsen C. App Surf Sci 1995;8788:264.

    [23] Al-Kassab T, Macht M-P, Wollenberger H. Appl Surf Sci

    1995;8788:327.

    [24] Al-Kassab T, Macht M-P, Naundorf V, Wollenberger H, Cham-

    breland S, Danoix F, et al. Appl Surf Sci 1996;9495:306.

    [25] Jeske T, Schmitz G, Kirchheim R. Mater Sci Eng A 1999;270:64.

    [26] Veiller L, Danoix F, Teillet J. J Appl Phys 2000;87(3):1379.[27] Schleiwies J, Schmitz G, Heitmann S, Hutten A. Appl Phys Lett

    2001;78:3439.

    [28] Schleiwies J, Schmitz G. Mater Sci Eng A 2002;327:94.

    [29] Jeske T, Schmitz G. Mater Sci Eng A 2002;327:101.

    [30] Lang Ch, Schmitz G. Mater Sci Eng A 2003;353:119.

    [31] Hasegawa N, Hono K, Okano R, Fujimori H, Sakurai T. Appl

    Surf Sci 1993;67:407.

    [32] Hono K, Maeda Y, Li J-L, Sakurai T. Appl Surf Sci 1993;67:386.

    [33] Hono K, Hasegawa N, Babu SS, Fujimori H, Sakurai T. Appl

    Surf Sci 1993;67:391.

    [34] Hono K, Maeda Y, Babu SS, Sakurai T. J Appl Phys

    1994;76:8025.

    [35] Walls JM, Southworth HN, Rushton GJ. Vacuum 1974;24:475.

    [36] Henjered A, Norden H. J Phys E 1983;16:617.

    [37] Hellsing M. Mater Sci Technol 1988;4:824.[38] Larson DJ, Russell KF, Cerezo A. J Vac Sci Technol B

    2000;18(1):328.

    [39] Barber DJ. J Mater Sci 1970;5:1.

    [40] Liddle JA, Norman A, Cerezo A, Grovenor CRM. J Phys

    1988;49-C6:509.

    [41] Larson DJ, Teng C-M, Camus PP, Kelly TF. Appl Surf Sci

    1995;8788:446.

    [42] Kvist A, Andren H-O, Lundin L. Appl Surf Sci 1996;9495:356.

    [43] Waugh AR, Payne S, Worrall GM, Smith GDW. J Phys 1984;45-

    C9:207.

    [44] Alexander KB, Angelini P, Miller MK. J Phys 1989;50-C8:549.

    [45] Larson DJ, Petford-Long AK, Cerezo A, Smith GDW, Foord

    DT, Anthony TC. Appl Phys Lett 1998;73(8):1125.

    [46] Larson DJ, Petford-Long AK, Cerezo A, Smith GDW. Acta

    Mater 1999;47:4019.

    [47] Larson DJ, Foord DT, Petford-Long AK, Cerezo A, Smith

    GDW. Nanotechnology 1999;10:45.

    [48] Larson DJ, Foord DT, Petford-Long AK, Liew H, Blamire MG,

    Cerezo A, et al. Ultramicroscopy 1999;79:287.

    [49] Schleiwies, J. Doctoral Thesis, Universitat Gottingen; 2001.

    [50] Larson DJ, Wisman BD, Villieux RJ, Martens RL, Gribb TT,

    Erskine HF, et al. Microsc Microanal 2001;7:24.

    [51] Ayon AA, Braff R, Lin CC, Swain HH, Schmidt MA. J

    Electrochem Soc 1999;146(1):339.

    [52] Martens RL, Larson DJ, Kelly TF, Cerezo A, Clifton PH, Tabat

    N. Microsc Microanal 2000;6:522.

    [53] Larson DJ, Cerezo A, Ma YQ, Godfrey TJ, Clifton PH, Smith

    GDW. Microsc Microanal 2003;9(Suppl 2):578.

    [54] Larson DJ, Cerezo A, Clifton PH, Petford-Long AK, Martens

    RL, Kelly TF, et al. J Appl Phys 2001;89(11):7517.

    [55] Larson DJ, Petford-Long AK, Cerezo A, Bozeman SP, Morrone

    A, Ma YQ, et al. Phys Rev B 2003;67:144420.

    [56] Ladwig PF, Olson J, Bunton J, Larson DJ, Bonsager MC, Ulfig

    R, et al. Microsc Microanal 2003;9(Suppl 2):48.

    [57] Larson DJ, Petford-Long AK, Cerezo A, Foord DT, Anthony

    TC, Miller MK. Microsc Microanal 1999;5:150.

    [58] Larson DJ, Clifton PH, Tabat N, Cerezo A, Petford-Long AK,

    Martens RL, et al. Appl Phys Lett 2000;77:726.

    [59] Petford-Long AK, Doole R, Cerezo A, Jakubovics JP, Conyers

    JS. J Magn Magn Mater 1993;26:117.

    [60] Baker AM, Cerezo A, Petford-Long AK. J Magn Magn Mater

    1996;156:83.

    [61] Zhou XW, Wadley HNG, Johnson RA, Larson DJ, Tabat N,

    Cerezo A, et al. Acta Mater 2001;49:4005.

    [62] Kittel C. Introduction to solid state physics. 6th ed. New York:

    Wiley; 1986.

    [63] Zhou XW. Private communication; 2002.

    [64] Egelhoff WF, Chen PJ, Powell CJ, Stiles MD, McMichael RD,

    Judy JH, et al. J Appl Phys 1997;82(12):6142.[65] Kubinski DJ, Parsons M, Hangas J. MRS Symp Proc

    2001;614:F.8.8.1.

    [66] Miura S, Tsunoda M, Takahashi MJ. Appl Phys

    2001;89(11):6308.

    [67] Egelhoff WF et al. Prog Surf Sci 2001;67:355.

    [68] Weast RC, editor. CRC handbook of chemistry and physics. 63rd

    ed. Boca Raton, FL: CRC Press; 1983.

    [69] Massalski TB, editor. Binary alloy phase diagrams. Materials

    Park, OH: Massalski; 1990.

    [70] Egelhoff WF, Steigerwald DA. J Vac Sci Technol A

    1989;73(3):2167.

    [71] Larson DJ, Petford-Long AK, Cerezo A. Mater Sci Eng A

    1999;270:69.

    [72] Smits JW, Luitjens SB, den Broeder FJA. J Appl Phys

    1984;55:2260.[73] Maeda Y, Hirono S, Asahi M. Jpn J Appl Phys 1985;24:L951.

    [74] Hono K, Yeh K, Maeda Y, Sakurai T. Appl Phys Lett

    1995;66(13):1686.

    [75] Nishimaki J, Hono K, Hasegawa N, Sakurai T. Appl Phys Lett

    1996;69(20):3095.

    [76] Ouchi K. IEICE Trans Electron 2001;E84C(9):1121.

    [77] Weller D, Doerner MF. Annu Rev Mater Sci 2000;30:611.

    [78] Greaves SJ, Petford-Long AK, Kim Y-H, Pollard RJ, Grundy PJ,

    Jakubovics JP. J Magn Magn Mater 1992;113:63.

    [79] Larson DJ, Ma YQ, Petford-Long AK, Cerezo A, Barnes JP,

    Nolan TP. Presented at Intermag 03, Boston; April 2003.

    2862 D.J. Larson et al.: Overview No. 138 / Acta Materialia 52 (2004) 28472862