Half Metals and Their Properties
Half Metals and Their Properties
Half Metals and Their Properties
March 2, 2010
This document was prepared as an account of work sponsored by an agency of the United States
government. Neither the United States government nor Lawrence Livermore National Security, LLC,
nor any of their employees makes any warranty, expressed or implied, or assumes any legal liability or
responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or
process disclosed, or represents that its use would not infringe privately owned rights. Reference herein
to any specific commercial product, process, or service by trade name, trademark, manufacturer, or
otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the
United States government or Lawrence Livermore National Security, LLC. The views and opinions of
authors expressed herein do not necessarily state or reflect those of the United States government or
Lawrence Livermore National Security, LLC, and shall not be used for advertising or product
endorsement purposes.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Publishers’ page
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Publishers’ page
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Publishers’ page
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Publishers’ page
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
To our families
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Preface
vii
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Acknowledgments
We are grateful for the support of the National Science Foundation with
Grants ESC-0225007 and ECS-0725902. Work at Lawrence Livermore Na-
tional Laboratory was performed under the auspices of the U.S. Department
of Energy under Contract DE-AC52-07NA27344. We thank Professor Kai
Liu for a critical reading of portions relating to experiments, Dr. R. Du-
mas for providing a diagram of the setup of the radio frequency sputtering
scheme, Dr. R. Rudd for a number of helpful comments throughout the
writing, and Dr. Michael Shaughnessy for a critical reading of Chapter 5.
We also thank the many research groups who have graciously allowed us to
include their results.
ix
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Contents
Preface vii
Acknowledgments ix
1. Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Classes of half metals . . . . . . . . . . . . . . . . . . . . 4
1.3 Half metallic devices . . . . . . . . . . . . . . . . . . . . . 10
xi
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
3. Huesler alloys 71
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2 Half-Heusler and full-Heusler alloys . . . . . . . . . . . . . 72
3.3 Methods of growing Heusler alloys . . . . . . . . . . . . . 75
3.3.1 Bulk Heusler alloys . . . . . . . . . . . . . . . . . 75
3.3.2 Thin films . . . . . . . . . . . . . . . . . . . . . . 76
3.4 Characterization of Heusler alloys . . . . . . . . . . . . . . 81
3.4.1 Bulk Heusler alloys . . . . . . . . . . . . . . . . . 81
3.4.2 Thin films . . . . . . . . . . . . . . . . . . . . . . 82
3.5 Physical properties of bulk Heusler alloys . . . . . . . . . 83
3.5.1 Magnetic moments and the Slater-Pauling rule . . 83
3.5.2 Insulating gap in half metallic Heusler alloys . . . 85
3.5.3 Polarization at EF . . . . . . . . . . . . . . . . . . 89
3.5.4 Magnetic moments . . . . . . . . . . . . . . . . . 90
3.5.5 Curie temperature . . . . . . . . . . . . . . . . . . 92
3.5.6 Other magnetic properties . . . . . . . . . . . . . 92
3.5.7 Disorder in Heusler alloys . . . . . . . . . . . . . . 95
3.6 Physical properties of Heusler alloys in thin-film form . . 100
3.6.1 NiMnSb . . . . . . . . . . . . . . . . . . . . . . . 100
3.6.2 Co2 MnSi . . . . . . . . . . . . . . . . . . . . . . . 104
3.6.3 Co2 FeSi . . . . . . . . . . . . . . . . . . . . . . . . 117
Contents xiii
Bibliography 245
Index 269
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Chapter 1
Introduction
1.1 Background
1
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Introduction 3
In all of the above cases, the transport properties are determined solely
by electrons in states in the vicinity of EF with a single spin polarization.
In some applications, both charge and spin transport can be envisioned.
Transport utilizing the spin degree of freedom provides completely new
prospects for information storage and transmission. It is, all but certainly,
only a matter of time to incorporate this type of transport, with all its
possibilities, into semiconductor technologies. However, there are concerns
about the disappearance of half metallicity at room temperature (RT) in
these materials. Those materials experimentally demonstrated to reach
over 95% spin polarization, such as CrO2 , cannot sustain their spin po-
larization above RT (Dowben and Skomski, 2004). Structural transitions,
collective excitations, e.g., spin waves and phonons, correlations associated
with onsite Coulomb interactions, and spin polarons (Katsnelson et al.,
2008) which show non-Fermi liquid behavior and are formed in the gap
near EF , can cause loss of half metallicity. In this monograph, we focus
manily on the design, growth, and basic understanding of the electronic and
magnetic properties of half metallic materials determined experimentally at
low temperature and predicted theoretically at or near T = 0 K. We will,
however, briefly comment on issues at higher temperature as appropriate.
Introduction 5
Fig. 1.2 L21 crystal structure of full-Heusler alloy (X2 YZ), such as Co2 CrAl. The
X=Co atoms are denoted by filled and open circles, the Y=Cr atoms by open triangles,
and the Z=Al atoms by open squares. The outermost cube edge has length a. If the
sites with open circles (X(2)) are unoccupied, the structure corresponds to a half-Heusler
alloy (XYZ) and is denoted by C1b (Galanakis et al., 2002a).
of states (DOS) based on one-electron properties. The DOS for each class
shows distinct features as a consequence of the atoms involved and distinct
crystal structure. A typical DOS exhibits s-like states from the non-TM
atoms, such as oxygens, pnictides, and Group IV elements, in the low en-
ergy region of the valence manifold. The d-states of the TM are split into
triply and doubly degenerate multiplets in a cubic or tetrahedral environ-
ment. In the cubic case, the triply degenerate states are labeled t2g and
have lower energy than the doubly degenerate states labeled eg because the
lobes of the dxy , dyz , and dzx states point toward neighboring atoms while
the lobes of the eg states point toward second-nearest neighbors.2 In the
2 The t2g and eg labels are not the only ones used to denote triply and doubly degenerate
states, especially in condensed matter physics. For example, the labels Γ25′ and Γ12′ are
also used in cubic cases (having inversion symmetry). In the ZB structure, Γ15 is used
to label the three-fold degenerate states. In recent years, however, the t2g /eg notation
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 1.4 Schematic DOS of a half-Heusler alloy. ∆↓ is the energy gap for the ↓ spin
states.
Coey and Venkatesan (2002) classified the DOS of half metallic oxides
into three types. They are IA , IB , and IIB and are shown in Figs. 1.5, 1.6,
and 1.7. The label A indicates that the ↑ spin channel is conducting. The
label B indicates that the ↓ spin channel is conducting. In all three types,
the lowest energy s-states are below the energy range shown. The next
has become more prevalent (even when not strictly applicable to the symmetry at hand)
and we shall use this convention here.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Introduction 7
group of states are oxygen 2p in nature. They are bonding states. Both
↑ spin and ↓ spin channels are equally occupied. They experience a small
effect from ferromagnetic exchange.
Fig. 1.5 Type IA DOS, ∆↓ is the insu- Fig. 1.6 Type IB DOS.
lating gap for the ↓ spin channel.
For the type IA HMs (Fig. 1.5), the next structure in the DOS is de-
rived from the d-states of the majority-spin channel. They are partially
occupied and the Fermi energy EF passes through the t2g states. The eg
states overlap with the t2g states and are located at higher energy. This
is a consequence of the octahedral crystal field. The next higher energy
structure is the unoccupied d-states of the minority-spin channel. Just as
for the occupied majority-spin channel, the eg states are at higher energy
and overlap with the t2g states. The insulating (semiconducting) gap (∆↓ )
is in the minority-spin channel between the t2g states and bonding oxygen
p-states. An example of this type of HM is CrO2 . As shown later, the
difference between this schematic DOS and the calculated DOS is that the
latter shows some d-p hybridization in the occupied states, with dominant
oxygen p.
Fig. 1.7 Type IIB DOS. Fig. 1.8 Type IIIA DOS.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
For the type IB HMs (Fig. 1.6), the next structure above the bonding
oxygen p-states is also derived from the d-states of the majority-spin chan-
nel. In this case, however, they are fully occupied. The next higher energy
structure is the overlapping t2g and eg states of the minority-spin channel
with EF passing through t2g states. The insulating gap is formed between
fully occupied majority-spin d states and anti-bonding oxygen s states. An
example of this type of HM is Sr2 FeMnO6 .
For type IIB HMs (Fig. 1.7), the d manifolds of majority- and minority-
spin states do not overlap. The lower-energy majority-spin states are fully
occupied. The mechanism of conduction differs from that of type IA and
IB HMs. The minority t2g electrons form polarons. Fe3 O4 is an example
of this type of compound. The half metallic properties of this type of HM,
like others, can be affected by impurities. For example, if the sublattice
occupied by TM elements is substituted by other elements, the sample can
become a Mott insulator.
Based on the mobility µ and the effective mass m∗ of the mobile carri-
ers, rather than integer magnetic moment/unit-cell, Coey and Venkatesan
(2002) considered two additional types of HM (types IIIA and IVA ). As
shown in the DOS (Fig. 1.8) for a type IIIA HM, both spin channels in-
tersect EF . However, the majority-spin electrons are localized while the
minority-spin electrons are delocalized. There is thus a large difference of
µ and m∗ for carriers of different spin. The electrons in the majority-spin
channel are essentially immobile; and so conduction is confined mainly to
the minority-spin channel.
The main features of types I and II HMs can be qualitatively understood
in terms of two key facts: (i) oxygen atoms have large electronegativity and
(ii) TM elements have tightly bound d-states. Consequently, the oxygen
atom essentially ionizes the electrons of neighboring cations to fill its 2p
states. The d states of the cation split into three-fold t2g and two-fold
eg states in the octahedral field of the surrounding oxygen atoms. The
exchange interaction shifts the energies of the minority-spin states up rel-
ative to the majority-spin states. The occupied d states remain localized
at the cation site with more or less atomic-like features. The p states of
the oxygen atoms show less hybridization with the d states than do the p
states of the pnictide or Group IV elements in the Heusler alloys due to the
large electronegativity of the oxygen atoms. There is no d-d hybridization
because the TM elements are surrounded by the oxygen atoms.
For a HM with the ZB structure, we first give a qualitative discussion of
the bonding and then comment on key features of the DOS. The anion can
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Introduction 9
Fig. 1.9 MnAs: As atoms (filled circles) are at the corners and face centers. A Mn
atom (open circle) is surrounded by four As atoms and is located at (1/4,1/4,1/4) along
the body diagonal. A schematic diagram of the d-p hybridization is shown at right.
The ordering of these states is shown in the schematic DOS of Fig. 1.11.
Only states in the vicinity of EF are shown. The eg states are comprised of
dz2 and dx2 −y2 states. They point toward to second neighbors rather than
nearest neighboring cations and form the non-bonding states or bands in
the crystal. They can overlap in energy with the d-p bonding states (p-t2g )
or can be separated from them to form a gap.
As shown in the ↓ spin channel, a gap exists between the bonding p-
t2g states and non-bonding eg states. EF passes through this gap. For
the majority-spin states, an overlap between the bonding p-t2g and non-
bonding eg states is typical, as shown in the figure. The bonding states
have significant anion-p character. To accommodate the total number of
valence electrons in the unit cell, the lowest-energy antibonding states in
the majority-spin channel are occupied due to the exchange splitting of
majority- and minority-spin states. These antibonding states are d-p hy-
brid states with predominantly transition-element d character. This partial
occupation in the majority-spin channel gives rise to the half metallicity.
Among the three classes of HMs, the d-states are dominant near EF in
the Heusler alloys and the oxides. Hybridization is the strongest in the ZB
HMs. The states at EF for these are d-p hybrid in character.
Introduction 11
substantially essentially all GMR based devices in current use and offering
a host of new possibilities. Hence, this has been an active field of research.
There have been many theoretical proposals, based on extensive quantum
mechanical calculations, as we detail in subsequent chapters. Here, we note
some recent progress on practical device realization.
Recent efforts in spintronic device fabrication have focused mainly in
the areas of spin valves and spin transistors. The latter was proposed by
Datta and Das (1990) and is similar to standard FETs. However, the re-
quirements for effective FETs, such as large polarized current, capability
of amplification, and simple material structure, are extremely demanding,
and there has not yet been success fabricating practical devices in this
form. There has been more progress in fabricating HM based spin-valves.
In these devices, spin is injected from HMs to semiconductors. It has been
identified by Wang and Vaedeny (2009) that four conditions should be met
for an effective spintronic device. Among the four, the most important is
efficient spin injection from the metallic materials normally serving as elec-
trodes into semiconductors. To maintain the signal, electrons in the semi-
conductor should have long lifetimes for conduction and spin relaxation.
In crystalline semiconductors, the existence of the spin-orbit interaction
can change the spin moment of the carriers. Wang and Vaedeny (2009)
used organic semiconductors instead of conventional semiconductors. They
made a device with La2/3 Sr1/3 MnO3 (LSMO) and Co serving as electrodes.
The organic semiconductor CVB (4,4’–bis–(ethyl–3–carbazovinylene)–1,1’–
biphenyl) was used to reduce the effect of the spin-orbit interaction. The
measured spin polarized carrier density was 60–20% in the temperature
range of 60–120 K. Kodama et al. (2009) grew the layered structure
Cr/Ag/Cr/Co2MnSi/Cu/Co2 MnSi/Co0.75 Fe0.25 /Ir0.22 Mn0.78 /Ru on MgO
(001) substrate. Cu was included between Heusler layers to facilitate the
transport of spin polarized carriers from one Heusler alloy to the other,
which would be impeded by Cr. The Ag layer between the Cr layers served
to enhance the perpendicular current. The ratio of magnetoresistance was
found to be 8.6 % at room temperature and 30.7% at T = 6 K. Hence, more
work is needed to get magnetoresistance and robustness to desired levels,
and much work continues.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Chapter 2
2.1 Introduction
13
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Because HM crystal structures are not always the stable bulk ground state
structures, and because desired magnetoresistance (MR) properties are of-
ten achieved by multilayer structures, the ability to grow HM structures
in thin-film form is of crucial importance. The molecular beam epitaxy
(MBE) method is an ideal way to grow quality thin films. It was originally
designed to grow layered semiconductors and semiconductor heterostruc-
tures (Chang and Ploog, 1985).
Its application to grow NiMnSb was carried out by Van Roy et al. (2000)
and Turban et al. (2002). Van Roy et al. (2000) used the Riber 32P chamber
while a commercial chamber was used by Turban et al. (2002). Ambrose
et al. (2000) also used a commercial chamber to grow Co2 MnGe. With the
MBE method, Akinaga et al. (2000a) were the first group to successfully
grow CrAs thin film with ZB structure and Zhao et al. (2001) grew ZB
CrSb with a single monolayer. Hereinafter, we will refer to these types of
samples as HMs with the ZB structure.
Fig. 2.1 A schematic diagram of the setup for the MBE method. The substrate is just
above the holder. Sample is shown in black horizontal rectangle.
a relatively smaller lattice constant which is 4.21 Å. There should be a large
lattice constant mismatch. This kind of substrates, in general, can be used
to grow polycrystalline samples in film forms (Bauer, 2010). Therefore,
as described above the primary consideration for the choice of a substrate
is determined by matching lattice constants between the sample and the
substrate. Alternative choices to grow single crystal thin films can be made
based on experience by adding additional buffer layers.
2.2.2.3 Temperature
The temperature of the substrate is an important factor to determine the
quality of the films. For growing NiMnSb on GaAs(001), the best films
are obtained when the temperature of the substrate Tsub is kept at 300 ◦ C
(Van Roy et al., 2000). Turban et al. (2002) demonstrate clearly that Tsub
at about 350 ◦ C is the best to grow NiMnSb by measuring the intensity
of the reflection high energy electron diffraction (RHEED) to be described
later. On the other hand, the substrate temperature is maintained at 175
◦
C for Co2 MnGe growths. To grow a better CrAs, the temperature of the
substrate cannot be maintained at 580 ◦ C required to have a flat surface
of the substrate. It is necessary to bring down the temperature to 200 ◦ C
and 300 ◦ C range under As pressure.
During or after a sample is grown, it is necessary to characterize its
quality and to measure its physical properties. In the following two sec-
tions, we will discuss methods of characterization, in particular samples in
thin-film forms and of determining various physical properties relevant to
spintronic applications. These methods are generally applied to all HMs.
The results, however, will be sample dependent and they will be discussed
specifically for each sample.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
RHEED pattern The pattern is in striped form and provides the struc-
tural information of the sample. The primary stripes are the constructive
interference pattern relevant to the fundamental structure of the sample
surface. Therefore, one of the features of RHEED is the capability to char-
acterize a surface structure. Before the growth, it is important to check the
pattern of the substrate. When the primary stripes coincide with those of
the substrate, lattice-constant matched growth can be inferred. If there are
additional weak streaks, it indicates the surface is reconstructed. A typical
RHEED pattern for a surface structure is shown in Fig. 2.2. The pattern
Fig. 2.2 A typical RHEED pattern of a surface with streaks (Wang et al., 2005a). The
primary ones are shown in dark regions. The streaks are shown in light stripes.
shown in Fig. 2.2 is for a clean GaAs with electron beam focused along the
[11̄0] crystal axis. The stripes with elongated dark regions are the primary
lines. Spacing between these primary lines can fill with other weak streaks
which are related to the reciprocal lattice vector in the crystallographic
direction of the surface for the sample with a reconstruction. This is one
way to extract lattice-constant information of the sample. In this figure,
there is clearly a reconstruction with the periodicity of six times the lattice
constant—each primary line is accompanied by five weak streaks.
Fig. 2.3 The oscillatory behavior of RHEED intensity at (00) varying with time (Wang
et al., 2005a).
Fig. 2.5 Measured intensity of specu- Fig. 2.6 Intensity of diffuse X-ray re-
lar X-ray reflection as a function of θ flection as a function of θ (Bahr et al.,
(Toney and Brennan, 1989). 1993). The Yoneda peaks are marked
by arrows.
Fig. 2.7 A schematic diagram of STM. h is the separation between the tip and surface.
An ammeter (A) is shown as a large circle. The voltage “V” is maintained between the
tip and surface.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 2.8 STM line profile along a line parallel to the surface. a/2 = 2.85 Å, where a is
the lattice constant (Wang et al., 2005a).
Fig. 2.9 Processes in AES. Left figure shows a core hole (open circle) being created by
an incident X-ray. Right figure describes the hole being annihilated by a valence electron
(dark circle) and another valence electron (gray circle) being excited above the vacuum
level.
Fig. 2.10 Peak-to-peak intensity as a function of thickness in Co2 MnSi at 450 K (Wang
et al., 2005a).
2.3.4.4 Remarks
Because of the hole created in the core, AES can involve significant many-
body interactions due to possible polarization effects induced by the hole.
For simply determining the presence of certain species of atoms, however,
it is possible to ignore such complications.
Fig. 2.11 A schematic setup for the SQUID magnetometer. A dc magnetic field H is
applied to the sample. The SQUID is shown as a loop. It measures B field from the
sample
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 2.14 (a) Hysteresis loop shown in Kerr rotation as a function of H. Results from
different thickness are shown in black and light gray lines. Black loop is for thicker and
light gray loop is for thinner samples. (b) Hysteresis loop shown in terms of the Kerr
intensity as a function of H (Bader, 1991).
to the thickness of the sample. Let us take Co2 MnSi grown on GaAs(001) as
an example. After growing three bilayers, the sample starts to exhibit an in-
plane uniaxial magnetic anisotropy having the easy axis pointing along the
[11̄0] direction. The experiment of measuring the hysteresis loop was carried
out to probe the magnetization along the hard axis—the [110] direction.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Between 4 and 20 bilayers, there are two easy axes in the (110) plane.
Hysteresis loops are developed which are shown in gray in Fig. 2.14(a) as a
function of positive and negative external magnetic fields. The remanence
field is zero. As the film grows thicker—over 35 bilayers, the sub-loops
disappear. The following empirical relation is found: let Hs1 and Hs2 be
the fields in the middle of sub-loops at positive and negative H, then
Hs = (Hs1 − Hs2 )/2. (2.1)
The empirical relation is Hs ≈ 1/d, where d is the thickness of the film.
When H is applied along the easy axis, a simple hysteresis loop is obtained.
In Fig. 2.14(b), the hysteresis loop is in terms of the Kerr intensity as a
function of H.
The Kerr rotation under the remanence field along the easy axis can
be measured as a function of temperature. If the Kerr rotation is assumed
to be proportional to the magnetization, M, then the measured results fit
well with the Bloch formula indicating the excitations are spin waves. The
Bloch formula is
M (T ) = M (0)(1 − bT 3/2 ). (2.2)
Because MOKE cannot measure the absolute magnetization, it can only
probe the magnetization in the surface region with penetration depth of the
order of 10 nm. For Heusler alloys, this is a real disadvantage, because a
complex layer stacking can happen to diminish magnetic properties at the
surface. Alternative methods to measure the hysteresis loop are the alter-
native gradient magnetometer (AGM) and vibrating sample magnetometer
(VSM). Since these two methods are specific for Heusler alloys, the related
discussions will be given in Chapter 3.
lines, where s is the spin angular momentum. These lines are called L-
edges. The XMCD signal is the normalized difference in absorption between
the right- and left-hand polarized X-rays. It contains the information of
exchange splitting and the spin-orbit coupling of the initial and final states.
By controlling the energy of the X-ray, it is possible to have XMCD probe
a particular element. After absorbing the X-ray, the core electron changes
its angular momentum due to the circularly polarized light. It is excited to
either an s- or a d-like state above the vacuum level. Spin-polarized bands
respond differently for different polarizations of the X-ray. Measurements of
photoelectrons with respect to the spin polarization of the sample provide
information about the magnetic moment of each TM element in the sample.
The processes are depicted in Fig. 2.15.
Fig. 2.15 Processes in XMCD. L2 and L3 are the core energy levels. The thick arrow
indicates the X-ray with the left and right polarization shown as circular arrows. Vertical
double arrows indicate possible transitions of core electrons. The spin-polarized density
of states (DOS) are shown in shaded areas (Stöhr, 1999).
Fig. 2.16 Simple schematic setup of XMCD. X-ray is incident from the left. The dashed
line is a Au mesh. The middle line at the right of the dashed line is the Al window.
Horizontal lines are bias rings. The gray rectangle is the sample. The black box is the
magnet with its south pole close to the sample and is separated from the sample by
a quartz window. Emitted electrons are collected by the photocurrent detector (gray
vertical line) (Stöhr, 1999).
Fig. 2.17 XMCD at L2,3 absorption edge of Mn for 45Å thick Co2 MnSi films at RT
(Wang et al., 2005a).
Schematic setup The schematic setup is shown in Fig. 2.18. The two
ferromagnetic thin films are indicated by gray and black regions and are in
contact with an ammeter denoted by an open polygon and a battery. The
mid region between the gray and black regions is the insulating region.
where R↑↑ and R↑↓ are the resistance in parallel and antiparallel configura-
tions in the two electrodes. In practice, measurements of TMR and MR are
complicated by the presence of insulating layers. Cautions in interpreting
the data are called for due to scatterings such as phonons.
These two kinds of results are given in Fig. 2.19. Figure 2.19(a) shows
TMR as a function of bias voltage and Fig. 2.19(b) shows MR as a function
of external field. The results also depend on the temperature and magne-
tization of electrodes. Figure 2.19(a) shows the RT case. The asymmetry
shown in Fig. 2.19(a) originates from the tunneling process dependent on
the direction of the bias voltage.
Fig. 2.19 (a)A typical plot of TMR as a function of bias voltage (Sakuraba et al., 2006),
and (b) a typical plot of MR as a function of external field (Gong et al., 1997)
.
where G↑↑ and G↑↓ are conductances of the parallel and antiparallel configu-
rations, respectively. n1 is the fractional number of majority-spin electrons
in the DOS at EF in electrode 1, and n2 is the corresponding fractional
number of electrons in electrode 2. The number of minority-spin electrons
is 1 − ni , i = 1, 2. G↑↑ is the sum of the conductance from electrons with
both electrodes having parallel orientations of majority- and minority-spin
channels. G↑↓ is for opposite spin orientations in the two electrodes. G↑↑
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
P1 = 2n1 − 1 (2.7)
P2 = 2n2 − 1. (2.8)
2.4.2.2 Resistivity
General discussion For bulk materials, samples are often in rod form
with axis along the growth direction. For samples in thin-film form, the re-
sistivity ρ is commonly measured in the plane of the sample. The resistivity
is measured as a function of temperature. The temperature dependence is
assumed to be:
ρ = ρo + cT n , (2.9)
Comments on setup For both bulk and thin-film samples, the standard
way of measuring resistivity is to attach four electrodes to the sample.
This is called the “four-probe” arrangement. In Fig. 2.20, a typical four-
probe arrangement is shown. One of the advantages of the four-probe
arrangement is that it reduces the effects of the contact resistance between
the sample and electrode (Pesavento et al., 2004; Bezryadin et al., 1998).
Fig. 2.20 A typical four-probe arrangement, (a) the side view and (b) the top view
(Pesavento et al., 2004).
Fig. 2.21 (a)A typical plot of Resistivity (ρ) as a function of temperature. (b) A typical
plot of ln(ρ − ρo ) as a function of ln(T ).
Fig. 2.22 The relative orientations of the magnetic field, current, and electric field. The
sample is shown in gray. Current Ix flows in the x direction. The magnetic field Hz is
in the z direction (into the paper shown by the cross). The electric field Ey is in the y
direction.
Fig. 2.23 Schematic diagram for Hall measurement. The sample is shown as the long
gray bar. The current is measured by the ammeter A. H is the strength of a magnetic
field pointing into the paper (shown as a cross). VH is the Hall voltage.
Fig. 2.24 (a) Longitudinal resistivity ratio (∆ρxx /ρxx ) as a function of H (Husmann
and Singh, 2006); (b) Transverse resistivity (ρxy ) as a function of H (Lee et al., 2004);
and (c) Magnetization M, as a function of H (Lee et al., 2004).
What is measured Two quantities, the longitudinal σxx and the trans-
verse σxy Hall conductivities as a function of H at different temperatures
are commonly measured. Husmann and Singh (2006) measured ∆ρxx /ρxx
as an alternative. The anomalous Hall coefficient RAH is then extracted
with the information from either longitudinal resistivity ratio ∆ρxx /ρxx as
a function of H (Fig. 2.24(a)) or transverse resistivity ρxy as a function of
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 2.25 Geometric configurations for (a) regular energy and angular resolved photoe-
mission, (b) longitudinal magnetization, and (c) transverse magnetization. θ and φ in
(a) are the polar and azimuthal angles.
cross-sections. If an electron passes to the left side of the target, the orbital
angular momentum reverses direction, so that the extra potential due to the
last term acting on the ↑ and ↓ spin states is also reversed. The phenomenon
is depicted schematically in Fig. 2.27.
Fig. 2.26 Different potentials experienced by the ↑ and ↓ spin states in Mott scattering.
Fig. 2.27 An electron passing to the left (top) (dark grey arrows for the spin and
velocity) and right (bottom) (light grey arrows) of an Au atom experiences different S-O
interactions due to the reversal of the effective B field. v is the velocity of the electron
and S is the spin of the electron.
mirror with focusing system, and other electron lens systems. At the end
of the chamber, a 90◦ spherical deflector deflects photoelectrons into the
Mott detector or Mott spin polarimeter by passing through another set of
lens systems. The spherical deflector serves also as an energy analyzer. At
its exit plane, an assembly of electron multiplier arrays are installed for
multichannel detections.
Fig. 2.29 Sherman function as a function of θ, the dashed curve is for the electronic
energy at 50 keV and the solid curve is for 120 keV (Gellrich and Kessler, 1991).
1
P∞ 2iηl
+ e2iη−l−1 Pl1 (cos θ)eiφ
g1 (θ, φ) = 2ik l=0 −e (2.33)
= g(θ)eiφ ,
and
where Pl (cosθ) is the Legendre polynomial and Pl1 (cosθ) is the associated
Legendre polynomial. For an incident wave with an arbitrary spin orienta-
tion, the spinor is:
1 ikz 0 ikz
ψ=A e +B e . (2.35)
0 1
The cross-section is
|a|2 +|b|2 2 2 ∗ iφ ∗ −iφ
σ(θ, φ) = |A|2 +|B|2
= (|f | + |g| ) + (f g ∗ − f ∗ g) −AB |A|
e +A Be
2
+|B|2
2 2 ∗ iφ ∗ −iφ (2.37)
= (|f | + |g| )(1 + S(θ)) −AB |A|
e +A Be
2
+|B|2
.
S(θ) is defined as
(f g ∗ − f ∗ g)
S(θ) = 2 2 . (2.38)
|f | + |g|
Fig. 2.30 Left panels: EDC and spin polarization for a normal emission from Ni(110);
Right panels: calculated spin-resolved EDCs (Raue et al., 1984).
Fig. 2.31 (a) The Al (superconducting) density of states denoted by solid curves are
shown in the upper panel. ∆ is half of the gap. The middle horizontal region is an
insulating barrier. The bottom panel illustrates the Ni density of states. When an
external magnetic field H is applied paralleled to the Al thin film, the superconductor
exhibits Zeeman splitting. Dashed curves are for the ↓ spin states while dotted curves
are for the ↑ spin states. (b) Ratio of conductance as a function of voltage for different
values of H (Meservey and Tedrow, 1994).
Fig. 2.32 (a) Energy as a function of DOS of tunneling between a normal metal and
a superconductor. (b) Energy as a function of DOS of tunneling between a HM and a
superconductor. Arrows mark the spin polarizations of states in the metal.
Fig. 2.33 Schematic of Andreev reflection method for detecting half metallic properties.
The gray rectangle is the sample. The Nb metal is in the tapered form serving as a point
contact. The circle “A” is the ammeter. “V” represents the battery.
What is measured The tunneling current, from the reading of the am-
meter, is the crucial quantity to be measured. To contrast characteristics
relevant to processes of the normal-metal-superconductor tunneling and
HM-superconductor tunneling, the corresponding tunneling currents as a
function of voltage are shown in Fig. 2.34. Note that in Fig. 2.34(a), the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
current is zero only when the bias voltage is zero. However, there is a range
of 2∆ of the voltage for the tunneling current to be zero in Fig. 2.34(b).
Fig. 2.34 (a) The tunneling I-V curve between a normal metal and a superconductor;
(b) the tunneling I-V curve between a HM and a superconductor.
Fig. 2.35 The band diagram of a normal (NM) or ferromagnetic metal and a supercon-
ductor (SC) in point contact (Blonder et al., 1982). 2∆ is the gap of the SC. I (small
filled circle) is the incident electron. T− and T+ are the possible transmitted electronic
states. The hole state labeled by A is for the Andreev reflection.
with
f→ (−E) = f0 (E + eV ), (2.43)
where f0 is the Fermi-Dirac distribution function and V is the poten-
tial energy due to an externally applied voltage. Substituting Eqs (2.41)
and (2.42) into Eq (2.40), the current from the normal metal to the super-
conductor is,
Z∞
IN M→SC = 2eAα Nσ (EF )υF σ {f0 (E−eV )−f0 (E)}{1+A(E)−B(E)}dE,
−∞
(2.44)
From this expression, the conductance can be obtained by taking the
derivative with respect to V. For an NM-NM contact, A(E) is zero be-
cause there is no Andreev reflection and D(E) is also zero due to no cross
Fermi surface scattering. With Eq. (2.42) and the sum rule Eq (2.39), the
conductance is
dIN M−N M
= 2eAα Nσ (EF )υF σ C(E) = Gn . (2.45)
dV
Gn is also the conductance for the NM-SC contact with V ≫ ∆. The
integrations in Eq. (2.44) were carried out by Blonder et al. (1982) for Z =
0.0, 0.5, 1.5, and 5.0, respectively. The results are shown in Fig. 2.36. For
Z=0 case, the normalized conductance is 2 for V < ∆. This is due to the
presence of a Cooper pair and Gn should correspond to one electron. For
|V | ≫ ∆, we have
G(V )
= 1. (2.46)
Gn
The electron density of states (DOS) and magnetic moment of half metal-
lic compounds are the primary interest. In addition, energetics for sta-
ble structures and the effects of dopings on half metallic properties at T
= 0 K are of interest. Theoretical methods based on density functional
theory (DFT) (Hohenberg and Kohn, 1964) in the local density approxi-
mation (LDA) (Kohn and Sham, 1965) or generalized gradient approxima-
tion (GGA) (Langreth and Perdew, 1980; Langreth and Mehl, 1983) for
exchange-correlation are the most often used. We shall give a brief review
of DFT and comment on popular methods of calculating the total energy
of HMs and corresponding electronic band structures.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
with
g(r, r′ )ρ(r′ ) 3 ′
Z
′
vxc [ρ(r)] = d r, (2.62)
|r − r′ |
where g(r, r′ ) is the pair correlation function. It is a non-local function
specifying the probability for a particle at r′ when there is a particle at r.
′
In LDA, vxc is approximated as follows: Let
Z
Exc [ρ] = εxc [ρ(r)]ρ(r)d3 r,
LDA
(2.63)
where εxc [ρ(r)] is the exchange-correlation energy per particle. εxc [ρ(r)] is
′
determined locally by the density at r. Thus vxc [ρ(r)] can be expressed as:
LDA
′ ∂Exc [ρ] ∂εxc [ρ]
vxc [ρ(r)] = = εxc [ρ(r)] + ρ . (2.64)
∂ρ ∂ρ ρ=ρ(r)
Eqs. (2.63) and (2.64) are the essence of LDA. εxc [ρ(r)] is the sum of the
exchange energy per particle, εx [ρ(r)], and correlation energy per particle,
εc [ρ(r)]. In terms of rs —the average spacing between particles—Perdew
and Zunger (1981) fit these energies to Monte Carlo results of Ceperley
and Alder (1980) to obtain
0.4582
εx [ρ(r)] = − , (2.65)
rs
and
− 1+1.05290.1423
√
rs +0.3334rs rs ≥ 1
εc [ρ(r)] =
−0.048 + 0.311 ln(rs ) − 0.0116rs + 0.002rs ln(rs ) rs < 1.
(2.66)
and Hedin (1972). Readers who are interested in this subject can read the
review by Zeller (2006). Here we only discuss the essentials of spin-polarized
Kohn-Sham equations (von Barth and Hedin, 1972) which are relevant to
half metallic properties. Let α denote the ↑ spin and β denote the ↓ spin
states. We follow the formalism in LDA by defining
X
∗
ραβ (r) = ψα,i (r)ψβ,i (r). (2.67)
i=occ
For diagonal terms,i.e. α = β, the two densities are denoted by either ρα (r)
or ρβ (r) and the Schrödinger equation is now in a coupled form,
X 2
−~ 2 X Z ργγ (r)
3 ′ ′
δµν ∇ + vext (r) + d r + vxc [ρ(r)]µν − λ ψν = 0,
ν 2m γ |r − r′ |
(2.68)
where µ, ν, and γ are either α or β. With the local spin density approxi-
mation (LSDA), the exchange-correlation functional ELSDA xc is
Z
LSDA
Exc [ραβ (r)] = εxc [ρα (r), ρβ (r)] (ρα (r) + ρβ (r)) d3 r (2.69)
LDA
′ ∂Exc [ρ]
vxc [ρ(r)]αβ = . (2.70)
∂ραβ
Note the matrix forms in Eqs. (2.69) and (2.70). Additional information
about the magnetic moment at a point inside a unit cell and the magnetic
moment per unit cell can be obtained,
and
Z
M= m(r)d3 r. (2.72)
Expressions for EcGGA and ExGGA For facilitating the numerical meth-
ods, PBE proposed analytic expressions for both energy functionals satis-
fying energetically significant conditions and having constant parameters.
We start with EcGGA ,
Z
GGA
Ec [ρα (r), ρβ (r)] = ρ(r) {εc [ρα (r), ρβ (r)] + C (rs , ρα−β (r), t)} d3 r,
(2.74)
where εc is the one used in Eq. (2.66) and
ρα−β (r) = (ρα (r) − ρβ (r))/(ρα (r) + ρβ (r)). (2.75)
p
t = |∇ρ(r)| /(2φkT F ρ) with kT F = (4kF )/(πaB ) as the Thomas-Fermi
screening wave number, where kF is the Fermi momentum and aB is the
Bohr radius. Let φ = (1/2) (1 + ρ)2/3 + (1 − ρ)2/3 , then
e2 3 β(1 + At2 )
2
C (rs , ρα−β (r), t) = γφ ln 1 + t , (2.76)
aB γ(1 + At2 + A2 t4 )
where A = (β/γ){exp(−εc /(e2 γφ3 /aB ) − 1}−1 , γ = (1 − ln 2)/π 2 and β =
0.066725. EcGGA is given as
Z
ExGGA [ρα (r), ρβ (r)] = ρ(r) {εx [ρ(r)]Fx (s)} d3 r, (2.77)
η |∇ρ|
where εx [ρ(r)] = −3e2 kF /(4π), Fx (s) = 1 + η − 1+0.21951s2 , s = 2k ρ , and
F
η = 0.804.
With the basic concept of DFT, we now discuss the essence of several
popular methods for calculating the electronic properties of various mate-
rials including HMs based on DFT.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Basic idea In order to make the method tractable, Slater (1951) em-
ployed an approximation that the potential has spherical symmetry near
an atom in a solid and is constant in the interstitial region: the so-called
muffin-tin (MT) approximation. A schematic diagram of the MT potential
is given in Fig. 2.38. The lattice is indicated by mesh. At each intersection,
there is a lattice point and an atom. The MT potential around each atom
is shown by a filled circle known as the MT sphere. Its radius is called the
MT radius, Rm . The arrows mark the interstitial region.
With the MT potential, the wave function inside the MT sphere can be
expanded in spherical harmonics, Ylm (θ, φ), where θ and φ are the stan-
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
dard angular variables. In the interstitial region the wave function can be
expanded in plane waves eiG·r . An augmented plane wave (APW) is then
defined as
P
l,m Alm Rl (r, E)Ylm (θ, φ) r ≤ Rm
ψAP W (r) = (2.78)
iG·r
e interstitial,
where Rl (r, E) is the solution of the radial equation with angular momen-
tum l—the quantum number—at energy E and is a function of both r and
E. Alm is determined by matching the expansion inside the MT sphere and
plane wave at Rm . The set of all such APW’s is then used as the basis for
the wave functions in the solid.
Remarks Since the basis functions of APW’s are used, the matrix ele-
ments of the Coulomb and exchange-correlation potentials can be readily
calculated. Over the years, a number of software implementations of the
method have been developed. One popular and versatile implementation is
the WIEN package1 . For a detailed discussion of the LAPW method, the
reader is referred to the excellent book of Singh and Nordström (2006).
three integers to specify the location of the n-th unit cell. At the origin of
the lattice, a Wigner-Seitz (WS) cell is defined. The MT sphere Sm and the
corresponding MT potentials are inside the WS cell. The Green’s function
is the response at a position r due to a unit source located at a point r′
and is scattered by the MT potentials at other lattice sites.
In order to facilitate the calculations, non-overlapping MT potentials
Σn Vm (r − Rn ) are used, where the summation runs over all lattice points
and Vm (r − Rn ) is the MT potential at Rn . The original KKR formulation
applied to solids is based on the work of Dupree and Beeby (1967). It has
been applied to half metallic compounds by Podloucky et al. (1980) and
Akai et al. (1985) in Germany and Japan, respectively.
The Green’s function associated with Eq. (2.79) satisfies the following
equation:
{H0 + E} G(r + Rn , r′ + Rn′ ; E) = −δ(r − r′ )δnn′ (r), (2.80)
where Ho is (−~2 /2m) ∇2 . By constructing a functional Λ and setting
∂Λ/∂Ψ∗(r) = 0, an integral equation is obtained, where
ψ ∗ (r) { n Vm (r − Rn )} ψ(r)d3 r
R P
Λ= RR
ψ ∗ (r′ ) { n Vm (r − Rn )} G(r, r′ ; E) { n′ Vm (r′ − Rn′ )} ψ(r′ )d3 rd3 r′
P P
+
(2.81)
and
Z nX o
ψ(r) = G(r, r′ ; E) ′
Vm (r′ − Rn′ ) ψ(r′ )d3 r′ . (2.82)
n
′
By changing the variable r and applying the Bloch theorem,
ψk (r + Rn ) = eik·Rn ψk (r), (2.83)
where k is a vector inside the first Brillouin zone (BZ), Green’s function is
now k-dependent and is given by:
X
Gk (r − r′ ; E) = e−ik·Rn Gk (r − r′ − Rn ; E). (2.84)
n
two resulting equations, and finally setting the surface integral term to be
zero:
∂ψ(r′ ) ′
Z
′ ∂Gk (r − r ; E)
Gk (r − r′ ; E) − ψ(r ) dS = 0. (2.85)
S ∂r′ ∂r′
To apply Eq. (2.85), an expression for G is necessary. Eq. (2.80) can be
easily solved for a single site,
1
G(r − r′ ; E) = . (2.86)
E − H0
By using the eigenfunctions of Ho —the plane waves—the spectral rep-
resentation of Green’s function with summation over BZ can be obtained.
After carrying out the sum, the real space representation of Green’s func-
tion is given as:
′
eiκ|r−r |
G(r − r′ ; E) = − , (2.87)
4π |r − r′ |
√ p
where κ is E if E is positive, and is i |E| if E is negative. Equation (2.87)
can be expanded in spherical harmonics, the so-called Neumann expansion,
′
eiκ|r−r | ′ ′ ∗ ′ ′ ′
P
− 4π|r−r ′| = κ l,m jl (κr)Ylm (θ, φ) {nl (κr ) − ijl (κr )} Ylm (θ , φ ) r < r ,
(2.88)
where (r, θ, φ) is the usual spherical coordinates. jl is the spherical Bessel
function of the first kind and nl is the spherical Bessel function of the second
kind. For r > r′ , one exchanges r and r′ at the right hand side of Eq (2.88).
Structure factor Since the KKR method is based on the idea that the
electron wave function at a point r is a consequence of interferences of
scattered waves from the potential centers, the important quantity is the
so-called structure factor which specifies phase relations of the potential
centers. For one atom per unit cell, Eq. (2.84) provides the basic form of
the structure factor. By substituting Eq. (2.87) into Eq. (2.84), one gets
cos(κ |r − r′ |) X
Gk (r−r′ ; E) = − + Γlm (k, E)jl (κ |r − r′ |)Ylm (θr−r′ , φr−r′ ),
4π |r − r′ | l,m
(2.89)
where θr−r′ is the polar angle and ϕr−r′ is the azimuthal angle of the vector
r − r′ , respectively. Γlm (k,E) is the structure factor and is given by
X δl0 δm0
Γlm (k, E) = κ e−ik·Rn {nl (κRn ) − ijl (κRn )} − iκ √ . (2.90)
n 4π
For the case of many atoms per unit cell, Eq. (2.90) can be easily modi-
fied to accommodate the situation. The most recent improvement including
the coherent potential method is given by Akai and Dederichs (1993).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
2.5.3.4 LDA+U
There are some concerns about electron-correlation effects on electronic
properties of Heusler alloys and oxides. Here we briefly discuss the essence
of electron-correlation treated in these two classes of HMs.
The most important electron-correlation effect due to localized d-
electrons in TM elements, such as Co, Cr, Fe, and Mn involved in Heusler
alloys and oxides, is the Coulomb repulsion experienced by localized elec-
trons. The LDA+U (including LSDA+U) scheme is designed to model lo-
calized states when on-site Coulomb interactions become important. Since
most modern DFT implementations such as VASP and WIEN have in-
cluded the LDA+U scheme, we explain the basic ideas and comment on
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Basic idea The concept of the U term characterizing the Coulomb inter-
action at an atomic site in a solid was introduced by Hubbard (1963). The
starting point is a many-electron Hamiltonian,
X ~2 X X
e2
H= − ∇2i + Vext (ri ) + , (2.94)
i 2m i j6=i |ri − rj |
(2.100)
where Ykq is a spherical harmonics.
Typically, U value is approximately 20 eV in a free TM element and is
on the order of a few eV in a solid.
and Andersen (1980) and Heine (1980). It is then extended by Oswald et al.
(1985) for systems with magnetic impurities. The theorem is frequently
used to compute exchange interaction parameters and adiabatic spin-wave
spectra of ferromagnets. The interest of this approach is the energy differ-
ence of different spin orientations at the atoms can be determined by the
change of the one-particle energy. It allows the calculation of the change
of the total energy to be carried out non-self-consistently. Thus, the theo-
rem provides a way to reduce the computational effort. Liechtenstein et al.
(1987) used the theorem for treating spin wave spectra of ferromagnetic
systems within mean field theory (MFT). Bruno (2003) and his collabora-
tors extended the applications to varieties of ferromagnetic metallic systems
and improved the MFT scheme proposed by Liechtenstein et al. (1987).
where Jij is the pairwise exchange interaction energy, Si is the spin moment
at site-i. If Jij is obtained, then the magnon energy ~ω(q)), where q is the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
where hi = gµ2B j6=i Jij Sj —the mean magnetic field at site i produced
P
Determination of Jij There are two ways to determine Jij and the
excitation energy. One is the real space scheme and the other is the q-
space scheme or the scheme of frozen magnon.
Real-space approach In Eq. (2.101), for each i one calculates Jij by consid-
ering the j-th shell of atoms. First-principles methods can be used
to determine Jij by calculating the change in energy when spin
moments at the i-th and j-th sites are rotated by equal and oppo-
site amounts with respect to the ground state (all spins aligned),
and subtracting the changes in energy for the isolated rotations at
the two sites, respectively. The tight-binding linear combination
of muffin-tin orbital (TB-LMTO) method has been used by Pajda
et al. (2001) with j up to 172-th shell for fcc structure and 195-th
shell for bcc lattice. In order to get converged ω(q), about 200
terms of Jij are needed.
Momentum-space approach In this approach, one specifies the spin mo-
ments on the neighboring atoms to have a spiral configuration and
carries out first-principles calculations to get the ~ω as a function of
q relations for the configuration. Where Jij is obtained by the in-
verse Fourier transform–an important contribution by Sandratskii
(1998) so large supercell is not necessary. The main relevant physics
is illustrated in Fig. 2.39 using a one-dimensional model.
In Fig. 2.39, each arrow makes a polar angle θ with the z-axis and
an azimuth angle ϕ with respect to the x-axis. The lattice vector
is Rn where n represents the n-th lattice site. The moment at the
n-th lattice site for a spiral structure with momentum q is given
by
Mn = |Mn | {cos(q · Rn + ϕn ) sin θn x̂ + sin(q · Rn + ϕn ) sin θn ŷ + cos θn ẑ} ,
(2.109)
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 2.39 One dimensional model showing the relative phase of the spin spiral structure
(Sandratskii, 1998).
where |Mn | is the magnitude of the spin moment at the n-th lat-
tice site. It is easy to see that when there is a lattice translation
Rn , correspondingly there is a spin moment rotation by an an-
gle of −q · Rn which leaves the invariance of the spiral structure.
By analogy to notations used in the space group, one can define
({q · Rn |e| Rn }) as a generalized translation of the spiral structure,
where e indicates that it does not have any real space coordinate
transformation, such as a rotation or a reflection and commutes
with the Hamiltonian of the spiral structure. Because of e, it has
no effect on the periodic part of the Bloch wave function—just a
unit operator. The effect of generalized translation on a spinor
wave function with electron momentum k is
Pk ({q · Rn |e| Rn }) ψk = U (αS ) e−ik·Rn ψk , (2.110)
where Pk is a unitary operator associated with the generalized
translation operator. The active point of view is adopted here
for the coordinate transformation on the wave!function.
i
e− 2 q·Rn 0
U (αS ) = i , (2.111)
0 e 2 q·Rn
Eq. (2.110) is the generalized Bloch theorem. For each k-point, the
eigenvalue εn (k) of the nonmagnetic case is changed to εn k − σq 2
, where σ is “+” or “-”, the spin projection along the z-axis. E(q)
can be obtained by summing all occupied state energies and sub-
tracting the energy of the reference state. From E(q) one can get
J(q), then Jij is obtained by inverse Fourier transform.
Comparing the two schemes, the k-space scheme is simpler owing
to this generalized Bloch theorem. However, the calculations are
still demanding in order to obtain accurate ω(q).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Chapter 3
Huesler alloys
3.1 Introduction
The Heusler alloys are expected to meet both of these criteria because their
structures are compatible with many semiconductors in both elemental and
compound forms. The TC ’s of Heusler alloys are well above RT. However,
only NiMnSb (Hanssen et al., 1990; Hordequin et al., 1996) has been found
to exhibit half metallic properties within 1/100 of an electron accuracy per
formula-unit experimentally. A number of these alloys have been predicted
to be half metallic theoretically (Wurmehl et al., 2005). These are excellent
candidates for spintronic materials.
71
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 3.1 L21 crystal structure with chemical formula, X2 YZ. X(1) sites are denoted by
filled circles, X(2) by open circles, Y by open triangles, and Z by open squares. This
structure corresponds to a FH alloy. The outermost cube edge has length a. If those
sites with open circles (X(2)) are unoccupied, the structure corresponds to a HH alloy
and is denoted by C1b (Galanakis et al., 2002a).
A great many Heusler alloys have been studied. However, no half metal
(HM) has been confirmed experimentally except NiMnSb (Hanssen et al.,
1990; Hordequin et al., 1996). They have been mostly predicted by first-
principles calculations. Tables 3.1 and 3.2 lists half-Heusler (HH) and
full-Heusler (FH) alloys, respectively, known to exhibit half metallic prop-
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Huesler alloys 73
Table 3.1 The magnetic moment/unit-cell and the measured Curie tempera-
ture (TC ) of half-Heusler alloys.
Compound Magnetic moment TC Reference
(µB /unit-cell) (K)
Exp. Theory Exp.
CoMnSb 4.0 490 Webster and Ziebeck (1988)
2.949 Galanakis (2005)
4.2 478 Otto et al. (1989)
FeMnSb 1.930 Galanakis (2005)
NiMnSb 4.0 Hanssen et al. (1990)
4.01±0.02 Hordequin et al. (1996)
4.02 Ritchie et al. (2003)
4.0 de Groot et al. (1983)
3.6 Clowes et al. (2004)
3.85 730 van Engen et al. (1983)
3.9±0.2 Turban et al. (2002)
3.991 Galanakis (2005)
4.0 Halilov and Kulatov (1991)
4.0 Block et al. (2004)
4.2 728 Otto et al. (1989)
PdMnSb 3.95 500 van Engen et al. (1983)
4.05 de Groot et al. (1983)
4.062 Galanakis (2005)
PtMnSb 3.97 582 van Engen et al. (1983)
4.14 572 Webster and Ziebeck (1988)
4.0 de Groot et al. (1983)
3.96 572 Otto et al. (1989)
3.997 Galanakis (2005)
4.003 Halilov and Kulatov (1991)
PtMnSn 3.42 330 van Engen et al. (1983)
3.60 de Groot et al. (1983)
3.51 Halilov and Kulatov (1991)
3.5 330 Otto et al. (1989)
Table 3.2 The magnetic moment/unit-cell and the measured Curie temperature (TC ) of ful-
l-Heusler alloys.
Compound Magnetic moment TC Reference
(µB /unit-cell) (K)
Exp. Theory Exp.
Co2 CrAl 4.811 Antonov et al. (2005)
2.999 Galanakis (2005)
0.53–0.86 310–330 Hirohata et al. (2005)
3.0 Block et al. (2004)
Co2 FeAl 5.29 (4K) Elmers et al. (2004)
4.9 Kelekar and Clemens (2004)
4.6–4.8 Hirohata et al. (2005)
4.996 Galanakis (2005)
Co2 FeSi 5.492 (RT),5.907(10.2K) >980 Niculescu et al. (1977)
5.492 (RT),5.97(5.0K) 6.000 1100 Wurmehl et al. (2005)
5.91 Hashimoto et al. (2005)
Co2 MnGe 5.11 Ambrose et al. (2000)
4.84 905 Brown et al. (2000)
5.00 Fuji et al. (1990)
5.012 Galanakis (2005)
4.93 Miyamoto et al. (2004)
5.0 Picozzi et al. (2002)
5.11±0.05 905±3 Webster (1971)
Co2 MnSi 4.96 985 Brown et al. (2000)
5.0 Fuji et al. (1990)
5.008 Galanakis (2005)
4.7(375◦ C) Kämmerer et al. (2004)
5.0 Kandpal et al. (2006)
5.0 Picozzi et al. (2002)
4.78 Ritchie et al. (2003)
4.95±0.25 Singh et al. (2004a)
5.1 Singh et al. (2006)
5.0 Wang et al. (2005a)
5.07±0.05 985±5 Webster (1971)
5.07 Westerholt et al. (2005)
Co2 MnSn 4.78 829 Brown et al. (2000)
5.03 Fuji et al. (1990)
5.089 Galanakis (2005)
5.0 Picozzi et al. (2002)
5.08±0.05 829±4 Webster (1971)
5.10 Zhang et al. (2005)
Co2 TiAl 0.70 148±2 Souza et al. (1987)
0.54–0.64 Ishida et al. (1982)
Co2 TiGa 0.82 128 Furutani et al. (2009)
0.80 130±2 Souza et al. (1987)
Co2 TiGe 1.94 2.0 380(5) Barth et al. (2010)
1.78 386(4) Carbonari et al. (1993)
Co2 TiSi 1.96 2.0 380(5) Barth et al. (2010)
1.10 375(4) Carbonari et al. (1993)
Co2 VAl 1.84 310(4) Carbonari et al. (1993)
Co2 VSn 1.20 105 Carbonari et al. (1993)
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Huesler alloys 75
The methods of growing Heusler alloys can be generally classified into two
categories: the growth of bulk samples and the growth of thin-film sam-
ples. Owing to the anticipation that spintronic device applications require
samples to be in thin-film forms especially, the MR properties are going to
be utilized, why is it interested in growing bulk alloys if most of the appli-
cations require samples having thin-film forms? One answer lies in the fact
that several thin-film growth methods, such as the radio frequency mag-
netron sputtering method and pulsed laser deposition method, use bulk
crystals or polycrystals (Giapintzakis et al., 2002; Caminat et al., 2004;
Shen et al., 2004) as sources. They add more incentives to grow bulk sam-
ples even polycrystals. Therefore, we decide to discuss methods of growing
both bulk samples and samples in thin-film forms.
Huesler alloys 77
We shall select a few examples for each of the three methods to illustrate
how the substrates are chosen and the effects of temperature.
added Si to obtain Co2 FeSi with fluxes of Fe and Co kept at the optimized
values. By varying the growth temperature between 100 ◦ C and 400 ◦ C,
they searched the optimal growth condition. This range of temperature
is considered low. Therefore, these authors adopted a low growth rate at
0.1nm/min in order not to degrade the quality of the crystal. These con-
ditions are essential to the growth of high quality crystal and interface.
The temperature of the Si cell (TSi ) varies between 1280 ◦ C and 1335 ◦ C
for the best stoichiometry. Characterization processes and magnetization
measurements confirm the quality of the single crystal and interface. Such
hybrid structures have been discussed in some detail by Hirohata et al.
(2006).
Huesler alloys 79
Fig. 3.2 Schematic setup of radio frequency sputtering method (Courtesy of Randy
Dumas).
not a major concern. For example, the lattice constant mismatch between
Co2 FeSi and MgO is 5.6%.
(2005a) to grow the FH alloy, Co2 MnSi. The targets for growing NiMnSb
were pressed and sintered polycrystalline pellets. Another possibility is to
obtain the pellets by radio frequency melting in Ar gas environment. For
Co2 MnSi, stoichiometric polycrystalline pellets were used.
The KrF excimer laser with the wavelength at 248 nm and pulse width
of 34 ns is used as the laser source for both growths. The laser beams are
incident on the rotating target at 45◦ . The energy of pulse is approximately
300 mJ and can have a maximum of 600 mJ. The repetition rate is 10 Hz.
Substrate The substrate for the growth of NiMnSb using PLD is poly-
crystalline InAs or Si(111) surface. For the growth of Co2 MnSi films,
GaAs(001) is used as an effective substrate. The substrates are cleaned.
One process for cleaning the GaAs(001) substrate is as follows: The sub-
strate is initially obtained from a commercial source. The first step is then
to degas in UHV at temperature up to 580 ◦ C. After that, the substrate is
sputtered by a 0.6 keV Ar+ ion beam with current density 4 µÅ/cm2 at 600
◦
C for 30 min, at an incident angle of 45◦ . The quality of the GaAs surface
is checked by AES and LEED to ensure absence of surface impurities and
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Huesler alloys 81
desired reconstruction.
Huesler alloys 83
The Heusler alloys are magnetic materials and many of them were pre-
dicted to show half metallic properties—they are the consequence of d-d
or d-p interactions. One property that is of central importance is the spin
polarization, P, at EF in the metallic channel. We therefore will focus on
magnetic moments of Heusler alloys, insulating gaps, and spin polarizations
of half metallic Heusler alloys. Disorder in a sample should also be con-
sidered because of the complication of atomic arrangements in a unit cell.
We divide the discussions of physical properties of Heusler alloys into three
main subjects:
Fig. 3.4 The Slater-Pauling curve for 3d TM elements and associated Heusler alloys,
filled circles are for Co2 -based alloys and filled grey circles are for other Heusler com-
pounds(Wurmehl et al., 2005). nv (nM ) is the average number of (magnetic) valence
electrons per atom.
Huesler alloys 85
where Nv is the total number of valence electrons per formula unit. Ex-
tending this argument to the FH alloys, there are 4 atoms/formula-unit for
a total 12 occupied bands in the ↓ spin channel, so that
MF H = Nv − 24. (3.5)
For example, half metallic Co2 FeSi has Nv = 30. The above rule then gives
MF H = 6µB , consistent with theory (Table 3.2).
One should bear in mind that the above rules (Eqs.(3.4) and (3.5))
determine only the magnetic moment/formula-unit for a known HM sample.
They do not predict whether a given sample is HM or not. Rather, the
values of n↓ are determined from a band structure for each alloy as suggested
by Kübler (1984).
(1990), Youn and Min (1995), and Halilov and Kulatov (1991). The spin-
orbit effect has been considered by Youn and Min (1995). Antonov et al.
(1997) extended the results of de Groot et al. (1983) to study the magneto-
optic Kerr effect in NiMnSb, PdMnSb, PtMnSb, and other ferromagnetic
ternary compounds. These results are listed in Table 3.3. Wang et al.
(1994) carried out calculations on NiMnSb and PtMnSb with the atomic-
sphere-approximation tight-binding LMTO method.
Most of these calculations of HH alloys show a general feature: the gap
is indirect between the occupied states at the Γ point and unoccupied state
at the X point. The valence states are predominantly the hybridizing p-
states of the Sb and d-states of the Ni (Pd, Pt) atom. The conduction state
at X is derived from antibonding d-states of the Mn atom. Therefore, p-d
hybridization plays a crucial role in determining half metallic properties
of HH alloys. The spin-orbit interaction does not diminish half metallic
properties in these alloys except for PtMnSb.
Fig. 3.5 PDOS of ↓-spin states at Γ-point of NiMnSb (Galanakis et al., 2002a).
Huesler alloys 87
sensitive to the U values and are derived from the antibonding states of
one of TM elements. Whether it is from the Co or Mn atom is not yet
clear. Similarly, the nature of the top of valence states is yet unclear.
In Table 3.7, we summarize reported values of the insulating gap in a
few Heusler alloys. For FH alloys, there are only theoretical values.
The nature of band-edge states of the minority-spin channel in Co2 MnSi
was examined by Galanakis et al. (2002b). They noted a stronger d-d
hybridization between the two Co atoms than between Co and Mn atoms.
Figure 3.6(a) shows the resulting states due to Co-Co interaction. The five
bonding states (lower group of five states in the middle of Fig. 3.6(b)) then
hybridize with the five states of the Mn atom, as shown in Fig. 3.6(b).
They form bonding and antibonding states. The five antibonding states of
the Co-Co interaction (top group of five states at the left of Fig. 3.6(b))
are located between the bonding and antibonding states resulting from the
hybridization of the Co and Mn atoms, and form nonbonding states between
Co and Mn hybridized states. These nonbonding states split into t1u and
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Huesler alloys 89
eu type states (states below and above EF in the middle of Fig. 3.6(b)).
EF falls in the gap between the t1u and eu states. Therefore, the gap states
are primarily from d-states of the Co atoms.
Fig. 3.6 Nature of gap states in the minority-spin channel: (a) the resulting states due
to Co-Co interaction, and (b) states around EF after Co-Co states interact with states
from the Mn atom (Galanakis et al., 2002b).
3.5.3 Polarization at EF
The polarization at EF , P, is a crucial quantity if any Heusler alloy is go-
ing to be used for spintronic applications. Table 3.8 lists the measured
and calculated values of P for bulk HH and Co-based FH alloys. A the-
ory of spin-polarized positron-annihilation on NiMnSb was put forward by
Hanssen and Mijnarends (1986).
Experiments were performed later by Hanssen et al. (1990) with the
positron momentum px in three different directions: [100], [110], and [111].
The [100] direction was analyzed at T=27 K and the other two directions
at 8 K. These experiments measured a P value of 100%, confirming half
metallic properties.
The point contact Andreev reflection method was recently applied to
Co2 Crx Fe1−x Si (Karthik et al., 2007). The measured P is 0.64 ± 0.01 at x
= 0.02. As compared to the value (0.57 ± 0.01) at x = 0, the presence of
Cr increases the P value. The authors suggest this increase is due primarily
to the doping effect that improves the L21 ordering structure as concluded
from their XRD and Mössbauer spectra.
Two groups have carried out first-principles calculations of these alloys.
Both groups used the KKR method with muffin-tin potentials. Fuji et al.
(1990) used the local spin density (LSD) approximation of von Barth-Hedin
(von Barth and Hedin, 1972) parametrized by Janak et al. (1975) for treat-
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Huesler alloys 91
compression of 1%, to aIII , places EF slightly above the gap. The effect of
lattice parameter on the local magnetic moment is determined within the
atomic sphere. For example, for NiMnSb, the lattice contraction increases
the hybridization between Ni and Mn atoms and increases the magnetic
moment of the Ni atom while decreasing the moment of the Mn atom. In
CoMnSb, which has a larger insulating gap, the moments on the Co and
Mn atoms are antiparallel. As a result, the transition of EF to the upper
edge of the insulating gap requires a large lattice contraction of ∼ 11%.
The magnetic moment at each atom is sensitive to the lattice parameter
while the total moment is relatively insensitive.
Huesler alloys 93
tra is the difference of intensities with the photon helicity (spin) parallel
(I+ ) and antiparallel (I− ) to the magnetization. The L2,3 excitations from
Mn 2p core states were measured.
Physically, when a core electron in the 2p level is excited, the absorption
process is governed by selection rules of the transition to 3d states of the Mn
atom. If the sample is half metallic, the final states of the absorption process
are majority-spin states at EF . The insulating minority-spin channel is not
expected to have any states to contribute. As the photon energy increases
to reach conduction states of the minority-spin channel, it is then possible
to obtain a finite intensity I. The difference I+ − I− can therefore reveal
half metallic properties. The XMCD Mn-L2,3 edge spectra of NiMnSb are
shown in Fig. 3.7. The negative part centred around 634.0 eV (L3 line) is
stronger than the positive contribution centred at 645.0 eV (L2 line).
Fig. 3.7 Experimental (Kimura et al., 1997) and theoretical (Galanakis et al., 2000)
absorption and XMCD spectra of Mn L2,3 lines in NiMnSb.
(3.6)
where j± mj± is an initial core state having energy Ej± , |nki is a final state
with energy Enk , p is the electron momentum, and e± are polarizations of
photons. The δ-function enforces conservation of energy. The calculated
peak positions agree well with experimental results. The two structures are
at 634 eV and 644.5 eV, respectively, (Galanakis et al., 2000) which agree
well with experimental results. The linewidth of the lower energy peak does
not agree due primarily to the fact that theoretical results were from ideal
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
crystals.
The emission process involves an electron near EF dropping into a core-
hole state. One measures the emitted X-ray. This process is more com-
plicated than absorption because the presence of a equilibrium hole in the
core region can affect energies of valence states. The corresponding struc-
tures in the measured spectra are at 640.5 eV and 652 eV, respectively
(Yablonskikh et al., 2000, 2001).
Another element-specific, surface-sensitive magnetic technique—the
spin resolved appearance potential spectroscopy (SRAPS)—has been ap-
plied to probe the surface magnetization of the NiMnSb(001) surface (Kolev
et al., 2005). The basic idea is related to core-core-valence Auger transi-
tions (Hörmandinger et al., 1988). In Fig. 3.8, the basic processes are
illustrated. The direct process has core-core transition between states 3
and 1 and valence-valence transition between states 4 and 2. The exchange
process is characterized by valence-core transition from state 4 to state 1
and core-valence transition from state 3 to state 2.
Fig. 3.8 Schematic energy level diagram for core-valence Auger transitions. Numbers
1 and 3 are core states; 2 and 4 are valence states (Hörmandinger et al., 1988).
Huesler alloys 95
dent on the location at the surface for the X-ray to create holes. The
measured surface SRAPS spectra are compared to bulk SRAPS spectra
in Fig. 3.9. The key feature in Fig. 3.9 is the ratio of amplitudes for the
majority- and minority-spin channels. The experimental result is approxi-
mately three times smaller than the theory based on the bulk model. This
indicates the reduction of spin polarization at the surface. This result agrees
with those of other experiments.
Fig. 3.9 Comparison of theoretical (bulk) and experimental (surface) APS spectra for
the ↓- and ↑-spin states (Kolev et al., 2005).
Fig. 3.10 Mössbauer spectra of Co2 Mn1−x Fex Sn (Zhang et al., 2005).
Huesler alloys 97
Fig. 3.11 (a) 55 Mn spin echo intensity as a function of frequency in Co2 Mn0.5 Fe0.5 Si.
The distribution of Fe atoms in the third coordination shell of the 55 Mn is also given; (b)
Relative area of resonance line vs. number of Fe atoms in the third coordination shell of
55 Mn. Shown are the results for the Co Mn
2 0.5 Fe0.5 Si compound with ideal stoichiometry
x = 0.5 (squares) and optimized stoichiometry x = 0.517 (triangles) (Wurmehl et al.,
2007).
Huesler alloys 99
Table 3.13 Three defect models studied by Orgassa et al. (1999, 2000).
Defect structure Type Site occupation
X(1) X(2) Y Z
X(1) ↔ Y 1 Ni1−x Mnx Nix Mn1−x Sb
X(1)Y ↔ X(2) 2 Ni1−x Nix Mnx Mn1−x Sb
YZ ↔ X(2) 3 Ni Mnx Sbx Mn1−x Sb1−x
where the first term is the total energy with disorder, the second term is
the total energy without disorder, ni is the number of atoms transferred to
and from a chemical reservoir of the i-th element, µ0,i is the corresponding
chemical potential, and X denotes Si or Ge. The chemical potentials of Mn,
Co and X were determined from fcc antiferromagnetic Mn, hcp ferromag-
netic Co, and diamond structure Si and Ge. The formation energies and
the total magnetic moments in the supercell are summarized in Table 3.15.
In the ideal (perfectly ordered) case, the total magnetic moment is 40 µB .
The Mn antisite has the smallest formation energy. They suggested that
it can be easily formed during growth, especially in the tri-arc Czochralski
method for Co2 MnSi at 1523 K. Experiments done by Raphael et al. (2002)
show there are 0.36x1022 cm−3 Mn antisites.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
3.6.1 NiMnSb
3.6.1.1 Stability of structure and half metallicity
Based on the theoretical studies of bulk properties of HH alloys, the
NiMnSb-type structure is the most stable among all HH alloys, with re-
spect to an interchange of atoms (Larson et al., 2000). Orgassa et al. (1999,
2000) showed that the half metallicity in NiMnSb is not destroyed by a few
% disorder. Galanakis et al. (2002b) attributed the gap in the non-metallic
channel to the d-d interaction between Ni and Mn atoms, which differs from
the explanation in terms of the d-p interaction between Mn and Sb atoms
proposed by de Groot et al. (1983). However, Galanakis et al. (2002b)
argued that the presence of Sb atoms is crucial to stabilize the structure
because NiMn does not favor any open structure.
Fig. 3.12 A schematic diagram of the DOS of a HM with the majority-spin channel
showing metallic behavior. The Fermi energy of the Heusler alloy is indicated as EF
and is located right below the conduction band edge by an spin-flip gap δ. ∆ is the
fundamental gap of the insulating channel.
the conduction band edge with their spin orientations flipped from the ↑
state to the ↓ spin state. As the conduction band of the ↓ spin channel is
occupied, the half metallicity is lost. The temperature for this to happen in
NiMnSb is 88 K (i.e., δ ≤ 88 K). Consequently, it is unlikely that NiMnSb
can be used for fabricating the RT spintronic devices.
Fig. 3.13 Surface band structure of the NiMnSb(111) surface (Jenkins, 2004). The
upper surface terminated by Mn is denoted as Aα , by Ni as Aβ , and by Sb as Aγ . The
bottom surface terminated by Sb is denoted by Bα .
that (100) interface diminished the spin polarization. On the other hand,
the (111) interface with interface composed of the Sb and S atoms as shown
in Fig. 3.14 restores half metallicity. It was also found that (100) and (111)
surfaces of NiMnSb do not exhibit half metallic properties.
Fig. 3.14 Structure of the half metallic NiMnSb(111)/CdS(111) interface (de Wijs and
de Groot, 2001).
also measured the Kerr rotation at RT on the same film using the MOKE
method. A laser source with a wavelength of 632.8 nm was incident on the
film surface at 45◦ . The magnetic field was swept between ±3.5 kOe. At
the limiting value of the magnetic field and in the polar configuration, the
Kerr rotation is 0.06◦ as compared to the bulk value of 0.10◦ ; It seems that
the film has magnetic properties close to those of the bulk.
Fig. 3.15 The hysteresis loop of a 220 nm NiMnSb film at 4.2 K (Kabani et al., 1990).
• For the Co atom, XAS shows lines at 776.5 eV and 791.5 eV corre-
sponding to 2p3/2 → 3d (first peak, L3 ) and 2p1/2 → 3d (second peak,
L2 ) transitions, respectively. The edges of L2,3 lines for the Mn atom
are at 638.5 eV and 651.4 eV.
• The line widths of the Mn atom are narrower than those of the Co
atom.
• The high energy line of Mn exhibits a doublet structure at 649.4 eV
and 650.8 eV.
• There is some variance with respect to the thickness of the films. For
example, the doublet structure of the Mn L2 line is smeared in a thicker
(45 Å) film.
• In contrast to the L3 line of the Co atom in other Heusler alloys such
as Co2 TiSn, Co2 ZrSn, and Co2 NbSn (Yamasaki et al., 2002), showing
multiple structures, this one does not exhibit any such structures.
to their environments.
The doublet structure in the higher energy line shows the atomic char-
acteristic of the Mn atom due to the fact that the Mn-Mn distance is larger
than the separation of the Co atoms and is attributed by Wang et al. (2005a)
to the interplay of (1) the exchange and Coulomb interactions between core
holes and unpaired valence electrons, and (2) the hybridization between 3d
orbitals and surrounding states. In a thicker film, the smearing out of the
structure according to Wang et al. (2005a) is due to the reduction of orbital
contributions as the thickness of the film increases. They also attributed
the absence of multiple structures appearing in Sn-based Heusler alloys to
more metallic Co d-states in Co2 MnSi.
The XMCD spectra are the normalized difference of the left- and right-
handed polarized X-rays. They probe the exchange splitting and spin-orbit
coupling of both initial core and final valence states.
Fig. 3.16 XAS Co-L2,3 lines (first and second peaks) of the Mn (a) and Co (b) atoms
from a Co2 MnSi film of 17 Å thick (Wang et al., 2005a).
Fig. 3.17 XMCD Co-L2,3 lines (first and second peaks) of the Mn (a) and Co (b) atoms
from a Co2 MnSi film of 45 Å thick (Wang et al., 2005a).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
• The structures in XMCD are in general 0.5 eV lower than the corre-
sponding peaks of XAS.
• For the Co atom, the L3 line is approximately a factor of two stronger
than the L2 line.
• For the Mn atom, the L3 line is more than a factor of three stronger than
the L2 line and about a factor of two stronger than the corresponding
Co line.
• Total intensities deduced from the sum rule correspond to 2.24 unoc-
cupied states for the Co atom and 4.52 for the Mn atom.
For the Mn atom, the large negative dichroism within the L3 region
(lower panel in Fig. 3.17) is attributed to the large separation between the
Mn atoms and smaller coordination numbers, resulting in weak Mn-Mn
magnetic coupling. The Mn atom, consequently, exhibits a large magnetic
moment. With 45 Å thick film, the sum rules provide information about the
spin moments: 1.04 µB per Co atom, comparing well with the predicted
value of 1.06 µB . However, one must bear in mind that the sum rules
underestimate both the effective spin moment and effective orbital moment.
The anisotropy in the plane of the film has been examined by measuring
the Kerr rotation in a 60 Å film under a magnetic field with different angles
between the easy [11̄0] and hard [110] axes. The geometry and results are
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 3.19 Geometric diagram of the Fig. 3.20 Kerr rotation as a function
magnetization and applied magnetic of magnetic field applied in the plane of
field (Wang et al., 2005a). a 60 Å film (Wang et al., 2005a).
shown in Fig. 3.19 and Fig. 3.20. Experiments were carried out at 80 K.
The important results shown in Fig. 3.20 are:
Fig. 3.21 Average spin-wave parameter b as a function of the inverse film thickness d
of the Co2 MnSi films grown on GaAs(001). The solid line is a linear fit (solid line) to
the experimental values (Wang et al., 2005b).
Fig. 3.22 SRPES of two Co2 MnSi thin films measured at RT. (a) and (c) are spin
resolved spectra. (b) and (d) are spin polarization (Wang et al., 2005b).
Fig. 3.23 ρ as a function of Tsub of a Co2 MnSi film (Singh et al., 2004b).
layer.
• Grow the upper magnetic electrode.
• Grow the upper current lead.
In practice, details of each step are important. During the growth, the
base pressure was 1x10−7 mbar. In order to have a large TMR, it is crucial
to have clean interfaces between Co2 MnSi and AlOx and between AlOx and
Co70 Fe30 —the upper magnetic electrode. This requires the whole growth
process to take place in a vacuum chamber without any interruption.
The steps for growing MTJs having SiO2 (substrate)/ V (42 nm)/
Co2 MnSi(100 nm)/ AlOx (1.8 nm)/ Co70 Fe30 (5.1 nm)/ Mn83 /Ir17 (10 nm)
are:
• Use dc magnetron sputtering to grow the V buffer, Co2 MnSi, and Al
layers.
• Oxidize the Al layer in pure oxygen plasma for 150 s.
• Anneal for 40 min at temperatures between 400 and 500 ◦ C.
The last step is an important one. It can result in ferromagnetically
textured Co2 MnSi with minimal disorder and homogenize the AlOx barrier.
It is tested by monitoring the magnetic moment/formula-unit, M, of the
ferromagnetic layer. After 1 hour at temperature above 250◦C, the value
of M is more than 4.0 µB . At 375◦ C, M reaches 4.7 µB . The annealing
helps the (110) texture formation in the Co2 MnSi layer as shown in the
X-ray diffraction pattern (Fig. 3.24). The existence of (110) texture was
evidenced by the gray (220) line (Kämmerer et al., 2004).
This line indicates that there is a periodic structure with periodicity
twice or four times longer than the distance between adjacent (110) planes.
When V is used, these two lines accompanied by the V(110) line and a
shoulder of the V(220) line are clearly seen (Fig 3.24).
• Oxidize the top AlOx layer for 50 seconds to clean the surface.
• Sputter Co70 Fe30 by dc magnetron sputtering to form the upper mag-
netic electrode.
• Use radio frequency magnetron sputtering to put an antiferromagnetic
Mn83 /Ir17 layer on top.
• Apply dc magnetron sputtering again to form a Cu/Ta/Au multilayer
lead.
• Impose an exchange bias between Mn83 /Ir17 and upper ferromagnetic
Co70 Ir30 electrode by annealing the multilayer structure for 1 hour at
275◦C under an external magnetic field with a strength of 100 mT.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
• Pattern the multilayer material using optical lithography and ion beam
etching to form a 200 µm quadratic (four different layers) MTJs.
The next issue is the determination of the TMR. Two quantities were
measured by Kämmerer et al. (2004) to determine the TMR of MTJs: The
resistance under a magnetic field and the change of resistance ∆R which is
defined as the difference of the resistances when the magnetization between
the two electrodes are parallel and antiparallel. The ratio ∆R/R(Hext )
of these two quantities is related to the TMR, where R is in general the
resistance measured when the two electrodes having parallel magnetization.
The ratio plotted against the external field of a is shown in Fig. 3.25.
Figure 3.26 explains schematically the tunnel processes (Sakuraba et al.,
2006). Note that in a case to be discussed, both the sample and the upper
electrode are made of the same materials. The upper panels are for the
parallel configuration of magnetizations in the sample (at right) and upper
electrode (at left). A bias potential e∆V is applied such that a positive
bias causes an electron to tunnel from the right to the left. The rectangular
area indicates the tunneling barrier. The density of states showing the half
metallicity of the sample and upper electrode are plotted on both sides
of the barrier. The gap in the minority-spin channel is denoted by EG .
δCB is the energy separation between the bottom of conduction bands and
EF , and δV B is the energy separation between EF and the top of valence
bands. When e∆V = 0, EF is aligned on both sides. An electron in the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
majority-spin channel can tunnel from the sample to the upper electrode.
When e∆V > EG , EF in the upper electrode is shifted down by e∆V. A
new tunneling channel between minority-spin states is open as the result.
Fig. 3.26 Schematic of electron tunneling process at finite bias voltage in the
Co2 MnSi/Al–O/Co2 MnSi system for parallel (upper panels) and antiparallel (lower pan-
els) configurations(Sakuraba et al., 2006).
field.
ATMR (T)
TMR(Hext ) = [1 − cos(θsample − θupper )], (3.12)
2
Msample cos(θsample ) Mupper cos(θupper)
M(Hext ) = + , (3.13)
[tsample /(tsample + tupper )] [tsample /(tsample + tupper )]
where AT MR is the measured amplitude of the ratio ∆R/R(Hext ). It is
called the TMR-effect amplitude and is a function of T. The measured
∆R/R(Hext ) as a function of external magnetic field for MTJs is shown
in Fig. 3.25. Using Eqs. (3.12) and (3.13), the corresponding TMR as
a function of T is shown in Fig. 3.27. The TMR-effect amplitude at low
temperature is 86%. More recently, Sakuraba et al. (2006) fabricated a
Co2 MnSi/AlO/Co2 MnSi MTJs and determined the TMR as a function of
T. At RT, the value of the TMR approaches 67%, while at 2K the value is
570%.
For MTJs fabricated by Sakuraba et al. (2006), its P value can also
be determined from Jullière’s formula. Using P = 89% for the Co2 MnSi
electrode and TMR-effect amplitude of 570%, the P value of the Co2 MnSi
sample is 83%. In addition, Schmalhorst et al. (2004) grew MTJs similar to
that of Kämmerer et al. (2004) except they replaced Co70 Fe30 by an anti-
ferromagnetic layer of Mn83 Ir37 . The P value for this sample is determined
to be 86% at 10 K.
By comparing the values of P determined from different magnetic mea-
surements, it is found that single crystalline Co2 MnSi films grown on
GaAs(001) substrates show the smallest P values. The possible reasons
for this discrepancy are: (i) there is another nonstoichiometric phase dis-
tributed inside the film, (ii) an nonmagnetic metallic phase in the sur-
face/interface regions could add an equal contribution to both spin com-
ponents and thus reduce the polarization. Wang et al. (2005a) suggested
that the possibility of the second reason is unlikely according to their Kerr
effect as a function of thickness measurements which indicate that the sur-
face/interface regions are magnetic.
To explain the 61% value of P which is substantially less than the pre-
dicted 100%, Schmalhorst et al. (2004) attributed the reduced value to spin
scattering of tunneling electrons from the paramagnetic Mn ions and atomic
disorder in the interface region. There is also the possibility of intermixing
of Co and Mn, and Co atoms occupying Mn sites. As shown by Picozzi
et al. (2004), with Mn sites occupied by the Co atoms, half metallicity per-
sists in Co2 MnSi. Neutron diffraction experiments (Raphael et al., 2002)
suggest that there are approximately 10–15% Co-Mn antisites exit in the
sample, with Mn sites occupied by the Co atoms.
Based on the above discussions, it should be clear that the TMR of
MTJs with the Co2 MnSi as the lower electrode (sample) depend critically
on the choices of the seed layer and materials for the upper electrode.
3.6.3.2 Characterizations
Bragg reflection of X-rays has been used to characterize the films (Schneider
et al., 2006). In Fig. 3.28, the Bragg lines of a Co2 FeSi film grown on
MgO(100) substrate are shown. The (200) and (400) lines are sharp with
widths of 0.3◦ . The results indicate that the growth in the out-of-plane
direction is good. For the films grown on Al2 O3 (112̄0) substrate, there are
three different epitaxial domains. Only Co2 FeSi films grown on MgO are
fully epitaxial (defect-free layers), as revealed by LEED pattern (Fig. 3.29).
The spots clearly show the four-fold symmetry.
Fig. 3.28 X-ray Bragg reflection lines Fig. 3.29 The LEED pattern identify-
from a Co2 FeSi film ing epitaxial growth of Co2 FeSi grown
grown on MgO(100) (Schneider et al., on MgO (Schneider et al., 2006). The
2006). primary electron beam energy is 318 eV.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Co2 FeSi (110) film on Al2 O3 (112̄0) substrate Two Co L2,3 lines deter-
mined by TEY are recorded in both XAS and XMCD experiments.
The results are shown in Fig. 3.31. Assuming that the luminiscence
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 3.31 XAS and XMCD spectra Fig. 3.32 XAS and XMCD spectra
of Co lines from a 68 nm Co2 FeSi of Fe lines from a 68 nm Co2 FeSi
film grown on a Al2 O3 (112̄0) substrate film grown on a Al2 O3 (112̄0) substrate
(Schneider et al., 2006). (Schneider et al., 2006).
For the Fe lines on the similar film sample, the two peaks in the XAS
spectra are located at 711 and 724 eV, respectively (Fig. 3.32). Only
a very weak shoulder at 715 eV is seen. The shifts of energies in the
XMCD spectra are relatively small compared to the Co case. Using
the sum rule with assumed number of d-holes for each element, nd (Fe)
= 3.4 and nd (Co) = 2.5 as reported for the pure elements (Wurmehl
et al., 2006b), magnetic moments of the Co and Fe atoms were deter-
mined. They are summarized in Table 3.17. Where TMD indicates
that the value was deduced from transmission data. The ratios of the
orbital and spin moments for Co and Fe are 0.12 and 0.04, respectively,
from the TEY results. The ratio for the Co atom (0.13) agrees well
with that in crystalline Co (Carra et al., 1993). However, for Fe, the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
All these results suggest that localized properties, such as the element
specific spin moment, are not affected by whether the sample is in thin-
film form or bulk. Theoretical predictions of spin moments of the Co
and Fe atoms based on LDA by Galanakis1 are 2.3 µB and 3.5 µB , re-
spectively. The spin moment calculated with LDA+U (Kandpal et al.,
2006) of Co is 1.53µB and that for Fe is 3.25 µB . The magnetic mo-
ment is 6.0 µB /unit-cell. The last value was confirmed experimentally
(Wurmehl et al., 2005). The deduced element specific spin moments
given in Tables 3.17 and 3.18 are smaller than the theoretical values,
in particular, the LDA results.
Fig. 3.34 (a) Relative orientations of the incident s-polarized light (S), magnetization
in the film (ML and MT ), and the external magnetic field H8 . (b) Relative orientations
of crystal axes, magnetization of the film, and the external magnetic field in different
orientations Hi . The electric field is perpendicular to the beam velocity (Hamrle et al.,
2007).
The MOKE loops in most Heusler alloy films are symmetric with re-
spect to angle α (Fig. 3.34). As experiments have progressed, evidences
have emerged that MOKE loops can be asymmetric. Postava et al. (2002)
attributed the asymmetry to the quadratic behavior of MOKE, QMOKE,
ML MT , and ML2 − MT2 .
Physically, the observation of the Kerr effect depends on both the spin-
orbit and exchange interactions. Unlike PMOKE and LMOKE, where the
magnetization M is parallel to the wave vector of the light k, QMOKE
requires M to be perpendicular to the plane of incidence.
Therefore, the first order contribution of the spin-orbit interaction does
not play a role. The second order contribution of the spin-orbit interaction
is smaller so QMOKE is a smaller effect. To show that QMOKE depends
on the quadratic contributions of ML and MT , we outline the macroscopic
derivations (Foner, 1956) as follows: The response function of a magnetic
medium to the light can be described by the dielectric tensor,
(0)
εij =εij +Kijk Mk +Gijkl Mk Ml +..., (3.15)
(0)
where εij is the linear dielectric tensor, Kijk and Gijk are the linear and
nonlinear magneto-optical tensors, respectively. Repeated indices are used
as the summation convention. The symmetries required by the Onsager
relation give the following conditions:
(0) (0)
εij (M) =εij (−M) = ε0 δij ,
Kijk = −Kjik , Kiik = 0, i 6= j 6= k, (3.16)
Gijkl = Gjikl = Gjilk Gijlk .
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 3.35 QMOKE Kerr rotation loops in a 21 nm Co2 FeSi film with ϕ = 0.5◦ and α
= –22.5◦ . The external magnetic field is in H8 k y direction. The filled circle at H = 0
is the QMOKE signal in saturation as determined by the 8-directional method for α =
–22.5◦ (Hamrle et al., 2007).
Fig. 3.36 SRPES spectra of (a) Co2 FeSi/Al2 O3 and (b) Co2 FeSi/MgO (Schneider et al.,
2006).
Fig. 3.37 Calculated DOS of bulk Co2 FeSi (Wurmehl et al., 2005).
over the whole energy range of measurements for both films. Therefore, the
spin polarization is positive. The maximum spin polarization is 4% for the
Co2 FeSi/Al2 O3 film and 12% for the Co2 FeSi/MgO film. P (EF ) is 0 for
the former and 6% for the latter. The reduction of the spin polarization is
not confined to the top surface layer. It extends 4-6 nm into the sample
for the Co2 FeSi/MgO film. Compared to XAS and XMCD, which measure
the integrated spectra, SRPES probes states near EF in a very thin region
of the surface. These spectra should therefore be more sensitive to disorder
in the surface region.
Fig. 3.38 M(T) as a function of temperature for a Co2 FeSi film grown on the MgO
substrate (Schneider et al., 2006).
law:
C
χ(T ) = . (3.21)
T − TC
The value of TC is 1100±20 K (Wurmehl et al., 2005).
Fig. 3.39 A schematic setup of VSM. The sample coil senses the signal proportional to
χ. The cup transmits the sound wave to the sample (Foner, 1956).
Fig. 3.40 Resistivity as a function of temperature in a typical Co2 FeSi film (Schneider
et al., 2006).
The measured results by Schneider et al. (2006) are shown in Fig. 3.41.
ρ⊥ (the top curve) is larger than ρk (the bottom curve). This is attributed
to the spin-orbit interaction. At H = 0, the spontaneous anisotropy can be
determined from the expression,
ρk − ρ⊥
1
. (3.22)
3 ρk + 2ρ⊥
Chapter 4
4.1 Introduction
Among the various oxides, a few are now known to show half metallic
properties. They are:
The last of these emerged from the study of high TC superconductors. All
three are appealing for spintronic applications because they are predicted
to have 100% spin polarization at EF . Experimentally, CrO2 exhibits P up
to 97% at 4 K (Ji et al., 2001). These materials are expected to show large
tunnel magnetoresistance (TMR) when used to form electrodes sandwiching
a thin insulating region, and so are good candidates for spintronic devices
such as spin valves and sensors.
The above three oxides are of comparable complexity to the Heusler
alloys with respect to growth and structure. In this chapter, we shall discuss
each one separately, considering the growth, characterization, and magnetic
and electronic properties for each in detail.
4.2 CrO2
131
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
4.2.1 Structure
CrO2 has the simplest structure among the three oxides discussed above.
It crystallizes in the rutile (TiO2 ) structure (Fig. 1.3).
The space group is D14 4h ( P42 /mnm), which is a nonsymmorphic group.
There are two formula units per unit cell. The Cr atoms form a body-
centered tetragonal lattice, as shown in Fig. 1.3. Each of the Cr atom is
surrounded by an octahedron of O atoms. The orientations of the two oc-
tahedra differ by a 90◦ rotation about the z-axis (parallel to the shorter
lattice vector, c). The positions of the Cr and O atoms are given in Ta-
ble 4.1. Denoting the longer lattice vector by a in the x-y plane, the c/a
ratio is 0.65958 (Schwarz, 1986). The lattice constant, a, is determined by
Thamer et al. (1957) to be 4.42 Å. Later, Porta et al. (1972) confirmed
these results.
Table 4.1 Positions of Cr and
O atoms in rutile structure ex-
pressed in reduced coordinates,
where δ is 0.3053.
Atom Coordinates (x,y,z)
Cr (0,0,0)
Cr (0.5, 0.5, 0.5)
O (δ, δ, 0)
O (-δ, -δ, 0)
O (0.5+δ, 0.5-δ, 0.5)
O (0.5-δ, 0.5+δ, 0.5)
4.2.2 Growth
Cr forms many competing oxides, namely Cr3 O4 , Cr2 O5 , CrO3 , and Cr2 O3 .
Crystalline CrO2 is the only oxide which is ferromagnetic at room tem-
perature (RT). But it is metastable at atmospheric pressure and is easily
decomposed to Cr2 O3 with heat. Therefore, it can be difficult to grow.
The successful growth techniques can be classified into two categories: the
“high-pressure” and “low-pressure” techniques. Thermal decomposition
(Ranno et al., 1997; Hwang and Cheong, 1997) is the high-pressure tech-
nique. Chemical vapor deposition (CVD) (Li et al., 1999; Ivanov et al.,
2001; Kämper et al., 1987; Gupta et al., 2000; Anguelouch et al., 2001) is
the low-pressure technique. More recently, pulsed laser deposition (PLD)
has been used to grow CrO2 (Shima et al., 2002). PLD is a low-pressure
technique and has been discussed in Chapter 2. In the following, we shall
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.1 The oxygen pressure as a function of temperature recorded by TPA for an 18
mg CrO3 sample under heating at 0.5 ◦ C/min (Ranno et al., 1997).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
4.2.3 Characterization
There are three techniques to characterize the quality of CrO2 in thin-film
form: RHEED (Li et al., 1999), X-ray diffraction (Gupta et al., 1999), and
AFM (Anguelouch et al., 2001).
Fig. 4.2 The RHEED patters obtained by Li et al. (1999). (a) On TiO2 (100), beam
< 001 >, (b) on TiO2 (100), beam < 110 >, (c) on Al2 O3 (0001), beam < 11̄00 >, (d) on
Al2 O3 < 0001 >, beam < 112̄0 >.
tion of 2θ of those three cases is given in Fig. 4.3. The first peak in each
case is associated with TiO2 and the second peak with CrO2 , because the
lattice constants of the sample and substrate are different. The lines for
TiO2 are very narrow for all three directions, and are reproducible, indicat-
ing that substrate surfaces were all properly cleaned. The lattice constant
mismatches are large in the three orientations. For (100) and (010) orien-
tations, it is 3.79%, and for (001) it is 1.48%. Therefore, there is a tensile
strain in the in-plane direction and a compressive strain in the perpendic-
ular direction. The peaks in case (c) show the largest shift to the higher
angle. The largest full width at half maximum for the rocking curve1 of the
second peak in case (b) is 0.4◦ , which is considered to be narrow.
Fig. 4.3 X-ray intensity as a function of 2θ (◦ ) with the substrate surfaces: (a) (200),
(b) (220), and (c) (002) (Gupta et al., 1999).
Fig. 4.4 AFM image of a 1050 Å CrO2 sample. The distinguishable lines correspond to
the atomic planes of CrO2 . The inset shows a sample cross section of the region marked
A with 4.4 Å steps (Anguelouch et al., 2001).
veloped for powdered forms of oxides, we shall discuss first this technique
and the associated powder magnetoresistance (PMR). We are particularly
interested in the half metallic properties of CrO2 . The experimental de-
termination of these properties will be discussed. Finally, the electronic
properties of this compound will be discussed based on photoemission ex-
periments and theoretical grounds.
M R ∼ (1 − Tb/d ). (4.5)
Fig. 4.5 Schematic of a point contact (Coey et al., 2002). The shaded triangular areas
are two magnetic domains. Their magnetizations are oriented relative to each other by
θ12 (not indicated). The dashed lines with arrows indicate a path of ballistic transport
while the solid lines show the diffusive transport from the left to the right.
Fig. 4.6 Basic schematic setup of point contact magnetoresistance measurement (Garcı́a
et al., 1999). M1 and M2 form the point contact. The battery connected to M2 applies
the bias voltage.
Measured results The PMRs of CrO2 for a pressed powder and for a
powder diluted with 75 wt% insulating Cr2 O3 particles with similar shape
were measured as a function of temperature. The results are shown in
Fig. 4.7. At 5 K, PMR is approximately 50% for the diluted powder com-
pact (Coey et al., 2002). The value of MR for undiluted CrO2 compact
(black squares) is approximately 30%. The corresponding polarizations are
62% and 82%, respectively. However it is only 0.1% at RT (Coey et al.,
2002).
Fig. 4.7 PMR results for CrO2 (black Fig. 4.8 Magnetization (solid curve)
squares) and 25% powder compact of and MR (dashed curve) at 5 K as func-
CrO2 (open circles) (Coey et al., 2002). tion of fields up to 60 kGauss (kG) for
a CrO2 powder compact (Coey et al.,
2002).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Tip and Probes The tips were obtained from wires with 0.03 inch in
diameter (Ji et al., 2001), treated first by mechanical polishing and then
electrochemical etching. Since the conductance as a function of voltage was
to be determined, a conventional four-probe scheme (e.g., Fig. 2.20) was
used.
Setup The setup is shown in Fig. 2.33. The tip was controlled by a
differential screw mechanism for microscopic movements toward the sample
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.10 (a) Photoemission EDC of polycrystalline CrO2 films measured at 300 K for
different times of sputter cleaning. (b) Spin-polarized spectrum of polycrystalline CrO2
films at 300 K after tsp = 120 s (Kämper et al., 1987).
summarized as follows:
• After tsp = 210 s, a shoulder developed at energy 2.3 eV below EF
for the total EDC. This feature agrees well with the earlier results of
Kämper et al. (1987). A similar feature is exhibited in the ↑ spin chan-
nel. The ↓ spin states do not show any such structure. The polarization
at EF is 95%.
• After tsp = 750 s, the shoulder of the total EDC is shifted to 2.0 eV
and its strength is increased.
• After tsp = 750 s and additional annealing for 12 hours at 150 ◦ C, the
shoulder feature persists. The polarization at EF is 85%.
Fig. 4.11 Photoemission EDC of CrO2 (100) surface for different sputter cleaning times
(Dedkov, 2004). The energy scale is referenced to EF .
CrO2 film grown by CVD on Al2 O3 (0001) substrates. The easy axis is in
the plane of the film. Both in-plane and out-of-plane magnetizations van-
ish at 386 K, as shown in Fig. 4.12(a), which led these authors to conclude
that TC is 385 K. An earlier experiment (Kouvel and Rodbell, 1967) on
compressed powdered stoichiometric CrO2 reported a TC of 386.5 K. These
two results agree well with each other.
Fig. 4.12 (a) M(T) as a function of T for thin-film form of CrO2 on Al2 O3 (0001)
substrates (Kuneš et al., 2002). (b) Hysteresis loops of CrO2 (100) films epitaxially
grown on a TiO2 (100) substrate with at T = 10 K (dots) and 300 K (open squares)
(Spinu et al., 2000).
Fig. 4.13 Band structure of CrO2 (Brener et al., 2000). (a) ↑ (majority) spin states;
(b) ↓ (minority) spin states.
Qualitatively, the three results given in Table 4.2 show a minimum in the
DOS at EF of the majority-spin channel. The presence of a minimum agrees
reasonably well with the photoemission spectra. However quantitatively,
the amplitude of the density of states differs by nearly a factor of eight.
Whereas the agreement of the gap values for minority-spin states is much
better.
All the theoretical calculations agree that the d-states of the Cr split
into a triply (t2g ) and a doubly (eg ) degenerate sets of states under the crys-
tal field of the O atoms. The bonding properties of CrO2 show both ionic
character and d-p hybridization of more covalent character. Korotin et al.
(1998) show that it is the eg states originating from the Cr atom which con-
tribute to the hybridization. The crystal field and hybridization combined
with the exchange interaction gives rise to the half metallic properties.
• The hole surface determined by Mazin et al. (1999) touches the A-Z
plane and forms hammerheads
• Mazin et al. (1999) find also needle shaped surfaces.
Fig. 4.15 The BZ and Fermi surfaces in the sections of ΓZAM and ΓZRX (Brener
et al., 2000). The grey region is the electron surface while the dashed regions are the
hole surfaces.
with respect to the LSDA results and to make the dip in the DOS at EF
of majority-spin channel more pronounced. Given these effects, it is still
controversial whether it is necessary and appropriate to determine the band
structure of CrO2 with LSDA + U.
4.3 Fe3 O4
4.3.1 Structure
The crystal structure above TV has been determined by X-ray diffraction
and its magnetic structure has been probed by neutron scattering (Shull
et al., 1951). It is the cubic spinel structure. The primitive lattice is fcc
with lattice constant of 8.396 Å. Associated with each fcc lattice point,
there are two smaller cubes as shown in Fig. 4.16. The space group of the
crystal is Fd3̄m. Each primitive fcc cell has the following configurations of
atoms:
• Fe2+ sites: The unit cell contains two Fe2+ ions, one of them is at the
origin of the unit cell—at the lower left corner (black polygon), and
the other one is at the center of one of the small cubes (open polygon).
They are at the tetrahedral (A) sites.
• Fe3+ sites: The other Fe ions are 3+ and are located at four corners of
the smaller cube (gray polygons) with a vacancy at its center. These
Fe ions are at the octahedral (B) sites.
• O sites: The O atoms occupy four corners of each small cube (dia-
monds).
In Table 4.3, we list the atom positions in reduced coordinates of the
lattice constant a. The magnetic structure of Fe3 O4 was first proposed
by Nèel (1948). All Fe atoms at B sites are ferromagnetically aligned
with each other and antiferromagnetically aligned with those at A sites.
This structure was experimentally confirmed by Shull et al. (1951).
Fe3+ ions on the A sublattice have magnetic moments of 5 µB while
Fe2+ ions have magnetic moments of 4 µB and 5 µB .
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.16 Part of the spinel structure associated with one lattice point (lower left corner)
of the conventional fcc structure. The tetrahedral (A) sites (black and open polygons) ,
the octahedral (B) sites (gray polygons) and the oxygen atoms (diamonds) in Fe3 O4 .
Fig. 4.17 The unit cell (outlined in thick lines) of Fe3 O4 at temperature below TV
(Wright et al., 2002).
4.3.2 Growth
Tiny single crystals and thin films of Fe3 O4 have been grown.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
tional cell of Fe3 O4 above TV . Before the growth, MgO was cleaved to MgO
(100) and annealed at 650 ◦ C for 1–2 hours under an oxygen pressure of
1.3×10−7 mbar. The base pressure in the growth chamber was maintained
at 1.3 ×10−10 mbar. The Fe atoms were emitted from a water-cooled ef-
fusion cell. They formed thin films of Fe3 O4 with the oxygen atoms under
pressure in the order of 6.7×10−7 mbar.
4.3.3 Characterization
Most layered forms of Fe3 O4 have been characterized by RHEED, TEM,
and STM methods. Since we have discussed the growth of Fe3 O4 by PLD,
we shall discuss the characterizations based on the same work.
The characterizations discussed by Reisinger et al. (2003) involve several
stages:
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
4.3.3.1 Si substrate
The Si substrate was heated to 600 ◦ C, then treated at Tsub = 900 ◦ C and
1150 ◦ C to remove amorphous Si oxides on Si(100), and eventually cooled
back down to 600 ◦ C for the growth of the buffer layers. RHEED patterns
were taken at the [110] azimuth. The sequence of the patterns is shown in
Fig. 4.20.
Figure 4.20(a) shows the RHEED pattern before the heat treatment.
It shows a homogeneous background and a few spots. The background is
due to the presence of amorphous Si oxides. The spots are the constructive
interference pattern of Si(001) surfaces. As the temperature increases to
about 900 ◦ C, it exhibits repeating long and short lines (Fig. 4.20(b)),
indicating a 2×1 reconstruction, with the lines somewhat streaky. The
homogeneous background is no longer visible. As the temperature increases
to 1150 ◦ C, the pattern characterizing the 2×1 reconstruction disappears.
This suggests that the dimmer bonds forming the 2×1 reconstruction have
dissolved. When the Si substrates are cooled down to 1000 ◦ C, sharp Laue
spots are clearly visible, indicating that the oxides have been removed. The
five spots indicate that a clean 2×1 surface has been obtained.
Fig. 4.20 The RHEED patterns of Si substrates during heat treatments (Reisinger
et al., 2003). (a) Tsub = 600 ◦ C, (b) after heating to 890 ◦ C, (c) 1150 ◦ C (the maximum
temperature), and (d) 1000 ◦ C .
Fig. 4.21 The RHEED pattern of TiN on a Si (100) substrate at Tsub = 600 ◦C
Fig. 4.22 The RHEED pattern of Fe3 O4 on a TiN buffer layer after 5000 pulses
(Reisinger et al., 2003).
Fig. 4.23 Spin polarization as function of binding energy for a 300-layer oxidized Fe
after annealing at 250 ◦ C (Dedkov et al., 2002). The P value is –(80±5)%. The thin
solid line through the data points was fitted by a three-point averaging fast Fourier
transformation procedure.
grown on MgO. The thicknesses of the films were 200, 50, and 15 nm,
respectively, with 10% uncertainty. The current was along the [100] axis
and the magnetic fields were applied along the [100] and [010] directions.
For the thickest films, additional measurements were made with the current
flowing in the [110] direction and magnetic fields oriented along the [110]
and [11̄0] axes. The four-point geometry with silver-paste contacts was
used to measure the resistivity. The MR was determined with respect to
the resistivity at the coercive field at which there is no magnetization in
the sample. A typical result of the MR showing the anisotropy is given in
Fig. 4.25. It is clear that the parallel and perpendicular MRs are different.
However, clear also is the anisotropy of each. Ziese and Blythe (2000)
attributed the anisotropy of the MR to the anisotropy of the samples.
Fig. 4.25 The magnetoresistance ratio Fig. 4.26 The magnetization as a func-
ρ/ρ0 of the Fe3 O4 single crystal and 200 tion of H of an Fe3 O4 /CoCr2 O4 /LSMO
nm thick film. The solid lines are mea- trilayer (Hu and Suzuki, 2002).
surements in longitudinal geometry and
the dashed lines in transverse geometry
(Ziese and Blythe, 2000).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.29 MR as a function of conductance at bias voltage of 0.3 V (filled circles) and
zero bias (open squares) for Fe3 O4 nanocontacts (Versluijs and Coey, 2001).
Fig. 4.30 MR as a function of magnetic field for Fe3 O4 grown on MgO(001) (a); and
for the two spin valves (b) and (c) (van Dijken et al., 2004).
the B sites. Two of them, Fe (B1 ) and Fe (B2 ), are at the centers of oxygen
octahedra. Fe (B3 ) and Fe (B4 ) are off center.
Resonance X-ray diffraction was used because the technique yields both
the site selective diffraction and the local absorption spectroscopy regarding
atomic species. Reflections in the range of tens of electron volts can be
recorded around the absorption edge of an element or elements where there
are evidence showing strong energy and angular dependencies (Nazarenko
et al., 2006). Physically, it involves virtual photon absorption-emission by
an electron initially occupying a core state and making a resonant transition
to some intermediate state near EF . This technique is particularly effective
for probing charge, orbital or spin orderings under distortions. For example,
different resonance frequencies can be correlated to the charge difference
at atoms suffering dissimilar distortions. By measuring the shift of the
absorption edges, the valences of the atoms were determined to be 5.38,
5.62, 5.40, 5.60 e for Fe (B1 ) to Fe (B4 ), respectively. On the other hand,
Wright et al. (2001) deduced from their neutron and X-ray measurements
the valences of the four Fe ions to be 5.6, 5.4, 5.4, and 5.6 e, respectively.
Therefore, there are some discrepancies for Fe (B1 ) and Fe (B2 ) between
these two measurements.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.31 (a) Intensity of photoemission spectra of opposite magnetization (dots and
solid lines) as a function of binding energy for different layer thicknesses (Kim et al.,
2000a). (b) The corresponding spin-resolved curves. The spectra for the pure Fe are
shown at the bottom of both panels. The binding energy is measured with respect to
EF .
Fe sample shows a narrow peak at around 53.5 eV. As the thickness in-
creases, the peak becomes broader because in Fe3 O4 the Fe ions have two
different charges, Fe2+ and Fe3+ . The difference spectra indicate antipar-
allel coupling between the oxide layers and the Fe substrate. This can be
seen in Fig. 4.32(b). For the Fe, the difference spectrum shows a positive
slope at 53.0 eV. The 900-layer case (top curve) shows negative slope at
around 54.0 eV. The authors confirmed this by growing more Fe on top
of Fe3 O4 . As the Fe film thickness increases (1.2 ML), the onset of MLD
changes its sign: i.e., the Fe metal and thick oxide overlayer have opposite
magnetization. Kim et al. (2000a) attributed this to an antiparallel mag-
netic coupling in the Fe3 O4 (111)/Fe(110) bilayer. This essential feature
can be seen in a schematic diagram (Fig. 4.33). As Compared to SPPS for
the bulk-terminated surface of Fe3 O4 , there are two important features:
Fig. 4.32 (a)The MLD spectra and (b) the corresponding differences of MLD signals
for four films of different thickness as a function of binding energy (Kim et al., 2000a).
polarized band structure of Fe3 O4 in 1991. They used the linear muffin-
tin orbital method within the atomic sphere approximation (LMTO-ASA)
(Andersen, 1975) and von Barth and Hedin exchange-correlation functional
(von Barth and Hedin, 1972) . The LMTO-ASA method is extensively dis-
cussed by Skriver (1983).
The Fe3 O4 structure is shown in Fig. 4.16. A lattice constant of 8.397
Å and 18 empty spheres were used. In Fig. 4.16, the Fe atoms are located
at the A and B sites. In terms of crystallographic terminology (Shepherd
and Sandberg, 1984), they are at 8a and 16d sites, respectively. The frozen
core approximation was applied to 3p-states of Fe and 1s-states of O. We
summarize the sites and muffin-tin sphere radii used in Table 4.5. Muffin-
tin orbitals of s, p, d on the Fe and O atoms and s and p type on the
empty spheres were used as basis functions. The total charge density was
constructed using a total of 56 k-points in the irreducible part of BZ.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.34 The spin-projected densities of states (DOS) and numbers of states for mag-
netites (NOS, dot lines) of the majority- and minority-spin channels in Fe3 O4 (Zhang
and Satpathy, 1991). The net magnetic moment is 52 − 44 = 8µB per unit cell which
consists of two formula units.
where nA A
0 (n ) is the total number of d-electrons at site A before (after)
the system exhibits magnetization. U A is the on-site Coulomb repulsion
energy at site A. A similar set of equations can be written for site B. Since
the two sites are not distinguished in Eq. (4.12), the present theory cannot
predict the relative orientation of the magnetizations at the two sites. The
Stoner conditions for the A and B sites are:
InA↑ n↓ N A (nA , mA ) = 1
(4.15)
InB↑ n↓ N B (nB , mB ) = 1.
By equating the two chemical potentials, µ↑ and µ↓ , at each site, one has:
εA A A A A A A B B B B B
↑ + ε↓ + 2(n − n0 )U + In↑ n↓ n = ε↑ + ε↓ + 2(n − n0 )U + In↑ n↓ n
B B
(4.16)
There is also the condition of conservation of total electron number with
and without the magnetization.
nA + 2nB = nA B
0 + 2n0 . (4.17)
The factor of 2 is due to the fact that there are twice as many Fe atoms
at B sites. The n’s and m’s are obtained by solving Eqs. (4.13) –(4.15)
self-consistently. Instead of carrying out self-consistent calculations, one
can estimate the magnetizations at the A and B sites. In Eq. (4.14), the
quantity U is the largest among all the terms. To satisfy Eq. (4.17), nA =
nA B B
0 and n = n0 are obtained. The Stoner parameters can be calculated by
using the scheme of Poulsen et al. (1976). They turn out to be between 60–
80 meV. The magnetizations at the two sites mA and mB are determined by
using the calculated N (n, m) for the two sites from the non-spin-polarized
band structure. mA and mB are approximately 3.5 µB .
conduction bands with the eg state just above EF . For the ↓-spin states,
the role of the A and B sites is reversed. Furthermore, the t2g states at the
B site form the bottom of the conduction band. Since the calculations were
intended to study charge ordering, the B sites were divided into B1 and B2
sites for Fe3+ and Fe2+ , respectively. At the B1 site, nB1 = n0 + δn and
B2 has nB2 = n0 − δn. The LSDA+U DOS shows that the t2g states of
the B2 site in the ↓-spin channel experience a d-d interaction so that they
split into bonding and antibonding states. All the other d-states including
states at B1 site are unoccupied. Furthermore, the t2g states of Fe at the
B1 site have energy below the antibonding states.
Fig. 4.35 (a) PDOS with LSDA, and (b) PDOS with LSDA + U (Anisimov et al.,
1996, 1997). A, tetrahedral coordinated Fe ions; B, octahedral Fe ions (B1 corresponds
to Fe3+ and B2 to Fe2+ ions. The Fermi level (EF ) is set at 0.
• With the refined low temperature structure of Wright et al. (2001), the
calculated results show not only the charge ordering consistent with
the results of Wright et al. (2001) but also the orbital ordering.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 4.36 Schematic energy level dia- Fig. 4.37 Orbital ordering in B-size
gram for the Fe ↓-spin d-orbitals un- sublattice, dyz and dzx are the orbitals
der distortion at B sites in Fe3 O4 (Jeng associated with the B1 atoms and dxy
et al., 2004). orbitals are with the B4 atoms (Jeng
et al., 2004).
4.4.1 Structure
The crystal structure of perovskite is shown in Fig. 4.38. The Mn atom is
surrounded by six O atoms forming an octahedron. This octahedraon is the
backbone of the ferromagnetic oxides. LSMO exhibits a structural phase
transition at TS = 280 K with zero external magnetic field. TS is a func-
tion of the external magnetic field and decreases to 220 K at magnetic filed
70 kG. At low temperature, LSMO has the orthorhombic structure with
space group Pnma. The high temperature phase has the rhombohedral
structure and belongs to the R3c space group. All these lattice parameters
and angles depend on the value of x. For example at x = 0.17, the lattice
constants are a = 5.547 Å, b = 7.790 Å, and c = 5.502 Å. The structure
consists of six formula units—six perovskite structures. Three of the oc-
tahedra stack up vertically and two of them align in the a-b plane. For
the rhombohedral phase, the lattice constant is 5.475 Å and the angle α
is 60.997◦. Paiva-Santos et al. (2002) refined earlier crystal parameters of
the La0.65 Sr0.35 MnO3 rhombohedral structure determined by Alonso et al.
(1997) using the Rietveld method (Rietveld, 1969). The method is now im-
plemented in a software package (Young et al., 1995). The lattice constant
a is 5.5032 Å and c is 13.3675 Å. The value of a agrees reasonably well with
the one obtained by Asamitsu et al. (1996). The interatomic distances of
La or Sr to the O and Mn atoms are summarized in Table 4.8.
Fig. 4.38 The unit cell of the perovskite structure. The Mn atoms (open circle at the
center) occupy the A sites and the La atoms (black circles at corners of the cube) occupy
the B sites. Oxygen atoms are shown as gray circles at face centers of the cube.
In fact, in La1−x Srx MnO3 the structural transition between the or-
thorhombic (O) and rhombohedral (R) phases depends on the temperature
and Sr concentration x. The phase diagram expressed in terms of temper-
ature and x is shown in Fig. 4.39, obtained from the neutron scattering
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
experiments (Kawano et al., 1996b). There are three sets of data points.
The circles and squares mark the onsets of antiferromagnetic and ferro-
magnetic to paramagnetic phase transitions, respectively. The filled and
open diamonds indicate the R → O transition temperature. The triangles
indicate the R → O structural phase transitions for x = 0.125.
The neutron scattering results show also that there are two O phases,
O′ and O∗ . The difference between these two phases was examined based
on the powder pattern using Rietvelt analysis. The O′ phase has the MnO6
octahedra√distorted due to the Jahn-Teller effect and its lattice constants
∗
√ 2 < c < a. The O phase exhibits the pseudocubic structure
satisfy b/
with b/ 2 ∼ c ∼ a and without any evidence of the Jahn-Teller effect.
Fig. 4.39 Magnetic (a) and structural (b) phase diagrams of La1−x Srx MnO3 with x ≤
0.17. CAF is canted antiferromagnetic, FM is ferromagnetic, and PM is paramagnetic.
The circles and squares indicate the onset of FM and AF components, respectively.
The diamonds indicate the R → O transition temperature. The triangles denote the
structural phase transitions for x = 0.125 (Kawano et al., 1996b).
2002). X-ray diffraction was used to characterize the crystals and probe the
possible refinement. Polycrystalline samples of LCMO were grown in two
ways. One was to first grow powered forms using similar mixtures as de-
scribed above. The mixture was calcined three times in the air at 1050 ◦ C
for 24 hrs with intermittent grinding. There is a slight difference in detailed
settings for the temperature and time (Asamitsu et al., 1996; Kawano et al.,
1996a). The resulted powders were pressed into rods under a hydrostatic
pressure of 1.4 ton/cm2 then fired at 1100 ◦ C in air for 38 hrs. Both groups
used X-ray powder diffraction to characterize the samples. Another way to
grow LCMO is to grow single crystals using the floating zone method. The
initial samples were prepared in rod form and were loaded into a flow zone
furnace. The feeding speed is 5–10 mm/hr. The single crystal samples were
then powered. To form a tunneling junction, the thin films of LCMO were
grown by PLD on LaAlO3 (100) substrates (Wei et al., 1997). The growth
temperature was 700 ◦ C and in an environment of 13.3 mbar of oxygen.
The samples were then annealed at 900 ◦ C in 1 atm oxygen for several
hours. XPS and STM were used to check the quality of the films. The XPS
results showed very low contamination indicating the surfaces of the films
were clean. The STM was operated at 1 nA constant-current mode with Pt
tip which was biased at about –2.0 eV. The images were taken at RT and
77 K in either high vacuum or ultra-pure He gas. They show atomically
smooth “rice-paddy” type of terraces with the step height of one unit cell.
curve) and 300 K (open circles) are shown in Fig. 4.40 for cases below and
above TC . The normalized conductances in Fig. 4.40 are proportional to
the DOS. In practice, it normalizes the inherent dependence of the STM
transmission probability on the voltage.
Fig. 4.40 Normalized conductance as a function of bias voltage of LCMO with x=1/3,
at T = 77 K (solid-dot curve) and T = 300 K (open circles) (Wei et al., 1997).
The important features are the two peaks at ±1.75 eV appear in the low
temperature results. These two peaks agree with most of the DOS derived
from the spin-polarized band structures of LSMO. The DOS results exhibit
the half metallic properties for LSMO (Singh and Pickett, 1998; Ma et al.,
2006). The –1.75 eV peak corresponds to the peak in the occupied states of
the majority-spin channel and the +1.75 eV peak corresponds to the peak
in the unoccupied minority-spin states above EF . The difference is the
strength of the exchange splitting. The half-metallic feature is explicitly
shown by the presence of the spin splitting peaks in the low temperature
results. At 300 K, there are no such peaks consistent with the paramagnetic
properties.
Fig. 4.41 (a) M(T)/M(T = 0 K) as a function of temperature and (b) a hysteresis loop
at T = 300 K for LCMO with x = 0.3 (Park et al., 1998b).
Fig. 4.44 A 20-atom model of LSMO (Ma et al., 2006). Sr is shown as a large gray
circle, filled black circles are La, small gray circles are Mn, and open circles are O.
In treating the LSMO alloy within the DFT calculations , Singh and
Pickett (1998) used virtual crystal approximation for La2/3 Ba1/3 MnO3 .
The size of the unit cell is that of LCMO with x = 1/3 exhibiting colossal
MR (CMR). Relaxations were performed in ferromagnetic and antiferro-
magnetic configurations. Their results are summarized as follows:
4.5.1 Superexchange
4.5.1.1 Model approach
First proposed by Kramers (1934), it was Anderson (1950) who simplified
Kramers’ general formalism and provided a simple picture. The effective
interaction based on Kramers’ formulation between two Mn atoms with a
common neighboring O atom is called superexchange. The essential idea of
the interaction is shown in Fig. 4.45. The O2− ion is located between two
Mn2+ ions. There are no free carriers in Anderson’s view of superexchange.
Fig. 4.45 Superexchange interaction. The dashed arrows indicate that the electrons
can hop.
The O2− ion has a filled p-shell so that it has a net charge of –2e. The
two electrons occupy one of the p-orbitals with spins pointing in opposite
directions. The Mn2+ ions have five d-electrons and align their local spins
according to Hund’s first rule. O p-electrons can hop to the neighboring
unoccupied Mn d-states while conserving spin. This delocalization reduces
the energy of the O2− ion. Because the spins of p-electrons are oppositely
oriented, this hopping can only occur if the neighboring d electrons are
antiferromagnetically aligned. The effective interaction was reformulated
in terms of spin operators and the orthogonality of the wave functions.
In reality, the wave functions at the neighboring atoms are not orthogo-
nal. This led Yamashita and Kondo (1958) to reconsider the superexchange
interaction. These authors pointed out that if the ionic model shown in
Fig. 4.45 is considered, then it is necessary to expand the total energy
to 4-th order in the overlap integral S in order to have the total energy
dependent on the spin arrangements. They argued that the ionic states
should include some mixing of electronic configurations for excited states
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
• The Slater Model (Slater, 1953). This model consists of [Mn2+ ; O2−
(2p)6 ](the ionic model) + [Mn2+ ; O2− (2p↑)2 (3s↑)(2p↓)3 ] +[Mn2+ ;
O2− (2p↑)3 (3s↓) (2p↓)2 ].
• The Kramer-Anderson model (Anderson, 1950). It has [Mn2+ ; O2− ] +
[Mn+ ;O− ].
• The Goodenough model (Goodenough, 1955). Two configurations are
[Mn2+ ; O2− ] + [Mn+ —O—Mn+ ].
• The band model (Yamashita and Kondo, 1958). The model also has
two configurations, [Mn2+ ; O2− ] + [Mn3+ —O2− —Mn+ ].
Fig. 4.46 Double exchange interaction. S1 and S2 are the spin moments of the Mn ions
contributed by the localized d-electrons. s is the spin of the itinerant electron.
are the intra-atomic exchange integral J between mobile electron and local
d-orbitals, and the magnitude of the transfer matrix element b characteriz-
ing the effective coupling between case 1 and case 2. The authors showed
that by coupling states from two Mn ions with b, the splitting of the Zener
levels is proportional to cos( 2θ ), where θ is the angle between classical spins
S1 and S2 at the two Mn ion sites and the effective transfer integral is
Θ (S0 + 21 )
tef f = b cos( )=b , (4.18)
2 2s + 1
where S0 is the sum of the spin moments of the two magnetic ions, and
s is the spin of the mobile electron. With b > J, they showed that the
ferromagnetic phase can occur through the mobile electron. It is possible
to conclude that double exchange favoring ferromagnetism should involve:
Consequently, the motion of the mobile electron between the two sites
is correlated. Most of the investigations of the magnetic properties of the
metal oxides have been based on either superexchange or double exchange
or both, depending on the particular oxide of interest.
4.5.3.1 CrO2
For CrO2 , there is general consensus that it is the double exchange mech-
anism that is responsible for its magnetic properties. Schlottmann (2003)
combined single-particle results with collective excitations to show that
double exchange in CrO2 depends critically on the distortion of the oc-
tahedron surrounding the Cr atom and therefore differs from the double
exchange mechanism in the manganites.
The physical picture is as follows: The Cr ion in the octahedron formed
by the O atoms has valence 4+. Only two d-electrons remain in valence.
Due to the cubic environment, the five d-states split into t2g (triply de-
generate) and eg (doubly degenerate) states, with the t2g states at lower
energy. Therefore, the t2g states are partially occupied and eg states are
unfilled. The t2g states are composed of dxy , dyz , and dzx orbitals. The
octahedron undergoes a Jahn-Teller distortion. This tetrahedral distortion
lifts the degeneracy of the t2g states with the dxy state having the lowest
energy. The other two orbitals form √12 (dyz +dzx ) and √12 (dyz –dzx ) states.
One of the two electrons occupies the dxy state and the other electron has
a 50% probability of occupying each of the combined orbitals. The level
scheme is shown in Fig. 4.47.
Fig. 4.47 Level diagram of the Cr4+ ion under cubic crystal field and tetragonal dis-
tortion. One of the two electrons occupying the dxy state is localized while the other
electron sharing √1 (dyz ± dzx ) states is itinerant.
2
Letting the hopping integral between the two sites, 1 and 2, be t, the
hopping Hamiltonian is
X †
Ht = −t c1ξσ c2ξσ + c†2ξσ c1ξσ + c†1ζσ c2ζσ + c†2ζσ c1ζσ , (4.19)
σ
where ζ and ξ label orbital angular momentum states and σ labels the spin.
The other term required is the on-site Coulomb energy U to model the
full Hamiltonian. To apply the model Hamiltonian to CrO2 , we note that
there are two Cr ion sites, with two electrons each. One of them occupies
the dxy localized state and the other is itinerant. From Hund’s first rule
at each site, the orbital wave function is odd under permutation of the
two electrons and the spin state is triplet. Combining the two sites, the
resultant spin momenta, S = 2, 1, 0, can be constructed and serve as the
states acted on by Ht . The S = 2 case consists of three states having an
orbital triplet and one state having an orbital singlet. The triplet is even
under permutation of two mobile electrons while the singlet is odd.
ζζ
ψeven (S = 2) = c†1ζ↑ c†2ζ↑ |1 ↑ 2 ↑i
(4.20)
ζξ
ψeven (S = 2) = √12 (c†1ζ↑ c†2ξ↑ + c†1ξ↑ c†2ζ↑ ) |1 ↑ 2 ↑i
ξξ
ψodd (S = 2) = c†1ξ↑ c†2ξ↑ |1 ↑ 2 ↑i
ζξ (4.21)
ψodd (S = 2) = √12 (c†1ζ↑ c†2ξ↑ − c†1ξ↑ c†2ζ↑ ) |1 ↑ 2 ↑i ,
4.5.3.2 Fe3 O4
The Fe ions in Fe3 O4 have two different valences, 2+ at the B2 and B3
sites and 3+ at the B1 and B4 sites. In addition, the Verwey transition
complicates investigations of the origin of the ferromagnetism in this com-
pound. At present, the origins of the ferromagnetism in Fe3 O4 and the
La oxide alloys are still somewhat controversial. In this section, we dis-
cuss a treatment considering the double exchange mechanism but taking
the antiferromagnetic ordering between A and B sites into account. The
approach differs from that of the preceding section, considering instead the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
where i and j are site indices, S is the total spin of the localized states, σss′
is the Pauli matrix, and the c’s are the creation and annihilation operators
for the itinerant electron. Recall that in the spinel structure of Fe3 O4 shown
in Fig. 4.16, there are tetrahedral (A) sites and octahedral (B) sites. The B
sites contribute to the local spins and itinerant electrons (Loos and Novák,
2002). The mean field approach was adopted to find an effective field acting
on an atom at a B site due to its neighbors.
Above the Verwey temperature, there are NB sites with each B site
having a local spin moment S of (5/2)~ and NB /2 itinerant electrons. Let
S ′ be the effective local spin moment resulting from the spin moment of
the itinerant electron coupled to the local spin through Hund’s first rule.
Grave et al. (1993) obtained a value of 2.9 for Fe3 O4 through their hy-
perfine measurements based on the double exchange formulation of Kubo
and Ohata (1972). This differs quite significantly from the ideal estimation
of 2.0. Loos and Novák (2002) addressed this discrepancy by including su-
perexchange between site A and site B. This is done by using effective fields
which are proportional to the z-components of the equilibrium magnetiza-
tions, mA,eq =< SA,z >eq and mB,eq =< SB,z >eq , acting respectively on
site B and site A ions. mA , therefore, can be expressed as
mA = SBS (λAB mB,eq + λAA mA ) , (4.23)
where BS is the Brillouin function. The λ’s characterize the superexchange
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
interactions, with
λij = βB SJij zj (i), (4.24)
where Jij is the strength of the superexchange interaction between sites i
and j, zj (i) is coordination number of the i-th atom with j nearest neigh-
bors, and βB is kB1T . Let λ, the mean field at T, be βSgµB Hef
z
f acting at
a B site; the magnetization at the same site is
z
mB (λ) = hSB (λ)i = (1 − x)SBS (λ) + xS ′ BS ′ (S ′ λ/S), (4.25)
where g is the Landé g-factor and x is the fraction of itinerant electrons
relative to the total at site B. S is the total local spin and S ′ is the resultant
spin at site B including the local spin S and the spin of the itinerant electron
coupled by Hund’s rule (Loos and Novák, 2002). The effective Hamiltonian
due to the superexchange and the free energy per B site fB can then be
given:
m2
1
hHSE i = − mB λBA mA,eq + λBB B (4.26)
βB S 2
and entropy,
(S)
X
λm/S
X λmB λm/S
S = kB (1 − x) ln( e ) + x ln(e ) − .
|m|≤ 52 |m|≤2 S
(4.30)
mB can then be determined by minimizing fB with respect to mB . The
theory has some shortcomings. If S ′ is set equal to 2, then TC is 1590
K, almost a factor of 2 larger than experiment. Loos and Novák (2002)
attributed this shortcoming to the mean field theory.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Chapter 5
5.1 Introduction
Both Heusler alloys and transition metal (TM) oxides are appealing for
spintronic applications. However, due to the possible large Coulomb corre-
lation at the TM sites, defects, phase transitions (Borca et al., 2001), and
surface and interface effects in thin-film samples discussed in Chapter 3, at
present, neither Heusler alloys nor TM oxides have been shown experimen-
tally to exhibit half metallic properties at room temperature (RT). Thus at
the beginning of this century, researchers began to search for half metals
(HM) with simpler structures, such as the zincblende (ZB) structure. Ex-
citing results were obtained by Akinaga et al. (2000a). They first predicted
CrAs in the ZB structure to be a HM by first-principles calculations then
grew it in thin-film form. In this new HM, the unit cell consists of just two
atoms, as in GaAs. It is especially appealing in light of the following:
191
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Etgens et al., 2004) and predict (Galanakis, 2002a; Pask et al., 2003; Xie
et al., 2003; Xu et al., 2002; Liu, 2003; Galanakis and Mavropoulos, 2003;
Continenza et al., 2001; Zhao et al., 2002; Şaşoǧlu et al., 2005a) new HMs
in the ZB structure. Since the growth of these compounds is the major
hurdle, due to the fact that the ZB structure is not the ground state struc-
ture (e.g., many have the hexagonal NiAs structure as the ground state,
in which they are not half metallic), we give the details of the growth
techniques adopted by various researchers to provide a foundation for fu-
ture developments. The interactions giving rise to the half metallicity in
these compounds will be discussed. We shall then describe the efforts to
grow (Mizuguchi et al., 2002; Akinaga and Mizuguchi, 2004) and design
quantum structures, including superlattices (Fong et al., 2004; Fong and
Qian, 2004), quantum dots (Qian et al., 2004a; Shirai, 2004), and digital
ferromagnetic heterostructures (DFH) (Sanvito and Hill, 2001; Qian et al.,
2006a; Wu et al., 2007; Zhu et al., 2008), and discuss their physical prop-
erties. Half metallic Si-based DFH’s will be elaborated in particular, due
to the wide availability of mature Si-related technologies. Finally, we shall
discuss attempts to design even simpler structures. In this respect, carbon
nanowires doped with TM elements are a promising direction (Dag et al.,
2005; Durgun et al., 2006).
5.2.1 Experiment
We shall focus on the experimental efforts on the growth, characterization,
and magnetic properties of HMs with ZB structure.
5.2.1.1 Growth
There are two forms for the growth: thin film and multilayers.
Thin films of CrAs In 2000, Akinaga et al. (2000a) used MBE to grow
CrAs on a GaAs(100) substrate. A molybdenum holder with indium solder
was used to hold the substrate. A 20 nm GaAs buffer layer was first grown
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
after the chamber was degassed. The CrAs was then grown on top at 0.02
nm/s. The temperature of the growth was 200 ◦ C. Finally, a 4.0 nm layer
of gold was used to cap the CrAs for preventing oxidations. The resulting
thickness of the CrAs layer with the ZB structure is 2 nm.
The results reported by Akinaga et al. (2000a) were not without con-
troversy. Etgens et al. (2004) investigated the structure of CrAs epilayers
grown on GaAs(001) and also used the MBE method to grow their sam-
ples. A 100 nm undoped GaAs buffer layer was first grown on GaAs(001)
substrate. The surface was annealed under As to improve the flatness of
the surface. CrAs was then grown at 200 ◦ C with a rate of 8 Å/min under
As rich conditions. At 295 K, CrAs has the orthorhombic MnP structure
with lattice constants, a = 5.637 Å, b = 3.445 Å, and c = 6.197 Å. Above
1100 K, it is grown in the hexagonal NiAs structure.
Thin films of CrSb Zhao et al. (2001) reported the growth of CrSb on
GaSb. They also used the MBE method to grow CrSb films on a GaAs(001)
substrate with three different buffer layers, GaAs, AlGaSb, and GaSb, re-
spectively. The layer configurations are shown in Fig. 5.1. The solid sources
of Ga, As, Sb, Al and Cr were provided. Before depositing the buffer layer,
the oxide layer on top of the substrate was cleaned at 580 ◦ C. The buffer
layers have different thickness as shown in Fig. 5.1. For example, case A has
500 nm thick GaAs and was grown at 560 ◦ C. During the growth on this
buffer layer, the As and Ga fluxes were terminated simultaneously at the
end of the growth. The Sb fluxes were immediately released to prevent the
sublimation of Group V elements from the surface. At this time, the tem-
perature of the substrate was lowered to grow CrSb. When the substrate
temperature reached 400 ◦ C, the flux of Sb atoms was stopped to avoid the
build-up of Sb on the surface. As the temperature of the substrate reached
250 ◦ C, both Cr and Sb fluxes were then turned on. Typically, the intensity
of the Cr flux was 2×1013cm−2 s−1 . The beam equivalent pressure ratio of
Sb/Cr was maintained at about 10. By monitoring RHEED patterns that
show ZB features in the sample, these authors specified the thickness in
terms of depositing time. Typically, the growth time was 60 s. The results
are summarized in Table 5.1.
Fig. 5.1 Layer configurations of growing CrSb. The time (seconds) of deposition for
the GaSb layers is also shown (Zhao et al., 2001).
Table 5.1 Summary of information for growing CrSb with the ZB structure on three different buffer
layers.
Sample Buffer Growth time Duration of Substrate temp. Cr flux
layer (s) ZB RHEED pattern (s) (◦ C) (1013 cm−2 s−1 )
A GaAs 60 40 250 2
B AlGaSb 64 60 250 3
C GaSb 64 60 250 2
Fig. 5.2 Schematic diagram showing the layer configurations, x, y, and z of CrAs/GaAs
multilayers. LT means low temperature (220 ◦ C) with respect to the buffer layer tem-
perature (580 ◦ C) (Mizuguchi et al., 2002).
Fig. 5.3 The RHEED patterns of CrAs films with the electron beam incident along
the [110] (a and c) and [100] (b and d) directions of a GaAs substrate (Akinaga and
Mizuguchi, 2004).
Fig. 5.4 The GIXD patterns for the three samples with different thicknesses (Etgens
et al., 2004). The axes are in reciprocal lattice units.
The b-axis of CrAs bCrAs is in the growth direction. From the spots, the
length of a differs from |aCrAs | by 0.23%. On the other hand, the one in
the direction of c has 9.5% mismatch. These basis vectors correspond to
the fully relaxed epilayer. Comparing the different GIXD patterns of the
thin to thick epilayers, the single phase can be due to a metastable phase
formed at the beginning of the growth. Consequently, the growth results
can depend on the film thickness. Since characteristics of the ZB structure
were not observed in these experiments, Etgens et al. (2004) attributed the
failure of finding the ZB structure to the lattice mismatching. However,
there is the possibility that this growth did not control the orientation of
CrAs and GaAs, so that CrAs is not in the ZB structure.
RHEED and HRTEM for CrSb epilayers The RHEED and high-
resolution cross sectional transmission electron microscopy (HRTEM) were
used to characterize the substrates and film forms of CrSb (Zhao et al.,
2001). The HRTEM spectra exhibit features of a film in real space.
The RHEED picture shows a streaky (1×3) pattern during the CrSb
growth on a Sb-terminated GaSb surface at the substrate temperature
ranging from 250 ◦ C to 400 ◦ C. This initial pattern became weaker and
disappeared after 40 seconds. For the growth on other substrates, such as
Al0.84 Ga0.16 Sb and GaAs, the streaky characteristic of the ZB GaAs buffer
layer disappeared after depositing CrSb for 60 seconds.
The results of HRTEM for a CrSb layer grown on GaAs for 30 s at the
substrate temperature of 250 ◦ C and with the Cr flux of 2×1013 cm−2 s−1
were obtained by Zhao et al. (2001). The film sample was capped by GaAs.
Both the limited and wider areas show 1 ML of CrSb without any defects,
such as dislocations, at the interfaces. The results provide an evidence of
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 5.5 XAS results of multilayer CrAs/GaAs samples of A, E, and F listed in Table 5.2.
Unfilled and filled ▽’s denote peak positions (Mizuguchi et al., 2002).
CrAs epilayers Akinaga et al. (2000a) measured the hysteresis loop for
an epitaxial film of CrAs grown on GaAs(001) at RT. The magnetic field
was applied parallel to the film plane. The saturation magnetization is 560
emu/cm3 . This is equivalent to 3.0 µB per formula-unit and agrees with
theoretical predictions (Akinaga et al., 2000a; Galanakis, 2002a; Pask et al.,
2003). For different samples under identical growth conditions, there is ap-
proximately 10% spread in values of the saturation magnetization. The
coercivity is small. Based on these RT measurements, the Curie tempera-
ture (TC ) is estimated to be greater than 400 K.
Fig. 5.6 The hysteresis loop of an epi- Fig. 5.7 Magnetization as a function
taxial CrAs grown on GaAs. The mea- of thickness of a CrAs film. The inset
surement was done at 300 K (Akinaga shows the hysteresis loops for the two
et al., 2000a). CrAs films with different thickness (Et-
gens et al., 2004).
did not discuss the methods employed. The main result is the saturation
magnetization as a function of thickness of the films shown in Fig. 5.7.
When the layers are ∼ 4.0 nm, the saturation magnetization of different
samples with the same thickness exhibits more than a factor of 3 fluctua-
tion with significant uncertainties. At 20 Å (2 nm), the saturation value is
1000 emu/cm3 , which gives ∼ 3µB per Cr atom for an orthorhombic CrAs,
assuming that only Cr atoms contribute to the magnetic moment. This
number agrees with that obtained by Akinaga et al. (2000a). As the layer
thickness increases, the magnetization reaches a value below 250 emu/cm3 .
The hysteresis loops for the two films with different thicknesses are shown
in the inset. The coercive field and remanent magnetization are also com-
parable to those given by Akinaga et al. (2000a).
CrSb epilayers Zhao et al. (2001) used SQUID to measure the hysteresis
loops of CrSb thin-film samples grown on three different buffer layers. They
also reported the remanent magnetization as a function of temperature.
The hysteresis loops at 300 K for the CrSb thin-film samples are shown
in Fig. 5.8. The contribution from the diamagnetic GaAs was subtracted.
Even the layer thickness was only 1 ML, the films exhibit ferromagnetic
properties at RT. The saturation magnetizations at 5 K are between 3
and 5 µB /formula-unit. The value of 3.0 µB /formula-unit agrees with the
theoretical calculations by Pask et al. (2003).
Fig. 5.8 Hysteresis loops of CrSb thin- Fig. 5.9 Temperature dependence of
film samples grown on three different the remanent magnetization of the three
buffer layers–GaAs (triangles), AlGaSb CrSb thin-film samples (Zhao et al.,
(circles), and GaSb (squares) (Zhao 2001).
et al., 2001).
grown on GaAs and GaSb are very close while the one grown on AlGaSb
alloy is higher. At 400 K, all the samples still exhibit finite remanent
magnetization indicating that the TC is higher than 400 K.
Fig. 5.10 The hysteresis loop of a multilayer structure: [(CrAs)2 /(GaAs)2 ]100 (Akinaga
and Mizuguchi, 2004).
5.2.2 Theory
Since the pioneering work of Akinaga and Mizuguchi (2004) on CrAs, much
theoretical work has been devoted to the prediction of new HMs with ZB
structure. Among them, Fong et al. (2004) and Fong and Qian (2004)
designed superlattices with HM properties. Later, Qian et al. (2004a) re-
ported an integer magnetic moment in a quantum dot composed of MnAs
(Qian et al., 2006b). Dag et al. (2005) and Durgun et al. (2006) designed
quantum wires showing half metallic properties. Sanvito and Hill (2001)
examined the half metallic properties in δ-layer-doped Mn in GaAs. Such
structures are now known as digital ferromagnetic heterostructure (DFH).
In this section, we shall review these efforts by starting with compounds.
We shall discuss the basic interactions causing the half metallicity in these
compounds. Then we shall discuss the associated quantum structures.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Despite the fact that the ZB structure is not the energetic ground state
structure for the TM compounds, many half metallic pnictides, a carbide,
and several chalcogenides have been examined theoretically. Table 5.3 lists
some of these theoretical efforts, including the bulk lattice constants, lattice
constants at which the compound exhibits half metallicity, and correspond-
ing magnetic moments.
The agreement among the theories is very good for all properties except
the magnetic moment; some show non-integer values. The discrepancies
are mainly caused by different methods. There was some doubt whether
the pseudopotential method would be suitable to treat these compounds.
As illustrated in Table 5.3, however the results obtained by the plane-
wave pseudopotential (PWPP) method agree very well with those of the
all-electron linearized augmented plane wave (LAPW) method.
Table 5.4 Compound, range of lattice constants a within ∼4% of the corre-
sponding equilibrium values over which half metallicity is predicted and the
suggested substrates for growth. Substrate lattice constants (Wyckoff, 1963)
are given in parentheses.
Compound Lattice constants (Å)
LAPW PWPP Substrate
CrP > 5.48 > 5.47 AlP (5.53)
CrAs > 5.51 > 5.50 GaAs (5.64), AlAs (5.62)
CrSb > 5.87 AlSb (6.13), GaSb (6.12), InAs (6.04)
MnAs > 5.75 > 5.77 InP (5.87)
MnSb > 6.06 > 5.96 AlSb (6.13), GaSb (6.12), InSb (6.48)
MnC > 4.26 > 4.20 Diamond (3.57), cubic SiC (4.35)
The energetic ground state structures of these compunds have also been
considered. Pask et al. (2003) compared NiAs and ZB structures using both
pseudopotential and LAPW methods with GGA exchange-correlation. The
NiAs structure is confirmed to be the ground state for the pnictides. Re-
alizing that the ZB structure is a metastable structure and thin-film forms
can be grown, a list of possible substrates (Table 5.4) for each compound
was suggested Galanakis and Mavropoulos (2003) also suggested GaAs as
a substrate to grow CrAs.
In the following, we focus first on a typical pnictide, MnAs, where
the underlying interactions associated with the half metallicity will be dis-
cussed. We then discuss a TM carbide, MnC, which shows some interesting
and opposite properties from the pnictides and chalcogenides.
5.2.2.1 MnAs
Several groups (Pask et al., 2003; Continenza et al., 2001; Şaşoǧlu et al.,
2005a) predicted MnAs in the ZB structure to exhibit half metallic prop-
erties. The ZB structure of MnAs is shown in Fig. 1.9, where the filled
grey circles are As atoms, and they are located at the corners and centers
of the cubic faces. Open circles are the Mn atoms. It is important to note
that the Mn and As atoms can form chains, for example, along the [110]
direction.
With the structure in mind, we now discuss the half metallic and elec-
tronic properties:
The upper panel shows the DOS of the majority- (↑) spin channel and the
lower panel is for the minority (↓) spin states. EF is set to be zero on the
energy scale. The features of the DOS are:
• At EF , a finite DOS, 0.77 states/eV-cell, appears in the ↑ channel.
• For the ↓ spin states, EF falls within a gap of 1.70 eV.
• P value at EF is 100%.
The position of EF in the insulating channel is lattice constant depen-
dent. When the lattice constant increases, EF drops toward the top of the
valence band. As a result, the half metallicity disappears when the lattice
constant reaches a critical value.
called the non-bonding d-states of the Mn atom. The lobes of the charge
distribution of these two states point toward the second neighbors of the
Mn atoms. In this cubic environment, it is possible to identify that they
are originated from the dz2 and dx2 −y2 states. In the ↑ spin channel, these
doubly degenerate states are occupied, while those in the ↓ spin channel
lie above EF . These bands separate from the lower energy Γ15 (t2g ) bands
by a gap. Therefore, the minority channel exhibits insulating behavior. In
the majority-spin (↑) states, we label another Γ15 (t∗2g ) above Γ12 . These
are antibonding states resulted from d-p hybridization. The broad fea-
ture of these bands make them partially occupied along the Γ–X direction
contributing to the metallic behavior of the majority-spin states. These
features show the half metallic properties from the band structure point of
view.
Fig. 5.12 Band structure of MnAs along symmetry directions of the BZ (Pask et al.,
2003). The BZ of the fcc lattice is shown on the right.
Fig. 5.13 The charge densities of the t2g (left) and eg (right) states for the majority-
spin states. The horizontal axis is in the [110] direction and the vertical axis is the z-axis
of a cube (Pask et al., 2003).
5.2.2.2 MnC
Half metallic phases Among all the compounds studied by Pask et al.
(2003), MnC shows interesting half metallic properties. The majority- and
minority-spin channels reverse their roles for the metallic and insulating
behaviors as compared to the pnictides and chalcogenides. The calculated
magnetic moment is only 1.0 µB /unit-cell which will be explained in sec-
tion 5.2.2.3. Qian et al. (2004b) found that MnC in fact has two different
ranges of lattice constants for which the half metallic properties exist. The
lattice constant dependent total energy and magnetic moment are shown
in Fig. 5.14. From the total energy curve (the lower panel), the optimized
lattice constant is found to be 4.39 Å. The interesting behavior of the mag-
netic moment/unit-cell is also shown. For lattice constant greater than 5.0
Å, the magnetic moment is 3.0 µB /unit-cell and can be accounted for by
the ionic model. For lattice constant between 3.28 and 4.23 Å, there is
another region with integer magnetic moment. The value of the magnetic
moment is 1.0 µB /unit-cell.
Fig. 5.14 Magnetic moment/unit-cell and total energy as a function of lattice constant
in MnC (Qian et al., 2004b).
states indicated by the energy levels at Γ(Fig. 5.15(b)) are the same as in
MnAs.
Fig. 5.15 (a) The band structure of MnC at a = 4.20 Å, (b) the energy levels at the
Γ-point as a function of lattice constant (Qian et al., 2004b).
BP, BN, and SiC have been suggested (Fong and Qian, 2004; Fong et al.,
2008). Based on the results predicted by Fong and Qian (2004), only a
monolayer MnC could be grown on a SiC(100) substrate.
Crystal field As shown in Fig. 1.9, each of the two atoms in a unit cell is
surrounded by four neighbors forming a tetrahedron. Neighboring ions
of the TM atom exert Coulomb fields on the d-states of the atom. The
effect is to split the five-fold degeneracy of the d-states into triply de-
generate t2g and doubly degenerate eg states. Therefore, a gap between
the t2g and eg states is formed.
d-p hybridization The triply degenerate t2g states of the TM element are
composed of dxy , dyz , and dzx states. Their linear combinations form
orbitals pointing toward the nearest neighbors of the metal element.
The sp3 orbitals of the non-metal atoms point their lobes toward their
nn. The orientations of a d-orbital and sp3 orbital are shown schemat-
ically in Fig. 1.9. The d- and p-states overlap and hybridize. This
hybridization forms the bonding and antibonding gap characterized by
the two Γ15 states at k = 0 in the band structure. These two interac-
tions can be summarized in terms of energy levels as shown in Fig. 1.10.
The Coulomb fields due to the neighboring TM elements cause the s-
and the p-states of the As atom to form sp3 directional orbitals. Under
this circumstance, the p-states remain degenerate.
Exchange interaction Up to this point, the bands in a compound are spin
degenerate. Now, the exchange interaction lifts the degeneracy of the
↑ and ↓ states and shifts the ↓ state energies upwards.
and the lowest energy Γ15 states can be used to specify the strength of the
crystal field. The difference of energies for the bonding and antibonding
Γ15 states can be used to estimate the magnitude of the hybridization. A
more refined approach is to determine first the average energies for the
states derived from the two Γ15 states and then take the difference. The
energy difference of the lowest Γ15 states for the majority- and minority-
spin channels is related to the strength of the exchange interaction.
To understand the smaller lattice constant of MnC in terms of the
strength of the three interactions, we compare the band structures shown in
Fig. 5.12 and Fig. 5.15(a). The major differences shown in Fig. 5.15(a) are:
(i) EF intersects the minority-spin states, (ii) the gap is formed within the
majority-spin states. The lattice dependent energy levels at the Γ-point
are shown in Fig. 5.15(b). For the majority-spin channel, the bonding-
nonbonding and bonding-antibonding gaps increase as the lattice constant
decreases. These results indicate that the strength of the crystal field and
p-d hybridization increase under compression. The minority-spin states
show similar behavior, except that the nonbonding (Γ12 ) state is lowered
with compression. This may be due to the fact that the Γ12 states are
not occupied. The increase of dispersion of the bonding and antibonding
states with respect to the case of larger lattice constant cause EF to shift.
The strength of the exchange interaction, gauged by the differences of the
lower Γ15 states of both ↑ and ↓ states, is lessened as the lattice constant
decreases.
Hund’s first rule Since d-states can accommodate five electrons in each
spin channel, the four electrons can align their spins as required by Hund’s
first rule. This is essentially the manifestation of Pauli’s exclusion principle
and the minimization of the Coulomb interaction considering the position
and spin as two dynamical variables to specify the states of the electrons.
Since the four d-electrons have the same spin states, they cannot occupy the
same spatial position. They are kept apart from one another. Consequently,
the effect is to reduce the Coulomb repulsion. The compound can therefore
have lower energy. The total spin moment of the compound is 2~. With
the g-factor of 2 for each electron, the resultant magnetic moment is 4.0
µB /unit-cell. The contribution from the orbital motion to the magnetic
moment is negligible due to electrons occupying 3d states.
It should be noted that for a HM the magnetic moment/unit-cell should
be an integer. The reasons are as follows:
• The total number of electrons is an integer in a compound.
• One of the spin channels, say the ↓ channel, is insulating. Therefore, it
is necessary to have an integer number of electrons, N↓ , to fill up the
top of the valence band.
• The number of electrons in the conducting channel, N↑ , must then be
an integer.
The magnetic moment/unit-cell is:
M = N↑ − N↓ . (5.2)
Therefore, for any theoretical prediction of a compound to be a HM, it is
necessary to have an integer value of the magnetic moment/unit-cell.
The above explanation applies to MnC having a large lattice constant.
Can we understand why the spin moment changes from 3.0 to 1.0 µB /unit-
cell in MnC as its lattice constant decreases? Let us start with the case
having a larger lattice constant, 5.0 Å. From the ionic model, the three elec-
trons remaining at the Mn atom align their spins. Two of them occupy the
Γ12 -related states and one occupies the low energy tail of the antibonding
state (see Fig. 5.15 for reference). At this lattice constant, there is ample
space for the electrons to spatially avoid each other consistent with Pauli’s
principle. As the lattice constant decreases, the charges of the d-states
around the Mn atom are pushed closer to the C atom as manifested by
the increase of the bonding (lower Γ15 states) and antibonding (upper Γ15
states) gap and the crystal field effect (Fig. 5.15(b)). As the lattice constant
decreases further, we take the value of 4.20 Å as an example, the space for
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
the three d-electrons to align their spins becomes so limited that one of the
three majority-spin electrons flips its spin. This effect is demonstrated in
Fig. 5.15(b) by the lowering in energy of the Γ12 states in the minority-spin
channel to accommodate this spin-flipped electron. To be more explicit,
the magnetic charge (spin density) distribution, defined as the difference
between the ↑ spin and ↓ spin charge densities, of the Γ12 related states
in the minority-spin channel is plotted in Fig. 5.16. The section of the
charge distribution is the (110) plane including the zig-zag chain of Mn and
C atoms. The dot-dash contour denotes the zero magnetic charge density.
The solid and the dotted contours exhibit the positive and negative values
of the density.
A possible method of detecting volume-dependent half metallic phases
is to carry out the measurement of the saturation magnetic moment of the
sample. If the saturation magnetization is over 500 emu/cm3 , then the
sample has 3 µB /unit-cell. If the measured result is around 180 emu/cm3 ,
the corresponding saturation magnetization is 1 µB /unit-cell.
Fig. 5.16 Magnetic charge densities of MnC at 4.20 Å in the (110) plane including the
zig-zag chain of the Mn and C atoms.
Total and partial density of states The total and partial DOS (PDOS)
of the (CrAs)1 /(MnAs)1 superlattice are shown in Fig. 5.18. The As occu-
pied s-states form bands below 10 eV from EF which is set to be zero. For
the occupied d-states, the Mn d-manifold is located about 2 eV lower than
those of the Cr d-states. This is expected because Mn has valence 7 while
Cr has 6. The eg states of both atoms are higher than the center of gravity
of the t2g states. As shown in Fig. 1.10, under the crystal field the t2g states
have higher energy than the eg states. The t2g states in the superlattice
are the hybridized bonding states. Another feature shown in the PDOS is
the eg states of both metallic elements in the minority-spin channel located
above EF . The valence states of this channel are predominantly the As p-
and Mn t2g -hybridized states.
Fig. 5.18 Total and partial DOS of the majority- and minority-spin channels. TDOS
are in units of states/eV-unit-cell.
for the two spin channels. The contours reflect the features of the DOS
and PDOS. In the majority-spin channel, the bond charges are explicitly
indicated by contours located between the Mn and As atoms. For the
minority-spin states, contours around the two TM elements exhibit the
four-lobe characteristic of the d-states.
Fig. 5.19 Total valence charge densities of (CrAs)1 /(MnAs)1 of majority-spin states
(upper panel) and minority-spin channel (lower panel) in a section containing the zig-
zag chain of the two TM elements and As atom.
where Ztotal is the total number of valence electrons, and Nmajority and
Nminority are the number of the majority- and minority- spin states, re-
spectively. For (CrAs)1 /(MnAs)1 , Ztotal = 23 and Nminority = 8 give M =
7 µB . Similarly, Eq. (5.3) gives a moment of 14 µB for (CrAs)2 /(MnAs)2 .
These values are confirmed by numerical calculations. They are the sum of
M for each constituent compound making up the superlattice. It is impor-
tant to note that the increased magnetic moment/unit-cell for larger cells
implied by Eq. (5.3) does not necessary mean an increase of the saturation
magnetization. The latter is defined as the magnetic moment density. Both
superlattices have the same saturation magnetization of 672.8 emu/cm3 .
Comparing the saturation magnetizations of CrAs (572.4 emu/cm3 , 3.0
µB )—which agrees well with the measured 560 emu/cm3 —and MnAs (763.2
emu/cm3 , 4.0 µB ), the calculated value for the superlattices is approxi-
mately the average of the two constituents. Therefore, it is not viable to
increase the saturation magnetization by growing superlattices. These re-
sults are summarized in Table 5.5 along with DOS at EF for the majority-
spin channel and energy gap (Eg ) of the minority-spin states.
The supercell consists of one layer of GaAs, one layer of MnAs, two layers
of CrAs, and two capping layers of GaAs. The supercell model is shown in
Fig. 5.20. The unit cell is outlined by the solid and dashed lines. The a-axis
is along
√ the [110] direction of the conventional cell in the ZB structure. |a|
= 2ao and |b| = 3a o , where ao =5.722 Å is the optimized lattice constant
for GaAs.
Fig. 5.20 A superlattice model exhibits the spin-polarized ballistic transport properties.
a is along the [110] direction of the conventional cubic cell (Qian et al., 2004b).
Total and partial density of states The total density of states (TDOS)
and the PDOS are shown in Fig. 5.21. The qualitative features are similar to
those of CrAs/MnAs superlattice discussed in section 5.3.0.4. The majority-
spin states exhibit metallic behavior (the top panel of Fig. 5.21). They are
primarily contributed by the d-states of the Cr atom (the third panel). The
occupied d-manifold of Cr is higher in energy than that of the Mn atom (the
second and the third panels). The eg states of the Cr atom are occupied.
The minority-spin channel exhibits a gap at EF . The gap value is 0.95 eV
and is nearly half of the values of CrAs (1.88 eV) and MnAs (1.74 eV). The
occupied d-states are the bonding t2g states and those eg states are above
EF .
Charge densities In Fig. 5.22, the charge densities are plotted in a sec-
tion consisting of a zig-zag chain of atoms. There are three panels. The
top panel shows the total valence charge density of ↑ spin states. In the
region between the Mn and Cr atoms, the distribution is similar to the
CrAs/MnAs superlattice. The As atoms are located between the labeled
atoms as their nn to form a chain. In the GaAs regions, bond charges
are formed between the Ga atoms and their nn. The middle panel shows
the charge distribution of ↓ spin states. Similar to the CrAs/MnAs case,
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 5.21 Total DOS and PDOS of a superlattice showing the ballistic transport (Qian
et al., 2004b).
the four lobes associated with d-states of Cr and Mn atoms are illustrated.
They are mainly the t2g type of bonding states.
Fig. 5.22 Total charge densities of ↑ spin states (top), ↓ spin states (middle), and the ↑
spin channel (bottom) in an energy ranging between EF and EF + 0.3 eV. The section
contains the zig-zag chain of the TM elements, Ga, and As atoms (Qian et al., 2004b).
where σ labels the spin channel. A is a finite cross section, e is the electronic
charge, Pz is the z-component of momentum operator. Ψkn,σ is the wave
function for band n and at a k point inside the BZ with energy εkn,σ . The
results of the conductance for the two spin channels are shown in Fig. 5.23.
The upper panel is for the ↑ channel and shows a finite conductance at
EF . The conductance of the ↓ states is shown in the lower panel. There
is zero conductance at EF . All these results illustrate that the superlattice
shown in Fig. 5.20 can have a spin-polarized ballistic transport. This kind
of superlattices can be used to fabricate spin filters.
Fig. 5.23 Ballistic conductance along the c-axis for the superlattice is shown as a func-
tion of energy (Qian et al., 2005). EF is set to be zero.
5.4.1 Experiment
Growths of quantum dots composed of TM pnictide, such as MnAs, have
been reported by Ono et al. (2002) and Okabayashi et al. (2004), respec-
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
tively. We discuss the growths first, then characterizations, and finally, the
experimental determination of the electronic properties.
Fig. 5.24 The TEM image and SAED patterns of nanoscale MnAs and GaAs substrates
(Ono et al., 2002).
Fig. 5.25 Size distribution of MnAs quantum dots determined by SEM (Ono et al.,
2002).
quantum dots but not for the film. It was therefore concluded that this is
another evidence showing that the quantum dots have the ZB structure.
Fig. 5.26 (a) The on-resonance (hν=50 eV) photoemission spectra of Ga1−x Mnx As
alloy, medium density quantum dots, high density quantum dots, and a NiAs-like MnAs
film. (b) The difference spectra between the on-resonance (hν=50 eV) and off-resonance
(48 eV) spectra can be obtained as a measure of the Mn 3d partial density of states
(Okabayashi et al., 2004).
Fig. 5.27 (a) Hysteresis loop, and (b) the temperature dependence of the magnetization
in MnAs quantum dots (Ono et al., 2002).
SAED pattern are distinct from GaAs but similar to MnAs quantum dots
grown on a GaAs(001) substrate without any passivation of the S atoms.
For the alloys, the low temperature growth on the 10 nm buffer layer shows
a 1 × 2 surface reconstruction in the RHEED pattern. The results differ
from those for the quantum dots.
5.4.2 Theory
5.4.2.1 MnAs quantum dot
Qian et al. (2006b) built a model of MnAs quantum dot in the ZB struc-
ture. The model, a cluster, is composed of a total of 41 atoms. They
calculated the electronic properties and magnetic moment using ab initio
pseudopotential method and DFT with GGA exchange-correlation.
Cluster model The cluster was built by choosing a center atom in the
cubic crystalline MnAs and by retaining up to the third shells of the
atoms. In terms of number atoms in each shell, the cluster is specified
as MnAs4 Mn12 As24 . It has a radius of 6.1 Å. To saturate the dangling
bonds at the edge of the cluster, 60 hydrogen atoms were used. Since these
H atoms are attached to the outer shell As atoms, their effective charge is
chosen at 0.75 e, where e is the electron charge. The cluster is put in a
cubic supercell having a lattice constant of 22.96 Å, four times the size of
of the bulk MnAs with the ZB structure. The nearest neighbor distance
between atoms in neighboring clusters is 4.06 Å.
tively. Note that the AFM is defined as the spin orientations of the center
Mn and the rest Mn atoms are opposite. The partial DOS are the projected
DOS on one atom. As usual, EF is set to be zero. The major difference
between FM and AFM is that in FM the minority-spin channel exhibits a
gap of 1.83 eV as compared to 0.97 eV in AFM. The calculated magnetic
moments for FM and AFM are 52 µB /cluster and 42 µB /cluster, respec-
tively. With the integer values of magnetic moments and gaps appearing in
the minority-spin channel, the cluster in either the ferromagnetic or antifer-
romagnetic phase can be a HM based on the criterion of having an integer
magnetic moment of the sample.
Fig. 5.28 TDOS and PDOS for the ferromagnetic phase (a) and antiferromagnetic phase
(b) of MnAs quantum dot (Qian et al., 2006b) .
5.5.1 Experiment
In the following, the growth and characterization of (Ga,Mn)As- and
(Ga,Mn)Sb-DFH will be discussed first, then focus will be on the asso-
ciated physical properties.
Fig. 5.29 X-ray diffraction result of Fig. 5.30 The TEM image of a
(10/0.5)100 DFH grown at 280 ◦ C (12/0.5)50 -DFH, the Mn atoms are con-
(Kawakami et al., 2000). tained in dark lines (Chen et al., 2002).
TEM image of a (12/0.5)50 DFH is shown in Fig. 5.30. The light and dark
lines are clearly identifiable. The dark lines are those containing the Mn
atoms.
Fig. 5.31 (a) The SQUID hysteresis loop for a (10/0.5)100 (Ga,Mn)As-DFH, and (b)
the remanent magnetization as a function of temperature (Kawakami et al., 2000).
the easy axis. The results are presented with the contribution of a linear
background from GaAs being subtracted. Solid lines are for the DFH, while
the dashed lines are for the alloy. There is a 100% remanence showing the
ferromagnetic ordering along an easy magnetization axis. The temperature
dependence of the remanent magnetization is shown in Fig. 5.31(b). The
magnetization M was initially achieved by applying an external magnetic
field of 1000 Oe at 5 K. Then, measurements of M(T) were carried out
at zero field as temperature increases. At about and beyond 50 K, the
magnetic moment of the DFH is zero. The alloy magnetization persists up
to about 70 K.
SQUID magnetometry was used by Luo et al. (2002) to measure the
magnetizations of the samples. There are some differences between the
samples grown by Luo et al. (2002) and those synthesized by Kawakami
et al. (2000). Instead of using M, Luo et al. (2002) defined the term called
“effective spin density” (ESD) from the saturation magnetization expressed
in terms of the total spin aligned by the field per unit-area. The experiments
were carried out at 5 K with magnetic fields up to 55 kGauss (kG). ESD
plotted as a function of Mn fraction is shown in Fig. 5.32, where ESD
exhibits a maximum value of 40 in units of 1013 /cm2 at Mn fraction 0.25.
At Mn fraction 0.45, ESD is only about 10 × 1013 /cm2 .
by DFH not by any MnSb precipitates. According to Abe et al. (2000), the
room temperature ferromagnetism has been observed in GaMnSb if there
are MnSb precipitates. But such samples show temperature independent
hysteresis loops with the coercive field at RT to be the same value as at 5
K. The results shown in Fig. 5.33 do not display such features in GaMnSb
with precipitates.
Fig. 5.33 (a) Hysteresis loops of (Ga,Mn)Sb measured at different temperatures. (b)
Magnetization as a function of temperature (Chen et al., 2002).
Fig. 5.34 Curie temperature (TC ) as a function of GaAs layer thickness for (10/0.5)100
and (10/0.25)100 DFH’s (Kawakami et al., 2000).
Table 5.6 TC ’s of (Ga,Mn)As- and (Ga,Mn)Sb-DFH’s. The thickness of the substrate or buffer
layers is represented in monolayers (ML).
DFH Thickness (ML) Growth temp. (◦ C) TC (K) Reference
(Ga,Mn)As
(10/0.5)100 10–15 280 50 Kawakami et al. (2000)
(10/0.5)100 100 280 40 Kawakami et al. (2000)
(10/0.5)100 200 280 33 Kawakami et al. (2000)
(10/0.25)100 100 280 22 Kawakami et al. (2000)
(10/0.5)100 260 19 Kawakami et al. (2000)
(10/0.5)100 240 5 Kawakami et al. (2000)
(9/0.25) > 100 275 33 Luo et al. (2002)
(9/0.4) > 100 275 17 Luo et al. (2002)
(Ga,Mn)Sb
(various/0.5)50 > 400 Chen et al. (2002)
out using the van der Pauw configurations in a cryostat having a 170
kG (kGauss) superconducting magnet and Hall bar configuration. All the
(Ga,Mn)As-DFH samples do not exhibit any metallic features. The sheet
resistance of a sample, (9/0.5), is fitted well with
1
ln R = ln Ro − (T/To ) 2 , (5.5)
where To is 61 K and Ro is 850 Ω. This exponential dependence of R with
respect to T is known for materials such as an n-type δ-doped GaAs where
the conduction is contributed by a variable range of hopping. On the other
hand, all the (Ga,Mn)Sb-DFH’s are metallic based on the MR results. This
is in distinct contrast to the (Ga,Mn)As-DFH case.
Anomalous Hall effect (AHE) is able to probe the information about
the interaction between mobile carriers and local spin moment of magnetic
elements at low external magnetic field. The expression of the transverse
(perpendicular to the external applied electric field) resistivity ρxy contains
a term induced by the magnetization M.
ρxy = Ro B + 4πRa M, (5.6)
where Ro is the normal Hall coefficient, B is the strength of the external
magnetic field, and Ra is anomalous Hall coefficient. The first term is the
normal Hall effect. The second term defines the anomalous Hall effect.
In Fig. 5.35, the sheet and Hall resistances of (Ga,Mn)As-DFH and the
sheet and anomalous Hall resistance for (Ga,Mn)Sb-DFH are shown. For
(Ga,Mn)As-DFH, the sheet resistance decreases with temperature. The
shapes do not change significantly as the temperature varies from 12 to 52
K. They are peaked at B = 0. With the four-terminal van der Pauw tech-
nique, the Hall resistance Rxy is equivalent to the Hall resistivity ρxy (Met-
alidis and Bruno, 2006). The magnitude of Rxy (RHall ) given in Fig. 5.35
for a (Ga,Mn)As-DFH decreases with temperature. Rxy changes from neg-
ative to positive value in a very narrow (∼ 20 kG) region when the positive
external magnetic field reverses its sign. The question of the physical ori-
gins causing the temperature behavior, the double structures in the sheet
resistances, and their small Rxy values compared to the large ones (one
order of magnitude larger) is not answered.
Since the above-mentioned question is not answered, we just point out
the difference of the magnetotransport properties of the two (Ga,Mn)-
pnictide-DFH’s. For (Ga,Mn)Sb-DFH, both the sheet and Hall resistances
behave quite differently from the case of (Ga,Mn)As-DFH. Let us first com-
pare the sheet resistances shown in Figs. 5.35(a) and 5.36(a). At low tem-
perature (T < 40 K), the shapes are somewhat similar. However, the
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
5.5.2 Theory
5.5.2.1 GaAs-based DFH
Structure Immediately after publication of the experimental work by
Kawakami et al. (2000), Sanvito and Hill (2001) reported their theoretical
study of an (NGaAs /1)∞ (Ga,Mn)As-DFH, where NGaAs is the number of
GaAs layers and ∞ stands for a superlattice. They used SIESTA algorithm
with the local spin density approximation (LSDA) (Sánchez-Portal et al.,
1997). The experimentally determined lattice constant of GaAs, 5.65 Å,
was used. These authors also examined antisite defects.
Fig. 5.37 Band structures near EF of the (a) majority- and (b) minority-spin channels
for a (15/1)∞ (Ga,Mn)As-DFH (Sanvito and Hill, 2001).
Fig. 5.38 Total and partial spin-polarized CIP and CPP conductances of a (15/1)∞
DFH (Sanvito and Hill, 2001).
HMs. The major obstacle is the growth of Heusler alloys, oxides, and TM
pnictides. Then, there is the T∗ (∼ 88 K) at which the half metallicity
disappears in a Heusler alloy due to spin-flip transitions and the fact that
P vanishes precipitously as the temperature approaches RT.
Qian et al. (2006a) proposed a DFH by doping a δ-layer of Mn atoms in
Si (Si-based DFH) motivated by the fact that Si technologies are the most
mature among the semiconductors so the growth problems can possibly
be eliminated. They used a supercell model of a (31/1)∞ DFH shown in
Fig. 5.39. The planewave pseudopotential method, and the GGA exchange-
correlation within DFT, were used to calculate the density of states of Si-
based DFH for the investigation of whether there is a T∗ in this DFH.
Table 5.7 Comparison of the unrelaxed and relaxed (31/1)∞ -DFH. m is the
magnetic moment per Mn atom, N(EF ) is the DOS of the conducting channel
at EF , ∆F A = EF M - EAF M , and Eg is the energy gap.
Structure m (µB ) N(EF ) (states/eV-cell) ∆F A (meV) Eg (eV)
Unrelaxed 3.0 1.06 –523.91 0.20
Relaxed 3.0 1.25 –442.38 0.25
Fig. 5.40 TDOS (states/eV-unit-cell) and PDOS (states/eV) for a (31/1)∞ DFH.
PDOS are for the nearest and second nearest neighbor Si atoms (Qian et al., 2006a).
The PDOS enable one to identify states around EF . With the energy
window between EF –1.0 and EF +2.0 eV, the majority-spin states are pri-
marily the hybridized p-states of the nearest neighbor Si atom (Si(I)) to
the Mn atom and the t2g states of the Mn atom (the second and the third
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
panels). As will be shown later, these are part of the antibonding states.
The strongest region of the hybridization is at 2.6 eV below EF . The eg
states of the Si(I) atom are centered around –2.5 eV with a width of about
1.0 eV. In the range between EF –1.0 and EF +2.0 eV, the second neighbor
Si(II) atom has hardly any significant contributions. Between EF –2.0 and
EF –4.0 eV, those p-states have a uniform distribution.
In the minority-spin channel, those states just below the valence band
maximum are the p(Si(I))-d(Mn(t2g )) hybridized bonding states. EF is
located near the top of the valence bands. The eg states of Mn atom are
not occupied. The conduction band minimum is contributed by the low
energy end of the eg states. Neither Si(I) nor Si(II) have contributions to
the states at the conduction band minimum.
Fig. 5.41 (a) The band structure of the majority-spin states and (b) the band structure
of the minority-spin channel (Qian et al., 2006a).
Fig. 5.42 (a) The Fermi surface in the X-Y section of the two dimensional BZ, and (b)
the hole charge distribution for the states near EF (Qian et al., 2006a).
↓ spin channel.
We now focus on the Si-based DFH. Its EF is explicitly labeled and
is located near the valence band maximum (VBM) of the ↓ spin states.
The gap of the insulating minority-spin states is denoted by ∆↓ . The
important gap is labeled as δ which is measured from EF to CBMin in the
minority-spin channel. Its value is approximately 0.25 eV (Table 5.7) which
is about 10 times larger than RT. Therefore, the probability is drastically
reduced for an electron in the majority-spin channel to make a spin-flip
transition from EF to CBMin of the minority-spin channel. However, one
can ask the following question: what about the possibility for an electron
from VBM to make a spin-flip transition to a state at EF ? As we see
from the PDOS (Fig. 5.40), the valence band is dominated by p-states of
Si(I). Consequently, one expects the corresponding charge density to be
concentrated near the Si atom. The overlap between the Si(I) p-states
and Mn d-states will be small. As shown in Fig. 5.40, the hole states
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Fig. 5.43 Schematic diagram of the DOS of a HM with the majority-spin channel show-
ing metallic behavior. The Fermi energy of the Heusler alloy is now indicated as “EF
of HA” and is right below CBMin. The Fermi energy of the Si-based DFH is labeled
as EF and is located close to the valence band maximum (VBM) of the DFH. δ is the
energy difference between CBMin and EF . ∆↓ is the fundamental gap of the insulating
channel.
Even though DFH’s behave as two dimensional systems, their basic struc-
tures are still three dimensional. To explore the possibility of finding half
metallicity in truly low dimensional systems, Dag et al. (2005) investigated
one dimensional systems by doping TM elements in carbon wires, Cn (TM).
They used a pseudopotential method within DFT with spin-polarized GGA
exchange-correlation.
spin polarization and have lower energies than the corresponding unpolar-
ized cases. The values of the magnetic moment/unit-cell (4.0 µB ) are larger
than the one for MnC with larger lattice constant. A simple explanation is
that each of the two neighboring C atoms transfers one electron from the
Cr atom to doubly occupy its p-orbital. It is interesting to note that the
half metallic behavior depends on n. The metallic feature alternates be-
tween the majority- and minority-spin states as n increases from 2 to 7. For
even n, the majority-spin channel is metallic. In addition to the electronic
properties, these authors also investigated the stability of the wires and the
effects of spin-orbit interaction. Based on their findings, they suggested a
way to grow these HMs.
bonds form alternately between C atoms and a longer bond between the
Cr atom and its neighbors. The distributions are C-1.25-C-1.33-C-1.25-C-
2.1-Cr. The different bond lengths between the Cr and C atoms result in
different band structures.
From the PDOS, the mσ bands are formed by Cr-3d states hybridized
with 2p-state of the neighboring C atoms. The vσ band is composed of s-
states and d-states of the Cr atom. The unoccupied cσ band is contributed
by the hybridized antibonding C p-states and Cr d-states.
Fig. 5.44 The spin-polarized band structures and density of states of (a) C3 Cr and (b)
C4 Cr (Dag et al., 2005).
Fig. 5.45 TDOS of the two spin channels with (solid lines) and without (shaded areas)
S-O interaction for C3 Cr (Dag et al., 2005).
ab initio molecular dynamics simulations with T set between 750 and 1000
K using Nosé thermostat. The atoms were displaced in random directions.
The results confirm the stability of the chains at small n.
There are still the possibilities of the Peierls instability and strain issues.
These authors expanded the models by a factor 2 and calculated the band
structures. The splitting of bands at EF should manifest the effect of the
Peierls instability. There is no splitting of metallic bands at EF in Cn Cr
chain structures. Axial strains were applied to examine whether strain can
destroy half metallicity. The half metallicity is robust for n = 4 under ǫz
= ±0.05. However, for n = 5, its half metallic properties remain only at
ǫz ≤0.05. With ǫz = -0.05, C5 Cr is transformed to a ferromagnetic metal.
At ǫz = 0.10, C3 Cr is a semiconductor while it becomes a ferromagnetic
metal having the magnetic moment of 3.1 µB /unit-cell when ǫz = -0.10.
Fig. 5.46 (a) Energetics of doping path for C5 Cr, and (b) a possible path of growing
C7 Cr wire (Dag et al., 2005).
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Growth In Fig. 5.46(a), the doping path gives the energetics of a process
of doping Cr in C5 . The steps are indicated as A, B, C, and D. From A to
D, there is a barrier of 1.86 eV. This is considered to be a large barrier to
overcome. Therefore, the doping process is not a viable method to grow a
half metallic wire. The growth path in Fig. 5.46(b) suggests a possible way
to grow a C7 Cr wire without a large barrier. The process A shows to first
grow a C7 chain, then a Cr atom is deposited at one end of the C7 . The
processes of growing C7 Cr wire by adding more C atoms to one end of Cr
atom are indicated as B and C. From the interaction energy with distance
d there is no significant barrier.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Appendix A
Appendix
Fig. A.1 The resistivity of a ferromagnetic metal vs. Hext . The resistivities at A and
B are used to define ∆ρ.
243
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography
Abe, E., Matsukura, F., Yasuda, H., Ohno, Y. and Ohno, H. (2000). Molecular
beam epitaxy of III–V diluted magnetic semiconductor (Ga,Mn)Sb, Physica
E: Low-dimensional Systems and Nanostructures 7, pp. 981–985.
Akai, H., Akai, M. and Kanamori, J. (1985). Electronic structure of impurities
in ferromagnetic iron. II. 3d and 4d impurities, J. Phys. Soc. Jpn. 54, pp.
4257–4264.
Akai, H. and Dederichs, P. H. (1993). Local moment disorder in ferromagnetic
alloys, Phys. Rev. B 47, pp. 8739–8747.
Akinaga, H., Manago, T. and Shirai, M. (2000a). Material design of half-metallic
zinc-blende CrAs and the synthesis by molecular-beam epitaxy, Jpn. J.
Appl. Phys. Lett. 39, p. L1118.
Akinaga, H. and Mizuguchi, M. (2004). Zinc-blende CrAs/GaAs multilayers
grown by molecular-beam epitaxy, J. Phys.: Condens. Mat. 16, pp. S5549–
S5553.
Akinaga, H., Mizuguchi, M., Ono, K. and Oshima, M. (2000b). Room-
temperature thousandfold magnetoresistance change in MnSb granular
films: Magnetoresistive switch effect, Appl. Phys. Lett. 76, pp. 357–359.
Albert, F. J., Katine, J. A., Buhrman, R. A. and Ralph, D. C. (2000). Spin-
polarized current switching of a Co thin film nanomagnet, Appl. Phys. Lett.
77, pp. 3809–3811.
Alonso, J. A., Martinez-Lope, M. J., Casais, M. T., Macmanus-Driscoll, J. L.,
de Silva, P. S. I. P. N., Cohen, L. F. and Fernandez-Diaz, M. T. (1997).
Non-stoichiometry, structural defects and properties of LaMnO3+δ with
high δ values (0.11 ≤ δ ≤ 0.29), J. Mater. Chem. 7, pp. 2139–2144.
Ambrose, T., Krebs, J. and Prinz, G. (2000). Epitaxial growth and magnetic
properties of single-crystal Co2 MnGe Heusler alloy films on GaAs(001),
Appl. Phys. Lett. 76, pp. 3280–3282.
Andersen, O. K. (1975). Linear methods in band theory, Phys. Rev. B 12, pp.
3060–3083.
Anderson, P. W. (1950). Antiferromagnetism. theory of superexchange interac-
tion, Phys. Rev. 79, pp. 350–356.
Anderson, P. W. (1956). Ordering and antiferromagnetism in ferrites, Phys. Rev.
245
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 247
Brener, N. E., Tyler, J. M., Callaway, J., Bagayoko, D. and Zhao, G. L. (2000).
Electronic structure and Fermi surface of CrO2 , Phys. Rev. B 61, pp.
16582–16588.
Brown, P. J., Neumann, K. U., Webster, P. J. and Ziebeck, K. R. A. (2000). The
magnetization distributions in some Heusler alloys proposed as half-metallic
ferromagnets, J. Phys.: Condens. Mat. 12, p. 1827.
Bruno, P. (2003). Exchange interaction parameters and adiabatic spin-wave spec-
tra of ferromagnets: A“renormalized magnetic force theore”, Phys. Rev.
Lett. 90, p. 087205.
Buschow, K. H. J. and van Engen, P. G. (1981). Magnetic and magneto-optical
properties of Heusler alloys based on aluminium and gallium, J. Magn.
Magn. Mater. 25, pp. 90–96.
Buschow, K. H. J., van Engen, P. G. and Jongebreur, R. (1983). Magneto-optical
properties of metallic ferromagnetic materials, J. Magn. Magn. Mater. 38,
pp. 1–22.
Bussmann, K., Prinz, G. A., Cheng, S.-F. and Wang, D. (1999). Switching of ver-
tical giant magnetoresistance devices by current through the device, Appl.
Phys. Lett. 75, pp. 2476–2478.
Büttiker, M., Imry, Y., Landauer, R. and Pinhas, S. (1985). Generalized many-
channel conductance formula with application to small rings, Phys. Rev. B
31, pp. 6207–6215.
Caballero, J., Park, Y., Cabbibo, A., Childress, J., Petroff, F. and Morel, R.
(1997). Deposition of high-quality NiMnSb magnetic thin films at moderate
temperatures, J. Appl. Phys. 81, pp. 2740–2744.
Cai, Y. Q., Ritter, M., Weiss, W. and Bradshaw, A. M. (1998). Valence-band
structure of epitaxially grown Fe3 O4 (111) films, Phys. Rev. B 58, pp. 5043–
5051.
Callen, H. (1963). Green function theory of ferromagnetism, Phys. Rev. 130, p.
890.
Caminat, P., Valerio, E., Autric, M., Grigorescu, C. and Monnereau, O. (2004).
Double beam pulse laser deposition of NiMnSb thin films at ambient tem-
perature, Thin Solid Films 453-454, pp. 269–272.
Carbonari, A. W., Pendl, W., Attili, R. N. and Saxena, R. N. (1993). Magnetic
hyperfine fields in the heusler alloys Co2 YZ (Y=Sc, Ti, Hf, V, Nb; Z=Al,
Ga, Si, Ge, Sn), Hyperfine Interactions 80, pp. 971–976.
Carra, P., Thole, B. T., Altarelli, M. and Wang, X. (1993). X-ray circular dichro-
ism and local magnetic fields, Phys. Rev. Lett. 70, pp. 694–697.
Celotta, R., Pierce, D., Wang, G., Bader, S. and Felcher, G. (1979). Surface
magnetization of ferromagnetic ni(110): A polarized low-energy electron
diffraction experiment, Phys. Rev. Lett. 43, p. 728.
Ceperley, D. and Alder, B. (1980). Ground state of the electron gas by a stochastic
method, Phys. Rev. Lett. 45, p. 566.
Chang, L. and Ploog, K. (1985). Molecular beam epitaxy and heterostructures,
NATO Adv. Sci. Inst. Ser. E87.
Chen, C. T., Idzerda, Y. U., Lin, H. J., Smith, N. V., Meigs, G., Chaban, E.,
H, G., Ho, Pellegrin, E. and Sette, F. (1995). Experimental Confirmation
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 249
of the X-Ray Magnetic Circular Dichroism Sum Rules for Iron and Cobalt,
Physical Review Letters 75, pp. 152–155.
Chen, X., Na, M., Cheon, M., Wang, S., Luo, H., McCombe, B. D., Liu, X.,
Sasaki, Y., Wojtowicz, T., Furdyna, J. K., Potashnik, S. J. and Schiffer,
P. (2002). Above-room-temperature ferromagnetism in GaSb/Mn digital
alloys, Appl. Phys. Lett. 81, pp. 511–513.
Cheng, S., Nadgomy, B., Bussmann, K., Carpenter, E., Das, B., Trotter, G.,
Raphael, M. and Harris, V. (2001). Growth and magnetic properties of
single crystal Co2 MnX (X=Si,Ge) Heusler alloys, IEEE Transactions on
Magnetics 37, pp. 2176–2178, doi:10.1109/20.951116.
Clowes, S., Miyoshi, Y., Bugoslavsky, Y., Branford, W., Grigorescu, C., Manea,
S., Monnereau, O. and Cohen, L. (2004). Spin polarization of the transport
current at the free surface of bulk NiMnSb, Phys. Rev. B 69, p. 214425.
Coey, J. M. D. and Venkatesan, M. (2002). Half-metallic ferromagnetism: Exam-
ple of CrO2 (invited), J. Appl. Phys. 91, pp. 8345–8350.
Coey, J. M. D., Versluijs, J. J. and Venkatesan, M. (2002). Half-metallic oxide
point contacts, J. Phys. D: Appl. Phys. 35, pp. 2457–2466.
Continenza, A., Picozzi, S., Geng, W. T. and Freeman, A. J. (2001). Coordination
and chemical effects on the structural, electronic, and magnetic properties
in Mn pnictides, Phys. Rev. B 64, p. 085204.
Şaşoǧlu, E., Galanakis, I., Sandratskii, L. M. and Bruno, P. (2005a). Stability
of ferromagnetism in the half-metallic pnictides and similar compounds: a
first-principles study, J. Phys.: Condens. Mat. 17, p. 3915.
Şaşoǧlu, E., Sandratskii, L. M., Bruno, P. and Galanakis, I. (2005b). Exchange
interactions and temperature dependence of magnetization in half-metallic
Heusler alloys, Phys. Rev. B 72, p. 184415.
Dag, S., Tongay, S., Yildirim, T., Durgun, E., Senger, R., Fong, C. Y. and Ciraci,
S. (2005). Half-metallic properties of atomic chains of carbon-transition
metal compounds, Phys. Rev. B 72, p. 155444.
Datta, S. and Das, B. (1990). Electronic analog of the electrooptic modulator,
Appl. Phys. Lett. 56, pp. 665–667.
Daughton, J. (1992). Magnetoresistive memory technology, Thin Solid Films 216,
pp. 162–168.
de Groot, R. A., Mueller, F., Engen, P. and Buschow, K. (1983). New class of
materials: Half-metallic ferromagnets, Phys. Rev. Lett. 50, pp. 2024–2027.
de Wijs, G. A. and de Groot, R. A. (2001). Towards 100% spin-polarized charge-
injection: The half-metallic NiMnSb/CdS interface, Phys. Rev. B 64, p.
020402.
Dedkov, Y. (2004). Spin-Resolved Photoelectron Spectroscopy of Oxidic Half-
Metallic Ferromagnets and Oxide/Ferromagnet Interfaces, Ph.D. thesis.
Dedkov, Y. S., Rüdiger, U. and Güntherodt, G. (2002). Evidence for the half-
metallic ferromagnetic state of Fe3 O4 by spin-resolved photoelectron spec-
troscopy, Phys. Rev. B 65, p. 064417.
Dieny, B., Humbert, P., Speriosu, V. S., Metin, S., Gurney, B. A., Baugart,
P. and Lefakis, H. (1992). Giant magnetoresistance of magnetically soft
sandwiches: Dependence on temperature and on layer thicknesses, Phys.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 251
Bibliography 253
Hordequin, C., Ristoiu, D., Ranno, L. and Pierre, J. (2000). On the cross-over
from half-metal to normal ferromagnet in NiMnSb, Euro. Phys. J. B 16,
pp. 287–293.
Hörmandinger, G., Weinberger, P., Marksteiner, P. and Redinger, J. (1988). The-
oretical calculations of core-core-valence Auger spectra: Applications to the
L3 M2,3 V transitions of Ti in nonstoichiometric Ti-C, Ti-N, and Ti-O, Phys.
Rev. B 38, pp. 1040–1046.
Hu, G. and Suzuki, Y. (2002). Negative spin polarization of Fe3 O4 in
magnetite/manganite-based junctions, Phys. Rev. Lett. 89, p. 276601.
Huang, D. J., Chang, C. F., Chen, J., Tjeng, L. H., Rata, A. D., Wu, W. P.,
Chung, S. C., Lin, H. J., Hibma, T. and Chen, C. T. (2002). Spin-resolved
photoemission studies of epitaxial Fe3 O4 (100) thin films, J. Magn. Magn.
Mater. 239, pp. 261–265.
Hubbard, J. (1963). Elecron correlations in narrow energy bands, Proc. R. Soc.
Lond. A 276, pp. 238—257, doi:10.1098/rspa.1963.0204.
Husmann, A. and Singh, L. (2006). Temperature dependence of the anomalous
Hall conductivity in the Heusler alloy Co2 CrAl, Phys. Rev. B 73, p. 172417.
Hwang, H. Y. and Cheong, S. W. (1997). Enhanced intergrain tunneling magne-
toresistance in half-metallic CrO2 films, Science 278, pp. 1607–1609.
Hwang, H. Y., Cheong, S.-W., Ong, N. P. and Batlogg, B. (1996). Spin-polarized
intergrain tunneling in La2/3 Sr1/3 MnO3 , Phys. Rev. Lett. 77, pp. 2041–
2044.
Iizumi, M., Koetzle, T. F., Shirane, G., Chikazumi, S., Matsui, M. and Todo, S.
(1982). Structure of magnetite Fe3 O4 below the verwey transition temper-
ature, Acta Crystall. B 38, pp. 2121–2133.
Inomata, K., Okamura, S., Miyazaki, A., Kikuchi, M., Tezuka, N., Wojcik, M.,
Jedryka, E. and et al, K. I. (2006). Structural and magnetic properties and
tunnel magnetoresistance for Co2 (Cr,Fe)Al and Co2 FeSi full-Heusler alloys,
J. Phys. D: Appl. Phys. 39, p. 816.
Ishida, S., Akazawa, S., Kubo, Y. and Ishida, J. (1982). Band theory of Co2 MnSn,
Co2 TiSn and Co2 TiAl, J. Phys. F: Met. Phys. 12, p. 1111.
Ishida, S., Masaki, T., Fujii, S. and Asano, S. (1998). Theoretical search for half-
metalliic films of Co2 MnZ (Z = Si, Ge), Physica B: Condensed Matter 245,
pp. 1–8.
Ivanov, P. G., Watts, S. M. and Lind, D. M. (2001). Epitaxial growth of CrO2
thin films by chemical-vapor deposition from a Cr8 O21 precursor, J. Appl.
Phys. 89, pp. 1035–1040.
Janak, J., Moruzzi, V. and Williams, A. (1975). Ground-state thermomechanical
properties of some cubic elements in the local-density formalism, Phys. Rev.
B 12, p. 1257.
Jeng, H.-T., Guo, G. Y. and Huang, D. J. (2004). Charge-orbital ordering and
Verwey transition in magnetite, Phys. Rev. Lett. 93, p. 156403.
Jenkins, S. J. (2004). Ternary half-metallics and related binary compounds: Sto-
ichiometry, surface states, and spin, Phys. Rev. B 70, p. 245401.
Ji, Y., Strijkers, G., Yang, F., Chien, C., Byers, J., Anguelouch, A., Xiao, G. and
Gupta, A. (2001). Determination of the spin polarization of half-metallic
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
CrO2 by point contact Andreev reflection, Phys. Rev. Lett. 86, p. 5585.
Jonker, G. H. and Santen, J. H. V. (1950). Ferromagnetic compounds of man-
ganese with perovskite structure, Physica 16, pp. 337–349.
Kabani, R., Terada, M., Roshko, A. and Moodera, J. S. (1990). Magnetic prop-
erties of NiMnSb films, J. Appl. Phys. 67, pp. 4898–4900.
Kallmayer, M., Schneider, H., Jakob, G., Elmers, H. J., Balke, B. and Cramm, S.
(2007). Interface magnetization of ultrathin epitaxial Co2 FeSi(110)/Al2 O3
films, J. Phys. D: Appl. Phys. 40, pp. 1552–1557.
Kämmerer, S., Heitmann, S., Meyners, D., Sudfeld, D., Thomas, A., Hütten,
A. and Reiss, G. (2003). Room-temperature preparation and magnetic
behavior of Co2 MnSi thin films, J. Appl. Phys. 93, pp. 7945–7947, doi:
10.1063/1.1556249.
Kämmerer, S., Thomas, A., Hütten, A. and Reiss, G. (2004). Co2 MnSi Heusler
alloy as magnetic electrodes in magnetic tunnel junctions, Appl. Phys. Lett.
85, pp. 79–81.
Kämper, K. P., Schmitt, W., Güntherodt, G., Gambino, R. J. and Ruf, R. (1987).
CrO2 –a new half-metallic ferromagnet? Phys. Rev. Lett. 59, pp. 2788–
2791.
Kandpal, H., Fecher, G., Felser, C. and Schöuml;nhense, G. (2006). Correlation
in the transition-metal-based Heusler compounds Co2 MnSi and Co2 FeSi,
Phys. Rev. B 73, p. 094422.
Kang, J.-S., Park, J.-G., Olson, C. G., Younq, S. J. and Min, B. I. (1995). Va-
lence band and Sb 4d core level photoemission of the XMnSb-type Heusler
compounds (X=Pt,Pd,Ni), J. Phys.: Condens. Mat. 7, pp. 3789–3800.
Karthik, S., Rajanikanth, A., Nakatani, T., Gercsi, Z., Takahashi, Y., Fu-
rubayashi, T., Inomata, K. and Hono, K. (2007). Effect of Cr substitution
for Fe on the spin polarization of Co2 Crx Fe1−x Si Heusler alloys, J. Appl.
Phys. 102, pp. 043903–6.
Katsnelson, M. I., Irkhin, V. Y., Chioncel, L., Lichtenstein, A. I. and deGroot,
R. A. (2008). Half-metallic ferromagnets: From band structure to many-
body effects, Rev. Mod. Phys. 80, pp. 315–378.
Kautzky, M. C., Mancoff, F. B., Bobo, J. F., Johnson, P. R., White, R. L. and
Clemens, B. M. (1997). Investigation of possible giant magnetoresistance
limiting mechanisms in epitaxial PtMnSb thin films, J. Appl. Phys. 81, pp.
4026–4028.
Kawakami, R. K., Johnston-Halperin, E., Chen, L. F., Hanson, M., Guebels, N.,
Speck, J. S., Gossard, A. C. and Awschalom, D. D. (2000). (Ga,Mn)As as a
digital ferromagnetic heterostructure, Appl. Phys. Lett. 77, pp. 2379–2381.
Kawano, H., Kajimoto, R., Kubota, M. and Yoshizawa, H. (1996a). Canted an-
tiferromagnetism in an insulating lightly doped La1−x Srx MnO3 with x ≤
0.17, Phys. Rev. B 53, pp. 2202–2205.
Kawano, H., Kajimoto, R., Kubota, M. and Yoshizawa, H. (1996b).
Ferromagnetism-induced reentrant structural transition and phase diagram
of the lightly doped insulator La1−x Srx MnO3 with x ≤ 0.17, Phys. Rev. B
53, pp. R14709–R14712.
Kelekar, R. and Clemens, B. (2004). Epitaxial growth of the Heusler alloy
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 255
Bibliography 257
M., Wang, S. M., Sasaki, Y., Liu, X. and Furdyna, J. K. (2002). Transport
and magnetic properties of ferromagnetic GaAs/Mn digital alloys, Physica
E: Low-dimensional Systems and Nanostructures 12, pp. 366–369.
Ma, C., Yang, Z. and Picozzi, S. (2006). Ab initio electronic and magnetic struc-
ture in La0.66 Sr0.33 MnO3 : Strain and correlation effects, J. Phys.: Con-
dens. Mat. 18, pp. 7717–7728.
Mackintosh, A. R. and Andersen, O. K. (1980). The electronic structure of tran-
sition metals, chap. 5.
Malozemoff, A., Williams, A. and Moruzzi, V. (1984). “Band-gap theory“ of
strong ferromagnetism: Application to concentrated crystalline and amor-
phous Fe- and Co-metalloid alloys, Phys. Rev. B 29, p. 1620.
Mancoff, F., Boboo, J., Richter, O., Bessho, K., Johnson, P., Sinclair, R., Nix,
W., White, R. and Clemens, B. (1999). Growth and characterization of
epitaxial NiMnSb/PtMnSb C1b Heusler alloy superlattices, J. Materi. Res.
14, p. 1560.
Marcus, P. M. and Moruzzi, V. L. (1988). Stoner model of ferromagnetism and
total-energy band theory, Phys. Rev. B 38, pp. 6949–6953.
Martin, R. M. (2004). Electronic Structure: Basic Theory and Practical Methods
(Cambridge University Press, Cambridge).
Matsubara, K., Anno, H., Kaneko, H. and Imai, Y. (1999). Electrical proper-
ties of half-metallic PtMnSb-based Heusler alloys, Eighteenth International
Conference on Thermoelectrics (1999). , pp. 60–63.
Mazin, I. I., Singh, D. J. and Ambrosch-Draxl, C. (1999). Transport, optical, and
electronic properties of the half-metal CrO2 , Phys. Rev. B 59, pp. 411–418.
McCormack, M., Jin, S., Tiefel, T. H., Fleming, R. M., Phillips, J. M. and
Ramesh, R. (1994). Very large magnetoresistance in perovskite–like La–
Ca–Mn–O thin films, Appl. Phys. Lett. 64, pp. 3045–3047.
McQueeney, R. J., Yethiraj, M., Montfrooij, W., Gardner, J. S., Metcalf, P.
and Honig, J. M. (2006). Investigation of the presence of charge order in
magnetite by measurement of the spin wave spectrum, Phys. Rev. B 73, p.
174409.
Meservey, R. and Tedrow, P. M. (1994). Spin-polarized electron tunneling, Physics
Reports 238, pp. 173–243.
Metalidis, G. and Bruno, P. (2006). Inelastic scattering effects and the Hall resis-
tance in a four-probe ring, Phys. Rev. B 73, p. 113308.
Miyamoto, K., Kimura, A., Iori, K., Sakamoto, K., Xie, T., Moko, T., Qiao, S.,
Taniguchi, M. and Tsuchiya, K. (2004). Element-resolved magnetic mo-
ments of Heusler-type ferromagnetic ternary alloy Co2 MnGe, J. Phys.:
Condens. Mat. 16, p. S5797.
Mizuguchi, M., Akinaga, H., Manago, T., Ono, K., Oshima, M., Shirai, M., Yuri,
M., Lin, H. J., Hsieh, H. H. and Chen, C. T. (2002). Epitaxial growth of
zinc-blende CrAs/GaAs multilayer, J. Appl. Phys. 91, pp. 7917–7919.
Moore, G. E. (1965). Cramming more components onto integrated circuits, Elec-
tronics 38, p. 114.
Moos, R., Menesklou, W. and Härdtl, K. H. (1995). Hall mobility of undoped n-
type conducting strontium titanate single crystals between 19 K and 1373
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 259
Bibliography 261
2351.
Ristoiu, D., Nozir̀es, J. P., Borca, C. N., Komesu, T., k. Jeong, H. and Dowben,
P. A. (2000b). The surface composition and spin polarization of NiMnSb
epitaxial thin films, EPL (Europhysics Letters) 49, p. 624.
Ritchie, L., Xiao, G., Ji, Y., Chen, T., Chien, C., Zhang, M., Chen, J., Liu, Z.,
Wu, G. and Zhang, X. (2003). Magnetic, structural, and transport prop-
erties of the Heusler alloys Co2 MnSi and NiMnSb, Phys. Rev. B 68, p.
104430.
Robinson, I. K. (1991). Handbook of synchrotron radiation, Vol. 3.
Ross, C. (2001). Patterned magnetic recording media, Ann. Rev. Mater. Res. 31,
pp. 203–235.
Rusz, J., Bergqvist, L., Kudrnovskỳ, J. and Turek, I. (2006). Exchange interac-
tions and Curie temperatures in Ni2−x MnSb alloys: First-principles study,
Phys. Rev. B 73, p. 214412.
Sacchi, M., Spezzani, C., Carpentiero, A., Prasciolu, M., Delaunay, R., Luning,
J. and Polack, F. (2007). Experimental setup for lensless imaging via soft
x-ray resonant scattering, Rev. Sci. Instrum. 78, pp. 043702–6.
Sakuraba, Y., Hattori, M., Oogane, M., Ando, Y., Kato, H., Sakuma, A.,
Miyazaki, T. and Kubota, H. (2006). Giant tunneling magnetoresistance
in Co2 MnSi/Al–O/Co2 MnSi magnetic tunnel junctions, Appl. Phys. Lett.
88, pp. 192508–3.
Salamon, M. B. and Jaime, M. (2001). The physics of manganites: Structure and
transport, Rev. Mod. Phys. 73, pp. 583–628.
Sánchez-Portal, D., Ordejón, P., Artacho, E. and Soler, J. M. (1997). Density-
functional method for very large systems with LCAO basis sets, Interna-
tional Journal of Quantum Chemistry 65, pp. 453–461.
Sandratskii, L. M. (1998). Noncollinear magnetism in itinerant-electron systems:
theory and applications, Advances in Physics 47, pp. 91– 160.
Sanvito, S. and Hill, N. (2001). Ab initio transport theory for digital ferromagnetic
heterostructures, Phys. Rev. Lett. 87, p. 267202.
Satpathy, S. and Vukajlović, Z. S. P. F. R. (1996). Electronic structure of the
perovskite oxides: La1−x Cax MnO3 , Phys. Rev. Lett. 76, pp. 960–963.
Savtchenko, L., Engel, B., Rizzo, N., Deherrera, M. and Janesky, J. (2003).
Method of writing to scalable magnetoresistance random access memory
element, Us patent.
Schlomka, J.-P., Tolan, M. and Press, W. (2000). In situ growth study of NiMnSb
films on MgO(001) and Si(001), Appl. Phys. Lett. 76, pp. 2005–2007.
Schlottmann, P. (2003). Double-exchange mechanism for CrO2 , Phys. Rev. B 67,
p. 174419.
Schmalhorst, J., Kämmerer, S., Sacher, M., Reiss, G., Hütten, A. and Scholl,
A. (2004). Interface structure and magnetism of magnetic tunnel junctions
with a Co2 MnSi electrode, Phys. Rev. B 70.
Schneider, H., Herbort, C., Jakob, G., Adrian, H., Wurmehl, S. and Felser, C.
(2007). Structural, magnetic and transport properties of Co2 FeSi Heusler
films, Journal of Physics D: Applied Physics 40, p. 1548.
Schneider, H., Jakob, G., Kallmayer, M., Elmers, H., Cinchetti, M., Balke, B.,
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 263
Bibliography 265
Yablonskikh, M., Yarmoshenko, Y., Grebennikov, V., Kurmaev, E., Butorin, S.,
Duda, L.-C., Nordgren, J., Plogmann, S. and Neumann, M. (2001). Origin
of magnetic circular dichroism in soft x-ray fluorescence of Heusler alloys
at threshold excitation, Phys. Rev. B 63, p. 235117.
Yamasaki, A., Imada, S., Suga, S., Kanomata, T. and Ishida, S. (2002). Magnetic
circular dichroism in the soft x-ray absorption spectra of Co-based Heusler
alloys, Surface Review and Letters 9, pp. 955–960.
Yamashita, J. and Kondo, J. (1958). Superexchange interaction, Phys. Rev. 109,
pp. 730–741.
Yanase, A. and Siratori, K. (1984). Band structure in the high temperature phase
of Fe3 O4 , J. Phys. Soc. Jpn. 53, pp. 312–317.
Yao, Y., Kleinman, L., MacDonald, A., Sinova, J., Jungwirth, T., sheng Wang,
D., Wang, E. and Niu, Q. (2004). First principles calculation of anomalous
Hall conductivity in ferromagnetic bcc Fe, Phys. Rev. Lett. 92, p. 037204.
Ye, J., Kim, Y., Millis, A., Shraiman, B., Majumdar, P. and Tescaronanovicacute,
Z. (1999). Berry pphase theory of the anomalous hall effect: Application to
colossal magnetoresistance manganites, Phys. Rev. Lett. 83, p. 3737.
Yoneda, Y. (1963). Anomalous surface reflection of x rays, Phys. Rev. 131, p.
2010.
Yoshimura, K., Miyazaki, A., Vijayaraghavan, R. and Nakamura, Y. (1985). Hy-
perfine field of the Co2 YZ Heusler alloy (Y = V, Cr, Mn and Fe; Z = Al
and Ga), Journal of Magnetism and Magnetic Materials 53, pp. 189–198.
Youn, S. and Min, B. (1995). Effects of the spin-orbit interaction in Heusler com-
pounds: Electronic structures and Fermi surfaces of NiMnSb and PtMnSb,
Phys. Rev. B 51, p. 10436.
Young, R. A., Sakthivel, A., Moss, T. S. and Paiva-Santos, C. O. (1995). DBWS –
9411 – an upgrade of the DBWS *.* programs for Rietveld refinement with
PC and mainframe computers, J. Appl. Cryst. 28, pp. 366–367.
Zeller, R. (2006). Computational nanoscience: do it yourself! p. 419.
Zener, C. (1951). Interaction between the d–shells in the transition metals. II.
ferromagnetic compounds of manganese with perovskite structure, Phys.
Rev. 82, pp. 403–405.
Zhang, F. M., Liu, X. C., Gao, J., Wu, X. S., Du, Y. W., Zhu, H., Xiao, J. Q. and
Chen, P. (2004). Investigation on the magnetic and electrical properties of
crystalline Mn0.05 Si0.95 films, Appl. Phys. Lett. 85, pp. 786–788.
Zhang, W., Jiko, N., Mibu, K. and Yoshimura, K. (2005). Effect of substitution
of Mn with Fe or Cr in Heusler alloy of Co2 MnSn, J. Phys.: Condens. Mat.
17, p. 6653.
Zhang, Z. and Satpathy, S. (1991). Electron states, magnetism, and the Verwey
transition in magnetite, Phys. Rev. B 44, pp. 13319–13331.
Zhao, J. H., Matsukura, F., Takamura, K., Abe, E., Chiba, D. and Ohno, H.
(2001). Room-temperature ferromagnetism in zincblende CrSb grown by
molecular-beam epitaxy, Appl. Phys. Lett. 79, pp. 2776–2778.
Zhao, Y.-J., Geng, W. T., Freeman, A. J. and Delley, B. (2002). Structural,
electronic, and magnetic properties of α– and β–MnAs: LDA and GGA
investigations, Phys. Rev. B 65, p. 113202.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 267
Zhu, W., Sinkovic, B., Vescovo, E., Tanaka, C. and Moodera, J. S. (2001). Spin-
resolved density of states at the surface of NiMnSb, Phys. Rev. B 64, p.
060403.
Zhu, W., Zhang, Z. and Kaxiras, E. (2008). Dopant-assisted concentration en-
hancement of substitutional Mn in Si and Ge, Phys. Rev. Lett. 100, p.
027205.
Ziebeck, K. R. A. and Webster, P. J. (1974). A neutron diffraction and magneti-
zation study of Heusler alloys containing Co and Zr, Hf, V or Nb, Journal
of Physics and Chemistry of Solids 35, pp. 1–7.
Ziese, M. and Blythe, H. J. (2000). Magnetoresistance of magnetite, J. Phys.:
Condens. Mat. 12, p. 13.
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Index
269
April 2, 2012 4:12 World Scientific Book - 9in x 6in HalfMetals
Bibliography 271