high-energy spectroscopic study of mn-based magnetic...

49
High-energy spectroscopic study of Mn-based magnetic semiconductors Master Thesis Yoshitaka Osafune Department of Physics, University of Tokyo January, 2006

Upload: ngothu

Post on 01-Jul-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

High-energy spectroscopic study

of Mn-based magnetic

semiconductors

Master Thesis

Yoshitaka Osafune

Department of Physics, University of Tokyo

January, 2006

Contents

1 Introduction 1

2 Principles of high-energy spectroscopy 9

2.1 Photoemission spectroscopy . . . . . . . . . . . . . . . . . . . . . 9

2.2 Resonant photoemission spectroscopy . . . . . . . . . . . . . . . . 12

2.3 X-ray absorption spectroscopy . . . . . . . . . . . . . . . . . . . . 13

2.4 X-ray magnetic circular dichroism . . . . . . . . . . . . . . . . . . 13

3 Experimental 15

4 Chalcopyrite-type magnetic semiconductor MnGeP2 thin films 17

4.1 Physical properties of MnGeP2 . . . . . . . . . . . . . . . . . . . 17

4.2 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . 20

4.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 21

5 Thermally diffused Mn/GaAs (001) thin films in depth profile 27

5.1 Physical properties of thermally diffused

Mn/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

5.2 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . 31

5.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 33

6 Summary 39

i

Chapter 1

Introduction

It is well known that electrons have the charge degrees of freedom correspond-

ing to the electric properties and the spin degrees of freedom corresponding to

the magnetic ones. These two characteristics have been separately applied to

various devices so far. For instance, the charges of electrons have been utilized

in semiconductor electronics such as integrated circuits, transistors, and lasers.

On the other hand, the spins of electrons have been utilized in magnetics such

as hard disks. Recently significant movement toward the coupling between the

charge and spin of electron has occurred and a new field, called “spintronics”

(spin electronics). The fascination of spintronics is that spintronics devices hold

the possibilities to show properties which have never been achieved by the in-

dependent use of devices in semiconductor electronics or magnetics. Diluted

magnetic semiconductors (DMS’s), where magnetic ions are doped into the non-

magnetic host semiconductors, are promising candidates for the application of

spintronics. In the early 1980s, II-VI-based DMS’s such as Cd1−xMnxTe have

been extensively studied, because Mn2+ ions substitute isovalently the II-site ions

in II-VI-based DMS’s, which enables high concentration of Mn doping. However,

most of II-VI-based DMS’s have few carriers, thus superexchange interaction

between the magnetic moments becomes predominant and show paramagnetic,

anti-ferromagnetic, or spin glass behavior. Although Cd1−xMnxTe is now used as

optical isolators [1], this has not been sufficient to demonstrate the potentiality

of spintronics.

The prosperity of spintronics has started by the breakthrough of crystal

growth technique in the late 1980s, that is, low temperature molecular beam

epitaxy (MBE), which is a sophisticated version of vacuum deposition and en-

ables us to grow single crystal sample epitaxially on a substrate in an ultrahigh

vacuum (UHV) by evaporating independently each element from each crucible.

1

Note that since this method is a non-equilibrium thermal growth technique, one

can incorporate transition metals into host semiconductors beyond its solubility

limit. Thus, III-V-based DMS’s, where solubility limit in equilibrium state is low

because heterovalent substitution of divalent ions for III-site ions takes place, has

attracted considerable interest. Using MBE method, Munekata et al. succeeded

in the synthesis of In1−xMnxAs [2], and Ohno et al. Ga1−xMnxAs [3]. Different

from II-VI-based DMS’s, they showed ferromagnetism, which made a big impact

in this research field. Ferromagnetism in III-V-based DMS’s is thought to be

caused by coupling between the local magnetic moments of magnetic ions medi-

ated by the charge carriers of the host semiconductors, that is, “carrier-induced

ferromagnetism”. Such ferromagnetism in III-V-based DMS’s initiates various

interesting properties. As for Ga1−xMnxAs, it was reported that magnetotrans-

port properties such as large negative magnetoresistance [4] and anomalous Hall

effect [5] appeared as shown in Fig. 1.1.

Here, we would like to introduce several proposed heterostructure devices

based on ferromagnetic III-V-based DMS’s. Figure 1.2 shows a light-emitting

p-i-n diode [6]. Under forward bias, the device emits circularly polarized light

through the recombination of spin-polarized holes and spin-unpolarized electrons,

which is the significant evidence of electrical spin injection. Figure 1.3 shows a

(In,Mn)As/GaSb heterostructure where ferromagnetism can be induced by pho-

togenerated carriers. The electron-hole pairs produced by the light irradiation

are split by the internal electric field, causing the holes to accumulate in the

(In,Mn)As layer at the surface. Since (In,Mn)As exhibits hole-induced ferromag-

netism, phase transition from paramagnetic state to ferromagnetic state occurs

in the device after light irradiation, as depicted in Fig. 1.3 (b), (c).

Search for high-TC ferromagnetic DMS’s, in particular, room-temperature fer-

romagnetic DMS’s have been heated up because the the range of application is

expected to be widely extended. Dietl et al. predicted the T C’s of various DMS’s

using mean-field approximation [8] as shown in Fig. 1.5(a). Inspired by this pre-

diction, many room-temperature ferromagnetic DMS’s have been fabricated such

as ZnO:Co [9], ZnV:O [10], GaN:Mn [11], GaN:Cr [12].

Recently, Mn-based II-IV-V2 chalcopyrite-type DMS’s showed ferromagnetism

above the room-temperature, and have attracted much attention of many re-

searchers. For instance, it has been reported that the T C’s of ZnGeP2:Mn and

CdGeP2:Mn are ∼350 K and ∼320 K, respectively, well above the room tem-

perature, as shown in Fig. 1.4(a)-(c) [13, 14]. In the II-IV-V2 DMS’s, isovalent

substitution of Mn2+ for the II-site ions become possible, which enables high

concentration Mn-doping. Carriers are expected to be doped by various kinds of

2

(a) (b)

(c)

(d)

Figure 1.1: Magnetotransport properties of Ga1−xMnxAs. (a) Temperature dependence ofresistivity for samples with x ranging from 0.015 to 0.071 [4]. (b) Magnetoresistance for asample with x = 0.071 at different temperatures ranging from 1.4 to 250 K [4]. (c) Magne-toresistance for a sample with x = 0.035 at different temperatures from 1.7 to 200 K [4]. (d)Dominant contribution of the anomalous Hall effect to the Hall coefficient RH (ρ/RH), whichis proportional to (T − TC), for a sample with x = 0.06 [5].

3

(a)

(b)

Figure 1.2: Electrical spin injection [6]. (a) Schematic diagram of device structure. Underforward bias, spin-polarized holes (h+) from p-(Ga,Mn)As recombine with spin-unpolarizedelectrons from n-GaAs substrate, and as a result, circularly polarized light (σ+) is emittedfrom the edge of the (In,Ga)As quantum well. (b) Dependence of the polarization ∆P of theemitted light on the magnetic field at each temperature.

(a)(b)

(c)

Figure 1.3: Photoinduced ferromagnetism [7]. (a) Device structure of (In,Mn)As/GaSb. (b)M -H curves at 5 K observed before (open circles) and after (solid circles) light irradiation.Solid line shows a theoretical curve. (c) Hall resistivity observed at 5 K before (dashed line)and after (solid line) light irradiation.

4

(a) (b)

(c) (d)

Figure 1.4: Magnetization curves of Mn-based II-IV-V2 chalcopyrite-type magnetic semicon-ductors. (a) M -H curves of CdGeP2:Mn at T = 300 K for magnetic field in two orientations(solid curves: in plane, open circles: perpendicular) [13]. (b) M -T curve of CdGeP2:Mn at H= 0 T [13]. (c) M -H curve of ZnGeP2:Mn at T = 350 K. The inset shows an enlarged plotindicating a hysteresis behavior [14]. (d) M -T curves of MnGeP2 and MnGeAs2 at H = 0.01T [15].

5

defects such as antisite IIIV ions and vacancies VII, VIV, and VV. Recent stud-

ies have revealed that 100 % Mn substitution in ZnGeP2, i.e., MnGeP2 can be

synthesized and shows p-type semiconducting properties (see Section 4.1) and

ferromagnetism at room-temperature as shown in Fig. 1.4(d) [15]. MnGeP2 also

shows magneto-optical effect such as Kerr effect [16] and application to various

magneto-optical devices can be considered.

Figure 1.5: Theoretical predictions of TC for various p-type semiconductors containing 5 %of Mn and 3.4 × 1020 holes per cm3 [8].

As the fabrication technique of DMS’s, thermal diffusion method as well as

MBE method has been in use. As a matter of fact, ZnGeP2:Mn and CdGeP2:Mn

mentioned above have been fabricated by thermal diffusion method [13, 14].

Schematic illustration of thermal diffusion is described in Section 5.1. Taking Mn

on ZnGeP2 substrate as an example, Mn is thermally diffused and reacted with

ZnGeP2 substrate by annealing the substrate. Then, by removing the surface

metallic Mn layer in some way such as noble gas ion sputtering, one can fabri-

cate the DMS ZnGeP2:Mn. In order to elucidate the origin of the ferromagnetism

of ZnGeP2:Mn, detailed electronic structure along the depth direction has been

provided by depth profile study using photoemission spectroscopy (PES) [17].

Mn doping into GaN thin film has also been achieved by the thermal diffusion

method [18, 19]. As noted above, Ga1−xMxAs (synthesized by MBE method)

is a remarkable DMS, and shows various magnetotransport properties. We at-

tempted to fabricate thermally diffused Mn/GaAs and compare the electronic

structure of it with that of Ga1−xMnxAs.

In this thesis, we have studied MnGeP2 thin films and thermally diffused

Mn/GaAs (001) thin films using high-energy spectroscopic method. As for

MnGeP2 thin films, we have focused on the elucidation of the electronic struc-

6

ture and magnetic properties. Photoemission spectroscopy (PES) is a powerful

technique to elucidate the core-level and valence-band electronic structure. X-

ray absorption spectroscopy (XAS) is also a useful technique to investigate the

electronic structure, and x-ray magnetic circular dichroism (XMCD) measure-

ments provide us with the information about element-specific spin and orbital

magnetic moments. We have adopted combined PES, XAS, and XMCD. As for

thermally diffused Mn/GaAs (001) thin films, we have combined PES with Ar+-

ion sputtering, which enabled us to perform depth profile analysis, which is a

suitable method for studying the electronic structure along the depth direction

and diffusion effect.

The present thesis is organized as follows. First, the principles of high-energy

spectroscopic method are explained in Chapter 2. The experimental setup is

described in Chapter 3. Detailed physical properties and experimental results

of MnGeP2 thin films and thermally diffused Mn/GaAs (001) thin films are

presented in Chapters 4 and 5, respectively. Finally, Chapter 6 is devoted to

summary.

7

Chapter 2

Principles of high-energy

spectroscopy

In this chapter, we describe the principles of high-energy spectroscopy, that is,

photoemission spectroscopy, resonant photoemission spectroscopy, x-ray absorp-

tion spectroscopy, and x-ray magnetic circular dichroism.

2.1 Photoemission spectroscopy

Photoemission spectroscopy (PES) is a useful technique to investigate the occu-

pied electronic structure in solids. Photoelectric effect, where an illuminated solid

emits electrons (photo-electrons), is utilized in PES. PES spectra are provided by

the measuring kinetic energy distribution of photo-electrons which escape from

the solid through the surface and overcome the vacuum level EV . Assuming that

when light with photon energy hν is illuminated on a solid, an electron with the

binding energy EB, which is referenced to the Fermi level EF, is emited with the

kinetic energy EVkin, which is referenced to EV . In this case, we can describe the

relationship between EB and EVkin using the energy conservation law:

EVkin = hν − φ −EB , (2.1)

where φ = EV − EF is the work function of the solid. The kinetic energy Ekin

which is reference to EF is practically observed in PES measurements, thus the

notation is rather simplified by

Ekin = hν − EB. (2.2)

Schematic diagram of PES is shown in Fig. 2.1. The density of states (DOS)

9

EB

Ekin hνphoton

photo-electron

EF

EV

E

DOS (N(E))

φ

EkinV

valence band

core levelIntensity (I(E))

Ekin

EF

PES spectra

Figure 2.1: Schematic diagram of photoemission spectroscopy. The density of states N(E) isobtained by measuring the photoemission spectra I(E).

N(E) is obtained by measuring the photoemission spectra I(E), which are broad-

ened by the resolution of light source and electron energy analyzer. In one-

electron approximation, the binding energy is equal to the negative Hartree-Fock

orbital energy with Bloch wave number k,

EB = −εk. (2.3)

Here, Koopmans’ theorem is used [20]. This assumption is valid when the wave

functions of both the initial and final states can be expressed by the single Slater

determinants of the n- and (n − 1)-electron systems, respectively, and the one-

electron wave functions do not change by the removal of the electron. If we apply

this approximation, the photoemission spectrum I(EB) can be expressed as:

I(EB) ∝∑

k

δ(EB + εk) ∝ N(−EB). (2.4)

Thus, when the one-electron approximation is valid, the photoemission spectrum

is proportional to the density of states of the occupied one-electron states N(E).

If the electron correlation effect is taken into account, one can no more con-

sider the electron system within the one-electron picture, because the relaxation

influences the photoemission final state such as screening of photo-holes by va-

lence electron. Thus, the energy difference between the n-electrons initial state

energy Eni and the (n− 1)-electrons final state energy En−1

f provides the binding

energy EB, that is,

10

Figure 2.2: Mean free path of electrons in solids as a function of electron energy. Dashedcurves indicate the approximate range of distribution [21].

EB = En−1f −En

i . (2.5)

Using Fermi’s golden rules, the PES spectrum, which now corresponds to the

single-particle excitation spectrum of the electron system, is expressed as:

I(EB) ∝∑

k

|〈Ψn−1f |ak|Ψn

i 〉|2δ[EB − (En−1f −En

i )], (2.6)

where Ψn−1f and Ψn

i denote the final and initial states, respectively, ak is the

annihilation operator of the electron occupying orbital k. Considering electron

correlation effect, the finite lifetime of quasi-particle also contributes to the spec-

tral broadening.

Figure 2.2 shows the energy dependence of the mean free path of electrons in

solids. The escape depth of electrons is described by the universal curve, roughly

independent of the material. Around the electron energy of 20-1000 eV, where

we perform the PES measurements, the escape depth is 5-10 A. This suggests

that PES measurement is quite surface-sensitive. Therefore, we should always

keep it in mind to eliminate the surface effect.

11

E

EF

final state

E

EF

final state

E

EF

direct photoemission

Fanoresonance

Auger decay

p6dn + hν p5dn+1 p5dn+1 p6dn-1 + e-

p6dn + hν p6dn-1 + e-

3p 3d absorption

3pvalence

EF

E

initial state

Figure 2.3: Schematic diagram of resonant photoemission spectroscopy. The case of 3p → 3dabsorption is taken as an example.

2.2 Resonant photoemission spectroscopy

Resonant photoemission spectroscopy (RPES) is an effective approach to extract

the PES spectrum for an impurity atom from the entire spectrum in the valence

band. RPES measurement is achieved by synchrotron radiation, where photon

energy hν can be continuously varied. Let us take 3p → 3d absorption as an

example. Schematic diagram of RPES is shown in Fig. 2.3. The direct PES

process of a valence 3d electron is described as:

p6dn + hν → p6dn−1 + e−. (2.7)

When the photon energy is equal to the absorption energy from the 3p core level

to the valence 3d state, 3p → 3d absorption and subsequent Auger decay, called

super Coster-Kronig decay occur.

p6dn + hν → p5dn+1 → p6dn−1 + e−. (2.8)

The final states of these two processes are the same electronic configurations,

and quantum-mechanically interfere with each other in consequence. Thus, the

photoemission intensity is resonantly enhanced and shows a so-called Fano profile

[22]. This enhancement helps detecting weak signals such as photoemission from

transition metal impurities in the valence band, which is difficult to obtain by

normal PES.

12

2.3 X-ray absorption spectroscopy

X-ray absorption spectroscopy (XAS) is a powerful technique to investigate the

unoccupied electronic structure in solids. The photo-absorption intensity by ex-

citation of a core-level electron into unoccupied states as a function of photon

energy hν is given by

Iµ(hν) ∝∑

f

|〈Ψf |Tµ|Ψi〉|2δ(Ef −Ei − hν), (2.9)

where T is the dipole transition operator, µ is the index of light polarization, and

Ei and Ef are the energies of the initial and final state, respectively. In the 3d

transition-metal compounds, transition-metal 2p (L2,3-edge) XAS spectra reflect

the electronic structure of the 3d states such as the spin state and the crystal-

field splitting. In order to interpret experimental spectra, various theoretical

calculations have been applied.

The measurement modes for XAS can be classified broadly into the trans-

mission mode and the total electron-yield mode. In the transmission mode, the

intensity of the x-ray is measure before and after the sample and the ratio of the

transmitted x-rays is counted. Transmission-mode experiments are standard for

hard x-rays, while for soft x-rays, they are difficult to perform because of the

strong interaction of soft x-rays with the sample and hence strong absorption.

In the present work, we have adopted the total electron-yield mode, because all

measurements have been performed in the region of soft x-rays.

2.4 X-ray magnetic circular dichroism

If circularly polarized light is used in XAS, the absorption intensity of magnetic

materials depends on the helicity of the incident light. This phenomenon has

been utilized in x-ray magnetic circular dichroism (XMCD), which is defined

as the difference in absorption spectra between right- and left-handed circularly

polarized x-rays when the helicity of x-rays are parallel and antiparallel to the

magnetization direction of the magnetic materials with a magnetic field.

The characteristic features of XMCD measurements are as follows.

• If the absorption region of an element does not overlap with other ab-

sorption region, we can study element-specific magnetic moments, which

enables us to investigate precisely the magnetism of particular orbitals of

each element.

13

• XMCD reflects the orbital and spin polarization of local electronic states.

Thus, using integrated intensity of the L2,3-edge XAS and XMCD spectra

of a transition-metal atom, one can separately estimate the values of orbital

[23] and spin [24] magnetic moments by applying XMCD sum rules.

XMCD sum rules provide the values of orbital and spin magnetic moments using

the following formulae,

Morb = −4∫

L3+L2(µ+ − µ−)dω

3∫

L3+L2(µ+ + µ−)dω

(10 − nd). (2.10)

M spin + 7MT = −6∫

L3(µ+ − µ−)dω − 4

∫L3+L2

(µ+ − µ−)dω∫

L3+L2(µ+ + µ−)dω

(10 − nd), (2.11)

where Morb and M spin are the orbital and spin magnetic moments in units of

µB/atom, respectively, µ+(µ−) is the absorption intensity for the positive (neg-

ative) helicity, nd is the d electron occupation number of the specific transition-

metal atom. L3 and L2 denote the integration range. MT is the expectation

value of the magnetic dipole operator, which is small when the local symmetry

of the transition-metal atomic site is high and is neglected here with respect to

M spin.

14

Chapter 3

Experimental

Ultraviolet photoemission spectroscopy (UPS) measurements were performed at

beamline BL-18A of Photon Factory (PF), Institute for Material Structure Sci-

ence, High Energy Accelerator Research Organization (KEK) using a VG CLAM

hemispherical analyzer. All the photoemission spectra were taken at room tem-

perature under an ultra high vacuum of 5.0 × 10−10 Torr. The Fermi level (EF)

was calibrated by the Fermi edge of a Cu metal in electrical contact with the sam-

ple. X-ray photoemission spectroscopy (XPS) measurements were performed at

University of Tokyo using a Gammadata-Scienta SES-100 hemispherical analyzer

for MnGeP2 and VSW125 analyzer for thermally diffused Mn/GaAs (001). The

total energy resolution was estimated to be ∼200 meV for UPS and ∼800 meV for

XPS including temperature broadening. In both UPS and XPS measurements,

photoelectrons were collected in the angle-integrated mode. XAS and XMCD

measurements at the Mn 2p → 3d (Mn L2,3) edge were performed at BL-11A of

PF. The XAS spectra were taken by the total electron yield mode with magnetic

fields applied perpendicular to the sample plane under an ultra high vacuum of

2.0 × 10−9 Torr. X-ray absorption spectra for right-handed (µ+) and left-handed

(µ−) circularly polarized x-rays were obtained by reversing the direction of mag-

netization. The difference between the µ+ and µ− spectra yields XMCD spectra.

Depth profile studies were performed by repeated cycles of sputter-etching and

subsequent PES measurement. Sputter-etching was done with Ar+-ion at 1.0 kV.

Here, it is necessary to eliminate possible surface effects such as oxidation of

the surface in ex situ treatment of the sample.

MnGeP2 thin films for the XAS, XMCD, and XPS measurements were coated

with a Ge-cap layer (∼3 nm). As for the UPS measurements, which are quite

surface-sensitive, the uncapped sample was used and cleaned repeatedly by Ar+-

ion sputtering at 0.8 kV with a vacuum of 4.0 × 10−5 Torr. Thermally diffused

15

Mn/GaAs (001) thin films were coated with an As-cap layer.

(a)

(b)

Figure 3.1: Schematic optical layout of (a) BL-18A at PF and (b) BL-11A at PF.

Manipulator

Photon source

Electron energy analyzer

Sample

Ion gun

Photo-electron

Ar+-ion

Figure 3.2: Schematic drawing of the experimental apparatus used for the photoemissionstudy.

16

Chapter 4

Chalcopyrite-type magnetic

semiconductor MnGeP2 thin

films

4.1 Physical properties of MnGeP2

MnGeP2 is a II-IV-V2 chalcopyrite-type room-temperature ferromagnetic semi-

conductor. The crystal structure of MnGeP2 is shown in Fig. 4.1(a). The lat-

tice parameters for MnGeP2 were estimated as a = 5.693 A, and c = 11.303

A and α = β = γ = 90 by reciprocal lattice mapping of x-ray diffraction (XRD)

shown in Fig. 4.1(b) [16], which is consistent with the experimental value of poly-

crystalline MnGeP2 [15]. Recently, Cho et al. have succeeded in synthesizing

MnGeP2 and MnGeAs2 thin films using MBE method and measured resistance

and Hall resistance as shown in Fig. 4.2(a), (b) [15]. The resistivity of MnGeP2

at room-temperature is ∼10−3 Ωcm, which is rather small for a semiconductor.

Hall measurements revealed that the carrier type of MnGeP2 and MnGeAs2 is

p- and n-type, respectively, with carrier concentration ∼1019 cm−3. Based on

the fact that various native defects such as group II vacancies and antisite de-

fects are present in II-IV-V2 chalcopyrites with carrier concentrations up to 1019

cm−3 [25, 26], p-type behavior of MnGeP2 may arise from point defects such as

cation Mn and Ge vacancies (VMn, VGe) and antisite defects MnGe. Anomalous

Hall effect was observed as shown in Fig. 4.2(b), suggesting the spin polarized

hole carriers exist in MnGeP2. In order to confirm the electrical properties, a p-n

junction has been fabricated by use of 100 A p-MnGeP2 and 100 A n-MnGeAs2

and measured I-V characteristics as shown in Fig. 4.2(c). Clear diode character-

istic was observed, indicating that MnGeP2 is certainly a p-type semiconductor.

17

Theoretical studies have also been carried out using full-potential linearized

augmented plane wave (FLAPW) method [27] in the local density approximation

(LDA). For instance, Zhao et al. derived the lattice constants of MnGeP2 as a =

5.673 A, c = 10.716 A [28], which is consistent with the value deduced by Sato

et al. [16]. Cho et al. applied total energy calculation to MnGeP2, revealing that

MnGeP2 is a semiconductor with an energy gap of ∼0.24 eV.

It is reported that MnP cluster was observed in polycrystalline Zn1−xMnxGeP2

by nuclear magnetic resonance measurement [29], and therefore we must be con-

cerned about the formation of MnP secondary phase in MnGeP2. MnP, which

is distorted from the NiAs-type structure with a = 5.916 A, b = 5.260 A, and c

= 3.173 A, is a ferromagnetic metal with T C ∼293 K [30, 31]. Hence, MnP also

contributes to the room-temperature ferromagnetism, however, a phase transi-

tion from the ferromagnetic to an antiferromagnetic state occurs at T N ∼47 K.

Therefore, the presence or the absence of MnP can be checked by investigating

the magnetization around T N.

18

Mn Ge P

a

c

a = 5.693 [A]

c = 11.303 [A]

(a) (b)

Figure 4.1: Chalcopyrite-type MnGeP2. (a) Crystal structure. (b) Reciprocal lattice mappingof x-ray diffraction for MnGeP2 grown on an InP (001) substrate [16], which provide the latticeconstants.

(c)

(a) (b)

(d)

Figure 4.2: Several physical properties and theoretical calculation for MnGeP2 and MnGeAs2thin films [15]. (a) Temperature dependence of the electrical resistances of MnGeP2 andMnGeAs2 thin films grown on GaAs (001) substrates. (b) Anomalous Hall resistances ofp-MnGeP2 and n-MnGeAs2 thin films. (c) I-V diode characteristics of a junction between 100A p-MnGeP2 and 100 A n-MnGeAs2 thin films grown on a Si (001) substrate. (d) Theoreticalcalculation of density of states (DOS) by the full-potential linearized augmented plane wave(FLAPW) method [27] in the local density approximation (LDA).

19

4.2 Sample preparation

Two single-crystal MnGeP2 thin films were fabricated by the molecular beam

epitaxy (MBE) method. One was a Ge-capped sample, and the other was without

capping. These samples were epitaxially grown on a Ge buffer layer at 435 C

which had been deposited on a GaAs (001) substrate at 380 C. It has been

reported that the Ge buffer layer enables a two-dimentional growth of MnGeP2

and improves the crystallinity of the sample [32]. We have done magnetization

measurements using a SQUID magnetometer (MPMS, Quantum Design, Co.,

Ltd.) on the MnGeP2 films prior to the spectroscopic measurements. Figure

5.9 shows magnetization curves for a magnetic field perpendicular to the sample

plane. The sample showed T C ∼320 K and thus was confirmed to be a room-

temperature ferromagnet as in the previous report [15]. Additionally the M-T

curve indicated that there was a Curie-Weiss component in the sample. This fact

suggested that a paramagnetic component existed in the sample. Details of the

sample fabrication are given in Ref. [32].

-1.5

-1.0

-0.5

0

0.5

1.0

1.5

Mag

netiz

atio

n ( µ

B/ M

n)

-2 -1 0 1 2Magnetic Field (T)

10K 100K 310K 330K

(a)

MnGeP2

0.8

0.6

0.4

0.2

0Mag

netiz

atio

n ( µ

B/ M

n)

300200100Temperature (K)

H = 0.01 T

(b)

Figure 4.3: Magnetization data of MnGeP2 thin film. (a) M -H curves at T = 10 K, 100 K,310 K, and 330 K. The linear component in the high magnetic field region has been subtracted.(b) M -T curve at H = 0.01 T. We obtained the M -H curve as follows. The sample was firstcooled down to 5 K without applying the magnetic field. Then the magnetic field was appliedperpendicular to the sample plane up to 1.0 T, and set it at 0.01 T and we started to measurethe sample with increasing temperature.

20

4.3 Results and discussion

Inte

nsit

y (a

rb. u

nits

)

15 10 5 0 -5Binding Energy (eV)

Mn Auger

h =60

57

55

54

53

52

515049484746 eV

difference (51 – 48 eV)

(a)MnGeP2

CIS

Int

ensi

ty (

arb.

uni

ts)

70656055504540Photon Energy (eV)

EB =15 eV

9.5 eV

7.5 eV

5.3 eV

2.6 eV

0.7 eV

(b)

Inte

nsit

y (a

rb. u

nits

)

12 8 4 0Binding Energy (eV)

MnGeP2

MnP

Ga0.931Mn0.069As

difference (51 eV – 48 eV)(c)

-1.0-0.500.51.0

Ga0.931Mn0.069As

MnP

MnGeP2

Figure 4.4: A series of valence-band spectra taken at various photon energies in the Mn 3p→ 3d core-excitation region. (a) Energy distribution curves. Vertical bars indicate a constantkinetic energy and represent the threshold (lower EB’s) and the peak position (higher EB’s) ofthe Mn Auger signals. The difference between the on-resonant (hν = 51 eV) and off-resonant(hν = 48 eV) spectra, which approximately represents the Mn 3d partial density of states, isshown at the bottom. (b) Constant-initial-states spectra at various binding energies. Arrowsshow the threshold and peak position of the Mn Auger signals. (c) Comparison of the difference(hν = 51−48 eV) spectra between MnP [33], Ga0.931Mn0.069As [34], and MnGeP2. Inset showsan enlarged plot around the Fermi level.

Figure 4.4(a) shows the valence-band spectra of the uncapped MnGeP2 thin

film taken at various photon energies in the Mn 3p→ 3d core-excitation region.

The intensities have been normalized to the photon flux. The vertical bars rep-

resent the threshold and the peak position of the Mn M2,3L4,5L4,5 Auger signals,

which is quite strong, indicating that there exist the itinerant nature of the Mn

3d states in MnGeP2. The intensity near EF was high as shown in Fig. 4.4(a),

but there was no Fermi edge in MnGeP2 as shown in Fig. 4.4(a). In contrast, a

clear Fermi edge has been observed in single crystal MnP [35]. Therefore, one can

conclude that MnP is not a dominant component of the sample. The resonantly

enhanced Mn 3d partial density of states (PDOS) was obtained by subtracting

the off-resonant (hν = 48 eV) spectrum from the on-resonant (hν = 51 eV) one as

shown at the bottom of Fig. 4.4(a). In order to properly carry out the subtraction,

we consider the photon energy dependence of the photoionization cross-section

of Ge 4p and P 3p. The difference spectrum showed a peak at binding energy

(EB) ∼2.6 eV. Another broad peak which appeared around EB = 6-15 eV can

be attributed to a charge-transfer satellite as in the case of Ga1−xMnxAs [34].

21

The existence of the charge-transfer satellite indicates that there exist strong

Coulomb interaction between the Mn 3d electrons and strong hybridization be-

tween the Mn 3d and other valence orbitals. However, it should be noted that

Mn M2,3L4,5L4,5 Auger signals exist in this region and overlaps with the satellite.

Figure 4.4(b) shows the constant-initial-state (CIS) spectra for various EB ’s in

the range of hν = 40-70 eV. Arrows represent the threshold and the peak posi-

tion of the Mn M2,3L4,5L4,5 Auger signals. All the CIS’s showed a peak at hν

∼51 eV (shown by a dashed line), whereas Mn M2,3L4,5L4,5 Auger overlaped with

the 51-eV peak for EB = 5-11 eV. Figure 4.4(c) shows comparison of the differ-

ence (hν = 51 − 48 eV) spectra between MnP [33], Ga0.931Mn0.069As [34], and

MnGeP2. The difference spectra of MnP show broader structure at EB = 1-4 eV

than that of MnGeP2, and there is a clear Fermi edge in the difference spectrum

of MnP, indicating that the Mn 3d states in MnP exhibit more itinerant behavior

than those in MnGeP2. On the other hand, the main peak of Ga0.931Mn0.069As

is sharper than that of MnGeP2, reflecting more localized nature of the Mn 3d

states in Ga0.931Mn0.069As than those in MnGeP2.

Inte

nsit

y (a

rb. u

nits

)

655 650 645 640 635Binding Energy (eV)

Mn 2p

2p3/22p1/2

(a)A

BA'B'

h = 1486.6 eV

MnGeP2

645 640 635Binding Energy (eV)

Mn 2p3/2

20º 60º

(b)A

B

Figure 4.5: XPS spectrum of the Mn 2p core level for MnGeP2. (a) Entire spectrum. (b)Mn 2p3/2 spectra for the take-off angle of 20 and 60 relative to the sample surface. Theintensity has been normalized to the peak height of structure B. The 20 spectrum was moresurface-sensitive than the 60 one.

Figure 4.5(a) shows the Mn 2p core level XPS spectra taken with the Al

Kα source. The spin-orbit doublet was observed and each spin-orbit component

has a main peak (A, A′) and a strong shoulder structure (B, B′), which is a

charge-transfer satellite located at a higher EB of the main peak. On the other

hand, structure A is fairly sharp, suggesting the itinerant nature of the Mn 3d

states. Figure 4.5(b) shows Mn 2p3/2 spectra for the measurement condition that

22

the sample surface forms an angle of 20 and 60 with the analyzer axis. The

20 measurement was more surface-sensitive than the 60 measurement. The

relative intensity of structure A to structure B in the 20 spectrum was weaker

than that in the 60 spectrum. This indicates that structure A derived not from

the surface region, but from the intrinsic origin of the bulk, meaning that the

itinerant character of the Mn 3d electrons is that of bulk MnGeP2 and not that

of its surface.

Figure 4.6(a), (b), and (c) show circularly polarized XAS spectra (µ+, µ−)

and XMCD spectra (µ+ − µ−) at the Mn L2,3 edge under various experimental

conditions. Circularly polarized XAS spectra have been normalized to the Mn

L3 peak height of the unpolarized XAS spectra [(µ+ + µ−)/2]. XMCD signal

for the remanence (H = 0 T), which represents the ferromagnetic component,

was clearly observed at 30 K, thus one can exclude contributions from MnP,

because a ferromagnetic-to-antiferromagnetic transition occurs at T N ∼47 K in

MnP [30, 31]. In Fig. 4.6(d), we compared the XAS spectra between MnAs [36],

Ga0.94Mn0.06As [36], and MnGeP2. Note that the XAS line shape of MnAs is

broad, which reflects the itinerant nature of the Mn 3d states in MnAs, while that

of Ga0.94Mn0.06As is narrower and exhibits multiplet structures, which reflects the

localized nature of the Mn 3d states. The XAS spectrum of MnGeP2 exhibits a

broad asymmetric line shape, which is similar to that of MnAs, and has multiplet

structures, suggesting that the Mn 3d states have both the itinerant and localized

character in MnGeP2.

In order to decompose the XMCD signals into the paramagnetic and ferro-

magnetic component, we subtracted the XMCD spectra at H = 0 T and T = 30

K (multiplied by the ratio of the saturation magnetization to the residual mag-

netization ∼1.77 at T = 30 K and ∼1.90 at T = 200 K) from those at H = 5.0

T and T = 30 K and 200 K in Fig. 4.7(a). It turns out that the spectral features

of the paramagnetic component at T = 30 K and 200 K [para (30 K), para (200

K)] are quite similar to each other, and are clearly different from those of the

ferromagnetic component at T = 30 K [ferro (30 K)]. The orbital (Morb), spin

(M spin), and total (M tot = Morb + M spin) magnetic moments of the paramag-

netic and ferromagnetic component were estimated by applying the XMCD sum

rules, and plotted as functions of temperature in Fig. 4.7(b). While Morb/M spin

for the paramagnetic component is ∼0.3-0.5, Morb/M spin for the ferromagnetic

component at 30 K is tiny. Therefore, a large orbital magnetic moment is thought

to be realized by the paramagnetic component.

23

-0.20-0.15-0.10-0.0500.05

XM

CD

(arb. units)

660640Photon Energy (eV)

1.00.80.60.40.2

0

Abs

orpt

ion

(arb

. uni

ts)

660640 660640

T = 200 KH = 5.0 T

+

+

T = 30 KH = 5.0 T

+

T = 30 KH = 0 T

+–

(a) (b) (c) Mn L2,3 edgeMnGeP2A

bsor

ptio

n (a

rb. u

nits

)

660655650645640635Photon Energy (eV)

MnGeP2

Ga0.96Mn0.04As

MnAs

Mn L2,3 edge(d)

Figure 4.6: Circularly polarized x-ray absorption spectra (top panels) and XMCD spectra(bottom panels) at the Mn L2,3 edge under various experimental conditions. (a) T = 200 K,H = 5.0 T. (b) T = 30 K, H = 5.0 T. (c) T = 30 K, H = 0 T (remanence). (d) Comparisonof the XAS spectra between MnAs [36], Ga0.94Mn0.06As [36], and MnGeP2.

24

Inte

nsit

y (a

rb. u

nits

)

665660655650645640635630Photon Energy (eV)

(a) MnGeP2

Mn L2,3 edge

ferro (30 K)

para (30 K)

para (200 K)

0.6

0.4

0.2

0

Mag

netic

Mom

ent (

B/M

n)

200150100500Temperature (K)

Morb(para)

Mspin(para)

Mtot(para)

(b)

Morb(ferro, residual)

Mspin(ferro, residual)

Mtot(ferro, residual)

Figure 4.7: Analysis of the XMCD spectra of MnGeP2. (a) Ferromagnetic component at 30K, and paramagnetic component at 30 K, 200 K of the XMCD spectra. The paramagneticcomponent has been deduced by subtracting the ferromagnetic component (H = 0 T, T = 30K) multiplied by the ratio of the saturation magnetization (M sat) to the residual magnetization(M res) from the XMCD spectra at H = 5.0 T, T = 30 K and 200 K. (b) Temperature depen-dence of the magnetic moments M spin, Morb, and M tot = Morb + M spin of the ferromagneticcomponent and the paramagnetic component estimated using the XMCD sum rules. The XASsignals were not decomposed into the paramagnetic and ferromagnetic component, thus thevalue of the vertical axis is rather underestimated.

25

Chapter 5

Thermally diffused Mn/GaAs

(001) thin films in depth profile

5.1 Physical properties of thermally diffused

Mn/GaAs (001)

Figure 5.1 shows a schematic illustration of the Mn incorporation into GaAs (001)

substrate by thermal diffusion. The thermal diffusion procedures can be classified

into “pre-annealing” procedure [Fig. 5.1(a)] and “post-annealing” [Fig. 5.1(b)]

one. We define “pre-annealing” as the procedure where Mn metal is deposited on

the annealed substrate. In contrast, “post-annealing” procedure means that Mn

metal is deposited on the unannealed substrate and then annealed. I would like

to introduce several reports for the thermally diffused Mn/GaAs (001) prepared

by both “pre-annealing” and “post-annealing” procedure.

Dong et al. have performed a depth profile study of thermally diffused Mn/GaAs

(001) thin films prepared by “pre-annealing” procedure using XPS [37]. Figure

5.2 shows relative core-level intensities as functions of sputtering time under var-

ious conditions. Note that there existed Mn diffusion layer, where percentage of

Mn decreased with increasing sputtering time at TS = 300 K, namely, the diffu-

sion occurred even at TS = 300 K (room temperature) as shown in Fig. 5.2(a),

suggesting a large diffusion coefficient. It was found that the metallic Mn region,

where the percentage of Mn was ∼100 % and those of Ga and As were ∼0 %,

that is, deposited Mn was not reacted with the GaAs substrate, existed at TS =

300 and 400 K [Fig. 5.2(a), Fig. 5.2(b)], while metallic Mn region did not exist

at TS = 450 K [Fig. 5.2(c)], indicating that deposited Mn was fully reacted with

the GaAs substrate.

The diffusion coefficients of the thermally diffused Mn(200 nm)/GaAs (001)

27

GaAs (001) substrate

Annealing temperature (T [oC])

deposited Mn

Mn diffusion into bulk

d [A]

GaAs (001) substrate

Substrate temperature (TS [oC])

Mn

Mn diffusion into bulk

d [A]

(b)

(a)

Annealing time (t [min])

Figure 5.1: Schematic illustration of the Mn incorporation into the GaAs (001) substrate bythermal diffusion. The procedure of the thermal diffusion can be classified broadly into twocategories. (a) Depositing Mn metal on the annealed substrate. (b) Depositing Mn metal onthe unannealed substrate and then annealing.

(c)

(b)(a)

Figure 5.2: Relative core-level intensities (Mn 2p, Ga 2p3/2, As 2p3/2) as functions of sput-tering time under various conditions [37]. (a) the thickness of Mn (d) is 12.5 nm, substratetemperature (TS) is 300 K. (b) d = 14.5 nm, TS = 400 K. (c) d = 8.5 nm, TS = 450 K.

28

(e)

(f)

(200 nm)

Figure 5.3: RBS spectra with a 2.3 MeV He+ ion beam for Mn(200 nm)/GaAs (001) thick filmsof as-grown sample(a), and sample with post-annealing up to 300 C for 1(b), 8(c), and 16 h(d)[38]. The simulations of the RBS spectra were done using Rutherford universal manipulationprogram (RUMP) [39]. The dashed curves are calculated spectra, which was derived on theassumption that the composition of the Mn diffusion layer is uniform (Mn0.6Ga0.2As0.2). Theindividual contributions of Mn, Ga, and As of the calculated spectra are also shown. Thevertical dashed lines represent the energies where each element reaches the sample surface. (e)Thickness of the Mn diffusion layer estimated by Fig. 5.3(a)-(d) as a function of square rootof annealing time at 275, 300, and 325 C. The horizontal dashed line represents the point ofcomplete Mn reaction [40]. (f) Temperature dependence of the diffusion coefficients determinedfrom the slope of (e) [40].

29

thick films were obtained by Rutherford backscattering spectroscopy (RBS) [38,

40]. Figure 5.3 shows the RBS spectra for Mn/GaAs (001) of as-grown sample(a),

and sample with post-annealing up to 300 C for 1(b), 8(c), and 16 hours(d).

The calculated spectra can be obtained using Rutherford universal manipulation

program (RUMP) [39] and decomposed into Mn-, Ga-, and As-derived spectra,

and one can see the time evolution of the Mn diffusion from the decomposed

calculated spectra. The thickness of the Mn diffusion layer was estimated by

Fig. 5.3(a)-(d) as a function of square root of annealing time at 275, 300, and

325 C as shown in Fig. 5.3(e). One can determine the temperature dependence

of the diffusion coefficients from the slope of Fig. 5.3(e) as shown in Fig. 5.3(f),

indicating that the diffusion coefficient exponentially increased with increasing

temperature.

30

5.2 Sample preparation

Thermally diffused Mn/GaAs (001) thin films were fabricated by the MBE method

as follows. First, an ordered GaAs buffer layer was grown on an epi-ready Si-

doped n+-GaAs (001) substrate with carrier concentration ∼1020 cm−3. The

GaAs buffer layer was obtained by heating the GaAs substrate up to 600 C

during the Ga and As depositions. Then, Mn metal was deposited on the GaAs

buffer layer at the rate of 0.029 nm/s as thick as ∼2 nm. The growth temper-

ature of Mn on the GaAs substrate was set at 50 C. After the Mn deposition,

Mn diffusion into the GaAs substrate was achieved by post-annealing the sam-

ple up to 600 C for 10 min. The growth procedure is illustrated in Fig. 5.4(a).

Surface reconstruction was monitored during the growth and the post-annealing

with reflection high-energy electron diffraction (RHEED), which showed (2×2)

pattern after the post-annealing as shown in Fig. 5.4(b). We prepared uncapped

samples of the same growth condition for PES to study the surface morphology

of the Mn/GaAs (001) thin film. Figure 5.5(a)-(c) shows the surface morphology

of the as-grown and thermally diffused (300 C, 600 C) Mn/GaAs (001) thin

films measured by atomic force microscopy (AFM). The AFM images suggested

that the level of the surface roughness decreased with increasing temperature. In

particular, a fairly flat surface was obtained for the thermally diffused (600 C)

Mn/GaAs (001) thin film.

31

(a)

(b)

Figure 5.4: The growth of thermally diffused Mn/GaAs (001) thin films. (a) The growthprocedure. (b) RHHED patterns in various growth steps. (2×2) surface reconstruction wasobserved down to 150 C.

(a) as-grown (b) 300 oC for 10 min (c) 600 oC for 10 min

Mn/GaAs (001)

Figure 5.5: Atomic force microscopy images for (a) as-grown sample. (b) Sample with post-annealing up to 300 C for 10 min. (c) Sample with post-annealing up to 600 C for 10 min.

32

5.3 Results and discussion

Figure 5.6(a) shows the core-level XPS spectra in a sputter-etching series, and

Fig. 5.6(b) shows their relative core-level intensities as functions of sputtering

time. The Mn 2p peak completely disappeared at t = 210 min, thus we consider

that Mn diffusion layer disappeared and the GaAs substrate appeared at t = 210

min. At this stage we set the ratio of Ga 2p3/2 to As 2p3/2 at 1.0. Throughout

the sputter-etching series, the Mn 2p spectra showed non-metallic signals, which

indicated that the Mn layer entirely reacted with the GaAs substrate and that

Mn was not metallic even in the first surface layer. In the early stage of the

sputtering (t = 0-10 min), one can see As-excess composition and a high EB of

the Mn 2p3/2 peak positions, which may be both attributed to the existence of

As-cap layer. One can also see that the As intensity rapidly decreased while the

Mn 2p3/2 position rapidly shifted to lower EB ’s for t = 0-20 min, indicating the

removal of the As-cap layer. We suspect that some kind of unexpected reaction

between As-cap layer and Mn layer took place in the surface region. We therefore

conclude that the peak energy shift of the Mn 2p3/2 core-level can be explained as

a change of the chemical species from the Mn-As compound found in the surface

region to the Mn diffusion layer. The ratio of the Ga 2p3/2 to As 2p3/2 intensities

became ∼1 after 20 min sputtering, and therefore the As-cap layer was thought

to be removed and the Mn diffusion layer appeared from this point. Except for

the surface region, the intensities of the Mn 2p3/2 changed so slowly that Mn

diffused into deep region where a dilute Mn phase existed. This is consistent

with the large diffusion coefficient for thermally diffused Mn/GaAs reported by

the Rutherford backscattering studies (see Section 5.1) [38,40].

Figure 5.6(c) shows the Mn 2p spectrum of Mn/GaAs (001) sputtered for

40 min, compared to that of Ga0.926Mn0.074As [41]. As for Mn/GaAs (001), the

spin-orbit doublet was observed and each spin-orbit component had a main peak

and a strong shoulder structure, which is a charge-transfer satellite located at

a higher EB of the main peak. The existence of the charge-transfer satellite

indicates that there are strong Coulomb interaction between the Mn 3d electrons

and strong hybridization between the Mn 3d and other valence orbitals. This

structure is quite similar to that of Ga0.926Mn0.074As (shown by vertical bars in

Fig. 5.6(c)), and therefore the Mn 3d state is expected to be basically localized

in the dilute Mn phase.

Figure 5.7 shows a series of valence-band spectra for various sputtering time

taken at various photon energies in the Mn 3p→ 3d core-excitation region. The

intensities have been normalized to the photon flux. The vertical bars represent

33

Inte

nsit

y (a

rb. u

nits

)

660 655 650 645 640 635

Mn 2p

t(min) = 0

5

10152030406090150210

(a)Mn/GaAs (001)

45 42 39

As 3d

1326 1322

As 2p3/2

1120 1115

Ga 2p3/2

21 19 17

Ga 3d

Inte

nsit

y (a

rb. u

nits

)

660 650 640Binding Energy (eV)

Ga0.926Mn0.074As

Mn/GaAs (001) (t = 40 min)

Mn 2p(c) 2p3/2

2p1/2Mn Auger1.0

0.8

0.6

0.4

0.2

0Rel

ativ

e C

hem

ical

Com

posi

tion

200150100500Sputtering Time (min)

(b)As 2p3/2

Ga 2p3/2

Mn 2p3/2×5

Mn/GaAs (001)

1.0

0.5

020151050

As 2p3/2

Ga 2p3/2

Mn 2p3/2×5

Binding Energy (eV)

Figure 5.6: Core-level XPS spectra in the sputter-etching series. (a) Mn 2p, Ga 2p3/2, As2p3/2, Ga 3d, and As 3d spectra. t(min) denotes the sputtering time. (b) Relative core-levelintensities as functions of sputtering time. Inset shows an enlarged plot for t = 0-20 min. Itshould be noted that the relative chemical composition of Mn 2p3/2 is rather approximate. (c)Comparison of the Mn 2p core-level spectra between Mn/GaAs (001) sputtered for 40 min andGa0.926Mn0.074As [41].

34

Inte

nsity

(ar

b. u

nits

)

15 10 5 0

difference (50 – 48 eV)

t = 0 (min)

47

48

495051

52

55 eV

Mn/GaAs (001)

h =

(a)

15 10 5 0

difference(50 – 48 eV)

t = 5 (min)

h =

47

484950

51

52

55 eV

As Auger

(b)

15 10 5 0

difference(50 – 48 eV)

t = 10 (min)

h =

47

48

49

50

51

52

55 eV

As Auger

(c)

15 10 5 0

t = 85 (min)

difference(50 – 48 eV)

h =

47

48

49

50

51

52

55 eV

As Auger

(d)

15 10 5 0

(e)

h =

55 eV53

51

5049

48

46

As Auger

difference(50 – 48 eV)

Ga0.931Mn0.069As

Binding Energy (eV)

Figure 5.7: A series of valence-band spectra in the sputter-etching series of as-grown(a), 5 min-sputtered(b), 10 min-sputtered(c), 85 min-sputtered Mn/GaAs (001)(d), andGa0.931Mn0.069As(e) [34] taken at various photon energies in the Mn 3p→ 3d core-excitationregion. Vertical bars represent the As Auger signal. The difference between the on-resonant(hν = 50 eV) and off-resonant (hν = 48 eV) spectra, which approximately represents the Mn3d partial density of states, is shown at the bottom of each panel.

Inte

nsity

(ar

b. u

nits

)

15 10 5 0

0

t(min)

5

=

10152535455585

h = 51 eV

(a)

Mn/GaAs (001)

12 8 4 0

difference@10 min (50 – 48 eV)

difference@51 eV (t = 10 – 85 min)

(b)

difference@51 eV (t = 15 – 85 min)

Binding Energy (eV)

Figure 5.8: Comparison of Mn 3d PDOS. (a) Valence-band spectra for hν = 51 eV in thesputter-etching series. The photon energy is fixed at hν = 51 eV. t(min) denotes the sputteringtime. (b) Comparison of Mn 3d PDOS deduced from the difference between the spectra at t =10 min and 85 min, the spectra at t = 15 min and 85 min, and the spectra at hν = 48 eV and50 eV.

35

the expected energy position of the As-derived Auger signals. Throughout the

depth profile study, the Fermi edge was not observed, reflecting the non-metallic

character of Mn 3d electrons, consistent with the XPS result. Although the

valence-band spectra drastically changed during the first 10 minutes, there was

little change after that. Hence, we conclude that the spectra of the first 10

minutes correspond to those of As-cap layer and the subsequent spectra represent

those of the dilute Mn phase. The resonantly enhanced Mn 3d partial density of

states (PDOS) has been obtained by subtracting the off-resonant (hν = 48 eV)

spectrum from the on-resonant (hν = 50 eV) one (shown at the bottom of each

panel). In order to properly carry out the subtraction, we considered the photon

energy dependence of the photoionization cross-section of As 4p. Throughout

the measurement, Mn-derived Auger was not observed and the Mn 3d PDOS

was suppressed near EF, indicating the localized nature of the Mn 3d states.

There was a main peak at EB ∼3.7 eV and a broad and strong charge transfer

satellite at EB = 5-13 eV in the Mn 3d PDOS at t = 10 min, whose feature was

similar to that of Ga1−xMnxAs [34]. One can see a slight energy difference in the

main peak between Mn/GaAs (001) (EB = 3.7 eV) and Ga1−xMnxAs (EB = 4.5

eV), possibly due to the different positions of the EF caused by the sputtering.

The Mn 3d PDOS completely disappeared at t = 85 min, and thus we considered

that the Mn diffusion layer disappeared and the GaAs substrate appeared at t =

85 min. The absence of Mn at t = 85 min was also confirmed by the absence of

Mn 2p core-level signal (not shown). We have also attempted to obtain the Mn

3d PDOS by subtracting the spectrum at t = 85 min (GaAs) from the one at t =

10 min, 15 min (Mn/GaAs in the dilute Mn phase) for the fixed photon energy

of hν = 51 eV as shown in Fig. 5.8(b). These difference spectra are similar to

those obtained from the RPES measurement.

We performed a magnetization measurement, using a SQUID magnetometer

(MPMS, Quantum Design, Co., Ltd.) for the sample which had been sputtered

for 30 min by the same ion gun as that used for the XPS measurements, that is,

the sputtered Mn/GaAs (001) in the dilute Mn phase. A tiny hysteresis was ob-

served as shown in Fig. 5.9(a), indicating that the sample exhibits ferromagnetic

behavior in the dilute Mn phase.

36

-1.0

0

1.0

Mag

netiz

atio

n (1

0-6em

u/m

m2 )

-1.0 -0.5 0 0.5 1.0Magnetic Field (T)

Mn/GaAs (001)

T = 5 K

(a)t = 30 (min)

-0.4

-0.2

0

0.2

0.4

.

0.01-0.01 0

0.05

0.04

0.03

0.02

0.01

0

Mag

netiz

atio

n (1

0-6em

u/m

m2 )

300200100Temperature (K)

Mn/GaAs (001)

H = 0.005 T

(b)

t = 30 (min)

Figure 5.9: Magnetization curves of thermally diffused Mn/GaAs (001) thin film sputteredfor 30 min. (a) M -H curve at T = 5 K. The linear component in the high magnetic field regionhas been subtracted. The inset shows an enlarged plot for H = −0.01-0.01 T. (b) M -T curvefor H = 0.005 T (zero-field cooled).

37

Chapter 6

Summary

In this thesis, we have studied the Mn-based magnetic semiconductors MnGeP2,

and thermally diffused Mn/GaAs (001) using high-energy spectroscopic tech-

nique.

In Chapter 4, we have investigated the electronic structure and the mag-

netic properties of the II-IV-V2 chalcopyrite-type room-temperature ferromag-

netic semiconductor MnGeP2 thin films by PES, XAS, and XMCD measure-

ments. All the spectra indicated intermediate electronic states between the itin-

erant and localized character of the Mn 3d states in MnGeP2. It has become

clear from the XMCD measurement that a paramagnetic component coexisted

with a ferromagnetic component and had a large orbital magnetic moment.

In Chapter 5, we have performed the depth profile study of the thermally

diffused Mn/GaAs (001) thin films using PES combined with Ar+-ion sputtering

to investigate the electronic structure of the sample along the depth direction.

We confirmed that Mn was thermally diffused into the GaAs substrate into the

deep region, and completely reacted with the the GaAs, consistent with the large

diffusion coefficient predicted by previous report [40]. The Mn 2p core-level and

Mn 3d valence-band spectra of thermally diffused Mn/GaAs (001) in the dilute

Mn phase are similar to those of Ga1−xMnxAs, indicating that the Mn 3d states

are well localized. Ferromagnetism was observed even in the dilute Mn phase.

39

Acknowledgements

It is my great pleasure to express my special gratitude to the following people

for their help concerning my master thesis.

First of all, I would like to express my heartfelt gratitude to Prof. Atsushi Fuji-

mori, who has given me a lot of attentive guidance and valuable advice throughout

this work. I have always got an impression by his knowledge and foresight in the

field of condensed matter physics. Thanks to his pertinent advice, I have been

able to go ahead with my investigation in an efficient manner so far. I also thank

to Prof. Takashi Mizokawa for his instructive advice about the interpretation of

the experimental results.

The experiments at Photon Factory were supported by a number of people.

I am indebted to the members of Kinoshita group, Dr. Taichi Okuda, Dr. Ayumi

Harasawa, Dr. T. Takanori Wakita, and Prof. Toyohiko Kinoshita, for their valu-

able technical support during the beamtimes at BL-18A of Photon Factory. I

also acknowledge Dr. Kazutoshi Mamiya, and Prof. Tsuneharu Koide for their

vital technical support and fruitful discussions about XMCD spectra during the

beamtimes at BL-11A of Photon Factory.

I am very grateful to the members of Sato Group. Mr. Kazuyuki Minami,

Dr. Takayuki Ishibashi, and Prof.Katsuaki Sato willingly provided me with the

high-quality samples of the MnGeP2 thin films.

I am deeply thankful to the members of Oshima Group. Mr. Ken Kanai,

Mr. Kotaro Kubo, Dr. Jun Okabayashi, and Prof.Masaharu Oshima, gave me

the invaluable opportunities to grow the thermally diffused Mn/GaAs (001) thin

films using MBE and perform AFM measurements.

I like to thank the members of Uchida Group for giving me helpful advice to

perform MPMS measurements. In particular, Dr. Kenji Kojima gave me a lot of

useful guidance for the maintenance of the SQUID magnetometer.

I would like to thank the members of Fujimori-Mizokawa Group. Mr. Yukiaki

Ishida always gave me suggestive advice about the interpretation for experimental

results. His attitude for research has inspired me to greater efforts. Mr. Jong-Il

41

Hwang taught me how to use and maintain photoemission instruments. I had

much to learn from his deep insight about DMS’s. Mr. Kazuaki Ebata gave me a

helping hand with my work when I was in trouble. Mr. Masaki Kobayashi gave

me valuable comments on my experimental results and helped me at Photon

Factory. Mr. Masaru Takizawa gave me a lot of useful advice about the neces-

sary things to analyze the experimental data. Mr. Yasuhiro Ooki helped me with

the maintenance of photoemission instruments and experiments. I also wish to

thank Dr. Teppei Yoshida, Dr. Kiyohisa Tanaka, Mr. Hajime Yagi, Mr. Hiroki Wa-

dati, Mr. Makoto Hashimoto, Mr. Masaki Ikeda, Mr. Takashi Maekawa, Dr. James

Quilty, Dr. Jin-Yong Son, Mr. Daishuke Asakura, Mr. ThangTrung Tran, Mr. Akira

Shibata, Mr. Yasuhiro Fujii, Mr. Kou Takubo, and Ms. Ayako Fukuya for their

cordial supports.

Finally, I would like to express my special thanks to my friends and family.

My friends have always encouraged me to bring my master thesis to completion.

My family has supported me mentally as well as financially all the time. I cannot

be too appreciative of their kindness.

Kashiwa, Chiba

January 2006

Yoshitaka Osafune

42

References

[1] K. Onodera, T. Matsumoto, and M. Kimura, Electron. Lett. 30, 1954 (1994).

[2] H. Munekata, H. Ohno, S. von Molnar, A. Segmuller, L. Chang, and L.

Esaki, Phys. Rev. Lett. 63, 1849 (1989).

[3] H. Ohno, A. Shen, F. Matsukura, A. Oiwa, A. Endo, S. Katsumoto, and Y.

Iye, Appl. Phys. Lett. 69, 363 (1996).

[4] A. Oiwa, S. katsumoto, A. Endo, M. Hirasawa, Y. Iye, H. Ohno, F. Mat-

sukura, A. Shen, and Y. Sugawara, Solid State Commun. 103, 209 (1997).

[5] A. V. Esch, J. D. Boeck, L. V. Blockstal, R. Bogaerts, F. Herlach, and G.

Borghs, J. Phys.: Condens. Mat. 9, L361 (1997).

[6] Y. Ohno, D. K. Young, B. Beschoten, F. Matsukura, H. Ohno, and D. D.

Awschalom, Nature 402, 790 (1999).

[7] S. Koshihara, A. Oiwa, M. Hirasawa, S. Katsumoto, Y. Iye, C. Urano, H.

Takagi, and H. Munekata, Appl. Phys. Lett. 78, 4617 (1997).

[8] T. Dietl, H. Ohno, F. Matsukura, J. Cibert, and D. Ferrand, Science 287,

1019 (2000).

[9] K. Ueda, H. Tabata, and Y. Kawai, Appl. Phys. Lett. 79, 988 (2001).

[10] H. Saeki, H. Tabata, and T. Kawai, 120, 439 (2001).

[11] S. Sonoda, S. Simizu, T. Sasaki, Y. Yamamoto, and H. Hori, J. Chrst.

Growth 237-239, 1358 (2002).

[12] M. Hashimoto, Y. Z. Zhou, M. Kanamura, H. Katayama-Yoshida, and H.

Asahi, J. Chrst. Growth 251, 327 (2003).

[13] K. Sato, G. Medvedkin, T. Nishi, Y. Hasegawa, R. Misawa, K. Hirose, and

T. Ishibashi, J. Appl. Phys. 89, 7027 (2001).

43

[14] G. Medvedkin, K. Hirose, T. Ishibashi, T. Nishi, V. Voevodin, and K. Sato,

J. Chrst. Growth 236, 609 (2002).

[15] S. Cho, S. Choi, G.-B. Cha, S. C. Hong, Y. Kim, A. J. Freeman, J. B.

Ketterson, Y. Park, and H.-M. Park, Solid State Commun. 129, 609 (2004).

[16] K. Sato, T. Ishibashi, K. Minami, H. Yuasa, J. Jogo, T. Nagatsuka, A.

Mizusawa, Y. Kanagawa, and A. Koukitu, J. Phys. Chem. Solid 66, 2030

(2005).

[17] Y. Ishida, D. D. Sarma, K. Okazaki, J. Okabayashi, J. I. Hwang, H. Ott, A.

Fujimori, G. A. Medvedkin, T. Ishibashi, and K. Sato, Phys. Rev. Lett. 91,

107202 (2003).

[18] M. L. Reed, N. A. El-Masry, H. H. Stadelmaier, M. K. Ritums, M. J. Reed,

C. A. Parker, J. C. Roberts, and S. M. Bedair, Appl. Phys. Lett. 79, 3473

(2001).

[19] X. M. Cai, A. B. Djurisic, M. X. Xie, H. Liu, X. X. Zhang, J. J. Zhu, and

H. Yang, Mat. Sci. Eng. B 117, 292 (2005).

[20] T. A. Koopmans, Physica 1, 104 (1933).

[21] S. Hufner, Photoelectron Spectroscopy (Springer-Verlag, Berlin, 2003).

[22] U. Fano, Phys. Rev. 124, 1866 (1961).

[23] B. T. Thole, P. Carra, F. Sette, and G. van der Laan, Phys. Rev. Lett. 68,

1943 (1992).

[24] P. Carra, B. T. Thole, M. Altatelli, and X. Wang, Phys. Rev. Lett. 70, 694

(1993).

[25] N. C. Giles and L. E. Halliburton, MRS Bull 23, 37 (1998).

[26] B. H. Bairamov, V. Y. Rud’, and Y. V. Rud’, MRS Bull 23, 41 (1998).

[27] E. Wimmer, H. Krakauer, M. Weinert, and A. J. Freeman, Phys. Rev. B 24,

864 (1981).

[28] Y.-J. Zhao, W. T. Geng, A. J. Freeman, and T. Oguchi, Phys. Rev. B 63,

201202 (2001).

[29] T. Hwang, J. H. Shim, and S. Lee, Appl. Phys. Lett. 83, 1809 (2003).

44

[30] T. Komatsubara, T. Suzuki, and E. Hirahara, J. Supercond. 28, 317 (1970).

[31] E. E. Huber, Jr., and D. H. Ridgley, Phys. Rev. 135, A1033 (1964).

[32] K. Minami, J. Jogo, Y. Morishita, T. Ishibashi, and K. Sato, J. Chrst.

Growth 278, 478 (2005).

[33] Y. Ishida, Master Thesis, University of Tokyo (2003).

[34] J. Okabayashi, A. Kimura, T. Mizokawa, A. Fujimori, T. Hayashi, and M.

Tanaka, Phys. Rev. B 59, R2486 (1999).

[35] A. Kakizaki, H. Sugawara, I. Nagakura, and T. Ishii, J. Supercond. 49, 2183

(1980).

[36] J. Okabayashi, M. Mizuguchi, M. Oshima, H. Shimizu, M. Tanaka, M. Yuri,

and C. T. Chen, Appl. Phys. Lett. 83, 5485 (2003).

[37] G. S. Dong, M. Xu, Y. Chen, X. F. Jin, and X. Wang, 24, 653 (1996).

[38] J. L. Hilton, B. D. Schultz, S. McKernan, and C. J. Palmstrom, Appl. Phys.

Lett. 84, 3145 (2004).

[39] L. R. Doolittle, Nucl. Instrum. Methods Phys. Res. B 9, 344 (1985).

[40] J. L. Hilton, B. D. Schultz, S. McKernan, S. M. Spanton, M. M. R. Evans,

and C. J. Palmstrom, 23, 1752 (2005).

[41] J. Okabayashi, A. Kimura, O. Rader, T. Mizokawa, A. Fujimori, T. Hayashi,

and M. Tanaka, Phys. Rev. B 58, R4211 (1998).

45