08.11.2014 Views

Magnetic and transport properties of ferromagnetic semiconductor ...

Magnetic and transport properties of ferromagnetic semiconductor ...

Magnetic and transport properties of ferromagnetic semiconductor ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Magnetic</strong> <strong>and</strong> <strong>transport</strong> <strong>properties</strong> <strong>of</strong> <strong>ferromagnetic</strong><br />

<strong>semiconductor</strong> multinary alloys<br />

Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As<br />

Izabela Kudelska<br />

PhD Dissertation<br />

Institute <strong>of</strong> Physics<br />

Polish Academy <strong>of</strong> Sciences<br />

Thesis Supervisor doc. dr. hab. W. Dobrowolski<br />

Warsaw 2004


Mojemu kochanemu Synkowi Olesiowi,<br />

który towarzyszył mi<br />

przez cały okres pisania pracy


PODZI ˛ EKOWANIA<br />

Pragnę serdecznie podziękować mojemu Promotorowi doc. dr hab. W. Dobrowolskiemu<br />

za stałą, wszechstronną pomoc i opiekę nad pracą, pouczające rozmowy i życzliwość<br />

w trakcie wykonywania pracy.<br />

Pr<strong>of</strong>. dr J. Furdynie bardzo dziękuję za inspirujące dyskusje i wprowadzenie w tematykę<br />

badań ferromagnetycznych kryształów Ga 1−x Mn x As.<br />

Pr<strong>of</strong>. dr M. Dobrowolskiej dziękuję bardzo za opiekę i pomoc w badaniach.<br />

Dr M. Arciszewskiej serdecznie dziękuję za pomoc przy wykonywaniu pracy i życzliwe<br />

zainteresowanie wynikami badań.<br />

Doc. dr hab. T. Wojtowiczowi dziękuję bardzo za wyhodowanie kryształów Ga 1−x Mn x As<br />

użytych do badań i pomoc przy powstawaniu pracy.<br />

Pragnę podziękować dr X. Liu za pomiary namagnesowania oraz wyhodowanie kryształów<br />

Ga 1−x Mn x As, W. Lim za pomoc w badaniach <strong>transport</strong>owych oraz dr Y. Sasaki za<br />

wyhodowanie kryształów Ga 1−x Mn x As badanych w pracy.<br />

Współpraca z pracownikami Laboratorium Silnych Impulsowych Pól Magnetycznych w<br />

Tuluzie była bardzo pomocna. W szczególności pragnę podziękować dr O. Portugall za<br />

pomoc w pomiarach magneto<strong>transport</strong>owych, dr M. Goiran za pomoc w pomiarach magnetooptycznego<br />

efektu Kerra, a także dr E.Haanappel, dr J.-M. Broto, dr H. Rakoto, dr B.<br />

Raquet za pomoc w pomiarach magneto<strong>transport</strong>owych.<br />

Pragnę także podziękować Pr<strong>of</strong>. dr V. Dugaev za cenne i inspirujące dyskusje.<br />

Pr<strong>of</strong>. dr W. Walukiewiczowi oraz dr K. M. Yu dziękuję za pomiary c-PIXE oraz c-RBS.<br />

Bardzo dziękuję dr J. Domagale za pomiary dyfrakcji rentgenowskiej.<br />

Dr E. I. Slynko oraz dr V. E. Slynko dziękuję za wyhodowanie kryształów<br />

Pb 1−x−y−z Mn x Eu y Sn z Te użytych do badań.<br />

Dr I. M. Fita dziękuję za pomiary namagnesowamia w ciśnieniu hydrostatycznym.<br />

Pragnę także podziękować Pr<strong>of</strong>. dr hab. R. Szymczak oraz dr Baranowi za pomiary namagnesowania.<br />

Wszystkim pracownikom i doktorantom z Odziału Fizyki Półprzewodników Półmagnetycznych<br />

dziękuję za stworzenie miłej i życzliwej atmosfery.<br />

Mojemu Mężowi Arkowi pragnę podziękować za nieustanne wspieranie mnie i cierpliwość.<br />

3


CONTENTS<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2. Experimental techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.2 The experimental technique <strong>of</strong> energy dispersive X-ray fluorescence<br />

(EDXRF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.3 The experimental setup for magnetotransort measurements . . . . . . . . 12<br />

2.4 AC/DC magnetometer . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.5 The measurements in the range <strong>of</strong> high pulsed magnetic fields . . . . . . 18<br />

2.6 The experimental setup for Kerr effect measurements . . . . . . . . . . . 19<br />

3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. . . . . . . . . . . . . 21<br />

3.1 Ferromagnetic bulk crystals <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te . . . . . . . . . 21<br />

3.2 Ferromagnetic layers <strong>of</strong> Ga 1−x Mn x As. . . . . . . . . . . . . . . . . . . . 21<br />

4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te . . . . . . . 27<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

4.2 Transport characterization <strong>of</strong> Pb 1−x−y−x Mn x Eu y Sn z Te samples . . . . . . 30<br />

4.3 <strong>Magnetic</strong> measurements <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals . . . 34<br />

5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As . . . . . . . . . . . . . . . 48<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.2 The procedure <strong>of</strong> annealing . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

5.3 The results <strong>of</strong> zero-field resistivity measurements . . . . . . . . . . . . . 51<br />

5.4 The results <strong>of</strong> magneto<strong>transport</strong> measurements . . . . . . . . . . . . . . 56<br />

5.5 The results <strong>of</strong> SQUID measurements . . . . . . . . . . . . . . . . . . . . 64<br />

5.6 The results <strong>of</strong> magnetooptical Kerr effect measurements . . . . . . . . . . 69<br />

5.6.1 Introduction - magnetooptical Kerr effect . . . . . . . . . . . . . 69<br />

5.6.2 The results <strong>of</strong> MOKE measurements for the as-grown <strong>and</strong> annealed<br />

GaMnAs epilayers . . . . . . . . . . . . . . . . . . . . . 71<br />

5.7 The channeling experiments - the results <strong>of</strong> c-RBS <strong>and</strong> c-PIXE measurements<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

5.8 The results <strong>of</strong> diffraction (HRXRD) measurements . . . . . . . . . . . . 90<br />

6. Conclusions <strong>and</strong> Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.1 Pb 1−x−y−z Mn x Eu y Sn z Te . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.2 Ga 1−x Mn x As . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.3 Suggestions for further studies . . . . . . . . . . . . . . . . . . . . . . . 101


Contents 5<br />

7. Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

7.1 Appendix 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

7.2 Appendix2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

List <strong>of</strong> Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

List <strong>of</strong> Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122


1. INTRODUCTION<br />

In this thesis, the results <strong>of</strong> a study <strong>of</strong> the magnetic <strong>and</strong> <strong>transport</strong> <strong>properties</strong> <strong>of</strong> III-V as<br />

well as IV-VI based <strong>ferromagnetic</strong> semimagnetic <strong>semiconductor</strong>s will be presented. Two<br />

systems <strong>of</strong> <strong>ferromagnetic</strong> <strong>semiconductor</strong> multinary alloys: PbSnMnEuTe <strong>and</strong> GaMnAs<br />

were systematically investigated.<br />

The common characteristic <strong>of</strong> two investigated semimagnetic materials,<br />

Ga 1−x Mn x As <strong>and</strong> Pb 1−x−y−z Mn x Eu y Sn z Te, is mediation <strong>of</strong> free holes in <strong>ferromagnetic</strong><br />

interactions <strong>of</strong> Mn ions. The mechanism <strong>of</strong> exchange interactions is well explored<br />

for Pb 1−x−y Mn x Sn y Te mixed crystals. The RKKY interaction is known to be responsible<br />

for the <strong>ferromagnetic</strong> <strong>properties</strong> <strong>of</strong> IV-VI semimagnetic <strong>semiconductor</strong>s. However,<br />

the origin <strong>of</strong> ferromagnetism in Ga 1−x Mn x As is still under active discussion. The<br />

experimental studies <strong>of</strong> physical <strong>properties</strong> are necessary to examination the origin <strong>of</strong><br />

ferromagnetism in this material.<br />

Recently, manipulation <strong>of</strong> the spin degree <strong>of</strong> freedom in <strong>semiconductor</strong>s has become<br />

a focus <strong>of</strong> interest. The main goal <strong>of</strong> spintronics is to make use <strong>of</strong> both charge <strong>and</strong> spin<br />

degrees <strong>of</strong> freedom in <strong>semiconductor</strong>s. In the context <strong>of</strong> spin electronics particularly<br />

interesting are <strong>ferromagnetic</strong> semimagnetic <strong>semiconductor</strong>s. In the case <strong>of</strong> Mn-based IV-<br />

VI, III-V <strong>and</strong> II-VI SMSC’s the ferromagnetism can be observed provided that the hole<br />

concentration is sufficiently high.<br />

Underst<strong>and</strong>ing <strong>of</strong> the carrier mediated ferromagnetism was initiated by a study <strong>of</strong><br />

ferromagnetism in IV-VI based SMSC’s. In this class <strong>of</strong> materials deviations from stoichiometry<br />

result in the carrier density sufficiently high to produce strong <strong>ferromagnetic</strong><br />

interactions between the localized spins. Ferromagnetic <strong>properties</strong> are observed for IV-<br />

VI semimagnetic materials with Mn <strong>and</strong> with the concentration <strong>of</strong> conducting holes p<br />

≥ 2-3 10 20 cm −3 . However, in the case <strong>of</strong> <strong>ferromagnetic</strong> IV-VI semimagnetic materials<br />

(such as PbSnMnTe), one have to admit that the <strong>ferromagnetic</strong> characteristics (magnetic<br />

anisotropy, coercive field) are not superior to other magnetic materials. Thus, applications<br />

related to <strong>ferromagnetic</strong> <strong>properties</strong> <strong>of</strong> these materials can be rather related to hybrid<br />

systems with IV-VI electronic devices incorporating the <strong>ferromagnetic</strong> element with controlled<br />

magnetic <strong>properties</strong>. An additional obstacle for these materials is the low temperature<br />

<strong>of</strong> <strong>ferromagnetic</strong> phase transition. However, two important features give IV-VI semimagnetic<br />

materials the distinguished position within the whole family <strong>of</strong> semimagnetic<br />

<strong>semiconductor</strong>s. First, the variety <strong>of</strong> magnetic <strong>properties</strong> observed in Mn based IV-VI<br />

SMSC’s. Second, the characteristic feature are semi-metallic electric <strong>properties</strong> with the<br />

well developed methods <strong>of</strong> control <strong>of</strong> carrier concentration. Pb 1−x−y−z Mn x Eu y Sn z Te are<br />

a unique compounds in which the interplay between magnetic <strong>and</strong> electronic <strong>properties</strong><br />

can be observed <strong>and</strong> studied. Particularly, the carrier induced paramagnet-ferromagnet<br />

<strong>and</strong> ferromagnet-spin glass transitions are present [1], [2].<br />

Recently, due to the rapid progress achieved in SMSC’s technology, namely nonequilibrium<br />

growth methods, the <strong>ferromagnetic</strong> phase transition was observed in III-V


1. Introduction 7<br />

semimagnetic materials. For the first time, by using low-temperature molecular beam<br />

epitaxy (LT-MBE; T < 300 0 C), a III-V based SMSC, In 1−x Mn x As layers were successfully<br />

grown in 1989 [3] <strong>and</strong> shown to exhibit hole-induced <strong>ferromagnetic</strong> ordering at<br />

low temperatures. Development <strong>of</strong> this technique resulted in the successful epitaxy <strong>of</strong><br />

Ga 1−x Mn x As [4], [5]. The III-V based compounds are prospective materials for spin<br />

electronics because are already in use in everyday electronics. The low temperature (LT)<br />

MBE grown Ga 1−x Mn x As, has become a favorite material for spintronics when it was<br />

shown [6] that the substitution <strong>of</strong> Mn for Ga in GaAs leads to ferromagnetism at temperatures<br />

as high as 110K. For a long time this was the highest Curie temperature T C<br />

observed in semimagnetic <strong>semiconductor</strong>s. For the spintronic applications, however, the<br />

temperature <strong>of</strong> the transition to the <strong>ferromagnetic</strong> phase has to be considerable increased.<br />

In this thesis, the systematic measurements <strong>of</strong> magnetic AC susceptibility, magnetization<br />

as well as <strong>transport</strong> characterization <strong>of</strong> bulk samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te<br />

multinary alloy were performed. The as-grown as well as annealed samples with different<br />

concentration <strong>of</strong> Mn x as well as Eu y were investigated. The paramagnet-ferromagnet as<br />

well as ferromagnet-spin glass transitions were observed <strong>and</strong> studied. The results <strong>of</strong> magnetic<br />

<strong>and</strong> <strong>transport</strong> studies showed that by introducing <strong>of</strong> two types <strong>of</strong> magnetic ions into<br />

IV-VI <strong>semiconductor</strong> matrix, one can change the magnetic <strong>properties</strong> <strong>of</strong> the investigated<br />

semimagnetic material.<br />

The variety <strong>of</strong> experimental techniques were used to explore the physical <strong>properties</strong> <strong>of</strong><br />

Ga 1−x Mn x As epilayers. The <strong>transport</strong>, magneto<strong>transport</strong>, magnetic, structural as well as<br />

channelling experiments were carried out <strong>and</strong> analyzed. The magnetooptical Kerr effect<br />

studies allowed to explore the magnetooptical <strong>properties</strong> <strong>of</strong> the investigated samples. The<br />

as-grown as well as annealed at various conditions samples were investigated. The main<br />

subject studied is the role <strong>of</strong> Mn interstitial in Ga 1−x Mn x As epilayers. The magnetic,<br />

<strong>transport</strong>, structural <strong>and</strong> channeling experiments indicated that the formation <strong>of</strong> interstitial<br />

Mn ions plays a crucial role in controlling the <strong>ferromagnetic</strong> transition in GaMnAs. For<br />

the first time it is shown that by a proper choice <strong>of</strong> annealing conditions (temperature,<br />

time, flow <strong>of</strong> the gas) the limit <strong>of</strong> the Curie temperature <strong>of</strong> Ga 1−x Mn x As epilayers (T C<br />

∼ 110K) can be shifted to much higher values (up to 127K for the sample with high Mn<br />

concentration x ∼ 0.08) [7]. The results presented in the thesis initiated further studies <strong>of</strong><br />

low temperature annealing in many laboratories. At present, Curie temperature exceeding<br />

160K was reported [8]. This progress has been achieved basically though optimization <strong>of</strong><br />

post-growth annealing time <strong>and</strong> temperature.<br />

The variety <strong>of</strong> experimental techniques were used in the studies presented in the thesis.<br />

The measurements <strong>of</strong> energy dispersive X-ray fluorescence (EDXRF), magneto<strong>transport</strong><br />

studies (up to 13T using superconducting magnet), AC magnetic susceptibility as well<br />

as DC magnetization (up to 9T) by use <strong>of</strong> 7229 LakeShore Susceptometer/Magnetometer<br />

setup were performed in Division ON-1 Physics <strong>of</strong> Semiconductors <strong>of</strong> Institute <strong>of</strong> Physics<br />

Polish Academy <strong>of</strong> Sciences in Warsaw.<br />

The results were obtained in collaboration with many groups. The XRD as well as<br />

RHEED measurements were carried out in Department <strong>of</strong> Physics, University <strong>of</strong> Notre<br />

Dame. High resolution X-ray diffraction studies as well as the st<strong>and</strong>ard powder X-ray<br />

measurements were performed in Laboratory <strong>of</strong> X-ray <strong>and</strong> Electron Microscopy Research,<br />

Group <strong>of</strong> Applied Crystalography, Institute <strong>of</strong> Physics Polish Academy <strong>of</strong> Sciences<br />

in Warsaw.<br />

The <strong>transport</strong> as well as magneto<strong>transport</strong> investigations in the static magnetic field


1. Introduction 8<br />

were carried out in Department <strong>of</strong> Physics, University <strong>of</strong> Notre Dame (up to 0.5T by use<br />

<strong>of</strong> classical electromagnet; up to 5T by use <strong>of</strong> superconducting magnet). The magneto<strong>transport</strong><br />

measurements in high pulsed magnetic fields (up to 55T) were performed in<br />

Laboratoire National des Champs Magnetiques Pulses Toulouse.<br />

The magnetization measurements under hydrostatic pressure were carried out in Division<br />

<strong>of</strong> Physics <strong>of</strong> Magnetism, Group <strong>of</strong> Phase Transitions, Institute <strong>of</strong> Physics Polish<br />

Academy <strong>of</strong> Sciences in Warsaw. The SQUID measurements were performed in Department<br />

<strong>of</strong> Physics, University <strong>of</strong> Notre Dame <strong>and</strong> Division <strong>of</strong> Physics <strong>of</strong> Magnetism, Group<br />

<strong>of</strong> Phase Transitions, Institute <strong>of</strong> Physics Polish Academy <strong>of</strong> Sciences in Warsaw. Indirect<br />

magnetization measurements were also performed. The magnetooptical Kerr Effect<br />

(MOKE) was measured in high pulsed magnetic fields (up to 25T) in Laboratoire National<br />

des Champs Magnetiques Pulses Toulouse.<br />

The channeling experiments i.e. channeling Rutherford backscattering (c-RBS) <strong>and</strong><br />

channeling particle-induced X-ray emission (c-PIXE) measurements were performed in<br />

Material Science Division Lawrence Berkeley National Laboratory.<br />

The thesis is organized in the following way. Chapter 1 describes selected experimental<br />

techniques used in the studies. The characterization <strong>of</strong> the samples is given in Chapter<br />

2. The results <strong>of</strong> <strong>transport</strong> <strong>and</strong> magnetic measurements <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te crystals<br />

are presented <strong>and</strong> analyzed in Chapter 3. The results <strong>of</strong> experimental studies <strong>of</strong> LT<br />

Ga 1−x Mn x As epilayers: the procedure <strong>of</strong> annealing, the results <strong>of</strong> zero-field resistivity,<br />

magneto<strong>transport</strong>, magnetic, channelling as well as structural measurements are presented<br />

<strong>and</strong> discussed in Chapter 4. The summary <strong>and</strong> outlook is given in Chapter 5.<br />

The results presented in this thesis were partially published in the papers:<br />

1. Kuryliszyn I., Arciszewska M., Abdel Aziz M.M., Slynko E.I., Slynko V.I., Dugaev<br />

V.K., "In quest <strong>of</strong> Mn-Eu interaction in IV-VI mixed crystals", Proc. 9th Int. Conf.<br />

On Narrow Gap Semiconductors, ed. by N. Puhlman, H.-U. Müller, M. Von Ortenberg,<br />

Humboldt University Berlin, p. 96-98 (2000).<br />

2. Yu K.M., Walukiewicz W., Wojtowicz T., Kuryliszyn I., Liu X., Sasaki Y., Furdyna<br />

J.K., "Effect <strong>of</strong> the location <strong>of</strong> Mn sites in <strong>ferromagnetic</strong> Ga 1−x Mn x As on its Curie<br />

temperature", Phys. Rev. B, vol.65, p. 201303(R), (2002).<br />

3. Kuryliszyn I., Wojtowicz T., Liu X., Furdyna J., Dobrowolski W., Broto J.-M., Goiran<br />

M., Portugall O., Rakoto H., Raquet B., Transport <strong>and</strong> magnetic <strong>properties</strong> <strong>of</strong> LT<br />

annealed Ga 1−x Mn x As", Acta Phys. Pol. A, vol.102, No 4-5, p.659, (2002).<br />

4. Kuryliszyn I., Wojtowicz T., Liu X., Furdyna J.K., Dobrowolski W., Broto J.-M.,<br />

Goiran M., Portugall O., Rakoto H., Raquet B., "Low Temperature annealing<br />

studies <strong>of</strong> Ga 1−x Mn x As", Journal <strong>of</strong> Supercondactivity (Incorporating Novel Magnetism),<br />

vol. 16, No 1, p. 63, (2003).<br />

5. Yu K.M., Walukiewicz W., Wojtowicz T., Kuryliszyn I., Liu X., Sasaki Y., Furdyna<br />

J.K., "Thermodynamic Limits to the maximum Curie Temperature in GaMnAs",<br />

Proceedings- 26th International Conference on the Physics <strong>of</strong> <strong>semiconductor</strong>s, Edinburgh<br />

2002.


1. Introduction 9<br />

6. Racka K., Kuryliszyn I., Arciszewska M., Dobrowolski W., Broto J.-M., Goiran M.,<br />

Portugall O., Rakoto H., Raquet B., " Anomalous Hall Effect in Sn 1−x−y Mn x Eu y Te<br />

Mixed Crystals", Journal <strong>of</strong> Supercondactivity, (Incorporating Novel Magnetism),<br />

vol. 16, No 2, p. 289, (2003).<br />

7. Furdyna J.K., Liu X., Lim W.L., Sasaki Y., Wojtowicz T., Kuryliszyn I., Lee S., Yu<br />

K.M. <strong>and</strong> Walukiewicz W., "Ferromagnetic III-Mn-V Semiconductors: Manipulation<br />

<strong>of</strong> <strong>Magnetic</strong> Properies by Annealing, Extrinsic Doping, <strong>and</strong> Multilayer Design",<br />

Journal <strong>of</strong> the Korean Physical Society, vol. 42, p. S579, (2003).<br />

8. Kuryliszyn-Kudelska I., Domagala J.Z, Wojtowicz T., Liu X., E. Łusakowska, Dobrowolski<br />

W., Furdyna J.K., "The effect <strong>of</strong> Mn interstitials on the lattice parameter<br />

<strong>of</strong> Ga 1−x Mn 1−x As", J. Appl. Phys. 95 (2) 603 (2004).<br />

9. Kuryliszyn-Kudelska I., Wojtowicz T., Liu X., Furdyna J.K., Dobrowolski W., Domagala<br />

J.Z., E.Łusakowska, M.Goiran, E.Haanappel, O. Portugall, "Effect <strong>of</strong> annealing<br />

on magnetic <strong>and</strong> magneto<strong>transport</strong> <strong>properties</strong> <strong>of</strong> Ga 1−x Mn x As epilayers",<br />

Journal <strong>of</strong> <strong>Magnetic</strong> Materials <strong>and</strong> Magnetism 272-276, p. e1575, (2004).<br />

10. Kacman P., Kuryliszyn-Kudelska I., "The role <strong>of</strong> Interstitial Mn in GaAs-based Dilute<br />

<strong>Magnetic</strong> Semiconductors", Lecture Notes in Physics (in press)


2. EXPERIMENTAL TECHNIQUES<br />

2.1 Introduction<br />

Chapter 2 includes briefly description <strong>of</strong> the selected experimental methods which were<br />

used by author.<br />

The chemical composition <strong>of</strong> the investigated samples was determined by use <strong>of</strong><br />

different techniques. The X-ray dispersive fluorescence analysis measurements (for<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples), X-ray diffraction (XRD) as well as high resolusion X-<br />

ray diffraction (HRXRD) studies <strong>and</strong> reflection high energy electron diffraction (RHEED)<br />

measurements (for Ga 1−x Mn x As epilayers) were employed to characterize the investigated<br />

samples. Section 1.1 contains briefly description <strong>of</strong> Tracor X-ray Spectrace 5000 -<br />

automated EDXRF analyzer.<br />

The structural characterization <strong>of</strong> both - Pb 1−x−y−z Mn x Eu y Sn z Te as well as<br />

Ga 1−x Mn x As epilayers was carried out. X-ray diffraction (Ga 1−x Mn x As samples), high<br />

resolution X-ray diffraction (Ga 1−x Mn x As samples) <strong>and</strong> st<strong>and</strong>ard powder X-ray measurements<br />

(Pb 1−x−y−z Mn x Eu y Sn z Te samples) were performed.<br />

The <strong>transport</strong> (resistivity) as well as magneto<strong>transport</strong> (Hall effect, anomalous Hall<br />

effect, magnetoresistance) <strong>properties</strong> <strong>of</strong> investigated <strong>ferromagnetic</strong> materials were studied<br />

by use <strong>of</strong> several experimental setups. The resistivity measurements <strong>of</strong> Ga 1−x Mn x As<br />

samples were performed in the temperature range between 10K <strong>and</strong> 300K by use <strong>of</strong> a<br />

helium flow cryostat. The DC six probe technique was used for the magneto<strong>transport</strong><br />

measurements. Section 2.2 describes the experimental setup for magneto<strong>transport</strong> measurements<br />

built by author in Department <strong>of</strong> Physics University <strong>of</strong> Notre Dame. The system<br />

together with the cryostat <strong>and</strong> superconducting magnet allows to perform the st<strong>and</strong>ard<br />

DC six probe technique measurements in magnetic field up to 5T <strong>and</strong> in the temperature<br />

range between 1.3 <strong>and</strong> 300K. Section 3 <strong>of</strong> this Chapter describes briefly pulsed magnetic<br />

fields.<br />

The magnetic studies (AC susceptibility <strong>and</strong> magnetization investigations) were performed<br />

by use <strong>of</strong> several techniques. The Pb 1−x−y−z Mn x Eu y Sn z Te samples were explored<br />

using AC/DC magnetometer - AC susceptibility as well as DC magnetization (up<br />

to 9T) were measured. The principle <strong>of</strong> operation <strong>and</strong> schematic figures <strong>of</strong> the 7229<br />

LakeShore Susceptometer/Magnetometer setup are shown in Section 2.4. Additionally,<br />

the magnetization measurements under hydrostatic pressure were performed using the vibrating<br />

sample magnetometer for Pb 1−x−y−z Mn x Eu y Sn z Te samples. For Ga 1−x Mn x As<br />

epilayers magnetization studies were performed. First, directly magnetization investigation<br />

were carried out by use <strong>of</strong> SQUID magnetometer. Second, magnetoptical Kerr effect<br />

(MOKE) were performed. The experimental setup for MOKE measurements is described<br />

in section 2.5.<br />

The channeling experiments - channeling particle-induced X-ray emission (c-PIXE)<br />

as well as channeling Rutherford backscattering (c-RBS) were performed for GaMnAs


2. Experimental techniques 11<br />

samples. The principle <strong>of</strong> channeling experiments is briefly described in Section 7 <strong>of</strong><br />

Chapter 5.<br />

2.2 The experimental technique <strong>of</strong> energy dispersive X-ray fluorescence<br />

(EDXRF)<br />

The chemical composition <strong>of</strong> IV-VI semimagnetic <strong>semiconductor</strong>s was determined by X-<br />

ray dispersive fluorescence analysis technique. This technique allows for determination <strong>of</strong><br />

chemical composition <strong>of</strong> the samples with uncertainty <strong>of</strong> 10 %. The experimental setup -<br />

X-ray beam geometry is shown schematically in Figure 2.1. The Tracor X-ray Spectrace<br />

5000 is an automated energy dispersive X-ray fluorescence (EDXRF) analyzer. It can be<br />

used for nondestructive elemental analysis <strong>of</strong> solids <strong>and</strong> liquids.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 2.1: The schematic view <strong>of</strong> the Tracor X-ray Spectrace 5000.<br />

The principle <strong>of</strong> operation is very simple. The primary X-rays from X-ray tube hit<br />

the sample <strong>and</strong> induce the emission <strong>of</strong> secondary X-rays by the elements contained in the<br />

sample.<br />

The relative intensity <strong>of</strong> an X-ray spectral line excited by monochromatic radiation<br />

can be computed for a given element <strong>and</strong> known spectrometer geometry using following<br />

equation:<br />

I L = I 0 ω A g L<br />

r A − 1<br />

r A<br />

dΩ<br />

4π<br />

C A µ A (λ pri ) csc(φ)<br />

µ M (λ pri ) csc(φ) + µ M (λ L ) csc(Φ)<br />

where:<br />

I L is the analyte line intensity<br />

I 0 is the intensity <strong>of</strong> the primary beam with effective wavelength λ pri<br />

λ pri is the effective wavelength <strong>of</strong> the primary X-ray<br />

λ L is the wavelength <strong>of</strong> the measured analyte line<br />

g L is the fractional value <strong>of</strong> the measured analyte line L in its series<br />

(2.1)


2. Experimental techniques 12<br />

r A is the absorption edge jump ratio <strong>of</strong> analyte A<br />

C A is the concentration <strong>of</strong> analyte A<br />

dΩ/4π is the fractional value <strong>of</strong> the fluorescent X-ray that is directed toward a detector<br />

µ A (λ pri ) is the mass absorption coefficient <strong>of</strong> analyte A for λ pri<br />

µ M (λ pri ) is the mass absorption coefficient <strong>of</strong> the matrix for λ pri<br />

µ M (λ L ) is the mass absorption coefficient <strong>of</strong> the matrix for analyte line λ L<br />

φ is the incident angle <strong>of</strong> the primary beam<br />

Φ is the take<strong>of</strong>f angle <strong>of</strong> fluorescent beam<br />

The Spectrace 5000 uses a low power X-ray tube (less than 50 watts) with a rhodium<br />

anode target. The X-ray tube generates the X-rays that are incident upon the sample so<br />

as to cause the sample to fluorescence. The X-rays emitted by the anode pass through a<br />

0.005" Beryllium window. Then the beam is collimated <strong>and</strong> filtered. The collimator <strong>and</strong><br />

X-ray tube define the illumination beam. To minimize scattering, the detector acceptance<br />

is 90 0 from the incident beam. The sample is placed at the intercept <strong>of</strong> the two beams as<br />

is shown in Figure 2.1. Liquid nitrogen inside a dewar cools the Si(Li) detector to reduce<br />

noise caused by the leakage current in the detector.<br />

2.3 The experimental setup for magnetotransort measurements<br />

The experimental setup for <strong>transport</strong> measurements presented in this section was built<br />

by author in Department <strong>of</strong> Physics University <strong>of</strong> Notre Dame. Figure 2.2 <strong>and</strong> Figure<br />

2.3 show schematically the principle <strong>of</strong> operation <strong>of</strong> the system. The system was built<br />

making use <strong>of</strong> Oxford Instruments optical cryostat <strong>and</strong> superconducting magnet. The<br />

main advantage <strong>of</strong> this system is possibility <strong>of</strong> carrying out both magnetooptical as well<br />

as magneto<strong>transport</strong> investigations. One <strong>of</strong> the application example are magneto<strong>transport</strong><br />

measurements with simultaneous optical excitation <strong>of</strong> the sample.<br />

The system allows to perform magneto<strong>transport</strong> studies in magnetic field up to 5T <strong>and</strong><br />

in the temperature range between 1.3K <strong>and</strong> 300K.<br />

The principle <strong>of</strong> operation involves the st<strong>and</strong>ard DC six probe technique. The 220<br />

Keithley Current Source serve as a source <strong>of</strong> DC current steering the sample, HP DMM<br />

is used for conductivity voltage measurements <strong>and</strong> 2001 Keithley DMM for Hall voltage<br />

measurements (see Figure 2.2). The Hall probe mounted inside the cryostat allows for<br />

detection <strong>of</strong> magnetic field. During the experiment the temperature is collected also.<br />

Additionally, the configuration with 7001 Keithley Scanner <strong>and</strong> 7065 Hall Card allows<br />

to study high resistivity samples as is shown in Figure 2.3. In this case the 2001 Keithley<br />

DMM is used to collect both the conductivity as well as Hall signal.<br />

In the present thesis the described experimental setup was used for magneto<strong>transport</strong><br />

investigations <strong>of</strong> GaMnAs epilayers with typical resistance <strong>of</strong> 1 kΩ. In this case the<br />

configuration shown in Figure 2.2 was used.<br />

Additionally, employing the second Lake Shore power supplier connected in parallel<br />

with Oxford Instruments power supplier (see Figure 2.4) allows for fluently reversion<br />

the direction <strong>of</strong> magnetic field (passing through zero <strong>of</strong> magnetic field). In particular, the<br />

hysteresis loops <strong>of</strong> Hall voltage were measured using two power suppliers.


2. Experimental techniques 13<br />

R<br />

Hall<br />

Probe<br />

HP DMM<br />

(Temperature)<br />

Temperature<br />

Controller<br />

Oxford<br />

Instruments<br />

Power<br />

Supplier<br />

Current<br />

Source<br />

HP DMM<br />

(<strong>Magnetic</strong> Field)<br />

Sample<br />

Cryostat<br />

&<br />

Superconducting<br />

Magnet<br />

HP DMM<br />

V σ<br />

Current<br />

Source<br />

Keithley 220<br />

Keithley DMM<br />

2001<br />

V H<br />

Fig. 2.2: The experimental setup for magneto<strong>transport</strong> measurements.<br />

2.4 AC/DC magnetometer<br />

The magnetic <strong>properties</strong> <strong>of</strong> IV-VI semimagnetic <strong>semiconductor</strong>s were investigated by use<br />

<strong>of</strong> 7229 LakeShore Susceptometer/Magnetometer system. The experimental setup presented<br />

schematically in Figure 2.5 allows to perform AC susceptibility as well as DC<br />

magnetization measurements. The specifications <strong>of</strong> the setup are shown in Table 2.1.<br />

The principle <strong>of</strong> operation <strong>of</strong> AC susceptometer involves subjecting the sample to a<br />

small alternating magnetic field. The flux variation due to the sample is picked up by a<br />

sensing coil surrounding the sample <strong>and</strong> the resulting voltage induced in the coil is detected.<br />

This voltage is directly proportional to the magnetic susceptibility <strong>of</strong> the sample.<br />

The alternating magnetic field is generated by a solenoid which serves as the primary in<br />

a transformer circuit. The solenoid is driven with an AC current source with variable<br />

amplitude <strong>and</strong> frequency. Additionally, a DC field may also be applied by supplying a<br />

DC current to the primary coil. Two identical sensing coils are positioned symmetrically<br />

inside <strong>of</strong> the primary coil <strong>and</strong> serve as the secondary coils in the measuring circuit. Figure<br />

2.6 shows a cross-sectional view <strong>of</strong> the coil assembly. The two sensing coils are<br />

connected in opposition in order to cancel the voltages induced by the AC field itself or<br />

voltages induced by unwanted external sources. Assuming perfectly wound sensing coils<br />

<strong>and</strong> perfect symmetry, no voltage will be detected by lock-in amplifier when the coil assembly<br />

is empty. When a sample is placed within one <strong>of</strong> the sensing coils, the voltage<br />

balance is disturbed. The measured voltage U is proportional to the susceptibility <strong>of</strong> the


2. Experimental techniques 14<br />

R<br />

Hall<br />

Probe<br />

HP DMM<br />

(Temperature)<br />

Current<br />

Source<br />

Temperature<br />

Controller<br />

Oxford<br />

Instruments<br />

Power<br />

Supplier<br />

HP DMM<br />

(<strong>Magnetic</strong><br />

Field)<br />

Sample<br />

Cryostat<br />

&<br />

Superconducting<br />

Magnet<br />

Keithley 7001<br />

Scaner<br />

+<br />

7065 Hall Card<br />

Current<br />

Source<br />

Keithley 220<br />

Keithley DMM<br />

2001<br />

V H , V σ<br />

Fig. 2.3: The experimental setup for magneto<strong>transport</strong> measurements (configuration with 7001<br />

Keithley Scanner <strong>and</strong> 7065 Hall Card).<br />

Coil<br />

HP<br />

DMM<br />

R<br />

Oxford<br />

Instruments<br />

Power<br />

Supplier I<br />

Lake<br />

Shore<br />

Power<br />

Supplier II<br />

Fig. 2.4: The configuration with two power suppliers (Oxford Instruments <strong>and</strong> Lake Shore power<br />

supplier) connected in parallel. This configuration allowed to reverse fluently the direction<br />

<strong>of</strong> magnetic field.


2. Experimental techniques 15<br />

Tab. 2.1: Specifications <strong>of</strong> 7229 LakeShore Susceptometer/Magnetometer system.<br />

Temperature<br />

AC/DC <strong>Magnetic</strong> Field (Primary Coil)<br />

AC Susceptibility Sensitivity<br />

DC Moment Sensitivity<br />

Superconducting Magnet Specifications<br />

Range: From 1.3 K to 325 K<br />

Accuracy: ±0.5% <strong>of</strong> T<br />

Stability: ±0.1K<br />

Range: from 0.00125 gauss to 20 gauss<br />

Accuracy: ±1.0%<br />

Stability: ±0.05%<br />

Frequancy: from 1Hz to 10kHz<br />

to 2 · 10 −8 emu<br />

9 · 10 −5 emu<br />

Field range: ±90000 gauss (±9 Tesla)<br />

Accuracy: ±1.0% <strong>of</strong> setting<br />

Accuracy: ±1.0% <strong>of</strong> setting<br />

Remnant field: 30 gauss<br />

(< 15 gauss after demagnetization cycle)<br />

sample <strong>and</strong> depends on a number <strong>of</strong> other experimental parameters:<br />

U = (1/α)mfBχ (2.2)<br />

where U is measured voltage, α is calibration coefficient, m is sample mass, f is frequency<br />

<strong>of</strong> AC field, B is magnetic field <strong>and</strong> χ is volume susceptibility <strong>of</strong> sample.<br />

The calibration coefficient is dependent on the sample <strong>and</strong> coil geometry <strong>and</strong> is experimentally<br />

determined by use <strong>of</strong> st<strong>and</strong>ard materials with a known susceptibility <strong>and</strong><br />

mass.<br />

The sample susceptibility has the following form:<br />

χ = αU/mfB (2.3)<br />

The absolute accuracy <strong>of</strong> the susceptibility depends on the accuracy with which all experimental<br />

parameters in the equation above can be determined.<br />

The AC susceptometer allows to measure both the real (in phase) χ’ as well as imaginary<br />

(out <strong>of</strong> phase) χ" component <strong>of</strong> susceptibility. As shown in Figure 2.5, the lock-in<br />

detector requires a reference signal which is at the same frequency <strong>and</strong> in phase with the<br />

current from AC current source. The reference signal serves two purposes. It tunes the<br />

lock-in amplifier to the frequency <strong>of</strong> the reference signal, <strong>and</strong> the lock-in amplifier provides<br />

an output E out which is sensitive to the phase difference Φ between the input signal<br />

E in <strong>and</strong> the reference signal:<br />

E out = E in cos(Φ) (2.4)


2. Experimental techniques 16<br />

The measurement has two contributions to the phase angle Φ. One contribution arises<br />

from the circuit itself. The second contribution to the phase shift arises from the signal<br />

due to the sample. Information about the phase angle Φ can be obtained through the phase<br />

adjust feature on the lock-in, that introduces a phase shift Θ in the reference channel <strong>of</strong><br />

the lock-in. The output is modified as follows:<br />

E out = E in cos(Φ − Θ) (2.5)<br />

"Phasing" a lock-in amplifier refers to the process <strong>of</strong> setting the phase shift Θ equal to Φ.<br />

However, the lock-in amplifier is most accurately phased by adjusting the phase for a zero<br />

output <strong>and</strong> then shifting the phase setting by 90 0 .<br />

The proper separation <strong>of</strong> χ’ <strong>and</strong> χ" requires that the phasing be performed with a<br />

test sample with a known χ"=0 (paramagnetic, insulating sample). Once this phase is<br />

determined, the lock-in amplifier signal measured at Θ will be proportional to χ ′ <strong>and</strong> the<br />

signal measured at Θ + 90 0 will be proportional to χ".<br />

In order to maintain consistency in the data acquisition <strong>and</strong> to guarantee that no information<br />

is lost for future analysis, all dual phase data are measured with the lock-in<br />

amplifier phase set to 0 0 <strong>and</strong> 90 0 . The phase angle Θ is then used in the data analysis to<br />

convert the measured voltages to the equivalent in phase <strong>and</strong> out <strong>of</strong> phase voltage signal:<br />

U ′ = U 0 cos(Θ) + U 90 sin(Θ) (2.6)<br />

U” = U 90 cos(Θ) − U 0 sin(Θ) (2.7)<br />

where U 0 is lock-in voltage at 0 0 , U 90 is lock-in voltage at 90 0 , U’ is in phase voltage<br />

reading for sample (voltage at phase angle Θ), U" is out <strong>of</strong> phase voltage reading for<br />

sample (voltage at phase angle Θ + 90 0 ).<br />

The voltage U’ is then used to determine the measured susceptibility χ ′ :<br />

χ ′ = αU ′ /mfB (2.8)<br />

The imaginary component <strong>of</strong> the measured susceptibility is determined from the following<br />

relationship:<br />

χ” = −αU”/mfB (2.9)<br />

The sign difference arises from the phasing conventions used in the 7229 LakeShore Susceptometer.<br />

The measurement <strong>of</strong> the magnetic moment is performed by using what has traditionally<br />

been called an extraction technique. This terminology is used generally to describe<br />

any method which relies on detecting a flux change as the sample is removed (extracted)<br />

from a sensing coil. The change in flux is then related directly to the moment <strong>of</strong> the<br />

sample.<br />

The configuration <strong>of</strong> 7229 LakeShore Susceptometer is adapted to perform such an<br />

extraction measurement by disabling the AC current source <strong>and</strong> replacing Lock-in Amplifier<br />

with a high-speed integrating digital voltmeter (DVM) (look at the Figure 2.5).<br />

The stepping motor is then used to move the sample between the centers <strong>of</strong> the two secondary<br />

coils. Since the DVM can operate on a much faster time scale than the sample<br />

movement, the output voltage can be recorded. The integral <strong>of</strong> the voltage over time can<br />

then be determined <strong>and</strong> directly related to the moment <strong>of</strong> the sample:<br />

M = kl v (2.10)


2. Experimental techniques 17<br />

where M is magnetic moment, k is DC moment calibration coefficient, l v = ∫ vdt is<br />

voltage integral over time.<br />

The DC moment calibration coefficient (k) is closely related to the AC susceptibility<br />

calibration coefficient (α). Both coefficients relate the flux coupled between a magnetized<br />

sample <strong>and</strong> a sensing coil.<br />

k = πα (2.11)<br />

Multiple "scans" (the single scan is defined as a moving the sample from coil 1 to coil 2<br />

<strong>and</strong> then back to coil 1 again) can be performed <strong>and</strong> averaged to yield measurements with<br />

greater precision.<br />

The mass magnetization or the volume magnetization cam be determined by dividing<br />

the moment by the appropriate quantity.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 2.5: Experimental setup for AC susceptibility/DC magnetization measurements - 7229<br />

LakeShore Susceptometer/Magnetometer system.


2. Experimental techniques 18<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 2.6: The cross-sectional view <strong>of</strong> the coil assembly. The two sensing coils are connected in opposition<br />

in order to cancel the voltages induced by the AC field itself or voltages induced<br />

by unwanted external sources.<br />

2.5 The measurements in the range <strong>of</strong> high pulsed magnetic fields<br />

The Ga 1−x Mn x As epilayers were investigated in the range <strong>of</strong> high magnetic fields. The<br />

pulsed magnetic fields were used to perform DC magneto<strong>transport</strong> measurements (Hall<br />

effect as well as magnetoresistance measurements) as well as magnetization measurements<br />

- by use <strong>of</strong> magnetooptical Kerr effect (MOKE). The high magnetic fields experiments<br />

were performed in LNCMP (Laboratoire National des Champs Magnetiques<br />

Pulses) in Toulouse.<br />

Pulsed field magnetometry uses capacitive discharge to generate high magnetic fields<br />

in conventional resistive solenoids. In principle, this technique employs short measurement<br />

cycles - the duration <strong>of</strong> high magnetc field is short.<br />

All experiments described in the thesis were performed in long pulsed magnetic fields.<br />

The Toulouse facility allows to use the pulsed magnetic fields with the pulse duration ≤<br />

1 second. The magnet is a conventional compact solenoid with uniform current distribution<br />

[9]. The coil performance is determined by mechanical constrains (the accumulated<br />

stress due to the applied magnetic pressure) <strong>and</strong> is limited by heating. The use <strong>of</strong> proper<br />

conductor materials (see [9]) allows to meet the necessary requirements, i.e. maximum<br />

mechanical strength to guarantee the highest possible peak field <strong>and</strong> large specific heat<br />

combined with low resistivity to permit the longest possible pulse duration. The coils are<br />

driven by a 24 kV, 14MJ capacitor bank. The 600 capacitors are divided into 10 modules<br />

that can be used separetely <strong>and</strong> with different polarity. The capacitor bank is discharged<br />

into the coil via stacks <strong>of</strong> optically triggered thyristors. In the present thesis the MOKE


2. Experimental techniques 19<br />

measurements were performed in magnetic fields up to 25T, the magneto<strong>transport</strong> experiments<br />

up to 55T.<br />

2.6 The experimental setup for Kerr effect measurements<br />

The magnetooptical Kerr effect (MOKE) was measured in the high pulsed magnetic fields<br />

(up to 25T) by use <strong>of</strong> the experimental setup shown schematically in Figure 2.7. The<br />

method consists in detection <strong>of</strong> the intensity difference between two orthogonal components<br />

<strong>of</strong> linear polarized light −→ e x <strong>and</strong> −→ e y .<br />

The principle <strong>of</strong> the measurement is as follows. The incident beam <strong>of</strong> light from the<br />

laser (red HeNe laser: λ=632.8 nm, P=5mW or green HeNe laser λ=540.5 nm, P=0.1mW)<br />

first passes through a linear Glan-Taylor polarizator. Next, the mirror placed on the top<br />

<strong>of</strong> the sample holder sends the incident as well as reflected beam <strong>of</strong> light. The sample<br />

reflects the light that comes almost at the normal incidence. After reflection from the<br />

sample, the polarization axis turns about Kerr angle Θ K . Then, the light passes through<br />

the retardation plate ( λ ) that works as a compensator. The Wollaston biprisme separates<br />

2<br />

the beam <strong>of</strong> the light for two spatial orthogonal linear polarizations −→ e x <strong>and</strong> −→ e y . Finally,<br />

the intensity difference between two components: −→ e x <strong>and</strong> −→ e y is measured by means <strong>of</strong> two<br />

silicon photodiodes.<br />

In the Jones-vector representation (see Appendinx1) the linearly polarized (incident)<br />

wave (introduced here as a −→ e x ) has the following form:<br />

( )<br />

1<br />

E i =<br />

(2.12)<br />

After reflection from the sample, the polarization axis turns about Kerr angle Θ K :<br />

( )<br />

cos(Θ<br />

E r =<br />

K )<br />

(2.13)<br />

0<br />

sin(Θ K )<br />

The action <strong>of</strong> the λ retardation plate with the optical axis leaned at the φ angle to the<br />

2<br />

x axis can be written as:<br />

(<br />

) ( ) (<br />

) (<br />

)<br />

cos(φ) − sin(φ) 1 0 cos(φ) sin(φ) cos(2φ) sin(2φ)<br />

=<br />

(2.14)<br />

sin(φ)<br />

cos(φ)<br />

0 −1<br />

− sin(φ)<br />

cos(φ)<br />

sin(2φ)<br />

− cos(2φ)<br />

The Wollaston biprisme separates beam <strong>of</strong> the polarized light for two orthogonal linear<br />

polarizations −→ e x <strong>and</strong> −→ e y , thus<br />

( ) (<br />

)<br />

E x cos(2φ) cos(Θ<br />

=<br />

K )+sin(2φ) sin(Θ K )<br />

(2.15)<br />

E y<br />

sin(2φ) cos(Θ K )−cos(2φ) sin(Θ K )<br />

The two Si PIN photodiodes allow to measure the difference <strong>of</strong> intensities for two<br />

orthogonal linear polarizations −→ e x <strong>and</strong> −→ e y :<br />

∆V = C(|E x | 2 − |E y | 2 ) = C(cos(2Θ K ) cos(4φ) + sin(2Θ K )sin(4φ)) (2.16)<br />

The procedure <strong>of</strong> measurement is as follows. First, at zero magnetic field (i.e. Θ K =0),<br />

the λ plate is oriented (for φ = ± π, φ = ± 3π, φ = ± 5 π...) to obtain:<br />

2 8 8 8<br />

∆V = 0 (2.17)


2. Experimental techniques 20<br />

Next, after turning on the magnetic field the equation 5 has the following form:<br />

∆V = C(sin(2Θ K )) (2.18)<br />

Usually, (for most <strong>of</strong> compounds), the following condition is satisfied Θ K ≪1, then:<br />

Θ K = ∆V<br />

2C<br />

(2.19)<br />

For Θ K =0 <strong>and</strong> φ=0 (or φ = π 2 ) ∆V has maximum value: ∆V=∆V max=C, thus constant<br />

C can be determined. Finally, determined Kerr rotation angle Θ K has the following form:<br />

Θ K =<br />

∆V<br />

2∆V max<br />

(2.20)<br />

The sensitivity <strong>of</strong> the Kerr rotation is <strong>of</strong> order <strong>of</strong> 1·10 −3 deg. Being a reflectivity<br />

measurement, the MOKE is extremely sensitive to any movement <strong>of</strong> the sample <strong>and</strong> the<br />

main difficulty is to cancel all the mechanical vibrations generated by the pulsed magnet.<br />

<br />

LMHCHFGHFBN<br />

4567895:<br />

;


3. SAMPLES OF Pb 1−x−y−z Mn x Eu y Sn z Te AND Ga 1−x Mn x As.<br />

Two types <strong>of</strong> <strong>ferromagnetic</strong> mixed crystals were studied. The bulk crystals <strong>of</strong><br />

Pb 1−x−y−z Mn x Eu y Sn z Te were grown by use <strong>of</strong> the modified Bridgman method in Chernivtsy<br />

Department <strong>of</strong> the Institute <strong>of</strong> Materials Science Problems Ukrainian Academy<br />

<strong>of</strong> Sciences. The thin layers <strong>of</strong> Ga 1−x Mn x As were obtained using the non-equilibrum<br />

growth conditions <strong>of</strong> low-temperature molecular-beam epitaxy (LT-MBE) in Department<br />

<strong>of</strong> Physics University <strong>of</strong> Notre Dame.<br />

3.1 Ferromagnetic bulk crystals <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te<br />

The crystals <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te were grown by the modified Bridgman method.<br />

In the present work the samples coming from several technological processes were investigated.<br />

The chemical composition <strong>of</strong> the samples was determined by X-ray dispersive<br />

fluorescence analysis technique (see section 2 <strong>of</strong> Chapter 2). This technique allows to<br />

determine the chemical composition <strong>of</strong> the samples with uncertainty <strong>of</strong> 10%. Typically,<br />

the crystals were cut crosswise the growth axis to the 1 - 2mm thick slices. The change<br />

<strong>of</strong> the chemical composition along such area is very slight (1-2%). The results <strong>of</strong> chemical<br />

analysis <strong>of</strong> all investigated Pb 1−x−y−z Mn x Eu y Sn z Te samples are gathered in Table<br />

3.1. Figure 3.1 shows the typical chemical composition distribution along the growth<br />

direction <strong>of</strong> the Pb 1−x−y−z Mn x Eu y Sn z Te crystal.<br />

The st<strong>and</strong>ard powder X-ray measurements revealed that investigated samples are<br />

single-phase <strong>and</strong> crystallize in NaCl structure, similarly as a nonmagnetic matrix <strong>and</strong><br />

semimagnetic <strong>semiconductor</strong> - Pb 1−x−y Mn x Sn y Te. It was shown that introduction <strong>of</strong><br />

Mn ions into the nonmagnetic matrix <strong>of</strong> Pb 1−x Sn x Te leads to the decrease <strong>of</strong> the lattice<br />

constant <strong>of</strong> resultant Pb 1−x−y Mn x Sn y Te [10]. The measured values <strong>of</strong> the lattice constant<br />

for several Pb 1−x−y−z Mn x Eu y Sn z Te samples as well as the lattice constant values<br />

<strong>of</strong> Pb 1−x−y Mn x Sn y Te crystals with analogous content <strong>of</strong> Mn <strong>and</strong> Sn content [11] are collected<br />

in Table 3.2. The careful inspection <strong>of</strong> Table 3.2 shows that introduction <strong>of</strong> Eu<br />

ions to Pb 1−x−y Mn x Sn y Te lattice leads to the increase <strong>of</strong> the lattice constant <strong>of</strong> resultant<br />

compound.<br />

3.2 Ferromagnetic layers <strong>of</strong> Ga 1−x Mn x As.<br />

The second investigated system was III-V Mn based Semimagnetic Semiconductor -<br />

GaMnAs. The layers <strong>of</strong> Ga 1−x Mn x As studied in the thesis were grown by use <strong>of</strong> low<br />

temperature (LT) MBE with elemental sources Ga, Mn, As, without intentional doping.<br />

Semi-insulating epiready (100) GaAs wafers were used as the substrates. Typically, a<br />

buffer <strong>of</strong> GaAs was first grown at high temperature (600 0 C). The substrate was then<br />

cooled to a temperatures in the range 250 0 C - 285 0 C, <strong>and</strong> a layer <strong>of</strong> low temperature (LT)


3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. 22<br />

Tab. 3.1: The chemical composition <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples determined by means<br />

<strong>of</strong> X-ray dispersive fluorescence analysis technique.<br />

sample # <strong>of</strong> the sample x P b x Mn x Eu x Sn<br />

SnMnEuTe 841_14 - 0.116 0.011 0.873<br />

SnMnEuTe 841_18 - 0.131 0.134 0.735<br />

SnMnEuTe 842_4 - 0.063 0.0045 0.932<br />

SnMnEuTe 842_8 - 0.068 0.003 0.929<br />

SnMnEuTe 842_14 - 0.070 0.007 0.923<br />

SnMnEuTe 842_20 - 0.091 0.009 0.900<br />

SnMnEuTe 848_4 - 0.061 0.0115 0.927<br />

SnMnEuTe 848_10 - 0.064 0.012 0.924<br />

SnMnEuTe 848_16 - 0.050 0.011 0.939<br />

SnMnEuTe 848_22 - 0.065 0.018 0.917<br />

SnMnEuTe 848_24 - 0.051 0.019 0.930<br />

SnMnEuTe 848_26 - 0.074 0.023 0.903<br />

PbSnMnEuTe 809_2 0.116 0.031 0.0027 0.850<br />

PbSnMnEuTe 809_4 0.118 0.030 0.0016 0.850<br />

PbSnMnEuTe 809_10 0.187 0.030 0.0031 0.780<br />

PbSnMnEuTe 809_12 0.215 0.022 0.003 0.760<br />

PbSnMnEuTe 809_28 0.243 0.020 0.007 0.730<br />

PbSnMnEuTe 809_30 0.256 0.024 0.010 0.710<br />

PbSnMnEuTe 809_32 0.27 0.026 0.014 0.690<br />

PbSnMnEuTe 809_34 0.276 0.027 0.017 0.680<br />

PbSnMnEuTe 809_36 0.272 0.025 0.013 0.690<br />

PbMnEuTe 793_2 0.990 0.010 0.000 -<br />

PbMnEuTe 793_4 0.989 0.010 0.001 -<br />

PbMnEuTe 793_6 0.982 0.009 0.009 -<br />

PbMnEuTe 793_10 0.910 0.005 0.004 -<br />

PbMnEuTe 793_12 0.990 0.007 0.003 -<br />

Tab. 3.2: The lattice constant a 0 <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples determined by the st<strong>and</strong>ard<br />

powder X-ray measurements <strong>and</strong> the values <strong>of</strong> the lattice constant <strong>of</strong> Pb 1−x−y Mn x Sn y Te<br />

a [11] with similar content <strong>of</strong> Mn <strong>and</strong> Sn as for the samples investigated in the thesis.<br />

# <strong>of</strong> the sample x P b x Mn x Eu x Sn a 0 a ∆ a=a 0 -a ∆ a/a 0<br />

[Å] [Å] [Å] %<br />

809_2 0.116 0.031 0.0027 0.85 6.3130 6.2866 0.0264 0.42<br />

809_4 0.118 0.030 0.0016 0.85 6.3237 6.2876 0.0361 0.57<br />

809_12 0.215 0.022 0.003 0.76 6.3375 6.3113 0.0262 0.41<br />

809_28 0.243 0.020 0.007 0.73 6.3563 6.3309 0.0254 0.40<br />

809_36 0.272 0.025 0.013 0.69 6.3427 6.3311 0.0116 0.18


3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. 23<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

N<br />

Fig. 3.1: Chemical composition distribution along the crystal growth direction for the crystal <strong>of</strong><br />

PbSnMnEuTe.


3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. 24<br />

<br />

<br />

<br />

<br />

! <br />

" <br />

!#<br />

<br />

<br />

<br />

Fig. 3.2: RHEED oscillations observed during the growth <strong>of</strong> a Ga 1−x Mn x As film with x = 0.062.<br />

The first 7 periods correspond to LT-GaAs. The "jump" in the signal occurs at the point<br />

when the Mn shutter has been opened <strong>and</strong> the rate <strong>of</strong> oscillations increased.<br />

GaAs was grown to a thickness in the range between 2nm -100nm. Finally, Ga 1−x Mn x As<br />

layer in the same range <strong>of</strong> substrate temperatures to a thickness <strong>of</strong> the 105nm - 302nm was<br />

grown. No special precaution was needed at the start <strong>of</strong> Ga 1−x Mn x As growth. However,<br />

the <strong>properties</strong> <strong>of</strong> grown Ga 1−x Mn x As do depend on growth parameters as As overpressure<br />

<strong>and</strong> T S . The growth was monitored in situ by reflection high energy electron diffraction<br />

(RHEED).<br />

The determination <strong>of</strong> Mn content in Ga 1−x Mn x As epilayers is quite difficult task. The<br />

Mn concentration x was determined using two different methods. First, during the growth<br />

the x values were estimated from the change in the growth rate monitored by RHEED<br />

oscillations after the Mn shutter was opened. An example <strong>of</strong> such data is shown in Figure<br />

3.2.<br />

Note the rate <strong>of</strong> growth measured by RHEED oscillations is in terms <strong>of</strong> atomic layers<br />

per second, <strong>and</strong> after the Mn shutter is opened it increases in proportion to precisely that<br />

fraction <strong>of</strong> the Mn flux which is required to completion <strong>of</strong> atomic layers as the growth<br />

proceeds. It is thus assumed that RHEED oscillations provide a measure <strong>of</strong> the concentration<br />

<strong>of</strong> substitutional Mn cations Mn Ga , since only these only are required to complete<br />

the formation <strong>of</strong> atomic layers.<br />

And second, the Mn content was obtained from X-ray diffraction measurements by<br />

assuming that the GaMnAs layer is fully strained by the GaAs substrate. The Mn concentration<br />

x was calculated from measured relaxed layer lattice constant (XRD measurements


3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. 25<br />

Tab. 3.3: The parameters <strong>of</strong> the LT MBE growth <strong>of</strong> GaMnAs samples - substrate temperature T S<br />

<strong>and</strong> temperature <strong>of</strong> the Mn effusion cell T Mn , thickness <strong>of</strong> the GaMnAs layers d GaMnAs<br />

determined from RHEED oscillations <strong>and</strong> Mn composition <strong>of</strong> investigated samples x (determined<br />

from RHEED oscillations, X-ray diffraction <strong>and</strong> high resolution X-ray diffraction<br />

measurements<br />

# <strong>of</strong> the sample T S T Mn d GaMnAs x x x<br />

[ 0 C] [ 0 C] [nm] (RHEED) (XRD) (HRXRD)<br />

GaMnAs/GaAs<br />

00811A 285 780 302 - 0.01 -<br />

GaMnAs/GaAs<br />

00119C 275 880 269 - 0.032 -<br />

GaMnAs/GaAs<br />

10727C 270 870 131 - 0.027 0.027<br />

GaMnAs/GaAs<br />

10727D 270 900 149 0.062 0.061 0.056<br />

GaMnAs/GaAs<br />

10727E 265 920 105 0.086 0.084 0.078<br />

GaMnAs/GaAs<br />

10823C 265 920 111 0.07 0.082 -<br />

GaMnAs/GaAs<br />

10823E 250 925 115 0.093 0.085 -<br />

GaMnAs/GaAs<br />

10529A 275 820 300 0.014 - -<br />

GaMnAs/GaAs<br />

11127A 275 - 220 0.048 - -<br />

performed at the Notre Dame University) by use <strong>of</strong> the following equation [12]: a Lrelax<br />

= 5.6547 + 0.0002433*x. The presence <strong>of</strong> Mn interstitials atoms as well as antisite defects<br />

can be the reason for the observed expansion <strong>of</strong> the lattice constant <strong>of</strong> GaMnAs.<br />

The results <strong>of</strong> high resolution X-ray diffraction measurements (HRXRD) (performed in<br />

the Institute <strong>of</strong> Physics Polish Academy <strong>of</strong> Sciences) <strong>and</strong> the effect <strong>of</strong> Mn interstitials on<br />

the lattice parameter <strong>of</strong> Ga 1−x Mn x As will be discussed in details in Chapter IV. In fact, a<br />

method <strong>of</strong> determining the Mn concentration x based on the measurements <strong>of</strong> the lattice<br />

constant a 0 in Ga 1−x Mn x As is not very reliable. The Mn concentration determined from<br />

RHEED oscillations, the results <strong>of</strong> XRD <strong>and</strong> HRXRD <strong>and</strong> details <strong>of</strong> the growth conditions<br />

<strong>of</strong> Ga 1−x Mn x As are collected in Table 3.3.<br />

Additionally, the Mn concentration for the sample 10823C was confirmed by use <strong>of</strong><br />

systematic particle-induced X-ray emission (PIXE) measurements [13]. The PIXE measurements<br />

revealed that sample with x determined by RHEED as 0.07 has total <strong>of</strong> Mn<br />

content equal to 0.092. The PIXE results show the total Mn content - substitutional, interstitial,<br />

<strong>and</strong> in the form <strong>of</strong> r<strong>and</strong>om precipitates (Mn inclusions) <strong>and</strong> are higher than the<br />

values obtained from RHEED oscillations.<br />

The Mn concentration specified in the next Chapters comes from the RHEED oscillations<br />

measurements with the exception <strong>of</strong> three samples with the lowest Mn content for


3. Samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As. 26<br />

which the change in the RHEED oscillations was too small to be reliable. In this case, i.e.<br />

00811A, 00119C, 10727C samples the Mn content was determined only by use <strong>of</strong> X-ray<br />

diffraction technique.


4. TRANSPORT AND MAGNETIC INVESTIGATIONS OF<br />

FERROMAGNETIC Pb 1−x−y−z Mn x Eu y Sn z Te<br />

4.1 Introduction<br />

One <strong>of</strong> the purposes <strong>of</strong> the studies presented in the thesis were magnetic <strong>and</strong> <strong>transport</strong><br />

investigations <strong>of</strong> <strong>ferromagnetic</strong> Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals. In particular,<br />

the influence <strong>of</strong> the presence <strong>of</strong> two types <strong>of</strong> magnetic ions incorporated into <strong>semiconductor</strong><br />

matrix on magnetic <strong>properties</strong> <strong>of</strong> resultant semimagnetic <strong>semiconductor</strong> is analyzed.<br />

There are several reasons for which the semimagnetic <strong>semiconductor</strong>s (SMSC’s)<br />

based on lead chalcogenides are ideal materials for such kind <strong>of</strong> investigations. The<br />

variety <strong>of</strong> magnetic <strong>properties</strong> occurring in IV-VI SMSC, e.g., the carrier concentration<br />

induced paramagnet-ferromagnet <strong>and</strong> ferromagnet-spin glass transition observed in<br />

Pb 1−x−y Mn x Sn y Te [1], [2] makes this system particularly attractive for such purposes.<br />

The non-trivial advantage <strong>of</strong> IV-VI materials is also relative simplicity <strong>of</strong> crystal growing<br />

<strong>and</strong> carrier concentration controlling – the latter may be achieved by means <strong>of</strong> either<br />

doping or isothermal annealing. The magnetic <strong>properties</strong> <strong>of</strong> these compounds depend<br />

not only on the concentration <strong>of</strong> manganese ions, but also on the density <strong>of</strong> free carriers<br />

[14]. This behaviour is due to the combination <strong>of</strong> an RKKY type <strong>of</strong> interaction between<br />

the magnetic ions as well as the possibility to manipulate the free carrier concentration.<br />

The additional advantage is that for Pb 1−x−y Mn x Sn y Te crystals are very well known parameters<br />

<strong>of</strong> crystal <strong>and</strong> energy structure. In order to simplify theoretical description <strong>of</strong><br />

investigated magnetic system, two types <strong>of</strong> magnetic ions were choosen with spin-only<br />

ground state: substitutional Mn 2+ possesses S = 5/2, while Eu 2+ , the second ion in our<br />

samples, has S = 7/2.<br />

Practically all IV-VI semimagnetic <strong>semiconductor</strong>s crystallize in rock salt crystal<br />

structure. The lattice parameter a 0 changes linearly with the content <strong>of</strong> magnetic ions<br />

following the Vegard law.<br />

In general, all IV-VI semimagnetic <strong>semiconductor</strong>s show metallic type <strong>of</strong> conductivity<br />

with a very large, temperature independent, concentration <strong>of</strong> carriers. However, under<br />

special conditions, IV-VI based semimagnetic <strong>semiconductor</strong>s can exhibit insulating<br />

<strong>properties</strong> – recently, the Eu composition induced metal-insulator transition was observed<br />

in epitaxial layers <strong>of</strong> Pb 1−x Eu x Te [15]. Carriers are generated by metal vacancies, <strong>and</strong><br />

their concentration can be controlled by thermal annealing or doping. In semimagnetic<br />

lead chalcogenides with Mn or with Eu the range <strong>of</strong> carrier concentration <strong>and</strong> the methods<br />

to control it are quite similar to the case <strong>of</strong> appropriate IV-VI <strong>semiconductor</strong>s. The presence<br />

<strong>of</strong> even 10 at. % <strong>of</strong> Mn or Eu ions has practically no effect on carrier concentration.<br />

Mn <strong>and</strong> Eu ions are electrically inactive in semimagnetic lead chalcogenides.<br />

IV-VI materials are narrow gap <strong>semiconductor</strong>s. Qualitatively, the electron b<strong>and</strong> structure<br />

is analogous to the b<strong>and</strong> structure <strong>of</strong> non-magnetic counterpart materials. A b<strong>and</strong>structure<br />

model based on the consistent interpretation <strong>of</strong> <strong>transport</strong>, optical, <strong>and</strong> magnetic


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 28<br />

Fig. 4.1: The b<strong>and</strong> structure model <strong>of</strong> Pb 1−x−y Mn x Sn y Te mixed crystals<br />

experimental data [16], [17], [18] is presented in Figure 4.1.<br />

The b<strong>and</strong> <strong>of</strong> electrons <strong>and</strong> the b<strong>and</strong> <strong>of</strong> light holes (the presence <strong>of</strong> further L b<strong>and</strong>s is<br />

not included here) are separated by a direct b<strong>and</strong> at the L point <strong>of</strong> Brillouin zone. The<br />

energy dispertion relations <strong>of</strong> electrons <strong>and</strong> holes are nonparabolic <strong>and</strong> anisotropic. There<br />

are four equivalent valleys <strong>of</strong> both the b<strong>and</strong> <strong>of</strong> electrons <strong>and</strong> the b<strong>and</strong> <strong>of</strong> light holes.<br />

Due to the narrow energy gap the energy dispertion relation is nonparabolic <strong>and</strong> usually<br />

described within Dimmock model [19]. The energy gap <strong>of</strong> lead chalcogenides increases<br />

rapidly with the content increase <strong>of</strong> Mn <strong>and</strong> Eu [20], [14]. In most <strong>of</strong> IV-VI semimagnetic<br />

<strong>semiconductor</strong>s the composition dependence <strong>of</strong> other b<strong>and</strong> parameters can be neglected.<br />

An increase <strong>of</strong> the energy gap with increasing temperature is observed, similarly to lead<br />

chalcogenides. Approximetely E Σ = 0.2 – 0.4 eV below the top <strong>of</strong> the b<strong>and</strong> <strong>of</strong> light holes<br />

there is a second valence b<strong>and</strong> <strong>of</strong> heavy holes. The top <strong>of</strong> this b<strong>and</strong> is located at the Σ<br />

point <strong>of</strong> the Brillouin zone <strong>and</strong> there are 12 equivalent energy valleys <strong>of</strong> this b<strong>and</strong> (Σ<br />

b<strong>and</strong>). Since the direct energy gap at the Σ point <strong>of</strong> the Brillouin zone is quite large, the<br />

heavy hole b<strong>and</strong> is expected to be parabolic. The electronic <strong>properties</strong> <strong>of</strong> p-type IV-VI<br />

semimagnetic <strong>semiconductor</strong>s with very high concentration <strong>of</strong> carriers (p ≥ 5·10 19 cm −3 )<br />

are influenced by the presence <strong>of</strong> the b<strong>and</strong> <strong>of</strong> heavy holes. The Σ b<strong>and</strong> is essential for<br />

the underst<strong>and</strong>ing <strong>of</strong> the correlations between magnetic <strong>and</strong> <strong>transport</strong> <strong>properties</strong> <strong>of</strong> IV-VI<br />

semimagnetic <strong>semiconductor</strong>s. The effective mass <strong>of</strong> the carriers in the Σ b<strong>and</strong> is much<br />

higher than that in the L b<strong>and</strong> (m ∗ Σ ≈ 1.7m e [18], m ∗ L ≈ 0.05m e [21]).<br />

Mn-based IV-VI semimagnetic <strong>semiconductor</strong>s can be divided in two groups. The<br />

first group consists <strong>of</strong> the materials with relatively low carrier concentration <strong>of</strong> free carriers<br />

[22] (in the range 10 17 – 10 19 cm −3 ), for instance Pb 1−x Mn x Te. From a magnetic point<br />

<strong>of</strong> view these materials are paramagnets above T=1K. Their magnetic behaviour closely<br />

resambles that <strong>of</strong> the Mn containing II-VI SMSC’s <strong>and</strong> can also be attributed to anti<strong>ferromagnetic</strong><br />

interactions <strong>of</strong> the superexchange type, although the interactions are much<br />

weaker than in II-VI semimagnetic <strong>semiconductor</strong>s. Other interspin interaction mechanisms<br />

(e.g. direct exchange or the RKKY interaction) are expected to be negligible due


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 29<br />

Fig. 4.2: Curie-Weiss temperature (Θ) versus free carrier concentration in Pb 1−x−y Mn x Sn y Te.<br />

to the large mean interspin distances <strong>and</strong> low concentration <strong>of</strong> carriers. Below T=1K a<br />

spin-glass phase was reported in Pb 1−x Mn x Te [23]. The second group <strong>of</strong> IV-VI SMSC’s<br />

consists <strong>of</strong> the materials with relatively high charge carrier concentrations, <strong>of</strong> order <strong>of</strong><br />

10 20 – 10 21 cm −3 , e.g. Pb 1−x−y Mn x Sn y Te (with low Pb content). These compounds<br />

exhibit a <strong>ferromagnetic</strong> phase transition at low temperatures [1]. The <strong>ferromagnetic</strong> interactions<br />

can be explained by RKKY interactions, made effective by the high carrier<br />

concentration <strong>and</strong> dominating over the superexchange interactions. The <strong>ferromagnetic</strong><br />

phase occurs once the holes start to occupy site b<strong>and</strong>s with a large effective mass. The<br />

RKKY interaction [24], [25], [26] is an indirect interaction between the magnetic ions,<br />

which is mediated by the free charge carriers. The interaction strength can be written as:<br />

J RKKY (R ij ) = N m∗ J 2 sd a6 0k 4 F<br />

32π 3¯h 2 [ sin(2k F R ij ) − 2k F R ij cos(2k F R ij )<br />

(2k F R ij ) 4 ] (4.1)<br />

where k F is the Fermi wave number, m ∗ the effective mass <strong>of</strong> the carriers, J sd the Mn ionelectron<br />

exchange integral, a 0 the lattice constant, N the number <strong>of</strong> valleys <strong>of</strong> the valence<br />

b<strong>and</strong>, R ij the distance between the magnetic ions.<br />

Story et. al. [1] showed that magnetic behaviour <strong>of</strong> Pb 1−x−y Mn x Sn y Te strongly depends<br />

on the concentration <strong>of</strong> free carriers. Figure 4.2 shows the Curie-Weiss temperature<br />

(Θ) versus free carrier concentration as reported by Story et. al. for Pb 0.25 Mn 0.03 Sn 0.72 Te.<br />

The nonzero Curie-Weiss temperature is proportional to the sum <strong>of</strong> all magnetic interactions<br />

present in the material. The characteristic feature <strong>of</strong> the T c (p) dependence is<br />

the existance <strong>of</strong> a certain threshold carrier concentration p = p t ≃ 3·10 20 cm −3 , above<br />

which the IV-VI semimagnetic <strong>semiconductor</strong>s show <strong>ferromagnetic</strong> <strong>properties</strong>. For carrier<br />

concentration lower than the threshold value p t , the crystals exhibit paramagnetic<br />

<strong>properties</strong> (similarly to low carrier concentration materials like PbMnTe). The observation<br />

<strong>of</strong> concentration dependence <strong>of</strong> Curie temperature has found an interpretation within<br />

the frames <strong>of</strong> the RKKY mechanism <strong>and</strong> the two valence b<strong>and</strong> model <strong>of</strong> the b<strong>and</strong> structure<br />

<strong>of</strong> PbSnMnTe <strong>and</strong> SnMnTe [16], [17]. The strength <strong>of</strong> the RKKY interaction scales<br />

with the effective mass <strong>of</strong> carriers. The RKKY interaction is expected to become strongly<br />

enhanced for p ≥ p t , when the Fermi level enters the Σ b<strong>and</strong> <strong>and</strong> heavy holes start to


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 30<br />

participate in charge <strong>transport</strong> <strong>and</strong> in RKKY interaction. Next to the effective mass <strong>of</strong> the<br />

carriers, the degeneracy <strong>of</strong> the valence b<strong>and</strong> maxima is also important (as a prefactor <strong>of</strong><br />

RKKY interaction). The L b<strong>and</strong> is four-fold degenerate (N L =4), whereas the Σ b<strong>and</strong> is<br />

twelve-fold degenerate (N Σ =12)<br />

The RKKY interaction is also responsible for the magnetic behaviour <strong>of</strong> canonical<br />

metallic spin-glasses like CuMn. Because <strong>of</strong> the high carrier concentration <strong>of</strong> carriers in<br />

these materials (10 23 cm −3 ) the interaction rapidly oscillates between <strong>ferromagnetic</strong> <strong>and</strong><br />

anti<strong>ferromagnetic</strong> as a function <strong>of</strong> the distance between two Mn ions. The period <strong>of</strong> this<br />

oscillation is short compared to the average Mn-Mn distance. Due to the r<strong>and</strong>om position<br />

<strong>of</strong> the Mn-ions, the interaction will be <strong>ferromagnetic</strong> or anti<strong>ferromagnetic</strong> at r<strong>and</strong>om. This<br />

causes a frustration <strong>of</strong> the spins resulting in a spin-glass state. The ferromagnet/spinglass<br />

phase transition is also observed in the case <strong>of</strong> IV-VI semimagnetic <strong>semiconductor</strong>s<br />

(Pb 1−x−y Mn x Sn y Te, Sn 1−x Mn x Te) [27], [28]. This effect can be explained in frame <strong>of</strong><br />

RKKY interaction. For the ferromagnetism the following condition should be fullfiled:<br />

R ≤ R 0 , where R is average interspin distance <strong>and</strong> R 0 is characteristic distance R 0 ∼<br />

1/k F ∼ 1/p 1/3 (in this case R 0 corresponds to the first switch <strong>of</strong> the RKKY interaction<br />

from <strong>ferromagnetic</strong> to anti<strong>ferromagnetic</strong>). For R ≥ R 0 the oscillatorty character <strong>of</strong> the<br />

RKKY interaction is expected <strong>and</strong> leads to the spin-glass order.<br />

All Eu based IV-VI semimagnetic lead chalcogenides with low carrier concentartions<br />

(n,p ≤ 10 19 cm −3 are paramagnetic down to about T = 1K (similarly to Mn based compounds).<br />

The very localized character <strong>of</strong> 4f orbitals <strong>of</strong> rare earth results in very weak<br />

exchange intreractions both between magnetic ions <strong>and</strong> between magnetic ions <strong>and</strong> free<br />

carriers. The crystals <strong>of</strong> Sn 1−x Eu x Te are not <strong>ferromagnetic</strong>. The reason <strong>of</strong> lack <strong>of</strong> ferromagnetism<br />

in this material is the very small sp–f exchange integral. It was experimentally<br />

established that the J sf carrier – magnetic moment exchange constants in SnTe<br />

with rare earths ions are related to the exchange constant for Mn ions in the following<br />

way: Jsf Gd Mn<br />

/Jsd<br />

= 1/5 <strong>and</strong> Jsf Eu Mn<br />

/Jsd<br />

= 1/8 [29]. It results in 1/25 <strong>and</strong> 1/64 reduction <strong>of</strong><br />

the strength <strong>of</strong> the RKKY interaction for Gd <strong>and</strong> Eu based IV-VI SMSC’s making this<br />

interaction negligible in these materials.<br />

4.2 Transport characterization <strong>of</strong> Pb 1−x−y−x Mn x Eu y Sn z Te samples<br />

All the investigated Pb 1−x−y−z Mn x Eu y Sn z Te samples were characterized by means <strong>of</strong> low<br />

magnetic field <strong>transport</strong> measurements. The aim <strong>of</strong> the <strong>transport</strong> characterization was to<br />

obtain information about the elementary electric <strong>properties</strong> <strong>of</strong> the investigated samples:<br />

type as well as concentration <strong>of</strong> free carriers <strong>and</strong> their mobility. The Hall voltage V H<br />

as well as conductivity voltage V σ were measured. In the case <strong>of</strong> IV-VI semimagnetic<br />

<strong>semiconductor</strong>s the carrier concentration is important parameter since the change <strong>of</strong> the<br />

carrier concentration influences the magnetic behavior <strong>of</strong> the material.<br />

The Hall bar samples with typical dimensions <strong>of</strong> 8mm × 2mm × 1mm were used for<br />

the <strong>transport</strong> measurements. The electrical contacts were prepared always in the same<br />

way. First the surface <strong>of</strong> the specimens was etched using the solution <strong>of</strong> Br 2 <strong>and</strong> HBr<br />

in the proportion 1 : 20. Next, the gold contacts were deposited by use <strong>of</strong> gold chloride<br />

water solution on the polished surface <strong>of</strong> the samples. Finally, the electrical contacts were<br />

made using indium solder <strong>and</strong> gold wires. Typical resistance <strong>of</strong> the samples was equal to<br />

1 mΩ. This allowed to apply relatively large current (up to 300mA). The Hall as well as


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 31<br />

conductivity measurements were performed at the room <strong>and</strong> liquid nitrogen temperature.<br />

The st<strong>and</strong>ard DC six probe technique at the static magnetic field up to 1T was used. All<br />

the investigated samples occurred to be p type.<br />

In the present thesis the nominal hole concentration was determined:<br />

p = 1<br />

(4.2)<br />

eR H<br />

where R H is the Hall constant.<br />

The nominal value <strong>of</strong> the hole concentrations results from the value <strong>of</strong> the Hall constant<br />

assuming that Hall coefficient r H is equal to 1. This assumption, widely applied in<br />

the literature is not precise <strong>and</strong> the nominal Hall concentration p is very <strong>of</strong>ten not equal<br />

to the real Hall concentration p 0 :<br />

p 0 =<br />

r H<br />

(4.3)<br />

eR H<br />

where r H is the Hall scattering factor.<br />

It is well known (see e.g. [19], [21], [30]) that the carrier <strong>transport</strong> in the IV-VI<br />

materials is served through the two hole types: light holes p l <strong>and</strong> heavy holes p h . The<br />

b<strong>and</strong> structure model that is usually used to analyze the <strong>transport</strong> effects in PbTe, SnTe<br />

<strong>and</strong> PbMnSnTe crystals with the high carrier concentration is schematically shown in<br />

Figure 4.1 <strong>and</strong> described in section 1 <strong>of</strong> this Chapter. The low content <strong>of</strong> europium ions<br />

in the studied samples (see Table 3.1) allows to assume that this model also describe b<strong>and</strong><br />

structure <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te crystals. The light hole b<strong>and</strong> (L point <strong>of</strong> Brillouin<br />

zone) is separated from the conduction b<strong>and</strong> by the energy gap E 0 ∼ 300 meV. About E Σ<br />

∼ 200meV - 400meV below the top <strong>of</strong> the valence b<strong>and</strong> the heavy holes b<strong>and</strong> is located<br />

(Σ point <strong>of</strong> the Brillouin zone).<br />

Two-carrier <strong>transport</strong> is the reason for the discrepancy between the calculated nominal<br />

hole concentration (by use <strong>of</strong> Equation 4.2) <strong>and</strong> the real value <strong>of</strong> hole concentration in<br />

the investigated samples. In the case <strong>of</strong> the carrier <strong>transport</strong> via light <strong>and</strong> heavy holes,<br />

Hall constant is the function <strong>of</strong> both light <strong>and</strong> heavy holes carrier concentration, mobility<br />

<strong>and</strong> Hall scattering factor:<br />

R H = p le l r l µ 2 l + p he h r h µ 2 h<br />

(p l e l µ l + p h e h µ h ) 2 (4.4)<br />

where p l , p h - light <strong>and</strong> heavy hole concentration e l , e h - electric charge <strong>of</strong> carriers r l , r h -<br />

Hall scattering factors <strong>of</strong> carriers µ l , µ h - mobilities <strong>of</strong> carriers.<br />

If both the light <strong>and</strong> heavy holes carry the same elementary charge (+e), Equation 4.4<br />

has the following form:<br />

R H = r lp l + b 2 p h r h<br />

e(p l + bp h ) 2 (4.5)<br />

where b = µ h /µ l .<br />

The Hall scattering factors r l <strong>and</strong> r h in the Equation 4.4 depend on both statistical<br />

energy distribution <strong>of</strong> carriers <strong>and</strong> b<strong>and</strong> structure anisotropy:<br />

r H = r τ r a (4.6)<br />

where r τ - st<strong>and</strong>ard Hall factor that takes into consideration statistical energy distribution<br />

<strong>of</strong> carriers r a - term related to the anisotropy.


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 32<br />

In the strong degenerate systems energy <strong>of</strong> carriers that participate in <strong>transport</strong> is equal<br />

to Fermi energy (delta type <strong>of</strong> statistical energy distribution) <strong>and</strong> r τ =1. For the materials<br />

with anisotropic effective mass or anisotropic time relaxation:<br />

where<br />

r a =<br />

3K(K + 2)<br />

(2K + 1) 2 (4.7)<br />

K = (m ‖ /m ⊥ )(τ ⊥ /τ ‖ ) (4.8)<br />

Anisotropic contribution to the Hall factor (r a ) decreases from r a =1 for K=1 to r a =0.75<br />

for K≫1.<br />

The real hole concentration in the sample can be determined by use <strong>of</strong> the following<br />

equation:<br />

p 0 = r ∗ Hp (4.9)<br />

where<br />

rH ∗ = (p lr l + b 2 p h )(p l + p h )<br />

(4.10)<br />

(p l + bp h ) 2<br />

<strong>and</strong><br />

p 0 = p l +p h - real hole concentration in the sample<br />

p = 1/(e R H ) - nominal hole concentration<br />

b = µ h /µ l - heavy hole mobility to light hole mobility ratio<br />

p l , p h - light <strong>and</strong> heavy hole concentration<br />

r l = 3K(K+2)/(2K+1) 2 - Hall scattering factor for light holes, K = 10 - anisotropy coefficient<br />

for light hole b<strong>and</strong>.<br />

To determine r ∗ H coefficient, that is function <strong>of</strong> hole concentration in the sample, one<br />

needs to know the light as well as heavy hole concentration <strong>and</strong> ratio <strong>of</strong> their mobilities.<br />

In the present thesis only nominal Hall concentration was determined at the room<br />

<strong>and</strong> liquid nitrogen temperature <strong>and</strong> all experimental data are shown as a function <strong>of</strong><br />

nominal concentration (p). The nominal Hall concentration is commonly used in the<br />

characterization <strong>of</strong> IV-VI compounds <strong>and</strong> is unambiguous experimental parameter.<br />

All the investigated samples occurred to be p type with the high almost temperature<br />

independent hole concentration (in the range between 2·10 18 cm −3 <strong>and</strong> 2·10 21 cm −3 ).<br />

The obtained values <strong>of</strong> hole concentration, conductivity <strong>and</strong> mobility measured at the<br />

room <strong>and</strong> nitrogen temperature are shown in Table 4.1. Typical values <strong>of</strong> the Hall voltage<br />

were equal from several to several dozen microvolts. Simultaneously, the large values<br />

<strong>of</strong> the asymmetry voltage (resulting from non equipotential positions <strong>of</strong> the Hall probes),<br />

exceeding 100µV were observed. The asymmetry voltage as well as influence <strong>of</strong> the magnetoresistance<br />

on the Hall effect was eliminated by the st<strong>and</strong>ard averaging procedure <strong>of</strong><br />

results obtained for the combination <strong>of</strong> two current as well as magnetic field polarizations.<br />

Additionally, using <strong>of</strong> the Keithley 150B voltometer allowed to reset the asymmetry voltage<br />

at the magnetic field equal to zero. The values <strong>of</strong> carrier concentration, conductivity<br />

<strong>and</strong> mobility were determined with the uncertainty <strong>of</strong> 15 percent at the room temperature<br />

<strong>and</strong> 30 percent at the liquid nitrogen temperature. One <strong>of</strong> the investigated samples 809 − 30<br />

was isothermally annealed, to increase hole concentration. The procedure <strong>of</strong> annealing<br />

performed at the telluride atmosphere <strong>and</strong> temperature <strong>of</strong> 700 0 C for 48 hours allowed to<br />

increase hole concentration from 4.22·10 20 cm −3 to 8.25 10 25 cm −3 . The surface <strong>of</strong> the<br />

sample was polished after annealing.


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 33<br />

Tab. 4.1: The results <strong>of</strong> <strong>transport</strong> characterization <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples - hole concentration<br />

p [10 21 cm −3 ], conductivity σ[(Ωcm) −1 ], mobility µ [cm 2 /Vs]) measured at<br />

the room <strong>and</strong> liquid nitrogen temperature.<br />

sample x P b x Mn x Eu x Sn p σ µ p σ µ<br />

number 300K 300K 300K 77K 77K 77K<br />

841_14 - 0.116 0.011 0.873 1.40 3554 15.8 0.93 3029 20.7<br />

841_18 - 0.131 0.134 0.735 1.30 3683 15.5 1.04 5598 33.3<br />

842_4 - 0.063 0.0045 0.932 1.40 1172 5.2 - - -<br />

842_8 - 0.068 0.003 0.929 1.56 3500 14.0 - - -<br />

842_14 - 0.070 0.007 0.923 1.56 1933 7.7 1.25 2750 13.7<br />

842_20 - 0.091 0.009 0.9 1.21 3480 17.9 1.16 4561 24.6<br />

848_4 - 0.061 0.0115 0.927 1.77 4533 16.0 1.70 6648 24.7<br />

848_10 - 0.064 0.012 0.924 1.93 1696 5.5 - - -<br />

848_16 - 0.050 0.011 0.939 1.24 6435 32.3 1.48 9635 40.60<br />

848_22 - 0.065 0.018 0.917 1.37 3557 16.2 - - -<br />

848_24 - 0.051 0.019 0.93 1.57 3216 17.4 1.29 4842 23.5<br />

848_26 - 0.074 0.023 0.903 1.66 6370 23.9 - - -<br />

809_2 0.116 0.031 0.0027 0.85 1.01 5860 29 1.01 19050 76<br />

809_4 0.118 0.030 0.0016 0.85 0.501 4050 48.6 0.657 596 30<br />

809_10 0.187 0.030 0.0031 0.78 0.601 2920 29 3.0 4983 35<br />

809_12 0.215 0.022 0.003 0.76 0.401 3390 49 0.60 6764 69<br />

809_30 0.256 0.024 0.010 0.71 0.425 1310 19.2 0.762 2058 16.9<br />

809_30 0.256 0.024 0.010 0.71 0.822 3346 25.4 1.04 5307 31.8<br />

anneal.<br />

809_32 0.27 0.026 0.014 0.69 0.325 2777 53.5 0.49 4731 35.2<br />

809_34 0.276 0.027 0.017 0.68 0.449 2970 41.4 0.98 2373 15.2<br />

809_36 0.272 0.025 0.013 0.69 0.318 2823 55.4 0.75 4816 40<br />

793_2 0.990 0.010 0.000 - 0.002 202 540 0.003 1502 2641<br />

793_4 0.989 0.010 0.001 - 0.003 112 263 0.004 807 1399<br />

793_6 0.982 0.009 0.009 - 0.002 72 75 0.003 723 1329<br />

793_10 0.910 0.005 0.004 - 0.005 11 14 - - -<br />

793_12 0.990 0.007 0.003 - 0.005 1910 24.5 0.007 3076 24.4


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 34<br />

4.3 <strong>Magnetic</strong> measurements <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals<br />

In the present section the results <strong>of</strong> magnetic studies <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples<br />

will be presented. As was shown in the previous section all the investigated samples occured<br />

to be p-type with practically temperature independent concentration <strong>of</strong> carriers.<br />

Two groups <strong>of</strong> IV-VI semimagnetic <strong>semiconductor</strong>s with two types <strong>of</strong> magnetic ions<br />

(Mn 2+ <strong>and</strong> Eu 2+ ) were investigated. First, the samples <strong>of</strong> Pb 1−x−y Mn x Eu y Te (0.005≤<br />

x ≤0.010, 0≤ y ≤0.009) with relatively low free carrier concentration (2.30·10 18 cm −3<br />

≤ p ≤ 4.80·10 20 cm −3 at T=300K) were studied. Next, the second group <strong>of</strong> samples:<br />

Pb 1−x−y−z Mn x Eu y Sn z Te (0.022 ≤ x ≤0.031, 0.002 ≤ y ≤0.017, 0.0680 ≤ z ≤0.850 )<br />

as well as Sn 1−x−y Mn x Eu y Te (0.055 ≤ x ≤0.131, 0.003 ≤ y ≤0.023), characterized by<br />

substantially larger carrier concentration (3.20·10 20 cm −3 ≤ p ≤ 1.01·10 21 cm −3 , 1.21<br />

10 21 cm −3 ≤ p ≤ 1.93·10 21 cm −3 for Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Sn 1−x−y Mn x Eu y Te,<br />

respectively at T=300K) was investigated.<br />

AC magnetic susceptibility studies in the temperature range 1.3-150 K using a mutual<br />

inductance method as well as DC magnetization measurements in the magnetic field range<br />

0-90 kOe (0-9 T) at various temperatures by use <strong>of</strong> extraction technique were carried out.<br />

The susceptibility measurements were carried out in AC magnetic field <strong>of</strong> frequency in<br />

the range 7-10000 Hz <strong>and</strong> amplitude not exceeding 5 Oe (5·10 −4 T).<br />

Generally, in the range <strong>of</strong> high temperatures all IV-VI semimagnetic <strong>semiconductor</strong>s<br />

are Curie-Weiss paramagnets with the temperature dependence <strong>of</strong> the magnetic susceptibility<br />

described by the Curie-Weiss law:<br />

χ(T ) = C/(T − Θ) (4.11)<br />

where C = g 2 µ 2 B S(S + 1)N M is the Curie constant <strong>and</strong> k B Θ = 1/3S(S + 1)x ∑ z i I(R i )<br />

is the paramagnetic Curie temperature (Curie-Weiss temperature). Here, N M is the concentration<br />

<strong>of</strong> magnetic ions, z i is the number <strong>of</strong> magnetic neighbors on i–th crystalographic<br />

shell, I(R i ) is the exchange integral between the central ion <strong>and</strong> its i–th magnetic<br />

neighbors, S is the spin <strong>of</strong> the magnetic ion, g is the spin-splitting g factor, k B is the<br />

Boltzman constant, µ B is the Bohr magneton.<br />

For all the investigated samples the high temperature behaviour <strong>of</strong> the inverse lowfield<br />

susceptibility χ −1 was nearly linear <strong>and</strong> all data were fitted with Curie-Weiss law <strong>of</strong><br />

the form 1 :<br />

χ(T ) = C/(T − Θ) + χ dia (4.12)<br />

where χ dia is the susceptibility <strong>of</strong> the host lattice (all IV-VI <strong>semiconductor</strong>s without magnetic<br />

ions are st<strong>and</strong>ard diamagnetic materials with magnetic susceptibility around χ dia<br />

≃ -3·10 −7 emu/g). Figure 4.3 presents the high temperature behaviour <strong>of</strong> inverse magnetic<br />

susceptibility for a few Pb 1−x−y Mn x Eu y Te samples. The determined values <strong>of</strong> paramagnetic<br />

Curie temperature Θ, which is proportional to the total strength <strong>of</strong> exchange<br />

interactions between magnetic ions <strong>and</strong> Curie constant C are presented in Table 4.2.<br />

In the case <strong>of</strong> PbMnEuTe obtained results (the determined negative <strong>and</strong> relatively<br />

small values <strong>of</strong> paramagnetic Curie-Weiss temperature Θ) indicate that weak anti<strong>ferromagnetic</strong><br />

superexchange interaction occurs the dominant mechanism in PbMnEuTe crystals.<br />

The obtained results correspond to those reported in PbMnTe. Inspection <strong>of</strong> Table<br />

1 The Curie-Weiss law in the form <strong>of</strong> equation 4.12 describes system with one type <strong>of</strong> magnetic ions,<br />

i.e. PbSnMnTe.


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 35<br />

Tab. 4.2: . The results <strong>of</strong> magnetic measurements for IV-VI mixed crystals.<br />

sample x P b x y z C Θ T C T f<br />

number [emu/g] [K] [K] [K]<br />

841_14 - 0.116 0.011 0.873 0.00215 19.65 17.46 -<br />

841_18 - 0.131 0.013 0.856 0.00209 18.18 16.63 -<br />

842_4 - 0.063 0.0045 0.932 - - 10.90 -<br />

842_8 - 0.068 0.003 0.929 - - 13.07 -<br />

842_14 - 0.070 0.007 0.923 - - 15.23 -<br />

842_20 - 0.091 0.009 0.900 0.00230 17.98 17.10 -<br />

848_4 - 0.061 0.0115 0.927 0.00141 11.57 11.57 -<br />

848_10 - 0.064 0.012 0.924 - - 8.35 -<br />

848_16 - 0.050 0.011 0.939 0.00149 11.02 8.75 -<br />

848_22 - 0.065 0.018 0.917 - - 10.75 -<br />

848_24 - 0.051 0.019 0.930 0.00222 12.08 10.98 -<br />

848_26 - 0.074 0.023 0.903 - - 11.31 -<br />

809_2 0.116 0.031 0.003 0.850 0.00047 5.16 - 2.0<br />

(Re(χ)<br />

625Hz)<br />

809_4 0.118 0.030 0.002 0.850 0.00050 4.88 - -<br />

809_10 0.187 0.030 0.003 0.780 0.00050 4.74 4.13 -<br />

809_12 0.215 0.022 0.003 0.760 0.00052 4.55 4.09 -<br />

809_30 0.256 0.024 0.010 0.710 0.00080 3.02 2.98 -<br />

809_30 0.256 0.024 0.010 0.710 0.00080 4.16 3.55 -<br />

anneal.<br />

809_32 0.27 0.026 0.014 0.690 0.00082 2.89 2.81 -<br />

809_34 0.276 0.027 0.017 0.680 0.00090 2.60 2.60 -<br />

809_36 0.272 0.025 0.013 0.690 0.00077 3.13 3.12 -<br />

793_2 0.99 0.010 0.000 - 0.00039 -0.93 - -<br />

793_4 0.989 0.010 0.001 - 0.00038 -0.69 - -<br />

793_4 0.989 0.010 0.001 - 0.00038 -0.69 - -<br />

793_6 0.982 0.009 0.009 - 0.00051 -1.12 - -<br />

793_10 0.991 0.005 0.004 - 0.00053 -0.42 - -<br />

793_12 0.99 0.007 0.003 - 0.00059 -0.39 - -<br />

793_14 0.99 0.005 0.005 - 0.00086 -0.41 - -


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 36<br />

4.2 shows that no distinct trends in Θ <strong>and</strong> C dependence on Eu concentration can be observed.<br />

However, it should be stressed that the values analyzed are determined with rather<br />

large error related with considerable uncertainty <strong>of</strong> chemical compositions <strong>of</strong> the samples.<br />

Figures 4.4 <strong>and</strong> 4.5 present high temperature part <strong>of</strong> inverse AC susceptibility for<br />

several PbMnEuSnTe <strong>and</strong> SnMnEuTe samples. The fitting procedure (the Curie–Weiss<br />

law – the same as for PbMnEuTe samples) revealed the positive values <strong>of</strong> Curie–Weiss<br />

temperature in this group <strong>of</strong> IV-VI mixed crystals. This indicates the presence <strong>of</strong> <strong>ferromagnetic</strong><br />

interactions. The obtained values <strong>of</strong> Curie–Weiss temperature Θ <strong>and</strong> Curie<br />

constant C are shown in Table 4.2.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

χ <br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 4.3: Inverse AC susceptibility versus temperature measured for Pb 1−x−y Mn x Eu y Te samples<br />

(793 − 2: x=0.010, y=0; 793 − 4: x=0.010, y=0.001; 793 − 10: x=0.005, y=0.004; 793 − 12:<br />

x=0.007, y=0.003; 793 − 14: x=0.005, y=0.005). The solid lines correspond to Curie -<br />

Weiss law fits.<br />

The careful inspection <strong>of</strong> Table 4.2 allows to notice significant changes <strong>of</strong> Curie-<br />

Weiss temperature with the Eu content. The decrease <strong>of</strong> the paramagnetic Curie temperature<br />

Θ with the increase <strong>of</strong> Eu concentration is clearly visible. The three samples <strong>of</strong><br />

Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 12, 809 − 30, 809 − 34 are characterized with very similar<br />

values <strong>of</strong> Mn content <strong>and</strong> concentration <strong>of</strong> free holes (see Table 4.1). For the Mn concentration<br />

equal to around x ≃ 0.02 <strong>and</strong> free hole concentration p ≃ 4 10 20 cm −3 increase<br />

<strong>of</strong> Eu content from y=0.003 to y=0.01 leads to the decrease <strong>of</strong> Curie–Weiss temperature<br />

from 4.55K to 3.02K <strong>and</strong> for y=0.017 paramagnetic Curie temperature is equal to 2.63K.<br />

In the case <strong>of</strong> Sn 1−x−y Mn x Eu y Te crystals such distinct tendency is not observed (see Table<br />

4.2). However, one needs to realize that obtained values <strong>of</strong> chemical composition as<br />

well as free carrier concentration are determined with quite large uncertainty.<br />

The low temperature studies revealed the presence <strong>of</strong> paramagnet/ferromagnet phase<br />

transition in the case <strong>of</strong> SnMnEuTe as well as the most <strong>of</strong> PbSnMnEuTe samples. Figures<br />

4.6 <strong>and</strong> 4.7 show the low temperature behaviour <strong>of</strong> real component <strong>of</strong> AC susceptibility<br />

Re(χ) for several samples <strong>of</strong> studied Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Sn 1−x−y Mn x Eu y Te<br />

mixed crystals. Typical behaviour <strong>of</strong> a ferromagnet is observed. Both components <strong>of</strong> the<br />

susceptibility dramatically increase at the Curie temperature T C . The Curie temperature


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 37<br />

<br />

<br />

<br />

<br />

<br />

χ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 4.4: The high temperature inverse AC susceptibility measured for several<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples (809 − 2: x=0.031, y=0.003, z=0.850; 809 − 12:<br />

x=0.022, y=0.003, z=0.760; 809 − 32: x=0.026, y=0.014, z=0.69; 809 − 34: x=0.027,<br />

y=0.017, z=0.680).The solid lines correspond to Curie -Weiss law fits.<br />

was determined by the maximum slope <strong>of</strong> dRe(χ) . The obtained values <strong>of</strong> T<br />

dT<br />

C are shown<br />

in Table 4.2. T C values are approximately equal to the Curie-Weiss temperature Θ determined<br />

from high temperature susceptibility measurements. The low temperature measurements<br />

confirmed the described above tendency for studied Pb 1−x−y−z Mn x Eu y Sn z Te<br />

crystals, i.e. the decrease <strong>of</strong> Curie temperature with the Eu content.<br />

Additionally performed DC magnetization measurements (up to 9T at various temperatures)<br />

collaborate susceptibility investigations <strong>and</strong> confirm <strong>ferromagnetic</strong> ordering<br />

observed for most <strong>of</strong> PbSnMnEuTe samples. Figure 4.8 presents an example <strong>of</strong> performed<br />

magnetization studies for 809 − 10 (x=0.030, y=0.003, z=0.78) <strong>ferromagnetic</strong> sample sample<br />

with p=6·10 20 cm −3 .<br />

For the sample <strong>of</strong> PbSnMnEuTe (809 − 2) the <strong>ferromagnetic</strong> to spin glass phase transition<br />

is observed. Figure 4.9 presents characteristic behaviour <strong>of</strong> low temperature part<br />

<strong>of</strong> real as well as imaginary component <strong>of</strong> susceptibility for spin glass (809 − 2) as well<br />

as <strong>ferromagnetic</strong> (809 − 12) samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te. In the case <strong>of</strong> <strong>ferromagnetic</strong><br />

sample (with the concentration <strong>of</strong> free holes equal to 4.0·10 20 cm −3 ) the sharp<br />

transitions in both real <strong>and</strong> imaginary components <strong>of</strong> susceptibility occur. For spin glass<br />

sample (characterized by higher free hole concentration p=1·10 21 cm −3 ) a cusp in Re(χ)<br />

is visible at the freezing temperature T f . The magnitude <strong>of</strong> the susceptibility at this cusp<br />

is much lower than the susceptibility <strong>of</strong> the <strong>ferromagnetic</strong> sample. A corresponding maximum<br />

in the out <strong>of</strong> phase <strong>of</strong> susceptibility Im(χ) is observed at slightly lower temperature.<br />

The 809 − 2 PbMnEuSnTe sample shows an obvious characteristics <strong>of</strong> spin glass–like<br />

phase. The cusp observed in susceptibility χ versus temperature T shifts to higher temperatures<br />

when the frequency f <strong>of</strong> the applied AC field is increased. This feature – the<br />

increase <strong>of</strong> the freezing temperature when the frequency is higher – was observed in the


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 38<br />

χ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 4.5: The high temperature inverse AC susceptibility measured for several Sn 1−x−y Mn x Eu y Te<br />

samples (841 − 18: x=0.131, y=0.013; 841 − 14: x=0.116, y=0.011; 842 − 20: x=0.091,<br />

y=0.009; 848 − 4: x=0.061, y=0.0115; 848 − 16: x=0.050, y=0.011; 848 − 24: x=0.051,<br />

y=0.019).<br />

well-known canonical spin glass systems [31], [32], [33], [34]. The increase <strong>of</strong> T f per<br />

decade <strong>of</strong> frequency is approximately constant <strong>and</strong> frequency dependence occurs in both<br />

real <strong>and</strong> imaginary part <strong>of</strong> AC magnetic susceptibility. Figures 4.10 <strong>and</strong> 4.11 present the<br />

frequency dependence <strong>of</strong> low temperature parts <strong>of</strong> real <strong>and</strong> imaginary components <strong>of</strong> susceptibility<br />

for 809 − 2 sample <strong>of</strong> PbMnEuSnTe. The relative shift <strong>of</strong> freezing temperature<br />

T f per decade <strong>of</strong> frequency R = ∆T f /T f ∆logf is equal to 0.021. The rate <strong>of</strong> the changes<br />

corresponding to maximum in imaginary part <strong>of</strong> susceptibility is higher: R=0.048.<br />

The values <strong>of</strong> R reported for known spin glass systems range from 0.005 (Cu) to<br />

0.11 (La 1−x Gd x Al 2 [35]) <strong>and</strong> the rate <strong>of</strong> the change in Im(χ) is the same as the rate <strong>of</strong><br />

the change in Re(χ). The values <strong>of</strong> R reported for Sn 1−x Mn x Te are equal: R=0.027<br />

(x=0.04) [27], R=0.022 (x=0.008) [36], R=0.027 (x=0.022) [36]. It appears that in<br />

the case <strong>of</strong> Sn 1−x Mn x Te mixed crystals the character <strong>of</strong> the spin glass phase does not<br />

depend on the manganese concentration. In the case <strong>of</strong> studied in the present thesis<br />

Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals, comparable with Mn-based IV-VI semimagnetic<br />

<strong>semiconductor</strong>s, the significant difference is visible in the inequality <strong>of</strong> frequency shift in<br />

Re(χ) <strong>and</strong> Im(χ). It has to be noted that Im(χ) in conducting media is distorted because<br />

<strong>of</strong> the eddy currents induced by AC magnetic field. Nevertheless, obtained values differ<br />

from those obtained for analogous materials (Sn 1−x Mn x Te), in particular the difference<br />

in the rate <strong>of</strong> frequency shift <strong>of</strong> cusp in real <strong>and</strong> imaginary part <strong>of</strong> susceptibility seems to<br />

be significant.<br />

Figure 4.12 presents magnetic phase diagram for PbMnSnTe <strong>and</strong> PbMnEuSnTe samples.<br />

The magnetic phase diagram for PbMnSnTe mixed crystals was taken from the Ref.<br />

[27] <strong>and</strong> data for PbMnEuSnTe samples studied in the present thesis were included.<br />

The magnetic phase <strong>of</strong> the samples at low temperatures is indicated as a function<br />

<strong>of</strong> both: manganese concentration <strong>and</strong> free hole concentration. Three regimes can be


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 39<br />

χ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$"%<br />

Fig. 4.6: The low temperature behaviour <strong>of</strong> real part <strong>of</strong> susceptibility for several<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples (809 − 12: x=0.022, y=0.003, z=0.760; 809 − 36:<br />

x=0.025, y=0.013, z=0.690; 809 − 30: x=0.024, y=0.010, z=0.710 ;809 − 32: x=0.026,<br />

y=0.014, z=0.690; 809 − 34: x=0.027, y=0.017, z=0.680).<br />

distinguished: a <strong>ferromagnetic</strong> regime for high manganese <strong>and</strong> high carrier concentration;<br />

a spin glass regime for low manganese <strong>and</strong> high carrier concentrations; <strong>and</strong> a reentrant<br />

spin glass regime separating the former two. A sample is considered as a reentrant spin<br />

glass if in the temperature range two transitions can be observed (<strong>ferromagnetic</strong> as well<br />

as spin glass). It is clearly visible that the presence <strong>of</strong> Eu shifts spin glass regime towards<br />

lower carrier concentration p. In the case <strong>of</strong> Sn 1−x Mn x Te any changes after introducing<br />

Eu to <strong>semiconductor</strong> matrix are not observed in magnetic phase diagram.<br />

The most likely reason <strong>of</strong> the strong dependence <strong>of</strong> Curie temperature on Eu content is<br />

a variation <strong>of</strong> the b<strong>and</strong> parameters with the alloy composition. In the qualitative analysis<br />

<strong>of</strong> the above problem the following points should be considered.<br />

• The Eu atom is a magnetic impurity in PbMnSnTe matrix with spin-only ground<br />

state: Eu 2+ has S =7/2. The electrons <strong>of</strong> the half-filled f-shell, responsible for the<br />

magnetic Eu moment, are very weakly coupled to the b<strong>and</strong> electrons. Thus, one can<br />

not expect a substantial contribution <strong>of</strong> Eu magnetic ions to average magnetization.<br />

The coupling constant J s−f is much smaller than J s−d constant. However there<br />

exists a small contribution to the total energy from s–f coupling between Eu atoms<br />

<strong>and</strong> carriers. One can expect an increase <strong>of</strong> Curie temperature with the Eu content<br />

from this mechanism. Since the experiment shows the opposite behaviour, it can be<br />

assumed that the role <strong>of</strong> Eu as a magnetic impurity is negligible.<br />

• Since EuTe is an antiferromagnet, one can expect a transition from positive (Curie)<br />

to negative (Neel) temperature <strong>of</strong> magnetic ordering <strong>of</strong> PbMnEuSnTe when changing<br />

Eu content y from 0 to 1. For small Eu content it can lead to a decrease <strong>of</strong><br />

T C .


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 40<br />

<br />

χ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!<br />

Fig. 4.7: The low temperature behaviour <strong>of</strong> real part <strong>of</strong> susceptibility for several<br />

Sn 1−x−y Mn x Eu y Te samples (841 − 14: x=0.121, y=0.011; 842 − 20: x=0.090, y=0.009;<br />

841 − 18: x=0.128, y=0.014; 848 − 24: x=0.055, y=0.0175; 848 − 16: x=0.054, y=0.011;<br />

848 − 4: x=0.058, y=0.011).<br />

• Eu is a component <strong>of</strong> the complex PbMnEuSnTe alloy <strong>and</strong> one should consider a<br />

variation <strong>of</strong> the b<strong>and</strong> parameters as a function <strong>of</strong> Eu content. For the qualitative<br />

analysis, let’s assume here a simplified two-b<strong>and</strong> model with some phenomenological<br />

parameters. It should be stressed that the real spectrum <strong>of</strong> IV-VI compounds<br />

is rather complicated <strong>and</strong> one should account for nonparabolicity <strong>and</strong> anisotropy <strong>of</strong><br />

energy b<strong>and</strong>s.<br />

For a qualitative analysis, the following formula describing RKKY interaction strength<br />

can be used:<br />

J RKKY (R ij ) = N m∗ Jsd 2 a6 0kF<br />

4 sin(2k F R ij ) − 2k F R ij cos(2k F R ij )<br />

32π 3¯h 2 (4.13)<br />

(2k F R ij ) 4<br />

where k F is the Fermi wave number, m ∗ the effective mass <strong>of</strong> the carriers, J sd the Mn ionelectron<br />

exchange integral, a 0 the lattice constant, N the number <strong>of</strong> valleys <strong>of</strong> the valence<br />

b<strong>and</strong>, R ij the distance between the magnetic ions.<br />

If one will take R ij equal to the mean distance between the magnetic ions (Mn) R<br />

this formula will give the mean interaction energy between magnetic impurities which is<br />

roughly equal to the transition temperature T C ≈ J RKKY (R)/k B . It should be stressed<br />

here, that Equation 4.13 was obtained for a parabolic energy spectrum <strong>and</strong> does not take<br />

into account the anisotropy.<br />

It is assumed that the hole energy spectrum in L points has the form:<br />

<strong>and</strong>, respectively, in Σ points:<br />

E L (k) = (∆ 2 + v 2 k 2 ) 1/2 (4.14)<br />

E Σ (k) = ɛ 0 + ¯h2 k 2<br />

2m ∗ Σ<br />

(4.15)


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 41<br />

0.3<br />

PbSnMnEuTe 809_10<br />

M [emu/g]<br />

0.2<br />

0.1<br />

4.3K<br />

12K<br />

30K<br />

0.0<br />

0 2 4 6 8 10<br />

B [T]<br />

Fig. 4.8: Magnetization measured at various temperatures <strong>and</strong> magnetic fields up to 9T for<br />

Pb 1−x−y−z Mn x Eu y Sn z Te 809 − 10 sample with x=0.030, y=0.003, z=0.780 <strong>and</strong> hole<br />

concentration p=6 10 20 cm −3 .<br />

Here, v is b<strong>and</strong>-coupling constant <strong>and</strong> it is assumed that v=5 10 −8 eV [37], ∆ = E g /2<br />

<strong>and</strong> ɛ 0 parameters depend on alloy composition, m ∗ Σ = 3m 0, m 0 is free electron mass.<br />

It is commonly known that the energy spectrum <strong>of</strong> Pb 1−y Eu y Te alloy is very sensitive<br />

to the concentration <strong>of</strong> Eu. The energy gap E g depends strongly on Eu content - from<br />

189.7 meV for y=0 (PbTe) to 248 meV for y=0.013 at the temperature T =10K [37],<br />

dE g /dy=5.788 eV at T =10K for y


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 42<br />

χχ !"<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

χ<br />

<br />

χ<br />

χ<br />

<br />

<br />

χ<br />

<br />

<br />

#$%&$'"<br />

Fig. 4.9: The low temperature behaviour <strong>of</strong> both real Re(χ) <strong>and</strong> imaginary Im(χ) component <strong>of</strong><br />

susceptibility for two samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 2 (x=0.031, y=0.003,<br />

p=1·10 21 cm −3 ) <strong>and</strong> 809 − 12 (x=0.022, y=0.003, p=4·10 20 cm −3 ). The typical <strong>ferromagnetic</strong><br />

characteristics is observed for 809 − 12 sample <strong>and</strong> spin glass behaviour for the<br />

sample with higher free hole concentration.<br />

which gives the total hole concentration p 0 <strong>and</strong> takes into account the degeneracy <strong>of</strong> each<br />

valley. It is accepted here that p 0 is equal to 4·10 20 cm −3 .<br />

The Fermi momentum in the Σ b<strong>and</strong> kF Σ = (2m∗ Σ (E F − ɛ 0 )) 3/2 can be found as a solution<br />

<strong>of</strong> Equation 4.20. Next, Equation 4.13 with k F = kF Σ, R = ( 3a3 0<br />

4πx )1/3 , a 0 =6.5·10 −8<br />

cm, N=12, J sd =1 eV is used for T C calculations. The obtained T C dependence on Eu concentration<br />

y for various values <strong>of</strong> Sn content (0.6 ≤ z ≤ 1) is shown in Figure 4.13. It is<br />

clearly visible that the obtained results <strong>of</strong> calculations performed within simple two b<strong>and</strong><br />

model can explain the observed experimentally tendency <strong>of</strong> Curie temperature decrease<br />

with Eu content. The obtained values <strong>of</strong> Curie temperature are higher than determined<br />

experimentally. However, it should be stressed that not all phenomenological parameters<br />

were known precisely. Many assumptions are introduced to the model. Nevertheless, the<br />

calculated dependence <strong>of</strong> Curie temperature on Eu content very well reflects experimentally<br />

confirmed effect <strong>of</strong> the T C decrease with Eu concentration y.<br />

Since the effect <strong>of</strong> the decrease <strong>of</strong> Curie temperature with the increase <strong>of</strong> Eu content<br />

due to change <strong>of</strong> b<strong>and</strong> parameters is pronounced in PbSnMnEuTe mixed crystals, one<br />

should observe the change <strong>of</strong> T C under hydrostatic pressure.<br />

Magnetization measurements under hydrostatic pressure up to 12 kbar were performed<br />

in the temperature above 4.2K <strong>and</strong> magnetic fields up to 16 kOe (1.6 T) using<br />

the vibrating sample magnetometer. A miniature CuBe container with the inner diameter<br />

<strong>of</strong> 1.42 mm was used as a pressure cell <strong>and</strong> a mixture <strong>of</strong> mineral oil <strong>and</strong> kerosene<br />

was used as a pressure-transmitting medium. The pressure at low temperatures was determined<br />

using the pressure dependence <strong>of</strong> the superconducting transition temperature<br />

for the pure Sn probe placed near the sample. The experimental setup allows to per-


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 43<br />

χ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!" #<br />

Fig. 4.10: The frequency dependence <strong>of</strong> real component <strong>of</strong> susceptibility Re(χ) for the sample<br />

<strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 2: x=0.031, y=0.003, p=1·10 21 cm −3 . The shift <strong>of</strong><br />

the freezing temperature T f towards higher temperatures with the frequency increase is<br />

clearly visible.<br />

form magnetization studies under hydrostatic pressure in the temperatures above T =4.2<br />

K. Thus, it was possible to investigate only Sn 1−x−y Mn x Eu y Te samples with Curie temperatures<br />

higher than 4.2K. Unfortunately, all available Pb 1−x−y−z Mn x Eu y Sn z Te samples<br />

were characterized by lower than 4.2K Curie temperature. Two samples were measured:<br />

one <strong>of</strong> Sn 1−x−y Mn x Eu y Te (842 − 8 with x=0.068, y=0.003, p=1.6·10 21 cm −3 ) <strong>and</strong> one <strong>of</strong><br />

Sn 1−x Mn x Te (x=0.10). Figures 4.14 <strong>and</strong> 4.15 show the results <strong>of</strong> zero field cooled (ZFC)<br />

low-field magnetization, measured at ambient <strong>and</strong> equal to 11.2 kbar pressure for SnMnEuTe<br />

<strong>and</strong> 10.5 kbar for SnMnTe sample. In both cases the decrease <strong>of</strong> Curie temperature<br />

under hydrostatic pressure is observed. T C shifts towards low temperatures with the same<br />

pressure coefficient for two investigated samples: dT C<br />

∼ -0.03 K/kbar.<br />

dP<br />

The ZFC magnetization measurements performed as a function <strong>of</strong> magnetic field up<br />

to 16 kOe (1.6T) at low temperatures (presented in Figures 4.16 <strong>and</strong> 4.17) revealed<br />

the decrease <strong>of</strong> spontaneous magnetization in both cases with almost the same pressure<br />

coefficient: dM 0<br />

∼ -0.026 emu/g kbar for SnMnEuTe sample <strong>and</strong> dM 0<br />

∼ -0.03 emu/g kbar<br />

dP dP<br />

for SnMnTe sample. There is no pressure effect on the coercive field <strong>of</strong> the hysteresis<br />

loops measured for both SnMnEuTe as well as SnMnTe samples.<br />

In fact, performed magnetization measurements under hydrostatic pressure revealed<br />

that introducing Eu ions into <strong>semiconductor</strong> matrix <strong>of</strong> SnMnTe does not influence significantly<br />

Curie temperature. However, the magnetic studies <strong>of</strong> SnMnEuTe mixed crystals<br />

with much higher concentration <strong>of</strong> Eu are needed to form the final conclusions.<br />

The effect <strong>of</strong> T C dependence on hydrostatic pressure is well known in PbMnSnTe<br />

semimagnetic <strong>semiconductor</strong>. T. Story [10] observed three different regions <strong>of</strong> the Curie<br />

temperature variation with the pressure - corresponding to three various pressure coefficients<br />

dT C<br />

dP . For the Pb 1−x−yMn x Sn y Te samples with low free carrier concentration (about<br />

3·10 20 cm −3 ) very significant increase <strong>of</strong> Curie temperature (approximately 100%) after<br />

applying the hydrostatic pressure equal to 10kbar is observed. Next, the samples with<br />

slightly higher free carrier concentration (6-7·10 20 cm −3 ) reveal no effect in the pressure


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 44<br />

χ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!" #<br />

Fig. 4.11: The frequency dependence <strong>of</strong> imaginary part <strong>of</strong> susceptibility Im(χ) for the sample <strong>of</strong><br />

Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 2 (x=0.031, y=0.003, p=1·10 21 cm −3 ). The maximum<br />

<strong>of</strong> the observed cusp shifts towards higher temperatures with the frequency increase.<br />

dependence <strong>of</strong> Curie temperature – in this case the values <strong>of</strong> dT C<br />

are equal to zero. At<br />

dP<br />

least, for the samples with high carrier concentration p=1·10 21 cm −3 the decrease <strong>of</strong> T C<br />

with the pressure increase is observed ( negative values <strong>of</strong> dT C<br />

coefficient). Simultaneously,<br />

<strong>transport</strong> investigations [10] revealed that hydrostatic pressure does not change the<br />

dP<br />

total free hole concentration, but significantly changes distribution <strong>of</strong> carriers between<br />

light hole (L) <strong>and</strong> heavy hole (Σ) b<strong>and</strong>s. The experimental results showed that pressure<br />

coefficient dT C /dP (p) dependence reflects the threshold behaviour <strong>of</strong> T C as a function <strong>of</strong><br />

hole concentration p. It is presented [10] that for PbSnMnTe mixed crystals energy gap<br />

increases with the hydrostatic pressure dE g /dP =8meV/kbar <strong>and</strong> ɛ 0 energy decreases with<br />

hydrostatic pressure dɛ 0 /dP =-4meV/kbar. T. Story showed [10] that hydrostatic pressure<br />

induces the changes in the b<strong>and</strong> parameters <strong>of</strong> PbSnMnTe mixed crystals <strong>and</strong> this in turn<br />

leads to the observed changes in Curie temperature.


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 45<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 4.12: <strong>Magnetic</strong> phase diagram for Pb 1−x−y Mn x Sn y Te [27] <strong>and</strong> Pb 1−x−y−z Mn x Eu y Sn z Te<br />

samples. Red triangles correspond to PbMnEuSnTe ferromagnets, red circles to PbMnEuSnTe<br />

spin glasses, black circles to PbMnSnTe ferromagnets, green circles to PbMn-<br />

SnTe spin glasses, blue circles to reentrant spin glasses. Lines present model calculations<br />

<strong>of</strong> the phase boundary (see Ref. [27]): solid line presents geometric model, dashed<br />

<strong>and</strong> dot dashed lines correspond to Sherrington-Kirkpatrick model, dot line correspond<br />

to Sherrington-Southern model<br />

8<br />

Curie Temperature [K]<br />

7<br />

6<br />

z=1<br />

z=0.9<br />

z=0.8<br />

z=0.7<br />

z=0.6<br />

5<br />

0,000 0,005 0,010 0,015<br />

Eu content y<br />

Fig. 4.13: Curie temperature calculated for Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals as a function<br />

<strong>of</strong> Eu content y for various values <strong>of</strong> Sn concentration 0.6 ≤ z ≤ 1 <strong>and</strong> Mn concentration<br />

x=0.02


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 46<br />

0.7<br />

0.6<br />

SnMnEuTe 842_8<br />

P=0<br />

P=11.2 kbar<br />

0.5<br />

ZFC<br />

B=20 Oe<br />

M [emu/g]<br />

0.4<br />

0.3<br />

0.2<br />

∆T C<br />

/dP ~ - 0.03 K/kbar<br />

0.1<br />

0.0<br />

4 6 8 10 12 14 16 18 20 22 24 26<br />

T [K]<br />

Fig. 4.14: Zero field cooled magnetization measured as a function <strong>of</strong> temperature for<br />

Sn 1−x−y Mn x Eu y Te sample (842 − 8 with x=0.068, y=0.003, p=1.6·10 21 cm −3 ) at ambient<br />

<strong>and</strong> equal to 11.2 kbar pressure.<br />

0.8<br />

0.7<br />

SnMnTe<br />

x=0.10<br />

M [emu/g]<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

∆T C<br />

/dP ~ - 0.03 K/kbar<br />

P=0<br />

10.5 kbar<br />

ZFC<br />

B=20 Oe<br />

0.2<br />

0.1<br />

0.0<br />

2 4 6 8 10 12 14 16 18 20 22 24 26<br />

T [K]<br />

Fig. 4.15: Zero field cooled magnetization measured as a function <strong>of</strong> temperature for Sn 1−x Mn x Te<br />

sample with x=0.10 at ambient <strong>and</strong> equal to 10.5 kbar pressure.


4. Transport <strong>and</strong> magnetic investigations <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te 47<br />

10<br />

8<br />

6<br />

4<br />

SnMnEuTe 842_8<br />

P=0<br />

P=11.2 kbar<br />

M 0<br />

M [emu/g]<br />

2<br />

0<br />

-2<br />

ZFC<br />

T=5 K<br />

∆M 0<br />

/dP ~ - 0.026 emu/g/kbar<br />

-4<br />

-6<br />

-8<br />

-10<br />

-18 -16 -14 -12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12 14 16 18<br />

B [kOe]<br />

Fig. 4.16: Zero field cooled magnetization measured as a function <strong>of</strong> magnetic field at low temperature<br />

T =5K for Sn 1−x−y Mn x Eu y Te sample (842 − 8 with x=0.068, y=0.003, p=1.6·10 21<br />

cm −3 ) at ambient <strong>and</strong> equal to 11.2 kbar pressure.<br />

8<br />

M [emu/g]<br />

6<br />

4<br />

M 0<br />

∆M 0<br />

/dP ~ - 0.03 emu/g/kbar<br />

P=0<br />

10.5 kbar<br />

SnMnTe<br />

x=0.10<br />

ZFC<br />

T=4.2 K<br />

2<br />

0<br />

0 2 4 6 8 10 12 14 16<br />

B [kOe]<br />

Fig. 4.17: Zero field cooled magnetization measured as a function <strong>of</strong> magnetic field at low temperature<br />

T =4.2K for Sn 1−x Mn x Te sample with x=0.10 at ambient <strong>and</strong> equal to 10.5 kbar<br />

pressure.


5. LOW TEMPERATURE ANNEALING STUDIES OF Ga 1−x Mn x As<br />

5.1 Introduction<br />

In this chapter the results <strong>of</strong> experimental studies <strong>of</strong> low temperature (LT) Ga 1−x Mn x As<br />

will be presented. Ga 1−x Mn x As has become the focus <strong>of</strong> current interest because <strong>of</strong> its<br />

high Curie temperature <strong>and</strong> possible spin-electronics applications (see [38] <strong>and</strong> references<br />

therein). One <strong>of</strong> the issues investigated in the presented thesis was the effect <strong>of</strong> the low<br />

temperature (LT) annealing on the electronic, magnetic <strong>and</strong> structural <strong>properties</strong> <strong>of</strong> the<br />

Ga 1−x Mn x As. The purpose <strong>of</strong> this work was to systematically anneal <strong>and</strong> study the samples<br />

over a wide range <strong>of</strong> Mn concentration. The role <strong>of</strong> Mn interstitial in GaMnAs is<br />

explored. The as-grown as well as annealed samples were systematically investigated.<br />

This introduction includes briefly review <strong>of</strong> the physical <strong>properties</strong> reported for LT<br />

Ga 1−x Mn x As epilayers. Ga 1−x Mn x As is the subject <strong>of</strong> very intense interest, since it became<br />

a favorite material for spintronics. Many papers present in the literature force to<br />

give very brief <strong>and</strong> selective survey <strong>of</strong> the reported results.<br />

Typically, the GaMnAs is grown by use <strong>of</strong> low temperature molecular beam epitaxy<br />

(LT MBE, growth temperature


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 49<br />

<strong>and</strong> x, it has not been really understood. Particularly, the results <strong>of</strong> high resolution X-ray<br />

diffraction (HRXRD) presented in this thesis show that Mn interstitials are responsible for<br />

the observed expansion <strong>of</strong> the lattice constant. These measurements are in good agrement<br />

with theoretical predictions proposed by Mašek et al. [44].<br />

It is expected that there are three possible electronic states <strong>of</strong> the Mn impurity substituting<br />

a trivalent cation: A 0 (d 4 ) <strong>and</strong> A 0 (d 5 +h) for Mn 3+ <strong>and</strong> A − (d 5 ) for Mn 2+ . Here,<br />

A 0 denotes the neutral center, A − is negatively charged center, the notation in brackets<br />

is the electronic configuration <strong>of</strong> d electrons. In the case <strong>of</strong> the A 0 (d 4 ) center the hole<br />

resides in the 3d shell. However, strong Hund’s intra-site exchange interaction may favor<br />

a state having five d electrons <strong>and</strong> a loosely bound hole. This is the case <strong>of</strong> the A 0 (d 5 +h)<br />

configuration, where A 0 (d 4 ) center traps tightly an electron in the 3d shell forming high<br />

spin, S=5/2, 3d 5 configuration, <strong>and</strong> this negatively charged Mn ion binds the hole in an<br />

effective mass state. The variety <strong>of</strong> experimental results indicate that the ground state <strong>of</strong><br />

the Mn impurity in III-V compounds corresponds to A 0 (d 5 +h) configuration [38]. However,<br />

no signal <strong>of</strong> A 0 (d 5 +h) centers is usually detected in MBE grown LT Ga 1−x Mn x As.<br />

For LT MBE grown Ga 1−x Mn x As epilayers with x0.03), the picture is more complicated <strong>and</strong> the explanation has not<br />

been presented so far [45]. However, if the argument <strong>of</strong> reduced binding hole energy <strong>of</strong><br />

A 0 center is correct for small x, it should be even more relevant for the epilayers with<br />

higher x, since the hole concentration increases with x [6].<br />

The hole concentration p as well as Mn content x are important parameters for GaMnAs<br />

since Curie temperature <strong>of</strong> this material increases with the increase <strong>of</strong> x <strong>and</strong> p. The<br />

<strong>transport</strong> as well as magneto<strong>transport</strong> measurements revealed that the conduction is p-type<br />

in the case <strong>of</strong> GaMnAs (see e.g. [38] <strong>and</strong> references therein, [46]). The hole concentration<br />

influences all major <strong>properties</strong> <strong>of</strong> this material [47], [48], [49], [50]. However, the determination<br />

<strong>of</strong> carrier type <strong>and</strong> concentration is difficult in the case <strong>of</strong> GaMnAs due to the<br />

presence <strong>of</strong> anomalous Hall effect (AHE) [38], [51]. To avoid this problem the method <strong>of</strong><br />

the electrochemical capacitance voltage method (ECV) can be used (see e.g. [46], [52],<br />

[13]). The Ga 1−x Mn x As samples exhibit the negative magnetoresistance al low temperatures.<br />

Recently, F. Matsukura et al. [53] presented that the magnitude <strong>of</strong> resistance<br />

strongly depend on relative orientations <strong>of</strong> magnetization <strong>and</strong> current <strong>and</strong> their directions<br />

in respect to crystal axes. It was shown (see [38] <strong>and</strong> references therein) that in terms<br />

<strong>of</strong> metal-insulator transition (MIT), the temperature dependence <strong>of</strong> resistivity can be cost<br />

into two categories. Low- <strong>and</strong> high-Mn composition samples (x0.06) are on<br />

the insulator side <strong>of</strong> MIT, whereas the layers containing intermediate Mn concentrations<br />

(0.03≤x≤0.06) are metallic.<br />

Due to the presence <strong>of</strong> anomalous Hall effect (AHE), magneto<strong>transport</strong> measurements<br />

provide valuable information on magnetism in GaMnAs. Particularly, the anomalous Hall<br />

effect studies indicated that the direction <strong>of</strong> easy axis is mainly controlled by epitaxial<br />

strain in Ga 1−x Mn x As [5]. It has been presented that tensile <strong>and</strong> compressive strains induce<br />

in-plane <strong>and</strong> out-<strong>of</strong>-plane moment orientation respectively. Recently, SQUID measurements<br />

[54] revealed that GaMnAs epilayers show rich characteristics <strong>of</strong> magnetic


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 50<br />

anisotropy depending not only on epitaxial strain but also on temperature <strong>and</strong> hole concentration.<br />

These experimental results are corroborated by magnetization measurements<br />

by use <strong>of</strong> magnetooptical Kerr effect (MOKE) [55]. Theoretical models based on holemediated<br />

magnetic order [49], [56], [47] give a good account <strong>of</strong> the observed moment observations,<br />

<strong>and</strong> the formation <strong>of</strong> magnetic domains is expected [56]. For the out-<strong>of</strong>-plane<br />

magnetized films stripe domain patterns were reported earlier [57]. Recently, magnetic<br />

domain structure in Ga 1−x Mn x As were studied by means <strong>of</strong> high resolution magnetooptical<br />

imaging technique [58]. Large, well defined magnetic domains, on the scale <strong>of</strong><br />

hundreds <strong>of</strong> micrometers, were observed. The clear evidence for a temperature dependent<br />

in-plane anisotropy was observed.<br />

The magnetooptical studies <strong>of</strong> GaMnAs (see e.g. [59]) revealed that the absorption<br />

edge <strong>of</strong> GaMnAs is not sharp. This is probably due to the impurity b<strong>and</strong> formation caused<br />

by the high concentration <strong>of</strong> ionized Mn <strong>and</strong> compansating donors [60]. The magnetic circular<br />

dichroism (MCD) studies (see [38] <strong>and</strong> references therein) in the reflectivity mode<br />

indicate for a negative value <strong>of</strong> the p-d exchange integral N 0 β, similarly to the case <strong>of</strong> II-<br />

VI semimagnetic <strong>semiconductor</strong>s. On the other h<strong>and</strong>, a positive value <strong>of</strong> MCD deduced<br />

from absorption measurements appears to suggest that N 0 β is positive [59]. This surprising<br />

result is explained if a large Burstein-Moss shift due to the high hole concentration<br />

specific to III-V SMSC’s is taken into consideration [59].<br />

A detailed theory <strong>of</strong> ferromagnetism in GaMnAs has not been established yet, but<br />

some important <strong>properties</strong> such as the Curie temperature, the magnetic anisotropy field,<br />

the temperature dependence <strong>of</strong> the spontaneous magnetization may be derived by use <strong>of</strong><br />

mean field theory <strong>of</strong> ferromagnetism in zinc blende semimagnetic <strong>semiconductor</strong> [47],<br />

[48], [49]. A mean field model based on exchange interaction mediated by delocalized<br />

holes in the ensemble <strong>of</strong> localized spins has been developed by Dietl et al. [47], [48]. The<br />

broad range <strong>of</strong> experiments can be explain by use <strong>of</strong> this model - assuming an ideal system<br />

without taking into account disorder or formation <strong>of</strong> impurity b<strong>and</strong>. The model uses a<br />

parameterized hole-spin exchange interaction, an exchange integral N 0 β. The Curie temperature<br />

T C is determined by the minimum <strong>of</strong> the free-energy functional with respect to<br />

magnetization M at a given hole concentration p. The hole contribution is computed by<br />

solving a 6×6 Luttinger-Kohn Hamiltonian with the presence <strong>of</strong> exchange. The theory<br />

with N 0 β=-1.2±0.2eV taken from photoemission experiments [61] <strong>and</strong> a carrier-carrier<br />

interaction enhancement <strong>of</strong> 1.2 [62] explains the large magnitude <strong>of</strong> T C =110K [4], [6]<br />

for Ga 1−x Mn x As with x=0.053 <strong>and</strong> hole concentration p=3.5·10 20 cm −3 [63]. This meanfield<br />

model is also capable <strong>of</strong> explaining the anomalous magnetic circular dichroism observed<br />

in GaMnAs [64]. The model explains also the strain dependence <strong>of</strong> the magnetic<br />

easy axis. For experimentally relevant carrier concentrations, the model predicts an inplane<br />

easy axis for compressive strain <strong>and</strong> a perpendicular axis for tensile strain. It was<br />

established experimentally that Curie temperature in Ga 1−x Mn x As increases with increasing<br />

Mn concentration x (as long as MnAs precipitates are not formed) <strong>and</strong> with the hole<br />

concentration p (see e.g. [38] <strong>and</strong> references therein). These observations are consistent<br />

with the mean-field model <strong>of</strong> ferromagnetism proposed by Dietl et al., which predicts that<br />

T C = Cxp 1/3 (5.1)<br />

where C is a constant specific to the host material.<br />

Recently, T. Jungwirth et al. [65] predicted theoretical calculations <strong>of</strong> Curie temperature<br />

in GaMnAs based on a model with S=5/2 local moments exchange coupled to


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 51<br />

itinerant holes in the valence b<strong>and</strong> <strong>of</strong> <strong>semiconductor</strong> host. Going beyond the st<strong>and</strong>ard<br />

mean-field theory <strong>of</strong> this model the enhancement <strong>of</strong> the T C due to the exchange <strong>and</strong> correlation<br />

in the itinerant-hole system <strong>and</strong> T C suppression due to collective fluctuations <strong>of</strong><br />

the ordered moments were estimated. The estimated theoretical Curie temperature T Cest<br />

is in good agreement with the experimental transition temperature <strong>and</strong> very close to meanfield<br />

T CMF value justifying the mean-field description <strong>of</strong> ferromagnetism in this material.<br />

It has been established that the Curie temperature <strong>of</strong> as-grown GaMnAs epilayers can<br />

be further improved by heat treatment (low temperature annealing). The highest <strong>ferromagnetic</strong><br />

transition temperature, T C , observed in GaMnAs (as well as in semimagnetic<br />

<strong>semiconductor</strong>s), was for a long time equal to 110K. The results presented in this thesis<br />

showed for the first time that Curie temperature can be enhanced above this limit. The heat<br />

treatment (annealing) <strong>of</strong> the grown by low-temperature molecular beam epitaxy (MBE)<br />

Ga 1−x Mn x As was the subject <strong>of</strong> intense studies in a number <strong>of</strong> laboratories word-wide.<br />

First, it was reported that LT annealing improves the Curie temperature T C <strong>and</strong> magnetization<br />

M(T) <strong>of</strong> the annealed samples depending on both the annealing temperature <strong>and</strong><br />

the duration <strong>of</strong> the annealing process [66], [67]. Recently, higher <strong>and</strong> higher Curie temperatures<br />

were reported [7], [13], even exceeding 160K [68], [8]. It was also shown that a<br />

proper choice <strong>of</strong> the Mn content <strong>and</strong> the thickness <strong>of</strong> GaMnAs epilayer enables a further<br />

increase <strong>of</strong> T C upon annealing [68], [8].<br />

The Mn ion in the substitutional, cation position in GaAs lattice acts as a acceptor,<br />

but in all Ga 1−x Mn x As samples the hole concentration is substantially lower than the Mn<br />

content. For a long time this has been ascribed to the formation <strong>of</strong> arsenic antisites [38]. It<br />

was shown that the formation <strong>of</strong> interstitial Mn ions plays here a crucial role [13]. These<br />

kind <strong>of</strong> defects play a key role in controlling the <strong>ferromagnetic</strong> transition in GaMnAs.<br />

5.2 The procedure <strong>of</strong> annealing<br />

In the present thesis the Ga 1−x Mn x As/GaAs epilayers in a wide range <strong>of</strong> Mn concentration<br />

(0.01


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 52<br />

<br />

ρ Ω<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.1: The zero-field resistivity <strong>of</strong> the as-grown GaMnAs epilayers in the wide range <strong>of</strong> Mn<br />

concentration 0.01≤ x ≤0.093.<br />

ment was in each case the same. For all the samples the electrical contacts for the <strong>transport</strong><br />

measurements were prepared always in the same way. The Hall bar samples with typical<br />

dimensions <strong>of</strong> 2mm × 6mm were used. The electrical contacts were made using indium<br />

solder <strong>and</strong> gold wires. No special precautions were needed to prepare electrical contacts.<br />

Typical resistance <strong>of</strong> the as-grown samples at room temperature was equal from 1 k Ω to<br />

2 k Ω. The values <strong>of</strong> resistivity were obtained with the uncertainty <strong>of</strong> about 15%. Figure<br />

5.1 shows the temperature dependence <strong>of</strong> zero-field resistivity ρ for the as-grown samples<br />

in a wide Mn content 0.01≤x≤0.093.<br />

It is clearly visible that all measured samples reveal metallic type <strong>of</strong> conductivity.<br />

The temperature dependence <strong>of</strong> the zero-field resistivity shows a broad maximum around<br />

Curie temperature T C where negative magnetoresistance also peaks. The hump structure<br />

is known to appear at the temperature slightly above the value <strong>of</strong> T C obtained from magnetic<br />

measurements. Such critical behavior <strong>of</strong> resistivity in <strong>ferromagnetic</strong> GaMnAs was<br />

observed by many groups (i.e. see [38] <strong>and</strong> references therein) <strong>and</strong> suggest the presence<br />

<strong>of</strong> the critical scattering, in which carriers are scattered by magnetic fluctuation through<br />

exchange interaction. It was shown that hump around T C can be interpreted in terms <strong>of</strong> a<br />

critical scattering by packets <strong>of</strong> <strong>ferromagnetic</strong>ally coupled spins, whose correlation length<br />

is comparable to the wavelength <strong>of</strong> the carriers at the Fermi level [6], [63].<br />

The measurements <strong>of</strong> resistivity at zero magnetic field occurred to be very suitable<br />

method to determine the optimal annealing conditions. Magnetization measurements<br />

M(T) revealed good agreement between <strong>transport</strong> <strong>and</strong> magnetic results. For the 10823C<br />

epilayer with x=0.07 annealed for 1 hour at the temperature <strong>of</strong> 280 0 C <strong>and</strong> under flow <strong>of</strong><br />

N 2 gas (7.1·10 −4 m 3 /min), the value corresponding to the zero-field resistivity hump is<br />

equal to 115K, <strong>and</strong> the SQUID measurements on the same sample yielded T C =113K (see<br />

Figure 5.2).<br />

For all investigated samples good agreement, within 10%, between zero-field resistivity<br />

<strong>and</strong> magnetization results was observed. The optimal time <strong>of</strong> annealing was in the<br />

range between 1h <strong>and</strong> 1 1/2 h. Figure 5.3 presents the Curie temperature estimated from


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 53<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

<br />

ρΩ<br />

<br />

<br />

<br />

<br />

Fig. 5.2: The temperature dependence <strong>of</strong> the zero-field resistivity (a hump structure slightly<br />

above Curie temperature is visible) <strong>and</strong> magnetization versus temperature measured by a<br />

SQUID magnetometer in small magnetic field (B=0.01T) parallel to the sample surface.<br />

A good agreement between <strong>transport</strong> <strong>and</strong> magnetic data is visible.<br />

the maximum in the temperature dependence <strong>of</strong> the zero-field resistivity (T ρ ) as a function<br />

<strong>of</strong> the annealing temperature.<br />

All presented samples were annealed for 1 hour under the flow <strong>of</strong> N 2 gas equal to<br />

7.1·10 −4 m 3 /min. The large changes in Curie temperature in a very narrow range <strong>of</strong><br />

annealing temperatures close to the growth temperature are observed. In the case <strong>of</strong> all<br />

investigated epilayers the optimal annealing temperature T a occurred to be around 289 0 C.<br />

The values <strong>of</strong> Curie temperature estimated from the zero-field resistivity measurements<br />

corresponding to different annealing temperatures for all studied samples are shown in<br />

Table 5.1.<br />

Figure 5.4 shows typical temperature dependences <strong>of</strong> the zero-field resistivity for<br />

the 10727E Ga 1−x Mn x Asepilayer with x=0.086 annealed at various temperatures. The<br />

resistivity <strong>of</strong> the as-grown sample <strong>and</strong> for samples annealed at relatively low temperatures<br />

(260 0 C, 289 0 C, 300 0 C <strong>and</strong> 310 0 C) show typical metallic behavior.<br />

For samples annealed at 350 0 C, an insulating behavior <strong>of</strong> the resistivity is observed.<br />

An increase <strong>of</strong> both Curie temperature <strong>and</strong> conductivity is visible after annealing at the<br />

optimal conditions (T a =289 0 C, t=1h). For the 10727E sample with x=0.086 the annealing<br />

procedure at the optimal temperature shifts the T ρ from 88K for the as-grown sample<br />

to 127K, annealing at higher temperature (T a =350 0 C) leads to lower Curie temperature<br />

T ρ =30K. In general, the large changes <strong>of</strong> the Curie temperature in the narrow range <strong>of</strong> the<br />

annealing temperature near the growth temperature are accompanied by the large changes<br />

<strong>of</strong> the conductivity. It was found that for the low Mn concentration, x < 0.05, the influence<br />

<strong>of</strong> the annealing procedure on both the Curie temperature <strong>and</strong> conductivity is weak.<br />

Figure 5.5 presents typical zero-field resistivity measured for GaMnAs sample with low<br />

Mn content x=0.027 (10727C) annealed at different thermal conditions. Inset shows the<br />

obtained T ρ values versus annealing temperature. It is clearly visible that in this case<br />

annealing procedure leads to the slight changes in both Curie temperature as well as resistivity<br />

<strong>of</strong> the sample.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 54<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.3: Curie temperature T ρ estimated from the the zero-field resistivity measurements versus<br />

temperatures <strong>of</strong> annealing for GaMnAs samples with different Mn concentration 0.01≤<br />

x ≤0.093.<br />

Tab. 5.1: Curie temperature T ρ estimated from the zero-field resistivity measurements for asgrown<br />

(a.g.) <strong>and</strong> annealed at different temperatures T a for 1 hour in nitrogen atmosphere<br />

GaMnAs samples with different Mn composition (0.1≤x≤0.093) <strong>and</strong> various layer thickness<br />

d.<br />

GaMnAs GaMnAs GaMnAs GaMnAs GaMnAs GaMnAs GaMnAs<br />

00811A 00119C 10727C 10727D 10727E 10823C 10823E<br />

d=302nm d=269nm d=131nm d=149nm d=105nm d=111nm d=115nm<br />

x=0.01 x=0.032 x=0.027 x=0.062 x=0.086 x=0.07 x=0.093<br />

T a T ρ T a T ρ T a T ρ T a T ρ T a T ρ T a T ρ T a T ρ<br />

[C] [K] [C] [K] [C] [K] [C] [K] [C] [K] [C] [K] [C] [K]<br />

a.g. 41 a.g. 70 a.g. 53 a.g 67 a.g. 88 a.g. 68 a.g. 53<br />

260 41 278 72 260 59.5 260 91 280 125 280 115 280 101<br />

300 40. 300 72 280 59 280 95 289 127 289 116 289 102<br />

350 34 350 42 300 62.5 300 101 300 118 300 113 300 99<br />

- - - - 319 62 320 92 310 103 310 101 310 97<br />

- - - - 350 50 350 35 350 30 350 40 350 38


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 55<br />

<br />

<br />

<br />

<br />

ρΩ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.4: Temperature dependence <strong>of</strong> the zero-field resistivity <strong>of</strong> GaMnAs sample with high Mn<br />

concentration (x=0.086) annealed at various temperatures.<br />

<br />

<br />

<br />

!"<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

%&<br />

ρ <br />

<br />

<br />

<br />

ρΩ<br />

<br />

<br />

<br />

<br />

# <br />

<br />

<br />

<br />

<br />

<br />

<br />

$ <br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.5: The temperature dependence <strong>of</strong> zero-field resistivity for as-grown <strong>and</strong> annealed at various<br />

temperatures sample with low Mn concentration x=0.027. Inset shows the Curie<br />

temperature estimated from the resistivity measurements versus annealing temperature.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 56<br />

It was shown that effects <strong>of</strong> low temperature annealing (at 240 0 C for 1 hour in a<br />

nitrogen atmosphere) are more pronounced for the thin layers <strong>of</strong> Ga 1−x Mn x As[69]. Recently,<br />

K.C. Ku et al. [70] reported that the highest T C in both as-grown <strong>and</strong> annealed<br />

(post-growth annealing for 90 minutes at 250 0 C in a nitrogen atmosphere) GaMnAs layers<br />

occurs for sample thickness between 10nm <strong>and</strong> 50nm <strong>and</strong> that T C is suppressed for<br />

thicker layers. The decrease <strong>of</strong> T C with the increase <strong>of</strong> the layer thickness for the asgrown<br />

GaMnAs epilayers was observed also by Matsukura et al. [39] <strong>and</strong> B.S. Sorensen<br />

et al. [71]. In present thesis GaMnAs samples with the thickness in the range between<br />

105nm <strong>and</strong> 302nm were systematically annealed <strong>and</strong> measured. The samples with different<br />

thickness <strong>of</strong> GaMnAs layer were characterized simultaneously by different Mn<br />

concentration. This made impossible to verify above suggestions. Nevertheless, the careful<br />

inspection <strong>of</strong> Table 5.1 allows to notice that increase <strong>of</strong> the Curie temperature after<br />

optimal thermal annealing depends on the Mn concentration <strong>and</strong> is the most effective for<br />

the high Mn concentration. The observed in the same time enhancement <strong>of</strong> conductivity<br />

is also the most effective for the GaMnAs epilayers with high Mn concentration. This<br />

effect is illustrated in the Figure 5.6 where conductivity versus annealing temperature is<br />

shown for two samples with low (x=0.027, 10727C) <strong>and</strong> high Mn concentration (x=0.086,<br />

10727E) with similar thickness <strong>of</strong> GaMnAs layer (131nm <strong>and</strong> 105nm, respectively). The<br />

annealing procedure is more effective in the case <strong>of</strong> higher Mn concentration.<br />

5.4 The results <strong>of</strong> magneto<strong>transport</strong> measurements<br />

The magneto<strong>transport</strong> measurements were performed in the range <strong>of</strong> low as well as high<br />

magnetic fields. As grown <strong>and</strong> annealed at different conditions <strong>ferromagnetic</strong> GaMnAs<br />

samples were investigated in the static (up to 0.5T by use <strong>of</strong> classical electromagnet <strong>and</strong><br />

up to 13T using superconducting magnet) as well as pulsed (up to 55T) magnetic fields<br />

applied perpendicular to the plane <strong>of</strong> the film. Both, the anomalous Hall effect as well<br />

as magnetoresistivity (MR) measurements were performed. The DC six probe technique<br />

was used to perform magneto<strong>transport</strong> measurements in the case <strong>of</strong> low as well as high<br />

magnetic fields.<br />

Magneto<strong>transport</strong> measurements provide valuable information on magnetism <strong>of</strong> thin<br />

films <strong>and</strong> are <strong>of</strong> particular importance in the case <strong>of</strong> thin films <strong>of</strong> semimagnetic <strong>semiconductor</strong>s<br />

in which the magnitude <strong>of</strong> the total magnetic moment is small. Particularly, the<br />

anomalous Hall effect (known also as a extraordinary Hall effect or spin Hall effect) - if<br />

understood theoretically - can serve to determine the magnitude <strong>of</strong> magnetization. However,<br />

it should be stressed that the AHE does not provide information about magnetization<br />

<strong>of</strong> the whole sample but only about its value in the regions visited by the carriers. Particularly,<br />

near the metal-insulator boundary the carrier distribution is highly nonuniform <strong>and</strong><br />

direct magnetic measurements may provide different magnetization values.<br />

The Hall resistance R Hall <strong>of</strong> a magnetic film <strong>of</strong> the thickness d is empirically known<br />

to be a sum <strong>of</strong> ordinary <strong>and</strong> anomalous Hall terms [72]:<br />

R Hall = R o<br />

d B + R s<br />

d M ⊥ (5.2)<br />

where R o <strong>and</strong> R s are the ordinary <strong>and</strong> anomalous Hall coefficients respectively, d is the<br />

sample thickness, B is magnetic field <strong>and</strong> M ⊥ is the component <strong>of</strong> magnetization perpendicular<br />

to the sample surface.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 57<br />

<br />

<br />

<br />

<br />

Ω <br />

<br />

<br />

<br />

<br />

<br />

<br />

!! <br />

<br />

<br />

<br />

<br />

Ω <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"## <br />

Fig. 5.6: Conductivity versus annealing temperature for the samples with high Mn concentration<br />

x=0.086 <strong>and</strong> with low Mn concentration x=0.027.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 58<br />

The anomalous Hall coefficient is usually assumed to be proportional to R α sheet , where<br />

R sheet is the sheet resistance. The exponent α is either 1 or 2 depending on the origin <strong>of</strong><br />

the effect. For the skew scattering mechanism α=1, whereas for the side jump mechanism<br />

α=2 [72]. A comparison <strong>of</strong> Hall resistivity <strong>and</strong> SQUID magnetization data allows to<br />

identify the the dominating mechanism.<br />

Recently, Jungwirth et al. [73] developed theory <strong>of</strong> the AHE in p-type zinc-blende<br />

magnetic <strong>semiconductor</strong>s <strong>and</strong> presented numerical results for GaMnAs employing formula<br />

for the side-jump mechanism in the weak scattering limit.<br />

Moreover, the experimental results demonstrated by Dietl et al. [51] suggested that<br />

the side-jump mechanism gives the dominant contribution for metallic samples. Also the<br />

theory discussed by the authors [51] indicates that the side-jump mechanism accounts<br />

for AHE in the investigated <strong>ferromagnetic</strong> material. The authors show that there is a<br />

good agreement between their experimental <strong>and</strong> theoretical [73] magnitude <strong>of</strong> the Hall<br />

conductivity. The theory discussed by Dietl et al. [51] explains the sign <strong>of</strong> the AHE (the<br />

coefficients <strong>of</strong> the normal <strong>and</strong> anomalous Hall effects are expected to have the same sign)<br />

<strong>and</strong>, together with the results obtained by Jungwirth et al. [73] explains the magnitude <strong>of</strong><br />

the Hall conductance.<br />

The presence <strong>of</strong> the AHE makes the determination <strong>of</strong> the carrier concentration <strong>and</strong><br />

type <strong>of</strong> carriers very difficult. Determination <strong>of</strong> the free carrier concentration is complicated<br />

by the dominance <strong>of</strong> the anomalous Hall effect term. Generally, in order to obtain<br />

free carrier concentration, the magneto<strong>transport</strong> investigations should be performed at low<br />

temperatures <strong>and</strong> at magnetic fields sufficiently high so that the magnetization saturates.<br />

Then, the ordinary Hall coefficient can be determined from the remaining linear change<br />

<strong>of</strong> the Hall resistance in the magnetic field. However, the experimental data, i.e. the magnetization<br />

measurements performed by means <strong>of</strong> magnetooptical Kerr effect up to 25T<br />

(see section 6 <strong>of</strong> this Chapter) revealed that magnetization does not saturate even at the<br />

highest magnetic fields. This makes determination <strong>of</strong> free carrier concentration difficult<br />

<strong>and</strong> doubtful.<br />

In the present paper, an effort to determine free hole concentration in the as-grown as<br />

well as annealed Ga 1−x Mn x As layers was made. The Hall resistivity as well as magnetoresistivity<br />

measurements were performed simultaneously in the high pulsed magnetic<br />

fields up to 55T <strong>and</strong> low temperatures (T=4.2K) for the epilayers <strong>of</strong> GaMnAs with various<br />

Mn concentration 0.027≤x≤0.086. Fig 5.7 shows the typical anomalous Hall voltage<br />

signal as a function <strong>of</strong> magnetic field for two samples <strong>of</strong> 10727E GaMnAs epilayer with<br />

x = 0.086: as-grown <strong>and</strong> annealed at the optimal temperature 289 0 C.<br />

The obtained data indicates that the contribution from the ordinary Hall term is rather<br />

small in the displayed range <strong>of</strong> temperature <strong>and</strong> magnetic field - the Hall voltage reflects<br />

M(B) behavior <strong>and</strong> the significant component to the Hall voltage comes from the anomalous<br />

Hall effect. It should be stressed that negative magnetoresistance (MR) that is present<br />

in the GaMnAs samples <strong>and</strong> persists up to high magnetic fields (look in Figure 5.8) adds<br />

to the measured signal <strong>of</strong> anomalous Hall effect. This, in turn, makes the determination<br />

<strong>of</strong> free carriers difficult.<br />

The magneto<strong>transport</strong> measurements (Hall resistivity <strong>and</strong> conductivity) that were performed<br />

even in the higher range <strong>of</strong> magnetic fields (up to 55T) do not give a unique value<br />

for the hole concentration <strong>of</strong> the investigated epilayers. As was mentioned before, in the<br />

range <strong>of</strong> high magnetic fields where one may expect magnetization saturation the Hall<br />

voltage should be related to free carrier concentration. In the experiments performed in


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 59<br />

<br />

<br />

<br />

ρ <br />

<br />

" #$%<br />

! <br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

µ<br />

<br />

<br />

&"#<br />

Fig. 5.7: The Hall voltage versus magnetic field for two samples <strong>of</strong> Ga 1−x Mn x As with x=0.086,<br />

as-grown <strong>and</strong> annealed at the optimal conditions, measured at T=4.2K. Note that dominant<br />

contribution to the Hall voltage comes from the AHE term; the Hall voltage reflects<br />

the M(B) behaviour.<br />

the present thesis the difference between the slopes <strong>of</strong> the Hall voltage in high magnetic<br />

fields is very small <strong>and</strong> the data obtained up to 55T do not give the unique value for<br />

the hole concentration <strong>of</strong> the investigated epilayers. The results obtained by the electrochemical<br />

voltage capacitance voltage (ECV) method in the Lawrence Berkeley National<br />

Laboratory on the samples annealed <strong>and</strong> investigated in the present thesis (as-grown <strong>and</strong><br />

annealed at the optimal conditions samples <strong>of</strong> 10823C epilayer with x=0.07) [13] clearly<br />

show that optimal annealing increases the free hole concentration. The increase from 6<br />

10 20 cm −3 for the as-grown sample to 1·10 21 cm −3 for the annealed sample was observed.<br />

Moreover, further investigations <strong>of</strong> the as-grown <strong>and</strong> annealed at the optimal conditions<br />

GaMnAs samples (annealing conditions similarly to these established by author) with<br />

various Mn concentration by ECV method (see e.g. [52]) confirmed these results. Contrary<br />

to expectation that free hole concentration should increase after annealing at the<br />

optimal conditions (as was shown by electrochemical capacitance voltage (ECV) pr<strong>of</strong>iling<br />

method) the obtained <strong>transport</strong> data in the high magnetic fields indicate the decrease<br />

<strong>of</strong> the free carrier concentration. The reason is that the experimental data - Hall voltage<br />

versus magnetic field - reflects M(B) dependence rather then normal Hall effect term. It<br />

should be also stressed here that performed magnetic investigations - both: direct SQUID<br />

as well as magnetooptical Kerr effect (MOKE) investigations (see sections 5 <strong>and</strong> 6 <strong>of</strong><br />

this Chapter) indicate that even at low temperatures (T=4.2K) <strong>and</strong> high magnetic fields<br />

(MOKE measurements performed in pulsed magnetic fields up to 25T) the magnetization<br />

is far from being saturated. Also negative magnetoresistance that hardly saturates even<br />

at very high magnetic fields (the measurements <strong>of</strong> MR performed up to 55T) makes the<br />

determination <strong>of</strong> free carrier concentration in GaMnAs films difficult.<br />

The measurements <strong>of</strong> MR were performed in the static (up to 13T) as well as pulsed<br />

(up to 55T) magnetic fields applied perpendicular to the plane <strong>of</strong> the film.<br />

Recently, it has been found that magnetoresistance <strong>of</strong> Ga 1−x Mn x As depends on the


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 60<br />

<br />

<br />

∆ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

∆ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

<br />

∆ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

Fig. 5.8: Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0, for the epilayer<br />

with high Mn content x=0.07 measured at various temperatures 4.2K, 40K, 110K, 150K<br />

<strong>and</strong> 200K for as-grown sample annealed at the optimal conditions (T a =289 0 C) <strong>and</strong> after<br />

annealing at higher temperature (T a =350 0 C).


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 61<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ρ<br />

<br />

∆ <br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

<br />

ρ<br />

<br />

<br />

!"#!$%<br />

Fig. 5.9: Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0, for the epilayer<br />

with high Mn content x=0.086: as-grown, annealed at the optimal conditions (T a =280 0 C)<br />

<strong>and</strong> after annealing at higher temperature (T a =350 0 C).<br />

relative orientation <strong>of</strong> the current <strong>and</strong> magnetic field [74], [75], [76], [53]. The observed<br />

anisotropic magnetoresistance (AMR) is quite sizable <strong>and</strong> depends also on the direction <strong>of</strong><br />

the field <strong>and</strong> current with respect to crystal axis. The measurements <strong>of</strong> MR in the present<br />

thesis were performed in the configuration with magnetic field B applied perpendicular to<br />

the sample surface <strong>and</strong> electric current I flowing r<strong>and</strong>omly with respect to the crystal axis.<br />

The effects <strong>of</strong> AMR were not studied. One <strong>of</strong> the aims <strong>of</strong> presented MR measurements was<br />

to explore the influance <strong>of</strong> low temperature annealing on the magnetic field dependence<br />

<strong>of</strong> resistance.<br />

All investigated samples in a wide range <strong>of</strong> Mn concentration 0.027≤x≤0.086 exhibit<br />

negative magnetoresistance above 0.5T. Presently, it is common knowledge (see e.g<br />

[38]) that for the Ga 1−x Mn x As samples, the negative magnetoresistance peaks around T C ,<br />

where the temperature dependence <strong>of</strong> resistance shows a maximum. This effect was also<br />

observed for investigated epilayers in the present thesis <strong>and</strong> is shown in Figure 5.8. The<br />

magnetoresistivity measured up to 13T ((R-R 0 )/R 0 , where R 0 is resistivity at zero magnetic<br />

field) is shown for three samples: as-grown, annealed at the optimal conditions T a =289 0 C<br />

<strong>and</strong> higher temperature <strong>of</strong> 350 0 C <strong>of</strong> 10823C epilayer with high Mn concentration x=0.07.<br />

It is clearly visible that MR peaks at the temperature equal to Curie temperature. Further<br />

investigations at higher pulsed magnetic fields (up to 55T) revealed that the negative magnetoresistance<br />

is unsaturated up to the highest value <strong>of</strong> the investigated field. This effect is<br />

presented in Figures 5.9 <strong>and</strong> 5.10, where magnetoresistivity is shown for the as-grown <strong>and</strong><br />

annealed samples with high (10727E x=0.086) as well as low Mn concentration (10727C<br />

x=0.027).<br />

The author found that annealing <strong>of</strong> the GaMnAs epilayers with high Mn concentration<br />

leads to very significant changes in magnetoresistivity in the range <strong>of</strong> low temperatures.<br />

As it is clearly visible in Figure 5.9, annealing at the optimal conditions <strong>of</strong> the epilayers<br />

with high Mn concentration leads to significant decrease <strong>of</strong> MR. Annealing at higher temperatures<br />

(for the 10727E epilayer, the sample annealed at 350 0 C was investigated) leads


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 62<br />

<br />

<br />

∆ <br />

<br />

<br />

<br />

<br />

<br />

!"#<br />

ρ <br />

<br />

<br />

<br />

ρ <br />

<br />

<br />

<br />

<br />

Fig. 5.10: Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0, for the epilayer<br />

with low Mn content x=0.027: as-grown <strong>and</strong> annealed at the optimal conditions<br />

(T a =300 0 C).<br />

to the substantial increase <strong>of</strong> magnetoresistance. It should be recalled here (see section 2<br />

<strong>of</strong> this Chapter), that zero-field resistivity measurements (resistance versus temperature)<br />

revealed metallic type <strong>of</strong> conductivity for the samples annealed at the optimal conditions<br />

<strong>and</strong> insulating type <strong>of</strong> conductivity for the samples annealed at higher temperatures. A<br />

large negative magnetoresistance was observed <strong>and</strong> reported on the insulating side <strong>and</strong><br />

in the vicinity <strong>of</strong> the metal-insulator transition in GaMnAs samples [77], [78]. For the<br />

GaMnAs samples with lower Mn content (x=0.027) the MR is not affected by the heat<br />

treatment at the optimal conditions. It should be stressed that also the influence <strong>of</strong> annealing<br />

procedure on both the Curie temperature <strong>and</strong> conductivity is weak in this case. For<br />

10727C epilayer with x=0.027 the observed changes in magnetoresistance after annealing<br />

at the optimal conditions are very small.<br />

Recently, the negative high field magnetoresistance was studied experimentally <strong>and</strong><br />

discussed by F. Matsukura et al. [53] <strong>and</strong> T. Dietl et al. [51]. The authors show that the<br />

resistance maximum <strong>and</strong> the associated negative magnetoresistance in the vicinity <strong>of</strong> T C<br />

seem to result from effects <strong>of</strong> thermally disordered spins upon localization <strong>and</strong> scattering.<br />

As is underline in the papers, there is a number <strong>of</strong> effects that can produce a sizable<br />

magnetoresistance in magnetic <strong>semiconductor</strong>s, especially at the localization boundary<br />

[79]. In particularly, spin disorder scattering shifts the MIT (metal to insulator transition)<br />

towards higher carrier concentration. Since the magnetic field orders the spins, negative<br />

magnetoresistance occurs, sometimes leading to the field-induced insulator-to-metal transition<br />

[80]. Negative magnetoresisstance <strong>and</strong> resistance maximum at T C persist deeply<br />

in the metallic phase owing to critical scattering [63]. The negative magnetoresistance<br />

hardly saturates even at high magnetic fields <strong>and</strong> occurs also at low temperatures. In<br />

order to explain this observation, the authors note that the giant splitting <strong>of</strong> the valence<br />

b<strong>and</strong> makes both spin-disorder <strong>and</strong> spin-orbit scattering relatively inefficient. Under such<br />

conditions, weak localization magnetoresistance can show up at low temperatures, where<br />

inelastic scattering ceases to operate. The low-temperature negative magnetoresistance<br />

appears to do not be any spin phenomenon but an orbital effect resulting from the destruc-


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 63<br />

<br />

<br />

<br />

<br />

Ω<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#<br />

Fig. 5.11: Magnetoresistivity R for the sample with x=0.01 in low magnetic fields at T=20K. The<br />

magnetic field B was applied perpendicular to the film. Note that hysteretic behavior is<br />

visible. The arrows <strong>and</strong> numbers indicate the history <strong>and</strong> direction <strong>of</strong> applied magnetic<br />

field.<br />

tive influence <strong>of</strong> the magnetic field on interference <strong>of</strong> scattered waves.<br />

In the range <strong>of</strong> low magnetic fields the positive magnetoresistance is present. The MR<br />

curves exhibit hysteretic behavior for the configuration with the perpendicular direction<br />

<strong>of</strong> magnetic field <strong>and</strong> samples under compressive strain. This effect is shown in Figures<br />

5.11 <strong>and</strong> 5.12.<br />

The low magnetic field data can provide information on process <strong>of</strong> field-induced rotation<br />

<strong>of</strong> magnetization in respect to magnetic field direction, crystal <strong>and</strong> easy axis. Recently,<br />

[53] showed that behaviour <strong>of</strong> magnetoresistance in the low magnetic field range<br />

depends on the character <strong>of</strong> the strain <strong>and</strong> on the field <strong>and</strong> current directions. In particular,<br />

for the magnetic field pointed along the growth direction, the values <strong>of</strong> the magnetic<br />

field corresponding to resistance maxima are <strong>of</strong> order <strong>of</strong> the anisotropy field (∼0.2T) that<br />

aligns magnetization along the growth direction.<br />

The values <strong>of</strong> B corresponding to the maxima <strong>of</strong> magnetoresistance for the samples<br />

investigated in the thesis are <strong>of</strong> the same order as reported in Ref. [53].<br />

The hysteretic behaviour <strong>of</strong> low magnetic field MR is correlated with the low magnetic<br />

field behavior <strong>of</strong> SQUID magnetization - hysteresis loops <strong>of</strong> magnetization (see section<br />

5 <strong>of</strong> this Chapter). As the temperature is reduced below Curie temperature, an increase<br />

in the coercive field is observed that collaborates the low field hysteretic behavior <strong>of</strong> MR.<br />

With the decrease <strong>of</strong> temperature the low field hysteretic feature <strong>of</strong> the magnetoresistivity<br />

is much stronger (see Figures 5.12).<br />

Moreover, the LT annealing affects also the magnetoresistivity <strong>and</strong> hysteresis loops.<br />

The decrease <strong>of</strong> both the coercive field <strong>and</strong> negative magnetoresistivity is pronounced<br />

after annealing at the optimal conditions (the SQUID magnetization data are described<br />

<strong>and</strong> discussed in details in section 5 <strong>of</strong> this Chapter). Only speculative <strong>and</strong> phenomenological<br />

interpretation is possible at present. Recently, T. Fukumura et.al. [81] showed<br />

that the films <strong>of</strong> GaMnAs with in-plane magnetization has unconventional domain structure<br />

that show r<strong>and</strong>om arrangement <strong>of</strong> the domains. It is obvious that hysteresis loop


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 64<br />

Fig. 5.12: Low magnetic field magnetoresistivity for the Ga 1−x Mn x As sample with x=0.01 measured<br />

at various temperatures. The magnetic field B was applied perpendicular to the<br />

film. The arrows <strong>and</strong> numbers indicate the history <strong>and</strong> direction <strong>of</strong> applied magnetic<br />

field.<br />

features, specifically coercive field, shape, saturation magnetization are correlated with<br />

domain structure. The <strong>transport</strong> <strong>properties</strong> <strong>of</strong> the <strong>ferromagnetic</strong> materials can be modified<br />

by domain walls, particularly domain wall can increase or decrease the resistance <strong>of</strong><br />

the system [82]. Moreover, it was shown that the shape <strong>and</strong> positions <strong>of</strong> the peaks <strong>of</strong> MR<br />

depend on domain structure <strong>and</strong> squareness <strong>of</strong> the hysteresis loops [83]. The observed<br />

large changes <strong>of</strong> the MR <strong>and</strong> magnetization can be related with the change <strong>of</strong> domain<br />

structure <strong>of</strong> GaMnAs system.<br />

Matsukura et al. [53] assigned the effect <strong>of</strong> hysteretic resistance jumps to a large ratio<br />

<strong>of</strong> the anisotropy <strong>and</strong> coercive fields, which makes that even a rather small misalignment,<br />

<strong>and</strong> thus a minute in-plane field, can result in magnetization switching between in-plane<br />

easy-directions.<br />

5.5 The results <strong>of</strong> SQUID measurements<br />

<strong>Magnetic</strong> <strong>properties</strong> <strong>of</strong> Ga 1−x Mn x As thin layers can be measured by direct magnetization<br />

measurements as well as magnetooptical studies (Magnetooptical Kerr Effect - MOKE)<br />

<strong>and</strong> magneto<strong>transport</strong> investigations. In this section the results <strong>of</strong> direct magnetization<br />

measurements will be shown <strong>and</strong> discussed. Direct measurements <strong>of</strong> the magnetization<br />

M <strong>of</strong> GaMnAs epilayers as a function <strong>of</strong> magnetic field B <strong>and</strong> temperature T were performed<br />

by use <strong>of</strong> a commercially available superconducting quantum interference device<br />

(SQUID) magnetometer. The temperature independent diamagnetic response <strong>of</strong> GaAs<br />

substrate was determined from separate measurement <strong>of</strong> only the same GaMnAs substrate<br />

used for epitaxial growth. Next, the diamagnetic component was subtracted from


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 65<br />

!"# $<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!$<br />

Fig. 5.13: Magnetization versus temperature M(T) measured in small magnetic field B = 10Gs for<br />

the as-grown <strong>and</strong> annealed at 289 0 C GaMnAs sample with x=0.07. It is clearly visible<br />

that annealing procedure at the optimal conditions leads to the increase <strong>of</strong> saturation<br />

magnetization.<br />

the total response to obtain magnetization <strong>of</strong> the magnetic layer. The as-grown <strong>and</strong> annealed<br />

at different conditions samples were measured when magnetic field B was applied<br />

parallel to the GaMnAs layer.<br />

Two types <strong>of</strong> magnetization measurement were performed. First, the temperature<br />

dependence <strong>of</strong> magnetization M(T) was measured in small magnetic fields (i.e., 10 Gauss)<br />

after the sample has been magnetized at higher magnetic field (1000 gauss) parallel to<br />

the sample surface. These measurements allowed to determine paramagnet/ferromagnet<br />

phase transition temperature. As an example the magnetization curves versus temperature<br />

measured for two samples with x=0.07 (10823C epilayer), as-grown <strong>and</strong> annealed at the<br />

optimal conditions (T a =289 0 C) are shown in Figure 5.13.<br />

The magnetization measurements M(T) revealed good agreement between <strong>transport</strong><br />

<strong>and</strong> magnetic results. The values <strong>of</strong> Curie temperature determined from SQUID measurements<br />

are in good agreement, within 10% with the values estimated from the zero-field<br />

resistivity. It is clearly visible that, the saturation magnetization M S increases after heat<br />

treatment at the optimal conditions, indicating that annealing increases the concentration<br />

<strong>of</strong> magnetically active Mn ions. Such effect is observed for all investigated samples in the<br />

whole range <strong>of</strong> Mn concentration.<br />

The second aim <strong>of</strong> the SQUID measurements was to investigate the hysteresis loops<br />

<strong>of</strong> as-grown <strong>and</strong> annealed epilayers <strong>of</strong> GaMnAs. Figure 5.14 shows magnetization curves<br />

at several temperatures <strong>of</strong> as grown <strong>and</strong> annealed at the optimal conditions (T a =289 0 C,<br />

t=1h, under flow <strong>of</strong> N 2 ) samples <strong>of</strong> a 111nm thick GaMnAs layer with Mn content x =<br />

0.07 (10823C epilayer) grown on GaAs substrate. When magnetic field B is applied parallel<br />

to the sample surface, M(B) curves show a clear hysteresis below Curie temperature<br />

(T C =67K for as-grown sample, T C =112K for annealed sample - as indicated from M(T)<br />

measurements).<br />

The measurements <strong>of</strong> M(B) showed that the annealing process also affects the hystere-


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 66<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

& ! '<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!# $%"<br />

<br />

! "<br />

<br />

<br />

<br />

!# &$%%'<br />

<br />

<br />

Fig. 5.14: Magnetization versus temperature M(T) measured in small magnetic field B = 10Gs for<br />

the as-grown <strong>and</strong> annealed at 289 0 C GaMnAs sample with x=0.07. It is clearly visible<br />

that annealing procedure at the optimal conditions leads to the increase <strong>of</strong> saturation<br />

magnetization.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 67<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.15: The hysteresis loop M(B) for the as-grown <strong>and</strong> annealed at the optimal conditions sample<br />

with x=0.07 measured at T=5K. The decrease <strong>of</strong> coercive field is visible after optimal<br />

annealing.<br />

sis loops particularly, the coercive field decreases when the samples are annealed at the<br />

optimal temperatures around 289 0 C. This effect is shown in Figure 5.15. Simultaneously,<br />

the increase <strong>of</strong> the saturation magnetization M S is also observed after heat treatment at the<br />

optimal conditions, indicating that annealing increases the concentration <strong>of</strong> magneticallyactive<br />

Mn ions. Thermal annealing at higher temperature leads to increase <strong>of</strong> the coercive<br />

field <strong>and</strong> decrease <strong>of</strong> the saturation magnetization. Figure 5.16 presents M(T) curves for<br />

the three samples with Mn concentration x=0.061: as-grown, annealed at the temperature<br />

280 0 C <strong>and</strong> higher temperature 350 0 C measured at small magnetic field B=10Gs.<br />

Annealing at high temperature <strong>of</strong> 350 0 C leads to the significant decrease <strong>of</strong> both the<br />

Curie temperature as well as the saturation magnetization. This effect is accompanied<br />

by the large increase <strong>of</strong> the coercive field. This effect is shown in Figure 5.17, where<br />

hysteresis loops for as-grown <strong>and</strong> annealed at 280 0 C <strong>and</strong> higher temperature <strong>of</strong> 350 0 C<br />

samples with x=0.062 (10727D epilayer) are presented.<br />

It has been found by anomalous Hall effect studies [42], [5], that the direction <strong>of</strong> the<br />

easy axis is mainly controlled by epitaxial strain in Ga 1−x Mn x Assystem. The direct magnetization<br />

studies by use <strong>of</strong> SQUID magnetometer (see e.g. [38] <strong>and</strong> references therein)<br />

confirmed that for GaMnAs layer grown on GaAs substrate (i.e. under compressive biaxial<br />

strain) in-plane magnetic easy axis is observed. The hard magnetic axis lies in the<br />

perpendicular direction. In the present work the SQUID magnetization curves were measured<br />

applying magnetic field parallel to the sample surface. An effort to measure SQUID<br />

magnetization in the magnetic field perpendicular to the sample surface was made for the<br />

as-grown <strong>and</strong> annealed 10727E epilayer with x=0.086, but unfortunatelly the signal <strong>of</strong><br />

105 nm thick layer was too small in this case. Recently, it was demonstrated by SQUID<br />

magnetization measurements [54] that GaMnAs films can exhibit rich characteristics <strong>of</strong><br />

magnetic anisotropy depending not only to the epitaxial strain but being strongly influenced<br />

by the Mn as well as hole concentration <strong>and</strong> temperature. The temperature-induced<br />

reorientation <strong>of</strong> the easy axis from [001] to [100], <strong>and</strong> then to [110] or equivalent di-


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 68<br />

!"#$%&# '<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

#$"'<br />

Fig. 5.16: Magnetization versus temperature measured by SQUID magnetometer at small magnetic<br />

field B=10Gs for three samples with x=0.062 (10727D epilayer): as-grown, annealed<br />

at the temperature 289 0 C <strong>and</strong> 350 0 C.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.17: The hysteresis loops M(B) for the sample with x=0.062 as-grown, annealed at the temperature<br />

equal to 289 0 C <strong>and</strong> higher temperature 350 0 C.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 69<br />

y<br />

y<br />

θ K<br />

E i<br />

x<br />

b<br />

a<br />

x<br />

<strong>Magnetic</strong> material<br />

Fig. 5.18: The principle <strong>of</strong> magnetooptical Kerr effect method.<br />

rections occurs in films <strong>of</strong> (001) GaMnAs grown on GaAs substrate with appropriately<br />

low values <strong>of</strong> p. The data collected near T C revealed a non-equivalence <strong>of</strong> the [110] <strong>and</strong><br />

[-110] crystal directions, the latter corresponding to hard axis. This kind <strong>of</strong> anisotropy<br />

develops independently <strong>of</strong> free hole concentration <strong>and</strong> was reported by other groups [80],<br />

[55]. Finally Sawicki et al. reported, that for the samples with relatively high values <strong>of</strong><br />

free hole concentration the hard axis remains oriented along [-110] direction but the three<br />

other main directions ([100], [110] <strong>and</strong> [010]) become equally easy.<br />

5.6 The results <strong>of</strong> magnetooptical Kerr effect measurements<br />

The magnetooptical Kerr effect (MOKE) measurements were performed in order to study<br />

the magnetic <strong>properties</strong> <strong>of</strong> the investigated GaMnAs layers. The experiments were carried<br />

out in the polar configuration under pulsed magnetic fields up to 25T <strong>and</strong> at temperatures<br />

5K≤T≤250K for the as-grown <strong>and</strong> annealed at the optimal conditions epilayers.<br />

First, the briefly outline <strong>of</strong> the MOKE method <strong>and</strong> next, the results <strong>of</strong> the performed<br />

measurements are presented.<br />

5.6.1 Introduction - magnetooptical Kerr effect<br />

The magnetooptical Kerr effect corresponds to a change in the polarization <strong>of</strong> light reflected<br />

from a magnetic material. Reflection <strong>of</strong> a beam <strong>of</strong> linearly polarized light from a<br />

magnetized material causes the polarization to become elliptical, with the main axis rotated<br />

over a small angle θ K with respect to the incident light. Figure 5.18 shows schematically<br />

the principle <strong>of</strong> the magnetooptical Kerr effect.<br />

The elipticity <strong>of</strong> the reflected light is quantified by ɛ (ellipticity angle):<br />

ɛ = arctan(e) (5.3)<br />

where e (ellipticity) is equal to the ratio <strong>of</strong> the length <strong>of</strong> the semi-minor axis <strong>of</strong> the ellipse<br />

(b) to the length <strong>of</strong> its semi-major axis (a): e= b a .<br />

The change in the polarization <strong>of</strong> light is described by Kerr complex angle:<br />

Φ K = θ K + iɛ (5.4)


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 70<br />

where θ K is Kerr rotation angle <strong>and</strong> ɛ is ellipticity angle.<br />

The magnetooptical Kerr effect can be measured in three different configurations.<br />

• Longitudinal Kerr effect: the magnetic field is applied in the plane <strong>of</strong> the sample as<br />

well as in the plane <strong>of</strong> incidence.<br />

• Transverse Kerr effect: the magnetic field is applied in the plane <strong>of</strong> the sample <strong>and</strong><br />

perpendicular to the incidence plane.<br />

• Polar Kerr effect: the applied magnetic field is perpendicular to the sample surface.<br />

Both longitudinal as well as polar magnetooptical Kerr effect configurations have in common<br />

that the magnetization lies in the plane <strong>of</strong> incidence. This is not the case anymore in<br />

transverse configuration. No Kerr rotation <strong>and</strong> ellipticity are expected in this configuration.<br />

Figure 5.19 illustrates possible configurations <strong>of</strong> MOKE.<br />

Longitudinal<br />

Kerr effect<br />

Transverse<br />

Kerr effect<br />

Polar<br />

Kerr effect<br />

Fig. 5.19: Three configurations for magnetooptical Kerr effect measurements: longitudinal, transverse,<br />

polar.<br />

In the present thesis MOKE measurements were carried out in the polar configuration.<br />

A simple interpretation <strong>of</strong> MOKE can be achieved using a macroscopic point <strong>of</strong> view. The<br />

Kerr effect appears because left- <strong>and</strong> right-circularly polarized light waves propagates<br />

differently in the magnetic material. A linearly polarized beam <strong>of</strong> light can be thought as<br />

a superposition <strong>of</strong> such two kinds <strong>of</strong> waves.<br />

The constants pertaining to the MOKE (polar) in the special case <strong>of</strong> normal incidence<br />

have the following form [84]:<br />

<strong>and</strong><br />

θ K = Im{ N + − N −<br />

1 − N + N −<br />

} (5.5)<br />

ɛ = Re{ N + − N −<br />

1 − N + N −<br />

} (5.6)<br />

where: θ K is Kerr rotation angle, N is the complex index <strong>of</strong> refraction (N=n-ik; n=real<br />

index <strong>of</strong> refraction; k=extinction coefficient), N + <strong>and</strong> N − are the complex indexes <strong>of</strong> refraction<br />

for the right-circularly <strong>and</strong> left- polarized light respectively, ɛ is ellipticity angle.<br />

The magnetooptical Kerr effect arises from antisymmetric, <strong>of</strong>f-diagonal elements in<br />

the optical conductivity tensor σ xy [84]. The <strong>of</strong>f-diagonal elements <strong>of</strong> conductivity tensor<br />

lead to different indices <strong>of</strong> refraction for right- <strong>and</strong> left-circular light in a medium <strong>and</strong> are<br />

thus responsible for mantetooptical effects (for instance magnetooptical Kerr effect).<br />

MOKE is proportional to the net magnetization <strong>of</strong> <strong>ferromagnetic</strong> sample. In the case<br />

<strong>of</strong> a non<strong>ferromagnetic</strong> material MOKE is proportional to the external magnetic field.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 71<br />

The calculations performed by H. S. Bennet <strong>and</strong> E. A. Stern [85] revealed that for<br />

<strong>ferromagnetic</strong> material:<br />

where M is magnetization.<br />

For non<strong>ferromagnetic</strong> material:<br />

Φ ferro<br />

K<br />

∝ σ xy ∝ M (5.7)<br />

Φ non<strong>ferromagnetic</strong><br />

K<br />

∝ σ xy ∝ B (5.8)<br />

where B is magnetic field.<br />

Another difference between these two types <strong>of</strong> materials is the order <strong>of</strong> magnitude.<br />

The rotation for nonferromagnet is much smaller then for a <strong>ferromagnetic</strong> material.<br />

A real microscopic explanation <strong>of</strong> magnetooptical Kerr effect in <strong>ferromagnetic</strong> materials<br />

was initiated by Hulme in 1932 <strong>and</strong> completed by P. Argyres in 1955 [84]. Argyres<br />

showed that including the interaction between an electron <strong>and</strong> an effective field that it<br />

"feels" as it moves through a material leads to non-zero <strong>of</strong>-diagonal elements in the conductivity<br />

<strong>and</strong> polarizability tensors that determine the index <strong>of</strong> refraction. The key step<br />

was introduction <strong>of</strong> spin-orbit interaction. The calculations <strong>of</strong> Argyres show that tensors<br />

<strong>of</strong> conductivity <strong>and</strong> poarizability have non-zero <strong>of</strong>f diagonal elements which come directly<br />

from included spin-orbit term. Knowledge <strong>of</strong> these tensor elements allowed him to<br />

calculate the difference in the index <strong>of</strong> refraction in right- <strong>and</strong> left- circular modes. The<br />

Kerr constants θ K <strong>and</strong> ɛ can be obtained using the following relationship:<br />

N + − N −<br />

= 1 σ xy<br />

1 − N + N − ε 0 ω N(1 − N 2 )<br />

(5.9)<br />

5.6.2 The results <strong>of</strong> MOKE measurements for the as-grown <strong>and</strong> annealed GaMnAs<br />

epilayers<br />

In order to study the magnetic <strong>properties</strong> <strong>of</strong> GaMnAs layers the magnetooptical Kerr effect<br />

(MOKE) measurements were performed. The experiments were carried out in the<br />

polar configuration for four epilayers in a wide range <strong>of</strong> Mn concentration: as grown <strong>and</strong><br />

annealed 10727E samples with x=0.086, as grown 11127A sample with x=0.048 <strong>and</strong> as<br />

grown 10529A sample with x=0.014. The experimental setup used for these measurements<br />

is schematically shown in Figure 2.7 <strong>and</strong> described in details in Chapter 2.<br />

The MOKE measurements were performed for two wavelengths <strong>of</strong> incident light:<br />

λ=632.8 nm (red HeNe laser) <strong>and</strong> λ=540.5 nm (green HeNe laser). The magnetooptical<br />

studies were carried out under pulsed magnetic fields up to 25T <strong>and</strong> at temperatures<br />

ranging from 5K up to 250K.<br />

Figure 5.20 presents Kerr rotation angle θ K versus magnetic field up to 6T for the as<br />

grown epilayer with high Mn content x=0.086 <strong>and</strong> a Curie temperature T C =88K.<br />

The data were collected at various temperatures - below <strong>and</strong> above Curie temperature.<br />

The magnetic field dependence <strong>of</strong> Kerr rotation angle θ K at different temperatures for the<br />

same epilayer with x=0.086 annealed at the optimal conditions T a = 289 0 C is shown in the<br />

Figure 5.21.<br />

It is clearly visible that below B=1T <strong>and</strong> for T below T C , θ K (B) exhibits a non monotonic<br />

field dependence. This non monotonic behavior is observed in the Kerr rotation<br />

curves for all <strong>of</strong> the investigated samples: as grown <strong>and</strong> annealed 10727E epilayer with


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 72<br />

θ Κ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!" #<br />

!!<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.20: Kerr rotation angle θ K versus magnetic field up to 6T for the as grown epilayer with<br />

high Mn content x=0.086 <strong>and</strong> a Curie temperature T C =88K.<br />

<br />

θ Κ <br />

<br />

<br />

<br />

<br />

<br />

!<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.21: The magnetic field dependence <strong>of</strong> Kerr rotation angle θ K at different temperatures for<br />

the epilayer with x=0.086 annealed at the optimal conditions T a = 289 0 C<br />

, λ=632.8nm.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 73<br />

<br />

<br />

<br />

λ <br />

<br />

<br />

θ Κ<br />

/θ Κ <br />

<br />

<br />

<br />

! "<br />

<br />

! "<br />

#<br />

! "<br />

#<br />

! "<br />

<br />

<br />

<br />

Fig. 5.22: The results <strong>of</strong> MOKE obtained at T=5K for λ=632.8nm for all investigated samples<br />

as grown <strong>and</strong> annealed 10727E epilayer with x=0.086 <strong>and</strong> for epilayers with lower Mn<br />

concentration: 11127A (x=0.048) as well as 10529A (x=0.014).<br />

x=0.086 <strong>and</strong> for epilayers with lower Mn concentration: 11127A (x=0.048) as well as<br />

10529A (x=0.014). The results obtained at T=5K for two wavelengths <strong>of</strong> incident light<br />

λ=632.8nm <strong>and</strong> λ=540.5nm are presented in Figure 5.22 <strong>and</strong> Figure 5.23 respectively.<br />

These figures indicate very clearly that low magnetic field feature shifts towards higher<br />

magnetic fields after annealing. The low magnetic field dependence <strong>of</strong> Kerr rotation angle<br />

was measured systematically at different temperatures (below as well as above Curie<br />

temperature) for two samples - as grown <strong>and</strong> annealed at the optimal conditions 10727E<br />

epilayer. These measurements showed that the non monotonic low magnetic field behavior<br />

is related to the magnetic <strong>properties</strong> <strong>of</strong> the studied epilayers. The observed feature is<br />

present up to temperatures near T C - it disappears at the temperatures just below Curie<br />

temperature. The shift towards higher magnetic field with the increase <strong>of</strong> Mn content<br />

is also visible. The comparison <strong>of</strong> θ K (B) behaviour for two used wavelengths (look in<br />

Figures 5.24 <strong>and</strong> 5.25 where magnetic field dependence <strong>of</strong> Kerr rotation angle for two<br />

as grown epilayers with x=0.086 <strong>and</strong> x=0.014 for red as well as green laser lines is presented)<br />

indicates that the position in the magnetic field is the same for red as well as green<br />

laser, however the change in the shape <strong>of</strong> the low magnetic field feature is visible.<br />

To compare the measured magnetic field dependence <strong>of</strong> Kerr rotation angle with the<br />

absolute magnetization M, additionally the SQUID investigations were performed for as<br />

grown 10727E sample with x=0.086. The measurements were performed in the perpendicular<br />

configuration at low temperature T=5K. The diamagnetic component <strong>of</strong> diamagnetic<br />

GaAs substrate was subtracted from the total response to obtain magnetization <strong>of</strong><br />

the GaMnAs layer. As was mentioned in the previous section, the signal <strong>of</strong> the 105 nm


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 74<br />

<br />

<br />

<br />

λ <br />

<br />

θ Κ<br />

/θ Κ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"<br />

Fig. 5.23: The results <strong>of</strong> MOKE obtained at T=5K for λ=540.5nm for all investigated samples<br />

as grown <strong>and</strong> annealed 10727E epilayer with x=0.086 <strong>and</strong> for epilayers with lower Mn<br />

concentration: 11127A (x=0.048) as well as 10529A (x=0.014).<br />

<br />

<br />

<br />

<br />

<br />

<br />

θ Κ<br />

/θ Κ <br />

<br />

<br />

λ <br />

λ <br />

<br />

<br />

<br />

!" #$<br />

Fig. 5.24: The comparison <strong>of</strong> θ K (B) behaviour for two used wavelengths (two wavelengths <strong>of</strong><br />

incident light λ=632.8nm <strong>and</strong> λ=540.5nm) for as grown epilayer with x=0.086.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 75<br />

<br />

<br />

λ <br />

<br />

θ Κ<br />

/θ Κ <br />

<br />

<br />

<br />

λ <br />

<br />

<br />

<br />

<br />

<br />

!"#$<br />

Fig. 5.25: The comparison <strong>of</strong> θ K (B) behaviour for two used wavelengths (two wavelengths <strong>of</strong><br />

incident light λ=632.8nm <strong>and</strong> λ=540.5nm) for as grown epilayer with x=0.014.<br />

thick GaMnAs layer was small <strong>and</strong> noisy for this configuration (hard axis). Figure 5.26<br />

presents measured magnetization M versus magnetic field at 5K. Unfortunately, this data<br />

does not give the unique answer about the existence <strong>of</strong> low magnetic field feature in the<br />

SQUID magnetization.<br />

At present only very speculative interpretation <strong>of</strong> the measured data is possible. It<br />

is known (see previous section) that GaMnAs films exhibit rich characteristics <strong>of</strong> magnetic<br />

anisotropy [54], [80], [55]. The low magnetic field behavior <strong>of</strong> Kerr rotation angle<br />

described above can be related with the magnetic anisotropy observed in GaMnAs epilayers.<br />

However, up to now the MOKE measurements as a function <strong>of</strong> wavelength were<br />

not investigated.<br />

Nevertheless, the SQUID measurements performed in the perpendicular configuration<br />

together with MOKE studies indicate that even at low temperatures <strong>and</strong> high magnetic<br />

fields the measured magnetization is far from being saturated. In addition, comparison <strong>of</strong><br />

the two data sets: Kerr rotation angle with the SQUID magnetization M (look in Figure<br />

5.27) collected in the same field range indicates that the usual relation θ K ∝ M, observed<br />

e.g. for metallic magnetic metals, is not valid for GaMnAs.<br />

The magnetic field dependence <strong>of</strong> Kerr rotation angle was also studied in the range<br />

<strong>of</strong> high magnetic fields. Figure 5.28 presents θ K (B) curves measured up to 25T at low<br />

temperature equal to 5K.<br />

For all the investigated samples in the wide range <strong>of</strong> Mn concentration the oscillatory<br />

terms are clearly visible in the range <strong>of</strong> high magnetic fields. For high magnetic fields,<br />

where the following condition is satisfied: s=ω C τ ≫ 1 (ω C is cyclotron frequency <strong>and</strong> τ<br />

is life time), the quantum phenomena can be observed - in optics the resonance optical


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 76<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$<br />

%<br />

%&<br />

<br />

<br />

<br />

<br />

Fig. 5.26: SQUID magnetization M versus magnetic field at 5K. The magnetic field was applied<br />

perpendicular to the sample surface.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

""#$%<br />

&<br />

!&<br />

<br />

<br />

<br />

θ Κ <br />

<br />

<br />

!<br />

<br />

Fig. 5.27: The comparison <strong>of</strong> the two data sets: Kerr rotation angle with the SQUID magnetization<br />

M(B) collected at T=5K.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 77<br />

θ Κ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

λ <br />

<br />

!"<br />

<br />

#!"<br />

$#!"<br />

<br />

<br />

<br />

<br />

Fig. 5.28: The results <strong>of</strong> MOKE obtained at T=5K for λ=632.8nm in the range <strong>of</strong> high magnetic<br />

fields for all investigated samples as grown <strong>and</strong> annealed 10727E epilayer with x=0.086<br />

<strong>and</strong> for epilayers with lower Mn concentration: 11127A (x=0.048) as well as 10529A<br />

(x=0.014)


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 78<br />

transitions. The oscillatory terms visible in the Figure 5.28 can be related to the interb<strong>and</strong><br />

optical transitions between L<strong>and</strong>au levels, i.e. resonance interb<strong>and</strong> transitions from heavy<br />

holes, light holes, spin-split-<strong>of</strong> b<strong>and</strong>s to conduction b<strong>and</strong>. The similar oscillatory behavior<br />

<strong>of</strong> Kerr rotation angle in high magnetic fields for GaMnAs epilayers was observed by N.<br />

Négre [86]. The author studied MOKE under high pulsed magnetic fields (up to 32T) for<br />

GaMnAs epilayers with low Mn concentration - up to x=0.034. The performed Fourier<br />

transform analysis <strong>of</strong> θ K (1/B) curves measured at T=5K for GaMnAs epilayers as well<br />

as for low temperature epilayers <strong>of</strong> GaAs allowed to obtain the resonance frequencies.<br />

In the case <strong>of</strong> GaAs three frequencies <strong>of</strong> the measured oscillations were clearly visible:<br />

F 1 =276T, F 2 =182T <strong>and</strong> F 3 =42T. Additionally, the first harmonic <strong>of</strong> frequency F 3 equal<br />

to 80T was observed. These three frequencies author assigned to the following interb<strong>and</strong><br />

transitions: F 1 - from heavy holes b<strong>and</strong> to conduction b<strong>and</strong>, F 2 - from light hole b<strong>and</strong><br />

to conduction b<strong>and</strong> <strong>and</strong> F 3 - from spin-split-<strong>of</strong> b<strong>and</strong> to conduction b<strong>and</strong>. For GaMnAs<br />

epilayers the same frequencies were observed as for GaAs sample. Additional frequency<br />

F equal to 141T was visible for GaMnAs sample with the highest Mn concentration.<br />

In the present thesis all measured curves characterized by oscillatory terms at high<br />

magnetic fields were analyzed by use <strong>of</strong> Fourier transform. Figures 5.29 <strong>and</strong> 5.30 present<br />

Kerr rotation angle oscillatory terms as a function <strong>of</strong> inverse magnetic field measured at<br />

T=5K. The results <strong>of</strong> Fourier transform analysis are gathered in Figures 5.31 <strong>and</strong> 5.32.<br />

For the sample 10529A with x=0.014 (look in Figure 5.31 the following frequencies<br />

are distinct: F 1 =277±5T, F 2 =179±5T, F 3 =93±5T <strong>and</strong> F 4 =45T±4T. For the sample<br />

11127A with x=0.048 the frequencies: F 1 =286±5T, F 2 =186±5T, F 3 =72±5T <strong>and</strong><br />

F 4 =49T±4T are distinct (see Figure 5.31). In the case <strong>of</strong> as grown 10727E epilayer<br />

with the higest Mn concentration (x=0.086) clearly three frequencies can be separated:<br />

F 1 =172±5T, F 2 =93±5T, F 3 =38±4T (Figure 5.32). In the range <strong>of</strong> higher frequencies<br />

any distinct peak can be isolated from the background. Finally, for the annealed at the optimal<br />

conditions 10727E epilayer one can distinguish the following distinct frequencies:<br />

F 1 =278±5T, F 2 =179±5T,F 3 =71±5T <strong>and</strong> F 4 =53T±4T. (Figure 5.32)<br />

There is no distinct peak around 141T for all investigated GaMnAs samples in a wide<br />

range <strong>of</strong> Mn concentration. The determined values <strong>of</strong> oscillations frequencies are very<br />

close to these observed for GaAs by N. Négre. One needs to realize that in the performed<br />

experiment the incident beam <strong>of</strong> light can penetrate deeper then the thickness <strong>of</strong> GaMnAs<br />

epilayer <strong>and</strong> reflections from GaAs are also possible. The observed interb<strong>and</strong> transitions<br />

can correspond to GaMnAs thin film as well as GaAs. Nevertheless, the performed<br />

Fourier transform analysis does not definitively settle this question.<br />

Additionally, the shift <strong>of</strong> oscillating terms with magnetic field after thermal annealing<br />

is observed. This effect is presented in Figure 5.33.<br />

The shift towards higher magnetic fields is clearly visible. The high resolution X-<br />

ray diffraction measurements showed (see section 8 <strong>of</strong> this Chapter) that annealing at the<br />

optimal conditions changes strain relations. The strain is reduced by annealing process.<br />

The induced by annealing changes in the strain relations modify the b<strong>and</strong> structure <strong>of</strong><br />

GaMnAs (b<strong>and</strong>s splitting, anisotropy) <strong>and</strong> this in turn may be responsible for the observed<br />

shift <strong>of</strong> magnetic fields oscillating terms.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 79<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$%!<br />

Fig. 5.29: The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> inverse magnetic field<br />

measured at T=5K for two epilayers: 10529A (x=0.014) <strong>and</strong> 11127A (x=0.048).


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 80<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$ <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"# <br />

Fig. 5.30: The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> inverse magnetic field<br />

measured at T=5K for as-grown as well as annealed sample <strong>of</strong> 10727E epilayer with<br />

x=0.086.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 81<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$<br />

%<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!"#$<br />

%<br />

<br />

<br />

<br />

&<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.31: The results <strong>of</strong> Fourier transform analysis <strong>of</strong> Kerr rotation angle curves measured at<br />

T=5K for two epilayers: 10529A (x=0.014) <strong>and</strong> 11127A (x=0.048)


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 82<br />

<br />

<br />

'<br />

!"#$%<br />

&'<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

"<br />

#<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!<br />

Fig. 5.32: The results <strong>of</strong> Fourier transform analysis <strong>of</strong> Kerr rotation angle curves measured at<br />

T=5K for as-grown as well as annealed sample <strong>of</strong> 10727E epilayer with x=0.086.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 83<br />

<br />

<br />

<br />

λ <br />

<br />

!<br />

!"#<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.33: The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> magnetic field B measured<br />

at T=5K for as-grown as well as annealed sample <strong>of</strong> 10727E epilayer with x=0.086.<br />

The shift <strong>of</strong> oscillating terms towards higher magnetic field after thermal annealing is<br />

visible.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 84<br />

<br />

<br />

Ψ<br />

<br />

Fig. 5.34: Schematic view <strong>of</strong> the channeling <strong>of</strong> ions directed at an angle Ψ to a close-packed row<br />

<strong>of</strong> atoms in a crystal.<br />

5.7 The channeling experiments - the results <strong>of</strong> c-RBS <strong>and</strong> c-PIXE<br />

measurements<br />

The channeling experiments i.e. channeling Rutherford backscattering (c-RBS) <strong>and</strong> channeling<br />

particle-induced X-ray emission (c-PIXE) measurements were performed on the<br />

as-grown <strong>and</strong> annealed samples <strong>of</strong> Ga 1−x Mn x As.<br />

The results <strong>of</strong> channeling measurements [13] [52] indicated that low temperature annealing<br />

introduces a rearrangement <strong>of</strong> Mn sites in Ga 1−x Mn x As lattice. The combined<br />

c-RBS <strong>and</strong> c-PIXE studies revealed the important role <strong>of</strong> interstitial Mn atoms in <strong>ferromagnetic</strong><br />

epilayers <strong>of</strong> Ga 1−x Mn x As. The channeling experiments clearly established correlation<br />

between the arrangement <strong>of</strong> Mn sites in Ga 1−x Mn x As <strong>and</strong> experimental effects<br />

observed after low temperature annealing performed at the optimal conditions (described<br />

in the previous sections <strong>of</strong> this Chapter), i.e., the increase <strong>of</strong> the Curie temperature, the<br />

increase in the saturation magnetization observed at low temperatures, the increase <strong>of</strong><br />

the conductivity <strong>and</strong> the increase <strong>of</strong> the hole concentration (measured by means <strong>of</strong> ECV<br />

method [46], [13], [52], [87]).<br />

Mn atoms incorporated into <strong>ferromagnetic</strong> Ga 1−x Mn x As lattice can occupy three distinct<br />

types <strong>of</strong> lattice site: substitutional positions in Ga sublattice Mn Ga , where Mn ++<br />

ions act as acceptors <strong>and</strong> also contribute to uncompensated spins; interstitial positions<br />

Mn I , commensurate with the zinc-blende lattice structure, where they act as a donors <strong>and</strong><br />

tend to passivate Mn Ga acceptors; <strong>and</strong> r<strong>and</strong>om locations Mn ran incommensurate with the<br />

zinc-blende lattice, i.e. that form clusters <strong>of</strong> Mn or MnAs clusters.<br />

In Ga 1−x Mn x As only substitutional Mn ++ ions act as acceptors, so Mn concentration<br />

x <strong>and</strong> free hole concentration p are closely related.<br />

Channeling is the steering <strong>of</strong> a beam <strong>of</strong> energetic ions into open spaces (channels)<br />

between close-packed rows or planes <strong>of</strong> atoms in a crystal, as shown schematically in<br />

Figure 5.34 <strong>and</strong> Figure 5.35.<br />

The steering is the result <strong>of</strong> a correlated series <strong>of</strong> small-angle screened Coulomb collisions<br />

between an ion <strong>and</strong> the atoms bordering the channel. Thus, channeled ions do<br />

not penetrate closer than the screening distance <strong>of</strong> the vibrating atomic nuclei, <strong>and</strong> the<br />

probability <strong>of</strong> large-angle Rutherford collisions (back-scattering), nuclear reactions, or<br />

inner-shell X-ray excitation is greatly reduced compared with the probability <strong>of</strong> such in-


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 85<br />

χ<br />

<br />

<br />

<br />

Fig. 5.35: The normalized yield χ h <strong>of</strong> ions that are backscattered from host atoms (the RBS yield)<br />

shows a strong dip at Ψ=0. If 50% <strong>of</strong> solute atoms are displaced into the channel, the<br />

normalized yield χ s <strong>of</strong> ions backscattered from the solute atoms is approximately half<br />

the r<strong>and</strong>om yield; i.e., χ s =0.5 at Ψ=0 (broken curve). If displaced solute atoms are<br />

located near the center <strong>of</strong> the channel, a peak in yield may occur (dotted line).<br />

Ψ<br />

teractions from a non-channeled (r<strong>and</strong>om) beams <strong>of</strong> ions. For the ions incident at small<br />

angles Ψ to a close-packed direction, a large reduction in yield from such interactions<br />

with host atoms is observed (look in Figure 5.35). The normalized yield χ h from host<br />

atoms for such interactions is defined by the ratio <strong>of</strong> the yield for ions incident at an angle<br />

Ψ to the yield for a "r<strong>and</strong>omly" directed beam.<br />

The position <strong>of</strong> solute atoms in a crystal lattice can be determined quite directly <strong>and</strong><br />

precisely from channeling experiments by measuring the normalized yields χ h from host<br />

atoms <strong>and</strong> χ s from solute atoms for the same depth increment in the crystal [88], [89]. If<br />

solute atoms are in the same lattice sites as host atoms, then χ s<br />

∼ = χh for any angle Ψ. If<br />

solute atoms project into a given channel, the channeled ions interact with them, causing<br />

an increased yield χ s . If, for example, 50% <strong>of</strong> solute atoms project into a given channel,<br />

the value <strong>of</strong> χ s for that channel would be 0.5 if the ion flux distribution were uniform in<br />

a channel (look in Figure 5.35). Because <strong>of</strong> the steering action involved in channeling,<br />

the ion flux is peaked near the center <strong>of</strong> a channel. Thus, if the solute atoms are displaced<br />

to positions near the center <strong>of</strong> a channel, a peaking in χ s occurs near perfect alignment<br />

(Ψ=0). Such a peaking effect is unambiguous evidence that the solute atoms lie near the<br />

center <strong>of</strong> the channel. By comparing yields for different channels, the position <strong>of</strong> solute<br />

atoms can be determined rather accurately by a triangulation procedure.<br />

The Ga 1−x Mn x As samples studied in the present thesis were investigated by directly<br />

comparing the Mn K α X-ray signals (c-PIXE) with the c-RBS signals <strong>of</strong> GaAs from the<br />

Ga 1−x Mn x As film simultaneously obtained using a 1.95 MeV 4 He + beam. The location<br />

<strong>of</strong> Mn sites in Ga 1−x Mn x As lattice was studied by comparing <strong>of</strong> these two signals for<br />

different channels.<br />

The channeling experiments revealed that in LT MBE grown <strong>ferromagnetic</strong><br />

Ga 1−x Mn x As with high Mn concentration x a significant fraction <strong>of</strong> incorporated Mn


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 86<br />

<br />

Fig. 5.36: Schematic <strong>of</strong> the tetrahedral interstitial positions for a zinc-blende lattice along the various<br />

axial directions.<br />

atoms (14% for the as-grown 10823C GaMnAs epilayer with x=0.092 as determined by<br />

PIXE measurements) occupies well defined, commensurate with the GaAs lattice interstitial<br />

positions. The most important experimental result is the fact that low temperature<br />

annealing performed at the optimal conditions leads to the significant reduction <strong>of</strong> Mn<br />

interstitial atoms (to 7% <strong>of</strong> the total Mn content for Ga 0.908 Mn 0.092 )[13] [52]. An increase<br />

in Mn interstitials <strong>and</strong> a decrease in substitutional Mn acceptors is observed with the<br />

increase <strong>of</strong> Mn content [52].<br />

In the zinc-blende lattice structure there are two possible interstitial positions, i.e. the<br />

tetrahedral interstitial positions with the four nearest neighbors cations <strong>and</strong> the hexagonal<br />

interstitial positions with the six (three cations <strong>and</strong> three anions) nearest neighbors. Impurity<br />

atoms in the interstitial sites in a diamond lattice are shadowed by the host atoms<br />

in the 〈100〉 <strong>and</strong> 〈111〉 directions but are exposed in the 〈110〉 axial channel. They can<br />

be distinguished by studying angular scans around the 〈110〉 axial direction. Schematic<br />

<strong>of</strong> the tetrahedral interstitial positions for a zinc-blende lattice along the various axial<br />

directions are shown in Figure 5.36.<br />

The arrangement <strong>of</strong> the tetrahedral interstitial positions in a zinc-blende lattice along<br />

〈110〉 gives rise to a double-peak feature due to the flux peaking effect <strong>of</strong> the ion beam<br />

in the channel. Figures 5.37, 5.38, 5.39 show the PIXE <strong>and</strong> RBS angular scans about<br />

the 〈100〉, 〈110〉 <strong>and</strong> 〈111〉 axes for as-grown sample with x=0.092. The normalized<br />

yield for the RBS (χ GaAs ) or the PIXE Mn X-ray signals (χ Mn ) is defined as the ratio<br />

<strong>of</strong> the channeled yield to the corresponding unaligned yield. The higher (χ Mn ) for the<br />

as-grown film observed in the 〈110〉 direction as compared to those in 〈111〉 〈100〉 directions<br />

suggest that non-r<strong>and</strong>om fraction <strong>of</strong> Mn ions in these samples do not all sit in<br />

substitutional sites. A double-peak feature is observable in the 〈110〉 scan (indicated by<br />

arrows in Figure 5.38). These results indicate that a fraction <strong>of</strong> non-r<strong>and</strong>om Mn atoms<br />

is located at the tetrahedral interstitial sites in which the interstitial is surrounded by four<br />

nearest neighbors. The fraction <strong>of</strong> these interstitial Mn atoms can be roughly estimated<br />

to be ∼14%, assuming that flux peaking in the 〈110〉 channel <strong>of</strong> GaAs is ∼1.5 [90], [89].<br />

The PIXE <strong>and</strong> RBS angular scans taken about the 〈110〉 <strong>and</strong> 〈111〉 axes for annealed at<br />

the optimal conditions (T=289 0 C) sample (see Reference [13]) show that the double-peak<br />

feature is less prominent for the sample annealed at the optimal conditions, indicating a<br />

reduced concentration <strong>of</strong> interstitial Mn in this sample. This suggest that Mn interstitials<br />

are highly unstable. The interstitial Mn atoms are reduced to 7% <strong>of</strong> the total Mn content.<br />

For the sample annealed at high temperature equal to 350 0 C [13] the normalized yield<br />

(χ Mn ) values are high <strong>and</strong> nearly equal in all three channeling scans. This suggest that<br />

not only all <strong>of</strong> the interstitial Mn, but also a significant fraction <strong>of</strong> the Mn atoms originally<br />

at the substitutional positions leave their positions. It was shown [52] that Mn interstitials


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 87<br />

1.2<br />

x=0.092<br />

<br />

Normalized yield, χ<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

host (RBS)<br />

Mn (PIXE)<br />

0.0<br />

-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5<br />

tilt angle (degree)<br />

2.0<br />

Fig. 5.37: The PIXE <strong>and</strong> RBS angular scans about the 〈100〉 axes for as-grown sample with<br />

x=0.092.<br />

increase (from 5% to 14%) as the Mn content increases (from x=0.02 to x=0.092).<br />

Figure 5.40 shows the fraction <strong>of</strong> nonr<strong>and</strong>om Mn f nr , calculated by comparing normalized<br />

yields χ Mn <strong>and</strong> χ GaAs using the following relation [89]:<br />

f nr = 1 − χ Mn<br />

1 − χ GaAs<br />

(5.10)<br />

For both the as-grown GaMnAs epilayer <strong>and</strong> annealed at the optimal conditions<br />

(289 0 C), the f nr values are similar for the 〈100〉 <strong>and</strong> 〈111〉 projections but show a much<br />

lower value for 〈110〉 axial direction. This is due to the flux peaking effect, i.e. Mn<br />

atoms are visible in the 〈110〉 channel <strong>and</strong> X-rays coming from these interstitial Mn are<br />

enhanced. Thus, non-r<strong>and</strong>om fraction <strong>of</strong> Mn atoms f nr calculated from χ Mn does not<br />

represent the true non-r<strong>and</strong>om Mn fraction. For the sample annealed at 350 0 C the f nr<br />

values are similar in all axial directions. The significant decrease <strong>of</strong> f nr is visible after<br />

thermal annealing at higher temperature (higher then 300 0 C), suggesting that post-growth<br />

annealing promotes r<strong>and</strong>om precipitation <strong>of</strong> Mn. 35% <strong>of</strong> the Mn atoms is estimated as<br />

a forming r<strong>and</strong>om precipitates after thermal annealing at high temperature 350 0 C. It was<br />

reported that Mn removed from ordered sites forms MnAs inclusions after annealing at<br />

the temperatures higher then 300 0 C [91].<br />

The most important experimental fact is that in the as-grown sample ∼ 14% <strong>of</strong> the<br />

Mn atoms reside on the tetrahedral interstitial sites. One should realize that the interstitial<br />

Mn I donor is both positively charge <strong>and</strong> relatively mobile. It is therefore expected<br />

to drift toward the negatively charged Mn Ga acceptor centers, thus forming Mn Ga -Mn I<br />

pairs. The electrostatic attraction between positively charged Mn interstitials Mn I <strong>and</strong><br />

negative substitutional Mn Ga acceptors stabilizes highly mobile Mn I in intersitial sites<br />

adjacent to Mn Ga , i.e. the tetrahedral position between four cations is preferred. Re-


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 88<br />

Normalized yield, χ<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

x=0.092<br />

<br />

host (RBS)<br />

Mn (PIXE)<br />

0.0<br />

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5<br />

tilt angle (degree)<br />

2.0<br />

Fig. 5.38: The PIXE <strong>and</strong> RBS angular scans about the 〈110〉 axes for as-grown sample with<br />

x=0.092.<br />

Normalized yield, χ<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

x=0.092<br />

<br />

host (RBS)<br />

Mn (PIXE)<br />

0.2<br />

0.0<br />

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0<br />

tilt angle (degree)<br />

Fig. 5.39: The PIXE <strong>and</strong> RBS angular scans about the 〈111〉 axes for as-grown sample with<br />

x=0.092.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 89<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

$%<br />

$%<br />

$%<br />

<br />

<br />

! "#<br />

Fig. 5.40: The fraction <strong>of</strong> nonr<strong>and</strong>om Mn f nr , calculated by comparing normalized yields χ Mn<br />

<strong>and</strong> χ GaAs for the 〈100〉, 〈110〉 <strong>and</strong> 〈111〉 projections.<br />

cently, this was confirmed by a calculation <strong>of</strong> the total energy <strong>of</strong> Mn ions located at different<br />

interstitial sites, which has shown that the tetrahedral interstitial position between<br />

cations is energetically favorable in strongly p-type material [92]. However it should be<br />

mentioned that, K. W. Edmonds et al. claim that according to their calculations, in p-<br />

type samples the tetrahedral interstitial site between the four As anions should be favored<br />

[93]. Annealing the sample at the optimal conditions (at the temperatures only slightly<br />

above the growth temperature) breaks relatively weak Mn I -Mn Ga . The decrease <strong>of</strong> the<br />

compensating Mn I donors leads to increase <strong>of</strong> the number <strong>of</strong> electrically-active Mn G a,<br />

<strong>and</strong> thus also the hole concentration. The electrochemical capacitance-voltage pr<strong>of</strong>iling<br />

(ECV) method measurements [13], [52] showed an increase in the free hole concentration<br />

from 6·10 20 cm −3 observed in the as-grown sample to 1·10 21 cm −3 for the GaMnAs<br />

epilayer annealed at the optimal conditions. As is well known, the increase in the hole<br />

concentration will automatically result in an increase <strong>of</strong> T C . Because removal <strong>of</strong> Mn I<br />

by annealing is accompanied by an increase <strong>of</strong> saturation magnetization (as evidenced<br />

by magnetization studies), this suggest that the Mn Ga -Mn I pairs are coupled anti<strong>ferromagnetic</strong>ally<br />

i.e., the magnetic moment <strong>of</strong> Mn I "neutralizes" the contribution <strong>of</strong> Mn Ga to<br />

the magnetization. Removal <strong>of</strong> Mn I from such a pair should thus automatically render<br />

the substitutional Mn ++ magnetically-active, increasing the saturation magnetization, as<br />

is indeed observed experimentally. Theoretical predictions [94], [95] revealed that Mn<br />

interstitials make some <strong>of</strong> the substitutional Mn ions magnetically inactive by forming<br />

with them close pairs, with the spins anti<strong>ferromagnetic</strong>ally coupled by the superexchange<br />

mechanism.<br />

The lattice rearrangement <strong>of</strong> Mn atoms in Ga 1−x Mn x As system provides a clear explanation<br />

<strong>of</strong> the experimental effects, i.e. the increase <strong>of</strong> the conductivity, the increase <strong>of</strong> the<br />

Curie temperature, <strong>and</strong> the increase in the saturation magnetization observed at low temperatures<br />

for the samples annealed at the optimal conditions. The results <strong>of</strong> channeling<br />

experiments (c-PIXE, c-RBS) revealed that in the as-grown <strong>ferromagnetic</strong> Ga 1−x Mn x As


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 90<br />

samples a fraction <strong>of</strong> incorporated Mn atoms occupies interstitial tetrahedral positions, the<br />

more significant the higher the Mn content (∼14% for the epilayer with x=0.092, whereas<br />

5% for x=0.02 [52]). K. M. Yu et al. [46] uses thermodynamical arguments to explain<br />

why the amount <strong>of</strong> interstitials in the sample, <strong>and</strong> subsequently the sample’s reaction to<br />

the annealing, depend on the Mn content: the Fermi level is pushed towards or into the<br />

valence b<strong>and</strong> by the increasing number <strong>of</strong> Mn substitutional acceptors, the formation <strong>of</strong><br />

interstitiasls becomes energetically favorable.<br />

5.8 The results <strong>of</strong> diffraction (HRXRD) measurements<br />

In the present thesis the structural investigation <strong>of</strong> as-grown as well as annealed<br />

Ga 1−x Mn x As epilayers was carried out using high resolution X-ray diffraction (HRXRD)<br />

measurements for a wide range <strong>of</strong> Mn concentrations (0.027 ≤ x ≤ 0.086), with a special<br />

attention on how the interstitial Mn atoms (Mn I ) influence the lattice parameter <strong>of</strong> this<br />

material.<br />

It was experimentally established that the lattice constant <strong>of</strong> Ga 1−x Mn x As layers increases<br />

with the increase <strong>of</strong> Mn concentration [43].<br />

The lattice constant a 0 measurements have <strong>of</strong>ten been used as a method <strong>of</strong> determining<br />

the Mn concentration x in Ga 1−x Mn x As. While there clearly exists a phenomenological<br />

correlation between a 0 <strong>and</strong> x, based on the remarks just made it is now clear that this<br />

correlation is quite complex <strong>and</strong> is not really understood. For example, it was recently<br />

reported that epilayers <strong>of</strong> Ga 1−x Mn x As with the same composition but prepared using<br />

different growth parameters have quite different lattice constants [96]. Moreover, it was<br />

suggested that Ga 1−x Mn x As is an example <strong>of</strong> a system does not obey Vegard’s law in the<br />

traditional sense <strong>of</strong> a linear variation between the two end-point compounds, i.e., between<br />

zinc blende GaAs <strong>and</strong> (hypothetical) zinc blende MnAs [96].<br />

Recently, first-principles theoretical calculations [44] have predicted that the presence<br />

<strong>of</strong> Mn interstitials atoms can be the reason <strong>of</strong> the observed expansion <strong>of</strong> the lattice constant<br />

<strong>of</strong> Ga 1−x Mn x As. It is also well known that a high arsenic antisite concentrations<br />

(As Ga ) in low temperature GaAs (LT-GaAs) also leads to an increase in the lattice constant.<br />

The As Ga defects are expected to be present in Ga 1−x Mn x As layers, since these<br />

are grown at low temperatures similar to those used for growing LT-GaAs. It has been<br />

shown that the lattice constant <strong>of</strong> Ga 1−x Mn x As additionally depends in a sensitive way on<br />

the growth conditions, quite possibly due to excess As incorporation, the degree <strong>of</strong> which<br />

may in itself depend on the presence <strong>of</strong> Mn in the system [40], [96].<br />

One <strong>of</strong> the aims <strong>of</strong> the performed structural investigations was to explore the influence<br />

<strong>of</strong> Mn interstitials atoms on the lattice constant <strong>of</strong> Ga 1−x Mn x As which, as noted in<br />

Reference [44], is expected to strongly affect the lattice parameter.<br />

The as-grown as well as annealed at the optimal conditions samples (T a ≈ 289 0 C)<br />

with x=0.027 (10727C), x=0.062 (10727D), <strong>and</strong> x=0.086 (10727E), were studied.<br />

The measurements were performed using a Philips X’Pert-MRD diffractometer<br />

equipped with a parabolic X-ray mirror <strong>and</strong> four-bounce Ge 220 monochromator at the<br />

incident beam, <strong>and</strong> a three-bounce Ge analyzer at the diffracted beam.<br />

Figure 5.41 shows the ω/2Θ scans obtained for the symmetric (004) reflection for the<br />

as-grown samples with x=0.027, x=0.062 <strong>and</strong> x=0.083. Figure 5.42 shows ω/2Θ scan for<br />

the sample with x=0.083 before <strong>and</strong> after annealing.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 91<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ω <br />

ω<br />

Fig. 5.41: The ω/2Θ scan for the symmetric (004) Bragg reflection for as-grown samples with<br />

x=0.027, 0.062, <strong>and</strong> 0.086.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

!<br />

<br />

<br />

ω <br />

ω<br />

<br />

Fig. 5.42: The ω/2Θ scan for the symmetric (004) Bragg reflection for as-grown <strong>and</strong> annealed<br />

sample with x=0.086.


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 92<br />

Tab. 5.2: The measured values <strong>of</strong> perpendicular to the layer plane lattice parameter(a ⊥ ), in-plane<br />

lattice parameter (a ‖ ), the calculated values <strong>of</strong> the relaxed mismatch (a relax -a s )/a s , <strong>and</strong><br />

thickness <strong>of</strong> the GaMnAs epilayers determined from both RHEED oscilations <strong>and</strong> XRD<br />

measurements before <strong>and</strong> after annealing.<br />

sample a ‖ a ⊥ a relaxed ∆a/a d [nm] d [nm]<br />

x [Å] [Å] [Å] [ppm] XRD RHEED<br />

0.027 5.65348 5.66941 5.661243 1373 122 131<br />

as-grown<br />

0.027 5.65348 5.66829 5.660697 1277 123 -<br />

annealed<br />

0.062 5.65348 5.68431 5.668505 2658 140 149<br />

as-grown<br />

0.062 5.65348 5.68049 5.666643 2328 136 -<br />

annealed<br />

0.083 5.65348 5.6953 5.67386 3605 98 105<br />

as-grown<br />

0.083 5.65348 5.68783 5.67022 2961 101 -<br />

annealed<br />

For both as-grown <strong>and</strong> annealed samples the pr<strong>of</strong>iles <strong>of</strong> ω/2Θ scan exhibit clear interference<br />

fringes, attesting to the high structural perfection <strong>of</strong> the layers. It is clearly visible<br />

that annealing procedure performed at the optimal conditions does not change the structural<br />

quality <strong>of</strong> the investigated epilayers. These oscillations also provide a very direct<br />

method for determining the Ga 1−x Mn x As layer thickness. Thickness values estimated<br />

from the thickness fringes in the ω/2Θ curves are collected in Table 5.2. The values <strong>of</strong><br />

the layer thickness obtained by this method agree rather well with those obtained from<br />

RHEED oscillations, thus providing an added measure <strong>of</strong> internal consistency <strong>of</strong> the performed<br />

experiments.<br />

The good agreement <strong>of</strong> the experimental values <strong>of</strong> full width at half maxima (FWHM)<br />

observed for the (004) Bragg reflections (from 170 to 180 arcsec) with the FWHM values<br />

obtained from the simulated curves (∼ 187 arcsec) shown in Figure 5.43 also indicate<br />

that the crystalline quality <strong>of</strong> the Ga 1−x Mn x As layers is rather high. Finally, the lack <strong>of</strong><br />

asymmetry in the pr<strong>of</strong>iles indicates the absence <strong>of</strong> detectable strain gradients within the<br />

entire thickness <strong>of</strong> the Ga 1−x Mn x As film.<br />

The reciprocal lattice maps measured for the (004) <strong>and</strong> (224) reflections revealed high<br />

crystalline perfection <strong>of</strong> the as-grown Ga 1−x Mn x As epilayers <strong>and</strong> showed that annealing<br />

does not alter the crystalline quality <strong>of</strong> the investigated layers. Figures 5.44 <strong>and</strong> 5.45<br />

present the reciprocal lattice maps obtained for the (004) <strong>and</strong> (224) reflections, respectively.<br />

The very narrow in Q x direction peaks corresponding to the Ga 1−x Mn x As layers (unchanged<br />

by the annealing process), along with the presence <strong>of</strong> the interference peaks due<br />

to multiple reflections within the Ga 1−x Mn x As layer, indicate a sharp interface between<br />

the Ga 1−x Mn x As layer <strong>and</strong> the GaAs substrate. This, together with the relative sharpness<br />

<strong>of</strong> the Ga 1−x Mn x As peak, suggests that the Mn content is uniform (negligible gradient


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 93<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ω <br />

ω<br />

Fig. 5.43: Comparison <strong>of</strong> the measured ω/2Θ scan for the symmetric (004) Bragg reflection<br />

(dashed curve) with simulation results (solid curve) for the annealed sample with x=<br />

0.086.<br />

in x) throughout the epilayer. The fact that the value <strong>of</strong> Q x in the asymmetric reciprocal<br />

lattice maps (see Figure 5.45) is the same for the Ga 1−x Mn x As layer <strong>and</strong> for the GaAs<br />

substrate also reveal that the Ga 1−x Mn x As films (as-grown as well as annealed) are fully<br />

strained to the (100) GaAs substrate (i.e., fully pseudomorphic), with no detectable relaxation<br />

throughout the thickness <strong>of</strong> the film. The cross shown in Figure 5.45 indicates<br />

the position where the peak from the hypothetical relaxed Ga 1−x Mn x As would occur on<br />

the (224) reciprocal map, thus serving to illustrate the degree <strong>of</strong> tetragonal distortion uniformly<br />

experienced by the Ga 1−x Mn x As alloy along the growth direction.<br />

Measurements <strong>of</strong> the (004) Bragg reflections allowed to calculate the lattice parameters<br />

perpendicular to the layer plane (a ⊥ ) for all samples studied. The combination <strong>of</strong> this<br />

<strong>and</strong> the measurements <strong>of</strong> the asymmetric (224) Bragg reflections were then used to determine<br />

the in-plane lattice parameter (a ‖ ). The values <strong>of</strong> relaxed layer lattice parameter<br />

a relax (calculated from the measured values <strong>of</strong> a ⊥ <strong>and</strong> a ‖ ), <strong>and</strong> the calculated values <strong>of</strong><br />

the relaxed mismatch (a relax − a s )/a s before <strong>and</strong> after annealing (where a s is the lattice<br />

parameter <strong>of</strong> the GaAs substrate), are shown in Table 5.2.<br />

The lattice parameters a relax for relaxed Ga 1−x Mn x As were obtained using the relation<br />

a relax = a ⊥ + 2ba ‖<br />

(5.11)<br />

1 + 2b<br />

where b=C 11 /(C 11 +2C 12 ). Here C 11 <strong>and</strong> C 12 are elastic constants for GaAs<br />

(C 11 =11.82·10 10 Pa, C 12 = 5.326 10 10 Pa [97]).<br />

As seen from the reciprocal maps for the (224) reflection shown in Figure 5.45,<br />

a parallel is essentially identical to the lattice parameter <strong>of</strong> GaAs. Analogous results were


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 94<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.44: Reciprocal space maps <strong>of</strong> the symmetric (004) reflection for the as-grown <strong>and</strong> annealed<br />

sample with x=0.086. (Q x <strong>and</strong> Q y represent reciprocal space vectors (Q x is in the direction<br />

parallel to the surface, Q y is in the direction perpendicular to the surface),<br />

both given in λ/2d units, λ=0.15406 nm, d denotes the interplanar spacing).


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 95<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.45: Reciprocal space maps <strong>of</strong> the asymmetric (224) reflection for the as-grown <strong>and</strong> annealed<br />

sample with x=0.086. (Q x <strong>and</strong> Q y represent reciprocal space vectors (Q x is in the direction<br />

parallel to the surface, Q y is in the direction perpendicular to the surface),<br />

both given in λ/2d units, λ=0.15406 nm, d denotes the interplanar spacing).


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 96<br />

obtained for the samples with the lower Mn content, x=0.062 <strong>and</strong> x=0.027. As in the<br />

case <strong>of</strong> x=0.086, the ω/2Θ scans, the reciprocal lattice maps for the symmetric (004)<br />

<strong>and</strong> asymmetric (224) reflections indicate high crystalline perfection <strong>of</strong> the samples, <strong>and</strong><br />

systematically reveal a distinct decrease <strong>of</strong> the perpendicular lattice parameter after annealing.<br />

The primary result <strong>of</strong> the high resolution X-ray diffraction measurements is the observation<br />

that the Ga 1−x Mn x As lattice parameter decreases when the epilayers are annealed<br />

at the optimal conditions i.e. the interstitial Mn atoms are removed from the alloy [13],<br />

[7]. As has been noted, such a result has been foreseen by Mašek et al. [44].<br />

As was noted in the previous section <strong>of</strong> this Chapter the channeling studies: c-RBS<br />

<strong>and</strong> c-PIXE revealed that optimal annealing leads to the significant reduction <strong>of</strong> Mn interstitial<br />

Mn atoms (to 7% <strong>of</strong> the total Mn content for Ga 1−x Mn x As with Mn concentration<br />

x=0.07 as determined from RHEED oscillations) [13]. The c-PIXE measurements distinguish<br />

between contributions from Mn in substitutional, interstitial <strong>and</strong> r<strong>and</strong>om positions<br />

(i.e., those in the form <strong>of</strong> r<strong>and</strong>om precipitates, such as MnAs inclusions). The total Mn<br />

concentration determined from c-PIXE investigations is always higher by 15 to 20 % than<br />

the values obtained from RHEED in as-grown samples. As was noted in Chapter 3 it<br />

can be assumed that RHEED oscillations provide a measure <strong>of</strong> substitutional Mn cations<br />

Mn Ga . The PIXE measurements revealed that sample with x determined by RHEED as<br />

0.07 has total <strong>of</strong> Mn concentration equal to 0.092, <strong>of</strong> which 0.072 was in substitutional<br />

positions, 0.013 in the form <strong>of</strong> interstitials, <strong>and</strong> 0.007 occurred as r<strong>and</strong>om precipitates. If<br />

one will use the above results on the sample with x=0.086 (as determined from RHEED<br />

oscillations) <strong>and</strong> will ascribe x=0.086 to substitutional Mn, then the total concentration<br />

<strong>of</strong> Mn in this sample can be estimated as x tot ∼0.109, with the atomic fraction <strong>of</strong> Mn<br />

interstitials (x int ) estimated as 0.015 in as-grown material <strong>and</strong> 0.008 after the sample was<br />

annealed, for a change in interstitial concentration estimated at ∆x int =0.007. The calculation<br />

<strong>of</strong> Mašek et al. [44] indicate that the relaxed lattice parameter <strong>of</strong> Ga 1−x Mn x As has<br />

the following form (in Å):<br />

a = a 0 + 0.02x sub + 1.05x int + 0.69y (5.12)<br />

where a 0 is the lattice parameter <strong>of</strong> GaAs, x sub <strong>and</strong> x int are the concentrations <strong>of</strong> substitutional<br />

<strong>and</strong> interstital Mn, <strong>and</strong> y is the concentration <strong>of</strong> As antisites As Ga .<br />

It is safe to assume that after low-temperature annealing x sub <strong>and</strong> y remain unchanged<br />

(the temperature T


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 97<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5.46: Relaxed lattice parameter <strong>of</strong> Ga 1−x Mn x As plotted as a function <strong>of</strong> the Mn concentration<br />

x for as-grown <strong>and</strong> annealed samples.<br />

are two interesting features which emerge from Figure 5.46. First, when one extrapolates<br />

the as-grown <strong>and</strong> annealed points to x=1.0, one obtains, respectively, the values <strong>of</strong> 5.90<br />

Å<strong>and</strong> 5.86 Åfor the hypothetical zinc blende MnAs. It is interesting that a theoretical<br />

calculation <strong>of</strong> the a 0 for hypothetical zinc blende MnAs using covalent radii r c (for As<br />

r c =1.225 Å; for Mn r c = 1.326 Å) (see e.g. [98]), one obtains a value <strong>of</strong> 5.89Å. While<br />

this agreement may be to some coincidental, this is probably the reason why the increase<br />

<strong>of</strong> a 0 with x, along with the use <strong>of</strong> Vegard’s law, has been rather widely accepted for<br />

Ga 1−x Mn x As, <strong>and</strong> has been initially interpreted as the effect <strong>of</strong> substitutional Mn in this<br />

alloy.<br />

Figure 5.46 also reveals another interesting feature. Note that the decrease <strong>of</strong> a 0 after<br />

annealing appears to be proportional to x, suggesting that the annealing-induced drop in<br />

the concentration <strong>of</strong> Mn I increases with the Mn concentration. Since it is known from<br />

the c-PIXE results that annealing reduces the Mn I concentration by roughly a factor <strong>of</strong><br />

two, this would suggest that the difference between the lattice parameter <strong>of</strong> as-grown<br />

material <strong>and</strong> <strong>of</strong> the hypothetical material in which there are no Mn interstitials would<br />

be approximately twice as large as that seen in Figure 5.46 between the as-grown <strong>and</strong><br />

annealed cases. Extending this logic to x=1.0, we obtain an estimate <strong>of</strong> 5.83Åfor the<br />

lattice parameter <strong>of</strong> zinc blende MnAs with only substitutional Mn. This is <strong>of</strong> course in<br />

disagreement with Reference [44], in which it is predicted that the effect <strong>of</strong> substitutional<br />

Mn on the lattice constant <strong>of</strong> Ga 1−x Mn x As is practically negligible (in stark contradiction<br />

to the estimated value obtained by using covalent radii). It is possible that this points to<br />

very physical insights: if one assumes zinc blende Ga 1−x Mn x As with only substitutional<br />

Mn, one must take into account that every Mn produces an uncompensated hole, <strong>and</strong> for<br />

MnAs (or even Ga 1−x Mn x As with a large value <strong>of</strong> x) one automatically has a metal. It is<br />

likely that the Coulomb interaction between the hole gas <strong>and</strong> the positive ions experience<br />

a Coulomb attraction, exactly as in a metallic bond, which causes the lattice parameter<br />

to shrink. One would assume that this effect is present in the first-principles calculations<br />

discussed in Reference [44] (although it may be over-estimated in the calculations). The<br />

tendency for a 0 to decrease further with decreasing concentration <strong>of</strong> Mn I signaled by


5. Low temperature annealing studies <strong>of</strong> Ga 1−x Mn x As 98<br />

the results plotted in Figure 5.46 may thus be a qualitative indication <strong>of</strong> the physical<br />

processes implicitly taken into account in the first-principles calculation <strong>of</strong> Mašek et al.<br />

[44].<br />

The presented data indicate that the samples are uniformly strained. The degree <strong>of</strong><br />

strain, defined as (a relax − a s )/a s (where a s is the lattice parameter <strong>of</strong> the underlying<br />

substrate) is also listed in Table 5.2. As expected, the degree <strong>of</strong> strain in the specimens<br />

increases with the Mn content; but the strain is reduced by the annealing process.


6. CONCLUSIONS AND SUMMARY<br />

In this thesis the results <strong>of</strong> magnetic <strong>and</strong> <strong>transport</strong> study <strong>of</strong> multinary alloys:<br />

Pb 1−x−y−z Mn x Eu y Sn z Te <strong>and</strong> Ga 1−x Mn x As are reported. For lead chalcogenides, the<br />

presence <strong>of</strong> two types <strong>of</strong> magnetic ions (Mn <strong>and</strong> Eu) on magnetic <strong>properties</strong> <strong>of</strong> the resultant<br />

semimagnetic <strong>semiconductor</strong> was studied. The effect <strong>of</strong> the low temperature<br />

annealing on the <strong>transport</strong>, magnetic, magnetooptical <strong>and</strong> structural <strong>properties</strong> <strong>of</strong> the<br />

Ga 1−x Mn x As was investigated.<br />

6.1 Pb 1−x−y−z Mn x Eu y Sn z Te<br />

The <strong>transport</strong> measurements revealed p-type <strong>of</strong> conductivity for all investigated samples<br />

<strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te. The samples are characterized with high almost temperature<br />

independent hole concentration (in the range between 2·10 18 cm −3 <strong>and</strong> 2·10 21 cm −3 ). The<br />

paramagnet/ferromagnet as well ferromagnet/spin glass phase transitions were observed.<br />

The following most important results were obtained for Pb 1−x−y−z Mn x Eu y Sn z Te mixed<br />

crystals:<br />

1. The presence <strong>of</strong> two types <strong>of</strong> magnetic ions (Mn <strong>and</strong> Eu) in IV-VI <strong>semiconductor</strong><br />

matrix influences magnetic <strong>properties</strong> <strong>of</strong> the resultant semimagnetic <strong>semiconductor</strong>.<br />

The results <strong>of</strong> magnetic measurements show that Curie temperature T C as<br />

well as Curie - Weiss temperature Θ decrease with the increase <strong>of</strong> Eu content in<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples. The magnetic susceptibility measurements revealed<br />

also that Eu changes spin glass dynamics in this material. The difference<br />

in the rate <strong>of</strong> frequency shift <strong>of</strong> cusp in real <strong>and</strong> imaginary part <strong>of</strong> susceptibility<br />

is visible. Such behaviour was not observed for Mn-based IV-VI semimagnetic<br />

<strong>semiconductor</strong>s. <strong>Magnetic</strong> phase diagram <strong>of</strong> investigated Pb 1−x−y−z Mn x Eu y Sn z Te<br />

shows that the presence <strong>of</strong> Eu shifts glass regime towards lower carrier concentration<br />

p (as compared to Pb 1−x−y Mn x Sn y Te system).<br />

2. The qualitative analysis showed that a variation <strong>of</strong> the b<strong>and</strong> parameters with the<br />

alloy composition (i.e. shift <strong>of</strong> ɛ 0 as well as E g parameter with Eu concentration y<br />

- see Chapter 4 ) is responsible for the observed strong dependence <strong>of</strong> Curie temperature<br />

on Eu content. Introduced simple two b<strong>and</strong> model explains both the order<br />

<strong>of</strong> the transition temperature values as well as T C dependence on Eu concentration.<br />

The calculated dependence <strong>of</strong> Curie temperature on Eu content very well reflects<br />

experimentally confirmed effect <strong>of</strong> T C decrease with Eu concentration y.<br />

6.2 Ga 1−x Mn x As<br />

The annealing studies on GaMnAs in the wide range <strong>of</strong> Mn concentration (up to high<br />

values x∼0.09) were performed. The systematic <strong>transport</strong>, magneto<strong>transport</strong>, magnetic,


6. Conclusions <strong>and</strong> Summary 100<br />

magnetooptical, structural <strong>and</strong> channeling measurements were carried out. Significant<br />

annealing-induced changes in the magnetic, electronic as well as structural <strong>properties</strong><br />

<strong>of</strong> this semimagnetic material, that depend on the Mn concentration <strong>and</strong> on annealing<br />

conditions were observed. In summary:<br />

1. The optimal conditions (temperature, time, flow <strong>of</strong> N 2 gas <strong>of</strong> post-growth annealing<br />

were established for studied samples. A method <strong>of</strong> increasing T C , was developed.<br />

For the first time it is shown that by a proper choice <strong>of</strong> annealing conditions the limit<br />

<strong>of</strong> T C in GaMnAs (∼110K) [6] can be shifted to much higher values (up to 127K for<br />

x∼0.08). Simultaneously, an increase in saturation magnetization, conductivity <strong>and</strong><br />

free hole concentration is observed for samples annealed at the optimal conditions.<br />

2. The most important experimental fact is that low temperature annealing performed<br />

at the optimal conditions leads to the significant reduction <strong>of</strong> Mn interstitial atoms<br />

(7% <strong>of</strong> the total Mn content for GaMnAs epilayer with x=0.092). LT post-growth<br />

annealing leads to a rearrangement <strong>of</strong> the Mn ions in the host lattice, in particular<br />

to the removal <strong>of</strong> Mn from the interstitial sites. The presented channeling measurements<br />

showed that in the process <strong>of</strong> LT annealing the Mn I ions are moved<br />

to r<strong>and</strong>om, incommensurate with GaAs lattice positions (e.g. MnAs clusters), in<br />

which the Mn ions are electrically inactive, what increases the hole concentration,<br />

conductivity <strong>and</strong> T C . According to the ab initio calculations [93], this is due to<br />

the out diffusion <strong>of</strong> Mn intersitials towards the surface - the biding energy for Mn I<br />

in GaMnAs allows for significant diffusion <strong>of</strong> this effect at temperatures above<br />

∼150 0 C. Recently, it was presented [99], [95], [100] that spins in the formed Mn I -<br />

Mn Ga pairs are expected to be anti<strong>ferromagnetic</strong>ally aligned, thus cancelling the<br />

magnetic contribution <strong>of</strong> Mn Ga to the magnetization <strong>of</strong> the GaMnAs system as a<br />

whole. Removal <strong>of</strong> Mn I from such a pair should thus automatically render the substitutional<br />

Mn ++ magnetically-active, increasing the saturation magnetization, as is<br />

indeed observed experimentally <strong>and</strong> shown in the thesis. The magnetic <strong>properties</strong><br />

<strong>of</strong> the Mn ion in interstitial sites, i.e. the negligible kinetic exchange constant <strong>and</strong><br />

strong anti<strong>ferromagnetic</strong> superexchange with the adjacent substitutional Mn ion,<br />

act towards diminishing the transition temperature. Mn interstitial act against the<br />

ferromagnetism in Ga 1−x Mn x As, not only due to the compensating character <strong>of</strong> this<br />

defect, but also because <strong>of</strong> their distinct magnetic behaviour.<br />

Here, it should be mentioned that fact <strong>of</strong> inequality between the hole concentration<br />

<strong>and</strong> the Mn content (hole concentration is substantially lower than the Mn content<br />

in all Ga 1−x Mn x As samples) has been ascribed for a long time to the presence <strong>of</strong><br />

compensating donors, in particular to the formation <strong>of</strong> arsenic antisites (As Ga ) during<br />

the eptaxial growth <strong>of</strong> GaMnAs at As overpressure. The LT annealing induced<br />

changes <strong>of</strong> the hole concentration <strong>and</strong>, therefore, Curie temperature were attributed<br />

to the decrease <strong>of</strong> the concentration <strong>of</strong> arsenic antisites. These antisites, however,<br />

are relatively stable defects - it was shown that to remove As Ga from LT MBE<br />

grown GaAs annealing temperatures above 450 0 C are needed [101].<br />

3. It is shown that the lattice parameter decreases when the epilayers <strong>of</strong> GaMnAs<br />

are annealed at the optimal conditions, i.e. when the interstitial Mn atoms are removed<br />

from the alloy. The results <strong>of</strong> high resolution X-ray diffraction (HRXRD)


6. Conclusions <strong>and</strong> Summary 101<br />

are agreement with the theoretical predictions [44]. Using the values <strong>of</strong> lattice constant<br />

before <strong>and</strong> after annealing obtained from HRXRD measurements a change <strong>of</strong><br />

the interstitial concentration was estimated. The obtained values are smaller but<br />

<strong>of</strong> the same order <strong>of</strong> magnitude as the change in concentration <strong>of</strong> intersitial Mn<br />

estimated from RBS/PIXE measurements.<br />

4. <strong>Magnetic</strong> investigations (SQUID measurements) revealed a decrease <strong>of</strong> the coercive<br />

field after heat treatment at the optimal conditions. Thermal annealing at a<br />

higher temperature leads to an increase <strong>of</strong> the coercive field. This can be related to<br />

the domain structure in the investigated material. It is easy to imagine that proper<br />

annealing diminishes the coercive field. The MOKE measurements showed the<br />

non monotonic low magnetic field behaviour. The observed features are connected<br />

with the magnetic <strong>properties</strong> <strong>of</strong> studied epilayers. At present very speculative interpretation<br />

is possible. Such behaviour can be related with the magnetic anisotropy<br />

obsrerved in GaMnAs epilayers. The oscillatory terms in the range <strong>of</strong> high magnetic<br />

fields are visible in the range <strong>of</strong> high magnetic fields (up to 25T). The shift<br />

towards higher magnetic fields is pronounced after thermal annealing at the optimal<br />

conditions. The induced by annealing changes in the strain relations modify<br />

the b<strong>and</strong> structure <strong>of</strong> GaMnAs <strong>and</strong> this in turn may be responsible for the observed<br />

shift <strong>of</strong> oscillation terms.<br />

5. All investigated samples indicated unsaturated negative magnetoresistance (MR)<br />

up to the highest value <strong>of</strong> investigated field (55T). A sizable negative magnetoresistance<br />

in the regime <strong>of</strong> strong magnetic fields can be assigned to the weak localization<br />

effect [51]. It was found that annealing <strong>of</strong> the Ga 1−x Mn x As epilayers<br />

with high Mn concentration leads to very significant changes in magnetoresistivity.<br />

Particulary, a pronounced decrease <strong>of</strong> magnetoresistivity MR after annealing at<br />

the optimal conditions <strong>and</strong> a substantial increase <strong>of</strong> MR after annealing at a higher<br />

temperature was observed. For samples with lower Mn content (x∼0.03) the MR is<br />

not affected by the heat treatment at the optimal conditions. It was also shown that<br />

for low Mn concentration the influence <strong>of</strong> annealing procedure on both the Curie<br />

temperature <strong>and</strong> conductivity is weak. The annealing induced large changes <strong>of</strong> the<br />

magnetoresistivity can be related with the change <strong>of</strong> the domain structure <strong>of</strong> GaMnAs<br />

system. In the range <strong>of</strong> low magnetic fields the hysteretic behaviour <strong>of</strong> MR<br />

is observed. The low field hysteretic feature is much more pronounced with the<br />

temperature decrease.<br />

6.3 Suggestions for further studies<br />

1. <strong>Magnetic</strong> studies <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals were performed for<br />

samples with relatively low Eu concentration. More distinct effect <strong>of</strong> the influence<br />

<strong>of</strong> the second type <strong>of</strong> magnetic ion on the magnetic <strong>properties</strong> <strong>of</strong> the resultant material<br />

should be visible for higher Eu concentration. The ideal samples for these kind<br />

<strong>of</strong> investigation should be characterized by the same content <strong>of</strong> Mn <strong>and</strong> high scatter<br />

<strong>of</strong> Eu content.<br />

The measurements <strong>of</strong> magnetization under hydrostatic pressure were performed<br />

only for Sn 1−x−y Mn x Eu y Te samples. For samples <strong>of</strong> PbSnMnEuTe the very pro-


6. Conclusions <strong>and</strong> Summary 102<br />

nounced effect <strong>of</strong> the decrease <strong>of</strong> T C with the Eu concentration related to the variation<br />

<strong>of</strong> the b<strong>and</strong> parameters with the alloy composition was visible. One can expect<br />

that for these samples the change in the magnetization measured under hydrostatic<br />

pressure with the Eu content.<br />

2. The measurements <strong>of</strong> MOKE were not performed as a function <strong>of</strong> the wavelength.<br />

This kind <strong>of</strong> investigations could be helpful to explore the origin <strong>of</strong> oscillating terms<br />

in the high magnetic field range <strong>of</strong> MOKE curves as well as low magnetic field<br />

behaviour <strong>of</strong> MOKE.<br />

3. Magnetization measurements in the perpendicular configuration (magnetic field directed<br />

perpendicular to the sample surface) would be very helpful in the further<br />

studies <strong>of</strong> the magnetic <strong>properties</strong> <strong>of</strong> GaMnAs epilayers. Particularly, SQUID measurements<br />

in this configuration in the range <strong>of</strong> low magnetic fields could throw the<br />

light on the low magnetic field behaviour <strong>of</strong> magnetooptical Kerr effect. The performed<br />

in the thesis measurements revealed that usual relation θ K ∝ M is not valid<br />

in GaMnAs. Further studies are necessary to explore the relation between SQUID<br />

magnetization M <strong>and</strong> Kerr rotation angle θ K .


7. APPENDICES<br />

7.1 Appendix 1<br />

The Jones vector [102] provides a concise representation <strong>of</strong> the electric vibration <strong>of</strong> a<br />

transverse-electric (TE) plane wave. The two-component complex Jones vector carriers<br />

information about the amplitude A, absolute phase δ, azimuth θ <strong>and</strong> ellipticity angle ɛ <strong>of</strong><br />

the elliptic vibration <strong>of</strong> the electric-field vector <strong>of</strong> a uniform monochromatic TE plane<br />

wave <strong>of</strong> light (look in Figure 7.1).<br />

Let’s assume that uniform, TE travelling plane wave <strong>of</strong> arbitrary polarization propagates<br />

along the positive direction <strong>of</strong> the z axis <strong>of</strong> an xyz orthogonal, right-h<strong>and</strong>ed, Certesian<br />

coordinate system, then the electric vector <strong>of</strong> such a wave becomes:<br />

−→ E (z, t) = Ẽ x cos[ωt − 2π λ z + δ x)] −→ e x + Ẽy cos[ωt − 2π λ z + δ y)] −→ e y (7.1)<br />

,<br />

where Ẽx <strong>and</strong> Ẽy represent the amplitudes <strong>of</strong> the linear, simple-harmonic oscillations<br />

<strong>of</strong> the electric-field components along x <strong>and</strong> y axes; δ x <strong>and</strong> δ y represent the respective<br />

phases <strong>of</strong> these oscillations, −→ e x <strong>and</strong> −→ e y are unit vectors in the positive directions <strong>of</strong> the x<br />

<strong>and</strong> y axes.<br />

The Jones vector −→ E (0) (in the form <strong>of</strong> 2×1 column vector (matrix)) <strong>of</strong> the wave<br />

(Equation 7.1) has the following form:<br />

)<br />

−→ E (0) =<br />

(Ẽ xe iδ x<br />

(7.2)<br />

Ẽ ye iδ y<br />

The vector −−→ E(0) is the concise representation <strong>of</strong> a single plane wave which is known to be<br />

monochromatic, uniform <strong>and</strong> transverse-electric <strong>and</strong> contains complete information about<br />

the amplitudes <strong>and</strong> phases <strong>of</strong> the field components, so about the polarization <strong>of</strong> the wave.<br />

From the Jones vector the time <strong>and</strong> space dependence <strong>of</strong> the entire wave can be obtained<br />

by using the following equation:<br />

−→ E (z, t) = Re[<br />

−→ E (0)e<br />

i(ωt− 2πz<br />

λ ) ] (7.3)<br />

The full explicit expression <strong>of</strong> the wave (Equation 7.1) can be obtained by reintroducing<br />

the units vectors −→ e x <strong>and</strong> −→ e y . In simplified notation the Jones vector has the following<br />

form:<br />

(<br />

−→ E =<br />

E x<br />

(7.4)<br />

E y<br />

)<br />

where E x =Ẽxe iδ x<br />

<strong>and</strong> E y =Ẽye iδ y<br />

.<br />

For an elliptical vibration which amplitude A, phase δ, azimuth θ <strong>and</strong> ellipticity angle<br />

ɛ are given a Jones vector can be constructed.


7. Appendices 104<br />

<br />

δ<br />

θ<br />

ε<br />

<br />

<br />

Fig. 7.1: The parameters that define the ellipse <strong>of</strong> polarization in its plane: 1) the azimuth θ, that<br />

defines orientation <strong>of</strong> the ellipse in its plane (− π 2 ≤ θ ≤ π 2 ); 2) the ellipticity e= b a ; 3)<br />

the ellipticity angle ɛ=arctan(e) (− π 4 ≤ ɛ ≤ π 4 ); 4) the amplitude A = (a2 +b 2 ) 1 2 ); 5)<br />

the absolute phase δ defined by the initial electric field <strong>and</strong> auxiliary (dashed) circle (-<br />

π ≤ δ ≤ π), it determines the angle between the initial position <strong>of</strong> the electric vector at<br />

t=0 <strong>and</strong> major axis <strong>of</strong> the ellipse.<br />

The Cartesian Jones vector <strong>of</strong> a given elliptical polarization state has the following<br />

form: ( ) (<br />

)<br />

E x<br />

= Ae iδ cos(θ) cos(ɛ)−i sin(θ)sin(ɛ)<br />

(7.5)<br />

E y<br />

sin(θ) cos(ɛ)+i cos(θ) sin(ɛ)<br />

The circular Jones vector <strong>of</strong> a given elliptical polarized state has the following form:<br />

( ) (<br />

)<br />

E l<br />

√<br />

= Ae iδ<br />

2 cos(ɛ)−sin(ɛ)e iθ<br />

(7.6)<br />

E r cos(ɛ)+sin(ɛ)e −iθ<br />

When a polarized light passes through optical elements, its polarization state is in general<br />

modified. This can be described in the framework <strong>of</strong> the Jones formalism. Assuming,<br />

that the polarization response <strong>of</strong> optical elements such as the sample, mirror, polarizer,<br />

etc. is linear it can be expressed by a matrix product. Let the incident <strong>and</strong> the outgoing<br />

plane waves be described by their appropriate Jones vectors E −→ i <strong>and</strong> −→ E o , respectively.<br />

The following equation express the law <strong>of</strong> interaction between the incident wave <strong>and</strong> the<br />

optical system as a simple linear matrix transformation <strong>of</strong> the Jones vector <strong>of</strong> the wave:<br />

−→<br />

E o =<br />

(<br />

T 11 T 12<br />

T 21 T 22<br />

) −→Ei<br />

(7.7)


7. Appendices 105<br />

B=0<br />

E g<br />

k=0<br />

Fig. 7.2: 1 Energy b<strong>and</strong>s for a simple <strong>semiconductor</strong> for B=0.<br />

7.2 Appendix2<br />

Let’s consider a simple case <strong>of</strong> two parabolic energy b<strong>and</strong>s in the effective mass approximation<br />

(look in Figure 7.2, where the upper curve represents the energy wave vector<br />

relation <strong>of</strong> the conduction electrons with effective mass m c <strong>and</strong> the lower curve, the valence<br />

electrons, or holes, with effective mass m v ).<br />

The simplification follows from the effective mass approximation. The Schrödinger<br />

equation for electrons moving through the periodic potential V(r) <strong>of</strong> the lattice, i.e.,<br />

[ −→ p 2 /2m + V ( −→ r )]Ψ = EΨ (7.8)<br />

is replaced by the simpler equation:<br />

−→ p 2 Ψ/2m c = E c Ψ (7.9)<br />

where −→ p is the momentum operator. The solution <strong>of</strong> the above solution is:<br />

E c = ¯h 2−→ k 2 /2m c (7.10)<br />

For the valence b<strong>and</strong>, which is separated from the conduction b<strong>and</strong> by the energy gap, E g ,<br />

one obtains:<br />

E v = −E g − ¯h 2−→ k 2 /2m v (7.11)<br />

In a magnetic field, the Schrödinger equation, neglecting spin, for Bloch electrons is:<br />

[ 1<br />

2m (−→ p + e −→ A ) 2 + V ( −→ r )]Ψ j ( −→ r ) = E j Ψ j ( −→ r ) (7.12)<br />

where −→ A is the vector potential for the magnetic field. The solution here is:<br />

Ψ j ( −→ r ) = Φ j ( −→ r )F j ( −→ r ) + 1 ∑ p α ij<br />

m (E j − E i ) (p α + e c A α)F j ( −→ r )Φ i ( −→ r ) (7.13)<br />

i≠j


7. Appendices 106<br />

where Φ j ( −→ r )is the Bloch function u jk exp[i( −→ k ·−→ r ) at the bottom <strong>of</strong> the b<strong>and</strong>, i.e. u j0 ; the<br />

summation is over all the other b<strong>and</strong>s i, p α ij is the α component <strong>of</strong> the momentum matrix<br />

element, i.e.,<br />

∫<br />

p α ij ≡ u ∗ i0( −→ r )p α u j0 ( −→ r )dτ (7.14)<br />

<strong>and</strong> F( −→ r ) is obtained from the solution <strong>of</strong> the equation:<br />

1<br />

2m (−→ p + e −→ A ) 2 F ∗ j ( −→ r ) = E j F j ( −→ r ) (7.15)<br />

c<br />

Assuming the magnetic field B in the z-direction <strong>and</strong> −→ A = (0, Bx, 0), the latter equation<br />

becomes:<br />

1<br />

2m [−→ p 2 + 2 eB ∗ c xp y + e2 B 2<br />

x 2 ]F<br />

c 2 j ( −→ r ) = E j F j ( −→ r ) (7.16)<br />

Choosing a solution <strong>of</strong> the form:<br />

results in the simplified equation:<br />

F j ( −→ r ) = g(x)exp[i(k y y + k z z)] (7.17)<br />

− ¯h2 d 2 g<br />

2m ast dx + [ 1<br />

2 2m (¯hk ∗ y + eB c x)2 + ¯h2 kz<br />

2 ]g = Eg (7.18)<br />

2m∗ that represents the motion <strong>of</strong> a one-dimensional simple harmonic oscillator. The eigenvalues<br />

<strong>of</strong> above equation are:<br />

where n=0, 1, 2, 3..., <strong>and</strong><br />

E = (n + 1 2 )¯hω c + ¯h 2 k 2 z/2m ∗ , (7.19)<br />

g(x) = φ n (x − c¯hk y /eB) (7.20)<br />

the one dimensional simple harmonic oscillator wave function. Thus, to the first order:<br />

Ψ j ( −→ r ) = u j0 F j ( −→ r ) =<br />

u<br />

√ j0<br />

(Ly L z ) exp[i(k yy + k z z)]φ n (x − c¯h<br />

eB k y) (7.21)<br />

where L x , L y , L z are the crystal dimensions.<br />

In a magnetic field the energy levels <strong>of</strong> the electrons <strong>and</strong> holes can be represented as:<br />

E c = (n + 1 2 )¯hω c + ¯h 2 k 2 z/2m c (7.22)<br />

<strong>and</strong><br />

E v = −E g − (n ′ + 1 2 )¯hω v − ¯h 2 k 2 z/2m v (7.23)<br />

respectively. These energy levels are shown in Figure 7.3.<br />

The electromagnetic wave can be represented by a time-varying field, <strong>of</strong> frequency ω<br />

<strong>and</strong> polarization −→ ɛ , which has a negligible space variation since the wavelength <strong>of</strong> the<br />

radiation is much longer than the lattice spacing. The probability <strong>of</strong> electron transitions


7. Appendices 107<br />

B>0<br />

n=2<br />

n=1<br />

n=0<br />

hω c<br />

1/2 hω c<br />

n'=0<br />

n'=1<br />

n'=2<br />

1/2 hω v<br />

hω v<br />

k=0<br />

Fig. 7.3: Energy b<strong>and</strong>s for a simple <strong>semiconductor</strong> for B>0.<br />

from an initial state, i, to a final state, f, is proportional to the square <strong>of</strong> the matrix P if<br />

which to first order equals:<br />

∫<br />

P if = 〈Ψ ∗ fH ′ Ψ i 〉 = u ∗ f0( −→ r )Ff ∗ ( −→ r ) −→ ɛ · ( −→ p + e A )ui0 (<br />

c−→ −→ r )F i ( −→ r )dτ (7.24)<br />

where H’ is the perturbing Hamiltonian (H’= −→ ɛ · ( −→ p + e/c −→ A )sin(ωt)).<br />

A simplification <strong>of</strong> above expansion is possible because the harmonic oscillator functions<br />

F i ( −→ r ), F f ( −→ r ), as well as the vector potential −→ A <strong>of</strong> the wave, vary slowly compared<br />

to the periodic b<strong>and</strong> edge functions u i0 <strong>and</strong> u i0 . Consequently the former functions can be<br />

treated as constant over a unit cell so that:<br />

∫<br />

P if = [ Ff ∗ ( −→ r ) −→ ɛ · ( −→ p + e A )Fi (<br />

c−→ −→ ∫<br />

r )dτ u ∗ f0u i0 dτ+<br />

crystal<br />

∫<br />

Ff ∗ ( −→ r )F i ( −→ ∫<br />

r )dτ<br />

cell<br />

u ∗ f0−→ ɛ · −→ p ui0 dτ] (7.25)<br />

crystal<br />

For the case <strong>of</strong> interb<strong>and</strong> transitions (transitions between two b<strong>and</strong>s), the first term <strong>of</strong><br />

Equation 7.2 goes to zero because <strong>of</strong> the orthogonality <strong>of</strong> u i0 , u f0 <strong>and</strong> the second term is<br />

usually non-vanishing.<br />

The simple harmonic oscillator functions, i.e. F c <strong>and</strong> F v do not involve <strong>properties</strong> <strong>of</strong><br />

the energy b<strong>and</strong>s. Consequently, because <strong>of</strong> orthogonality, the selections rules are:<br />

cell<br />

∆k y = ∆k x = 0 (7.26)<br />

<strong>and</strong><br />

∆n = 0 (7.27)


7. Appendices 108<br />

The energy for a transition (the spin is neglected) is:<br />

∆E = E c − E v = E g + (n + 1 2 )¯h(ω c + ω v ) + ¯h 2 k 2 z/2µ (7.28)<br />

where µ ≡ m c m v /(m c + m v ) is the reduced effective mass.<br />

When spin is included, Equation 7.19 can be written as:<br />

E = (n + 1 2 )¯hω c + Mµ B gB + ¯h 2 k 2 z/2m ∗ (7.29)<br />

where µ B is the Bohr magneton; g is the effective spectroscopic spin factor, which is equal<br />

to 2 for a free electron; M is the component <strong>of</strong> angular momentum along the magnetic<br />

field, characterized by values ± 1 , corresponding to the two possible values <strong>of</strong> electron<br />

2<br />

spin.<br />

If spin is included, the momentum matrix term ( ∫ u ∗ c0−→ ɛ ·−→ p uv0 dτ) yields an additional<br />

selection rule ∆M=0, or ±1. For propagation <strong>of</strong> the electromagnetic wave parallel to<br />

the magnetic field, then −→ E ⊥ −→ B <strong>and</strong> ∆M=±1 corresponding to the senses <strong>of</strong> circular<br />

polarization. For propagation perpendicular to −→ B , then ∆M=0 for −→ E ‖ −→ B <strong>and</strong> ∆M=±1<br />

for −→ E ⊥ −→ B . These selectional rules are determined from the Bloch functions with spin<br />

included.<br />

The energy for a transition with spin is:<br />

∆E = E c − E v = E g + (n + 1 2 )¯h(ω c + ω v ) + ¯h 2 k 2 z/2µ + (M c g c − M v g v )µ B B (7.30)<br />

By the fixing the photon energy <strong>of</strong> the incident radiation at a value somewhat above the<br />

energy gap <strong>and</strong> varying the magnetic field B, one will obtain a series <strong>of</strong> peaks in the<br />

absorption (or in MOKE curves - as was observed in the present thesis) which are periodic<br />

in 1/B:<br />

1<br />

B<br />

= F =<br />

(n + 1/2) ¯he<br />

µc ± 1 2 (g c + g v )µ B<br />

∆E − E g<br />

(7.31)<br />

Figure 7.4 shows schematically the interb<strong>and</strong> transitions for L<strong>and</strong>au levels <strong>of</strong> two<br />

simple parabolic b<strong>and</strong>s, the conditions for resonance transitions are visible.


7. Appendices 109<br />

E<br />

4<br />

3<br />

2<br />

1<br />

0<br />

E c<br />

E G<br />

E v<br />

4<br />

0<br />

3<br />

2<br />

1<br />

B<br />

Fig. 7.4: The intrerb<strong>and</strong> transitions between L<strong>and</strong>au levels (n=0,1,2,3,4) <strong>of</strong> two simple parabolic<br />

b<strong>and</strong>s ( the left side <strong>of</strong> the picture represents simple energy b<strong>and</strong>s for B=0). The arrows<br />

indicate the conditions for resonance transitions.


BIBLIOGRAPHY<br />

[1] T. Story, R. R. Galazka, R. B. Frankel, <strong>and</strong> P. A. Wolf. Phys. Rev. Lett., 56:777,<br />

1986.<br />

[2] W. J. M. de Jonge, T. Story, H. J. M. Swagten, <strong>and</strong> J. T. Eggenkamp. Europhysics<br />

Letters, 17:631, 1992.<br />

[3] H. Munekata, H. Ohno, S. von Molnar, A. Segmuller, <strong>and</strong> L.L. Chang. Phys. Rev.<br />

Lett., 65:1849, 1989.<br />

[4] H. Ohno, A. Shen, F. Matsukura, A. Oiwa, A. Endo, S. Katsumoto, <strong>and</strong> Y. Iye.<br />

Appl. Phys. Lett., 69:363, 1996.<br />

[5] A. Shen, H. Ohno, F. Matsukura, Y. Sugawara, N. Akiba, T. Kuroiwa, A. Oiwa,<br />

A. Endo, S. Katsumoto, <strong>and</strong> Y. Iye. J. Crystal Growth, 175/176:1069, 1997.<br />

[6] F. Matsukura, H. Ohno, A. Shen, <strong>and</strong> Y. Sugawara. Phys. Rev. B, 57:R2037, 1998.<br />

[7] I. Kuryliszyn, T. Wojtowicz, X. Liu, J. K. Furdyna, W. Dobrowolski, J. M. Broto,<br />

M. Goiran, O. Portugall, H. Rakoto, <strong>and</strong> B. Raquet. Acta Phys. Pol. A, 102:659,<br />

2002.<br />

[8] K. Y. Wang, K. W. Edmonds, R. P. Campion, B. L. Gallagher, N. R. S. Farley, C. T.<br />

Foxon, M. Sawicki, <strong>and</strong> T. Dietl P. Bogusł awski. J. Appl. Phys., 95, 2004 in press.<br />

[9] O. Portugall, F. Ledcouturier, J. Marquez, D. Givord, <strong>and</strong> S. Askenazy. Physica B,<br />

294-295:579, 2001.<br />

[10] T. Story. Magnetyczne przejścia fazowe w kryształach PbSnMnTe. PhD thesis,<br />

Instytut Fizyki Polskiej Akademii Nauk, Warszawa, 1988.<br />

[11] E. Dynowska. unpublished data.<br />

[12] J. Sadowski, J. Z. Domagał a R. Mathieu, P. Svedlindh, J. Ba¸k-Misiuk, K. Świa¸tek,<br />

M. Karlsteen, J. Kanski, L. Ilver, H. Asklund, <strong>and</strong> U. Sodervan. Appl. Phys. Lett.,<br />

78:3271, 2001.<br />

[13] K. M. Yu, W. Walukiewicz, T. Wojtowicz, X. Liu I. Kuryliszyn, Y. Sasaki, <strong>and</strong> J. K.<br />

Furdyna. Phys. Rev. B, 65:201303(R), 2002.<br />

[14] T. Story. Semimagnetic <strong>semiconductor</strong>s based on lead chalcogenides. In M. O.<br />

Manosreh <strong>and</strong> Dmitriv Khoklov, editors, Optoelectronic <strong>properties</strong> <strong>of</strong> Semiconductors<br />

<strong>and</strong> Superlattices. Taylor <strong>and</strong> Francis Books, 2003.


Bibliography 111<br />

[15] G. Bauer <strong>and</strong> H. Pasher. Diluted magnetic <strong>semiconductor</strong>. page 339. World Scientific,<br />

Singapore, 1991.<br />

[16] T. Story, G. Karczewski, L. Świerkowski, <strong>and</strong> R. R. Galazka. Phys. Rev. B,<br />

42:10477, 1990.<br />

[17] H. J. M. Swagten, W. J. M. De Jonge, R. R. Galazka, P. Warmenbol, <strong>and</strong> J. T.<br />

Devreese. Phys. Rev. B, 37:9907, 1988.<br />

[18] G. Karczewski, L. Świerkowski, T. Story, A. Szczerbakow, J. Niewodniczańska-<br />

Blinowska, <strong>and</strong> G. Bauer. Semicond. Sci. Technol., 5:1115, 1990.<br />

[19] G. Nimtz <strong>and</strong> B. Schlicht. Narrow gap lead salts. In G. Höhler, editor, Narrow-<br />

Gap Semiconductors, volume 98 <strong>of</strong> Springer Tracts in Modern Physics. Springer-<br />

Verlag, Berlin, 1985.<br />

[20] R. R. Galazka <strong>and</strong> J. Kossut. Semimagnetic <strong>semiconductor</strong>s. In O. Madelung,<br />

editor, Semiconductors, volume 17b <strong>of</strong> L<strong>and</strong>olt-Börnstein, New Series, Group III,<br />

chapter 5, pages 302–318. Springer-Verlag, Berlin, Heidelberg, 1982.<br />

[21] M. Ocio. Phys. Rev. B, 10:4274, 1974.<br />

[22] G. Bauer, H. Pasher, <strong>and</strong> W. Zawadzki. Semicond. Sci. Technol., 7:703, 1992.<br />

[23] M. Escorne, A. Mauger, J. L. Tholence, <strong>and</strong> R. Triboulet. Phys. Rev. B, 29:6306,<br />

1984.<br />

[24] M. A. Ruderman <strong>and</strong> C. Kittel. Phys. Rev., 96:99, 1954.<br />

[25] T. Kasuya. Prog. Theor. Phys., 16:45, 1956.<br />

[26] K. Yosida. Phys. Rev., 106:893, 1957.<br />

[27] P. J. T. Eggenkamp. Carrirer concentration dependence <strong>of</strong> the magnetic <strong>properties</strong><br />

<strong>of</strong> SnMnTe. PhD thesis, Eindhoven University <strong>of</strong> Technology, Eindhoven, 1994.<br />

[28] C. W. H. M. Vennix. <strong>Magnetic</strong> correlations in the diluted magnetic <strong>semiconductor</strong><br />

SnMnTe -an experimental study. PhD thesis, Eindhoven University <strong>of</strong> Technology,<br />

Eindhoven, 1993.<br />

[29] P. Urban <strong>and</strong> G. Sperlich. Solid State Comm., 16:927, 1975.<br />

[30] T. Story, J.T. Eggenkamp, C. H. W. Swuste, W. J. M. de Jonge H. J. M. Swagten,<br />

<strong>and</strong> A. Szczerbakow. Phys. Rev. B, 47:227, 1993 <strong>and</strong> references therein.<br />

[31] J. A. Mydosh. Spin Glasses. Taylor & Francis, London, 1993.<br />

[32] J-L. Tholence. Solid State Comm., 35:113, 1980.<br />

[33] C. A. M. Mulder, A. J. van Duyneveldt, <strong>and</strong> J. A. Mydosh. Phys. Rev. B, 23:1384,<br />

1981.<br />

[34] D. Huser, A. J. van Duyneveldt, G. J. Nieuwenhuys, <strong>and</strong> J. A. Mydosh. J. Phys. C:<br />

Solid State Phys., 19:3697, 1986.


Bibliography 112<br />

[35] J. L. Dormann, A. Saifi, V. Cagan, <strong>and</strong> M. Nogues. Phys. Stat. Sol. B, 131:573,<br />

1985.<br />

[36] M. Godinho, J-L. Tholence, A. Mauger, M. Escorne, <strong>and</strong> A. Katty. In Proceedings<br />

<strong>of</strong> the 17th International Conference on Low Temperature Physics, page 647,<br />

Amsterdam, 1984. Elsevier.<br />

[37] R. R. Galazka, J. Kossut, <strong>and</strong> T. Story. Ii-vi <strong>and</strong> i-vii compounds; semimagnetic<br />

compounds. In U. Rössler, editor, Semiconductors, L<strong>and</strong>olt-Börnstein, New Series,<br />

Group III/41. Springer-Verlag, Berlin, Heidelberg, 1999.<br />

[38] F. Matsukura, H. Ohno, <strong>and</strong> T. Dietl. In K. H. J. Buschow, editor, H<strong>and</strong>book on<br />

<strong>Magnetic</strong> Materials, volume 14. Elsevier, Amsterdam, 2002.<br />

[39] F. Matsukura, A. Shen, Y. Sugawara, Y. Ohno T. Omiya, <strong>and</strong> H. Ohno. In Proc. 25th<br />

Int. Symp. Compound Semiconductors 12-16 October 1998, Nara, Japan, Institute<br />

<strong>of</strong> Physics Conference Series, No. 162, page 547. IOP Publishing Ltd, Bristol,<br />

1999.<br />

[40] G. M. Schott, W. Faschinger, <strong>and</strong> L. W. Molenkamp. Appl. Phys. Lett., 79:1807,<br />

2001.<br />

[41] J. Sadowski, R. Mathieu, P. Svedlindh, J. Z. Domagała, K. Świa¸tek J. Ba¸k-Misiuk,<br />

M. Karlsteen, J. Kański, L. Ilver, H. Asklund, , <strong>and</strong> U. Södervall. Appl. Phys. Lett.,<br />

78:3271, 2001.<br />

[42] H. Ohno, F. Matsukura, A. Shen, Y. Sugawara, A. Oiwa, S. Katsumoto, <strong>and</strong> Y. Iye.<br />

In M. Scheffler <strong>and</strong> R. Zimmermann, editors, Proceedings <strong>of</strong> the 23rd International<br />

Conference on Physics <strong>of</strong> Semiconductors, page 405. World Scientific, 1996.<br />

[43] H. Ohno. J. Magn. Magn. Mater., 200:110, 1999.<br />

[44] J. Mašek, J. Kudrnovski, <strong>and</strong> F. Máca. Phys. Rev. B, 67:153203, 2003.<br />

[45] A. Twardowski. Materials Science & Engineering B, 63:96, 1999.<br />

[46] K. M. Yu, W. Walukiewicz, T. Wojtowicz, W. L. Lim, X. Liu, U. Bindley, M. Dobrowolska,<br />

<strong>and</strong> J.K. Furdyna. Phys. Rev. B, 68:041308(R), 2002.<br />

[47] T. Dietl, H. Ohno, <strong>and</strong> F. Matsukura. Phys. Rev. B, 63:195205, 2001.<br />

[48] T. Dietl, H. Ohno, F. Matsukura, J. Cibert, <strong>and</strong> D. Ferr<strong>and</strong>. Science, 287:1019,<br />

2000.<br />

[49] J. Konig, J. Schliemann, T. Jungwirth, <strong>and</strong> A. H. MacDonald. Ferromagnetism in<br />

(iii,mn)v <strong>semiconductor</strong>s. In D. J. Singh <strong>and</strong> D. A. Papaconstantopoulos, editors,<br />

Electronic Structure <strong>and</strong> Magnetism <strong>of</strong> Complex Materials. Springer, Verlag, 2002.<br />

[50] S. Sanvito, P. Ordejón, <strong>and</strong> N. A. Hill. Phys. Rev. B, 63:165206, 2001.<br />

[51] T. Dietl, F. Matsukura, H. Ohno, J. Cibert, <strong>and</strong> D. Ferr<strong>and</strong>. In I.D. Vagner et al.,<br />

editor, Recent Trends in Theory <strong>of</strong> Physical Phenomena in High <strong>Magnetic</strong> Fields,<br />

page 197. Kluwer Academic Publishers, 2003.


Bibliography 113<br />

[52] K. M. Yu, W. Walukiewicz, T. Wojtowicz, I. Kuryliszyn, X. Liu, Y. Sasaki, <strong>and</strong><br />

J. K. Furdyna. In Proceedings <strong>of</strong> the 25th International Conference on Physics <strong>of</strong><br />

Semiconductors, Edinburgh, 2002.<br />

[53] F. Matsukura, M. Sawicki, T. Dietl, D. Chiba, <strong>and</strong> H.Ohno. Physica E, 21:1032,<br />

2003.<br />

[54] M. Sawicki, F. Matsukura, A. Idziaszek, T.Dietl, G. M. Schott, C. Ruester, G. Karczewski,<br />

G. Schmidt, <strong>and</strong> L. W. Molenkamp. Phys. Rev. B, in press, e-print:<br />

http://arXiv.org/cond-mat/0212511, 2002.<br />

[55] D. Hrabovsky, E. Vanelle, A. R. Fert, D. S. Yee, J. P. Redoules, J. Sadowski, J. Kanski,<br />

<strong>and</strong> L. Ilver. Appl. Phys. Lett., 81:2806, 2002.<br />

[56] T. Dietl, J. Konig, <strong>and</strong> A. H. McDonald. Phys. Rev. B, 64:241201, 2001.<br />

[57] T. Shono, T. Hasegawa, T. Fukumura, F. Matsukura, <strong>and</strong> H. Ohno. Appl. Phys.<br />

Lett., 77:1363, 2000.<br />

[58] U. Welp, V. K. Vlasko-Vlasov, X. Liu, J. K. Furdyna, <strong>and</strong> T. Wojtowicz. Phys. Rev.<br />

Lett., 90(16):927, 2003.<br />

[59] J. Szczytko, W. Mac, A. Twardowski, F. Matsukura, <strong>and</strong> H. Ohno. Phys. Rev. B,<br />

59:12935, 1999.<br />

[60] T. Kuroiwa, T. Yasuda, F. Matsukura, A. Shen, Y. Ohno, Y. Segawa, <strong>and</strong> H. Ohno.<br />

Electron. Lett., 34:190, 1998.<br />

[61] J. Okabayashi, A. Kimura, O. Rader, T. Mizokawa, A. Fujimori, A. Hayashi, <strong>and</strong><br />

M. Tanaka. Phys. Rev. B, 58:R4211, 1998.<br />

[62] T. Jungwirth, W. A. Atkinson, B.H. Lee, <strong>and</strong> A. H. MacDonald. Phys. Rev. B,<br />

59:9818, 1999.<br />

[63] T. Omiya, F. Matsukura, T. Dietl, Y. Ohno, T. Sakon, M. Motokawa, <strong>and</strong> H. Ohno.<br />

Physica E, 7:976, 2000.<br />

[64] B. Beschoten, P. A. Crowell, I. Machejovich, D. D. Awschalom, F. Matsukura A.<br />

Shen, <strong>and</strong> H. Ohno. Phys. Rev. Lett., 83:3073, 1999.<br />

[65] T. Jungwirth, J. Konig, J. Sinova, J. Kucera, <strong>and</strong> A. H. MacDonald. Phys. Rev. B,<br />

66:012402, 2002.<br />

[66] T. Hayashi, Y. Hashimoto, S. Katsumoto, <strong>and</strong> Y.Iye. Appl. Phys. Lett., 78:1691,<br />

2001.<br />

[67] S. J. Potashnik, K. C. Ku, S. H. Chun, J. J. Berry. N. Samarth, <strong>and</strong> P. Schiffer. Appl.<br />

Phys. Lett., 79:1495, 2001.<br />

[68] K. Edmonds, K. Y. Wang, R. P. Campion, A. C. Neumann, N. R. S. Farley, B. L.<br />

Gallagher, <strong>and</strong> C. T. Foxon. Appl. Phys. Lett., 81:4991, 2002.


Bibliography 114<br />

[69] R. Mathieu, B. S. Sorenson, J. Sadowski, J. Kanski, P. Svedlindh, <strong>and</strong> P. E. Lindel<strong>of</strong>.<br />

Phys. Rev. B, 68:184421, 2003.<br />

[70] K. C. Ku, S. J. Potashnik, R. F. Wang, M. J. Seong, E. Johnston-Halperin, R. C.<br />

Meyers, S. H. Chun, A. Mascarenhas, A. C. Gossard, D. D. Awschalom, P. Schiffer,<br />

<strong>and</strong> N. Samarth. Appl. Phys. Lett., 82:2302.<br />

[71] B. S. Sorensen, J. Sadowski, S. E. Andersen, <strong>and</strong> P. E. Lindel<strong>of</strong>. Phys. Rev. B,<br />

66:233313, 2002.<br />

[72] L. Chien <strong>and</strong> C. R. Westgate, editors. The Hall Effect <strong>and</strong> Its Applications. Plenum,<br />

New York, 1980.<br />

[73] T. Jungwirth, Qian Niu, <strong>and</strong> A.H. MacDonald. Phys. Rev. Lett., 88:207208, 2002.<br />

[74] D. Baxter, D. Ruzmetov, J. Scherschlight, Y. Sasaki, X. Liu, J. K. Furdyna, <strong>and</strong><br />

C. H. Mielke. Phys. Rev. B, 65:212407, 2002.<br />

[75] K. Y. Wang, K. W. Edmonds, R. Campion, L. X. Zhao, A. C. Neumann, C. T.<br />

Foxon, B. L. Gallagher, C. Main, <strong>and</strong> C. H. Marrows. In Presented at 25th International<br />

Conference on Physics <strong>of</strong> Semiconductors, Edinburgth, 2002.<br />

[76] T. Jungwirth, J. Sinova, K. Y. Wang, K. W. Edmonds, R. Campion, B. L. Gallagher,<br />

C. T. Foxon, Qian Niu, <strong>and</strong> A. H. MacDonald. Appl. Phys. Lett., 83:320, 2003.<br />

[77] A. Oiwa, S. Katsumoto, A. Endo, Y. M. Hirasawa, Iye, H. Ohno, F. Matsukura,<br />

A. Shen, <strong>and</strong> Y. Sugawara. Phys. Stat. Sol. B, 205:167, 1998.<br />

[78] Y. Iye, A. Oiwa, A. Endo, S. Katsumoto, F. Matsukura, A. Shen, H. Ohno, <strong>and</strong><br />

H. Munekata. Materials Science <strong>and</strong> Engineering, B63:88, 1999.<br />

[79] T. Dietl. Diluted magnetic <strong>semiconductor</strong>s. In T. S. Moss <strong>and</strong> S. Mahajan, editors,<br />

H<strong>and</strong>book on Semiconductors, volume 3, chapter 17, page 1251. Elsevier Science<br />

B.V., Amsterdam, 1994.<br />

[80] S. Katsumoto, A. Oiwa, Y. Iye, F. Matsukura H. Ohno, A. Shen, <strong>and</strong> Y. Sugawara.<br />

Phys. Stat. Sol. B, 205:115, 1998.<br />

[81] T. Fukumura, T. Shono, K. Inaba, T. Hasegawa, H. Koinuma, F. Matsukura, <strong>and</strong><br />

H. Ohno. Physica E, 10:135, 2001.<br />

[82] J. F. Gregg, W. Allen, K. Ounadjela, M. Hehn M. Viret, S. M. Thompson, <strong>and</strong><br />

J. M. D. Coey. Phys. Rev. Lett., 77:1580, 1996.<br />

[83] S. V. Barabash <strong>and</strong> D. Stroud. Appl. Phys. Lett., 79:979, 2001.<br />

[84] P. N. Argyres. Phys. Rev., 97:334, 1955.<br />

[85] E. A. Stern H. S. Bennet. Phys. Rev., 137:A448, 965.


Bibliography 115<br />

[86] N. Négre. Magnéto-optique et forts champs magnétiques : nouvelles possibilités<br />

d’étude du magnétisme des solides. Application aux couches minces ŕ anisotropie<br />

élevée TbFe2 et bicouches AF/F, au semi-conducteur semi-magnétique GaMnAs<br />

et aux complexes moléculaires du FeII ŕ transition de spin. PhD thesis, INSA,<br />

Toulouse, 2000.<br />

[87] J. K. Furdyna, X. Liu, W. L. Lim, Y. Sasaki, T. Wojtowicz, I. Kuryliszyn, S. Lee,<br />

K. M. Yu, <strong>and</strong> W. Walukiewicz. Journal <strong>of</strong> the Korean Physical Society, 42:S579,<br />

2003.<br />

[88] M. L. Swanson. Rep. Prog. Phys., 45:42, 1982.<br />

[89] L. C. Feldman, J. W. Mayer, <strong>and</strong> S. T. Picraux. Materials Analysis by Ion Channeling.<br />

Academic Press, New York, 1982.<br />

[90] K. M. Yu. J. Appl. Phys., 74:86, 1993.<br />

[91] J. De Boeck, R. Oesterholt, A. Van Esch, H. Bender, C. Bruynseraede, C. Van<br />

Ho<strong>of</strong>, <strong>and</strong> G. Borghs. Appl. Phys. Lett., 68:2744.<br />

[92] J. Mašek <strong>and</strong> F. Máca. cond-mat/0308568, 2003.<br />

[93] K. W. Edmonds, P. Bogusławski, K. Y. Wang, R.P. Campion, B. L. Gallagher,<br />

N. R. S. Farley, C. T. Foxon, M. Sawicki, T. Dietl, M. B. Nardelli, <strong>and</strong> J. Bernholc.<br />

Phys. Rev. Lett., 92:037201, 2004.<br />

[94] J. Blinowski <strong>and</strong> P.Kacman. Phys. Rev. B, 67:121204(R), 2003.<br />

[95] J. Blinowski, P. Kacman, K. M. Yu, W. Walukiewicz, T. Wojtowicz, X. Liu, <strong>and</strong><br />

J. K. Furdyna. In Proc. XV Int. Conf. on High <strong>Magnetic</strong> in Semicond. Phys. Oxford<br />

2002, 2002.<br />

[96] G. M. Schott, G. Schmidt, G. Karczewski, R. Jakiela L.W. Molenkamp, A. Barcz,<br />

<strong>and</strong> G. Karczewski. Appl. Phys. Lett., 82:4678, 2003.<br />

[97] In N. R. Brozel <strong>and</strong> G. E. Stillman, editors, Properties <strong>of</strong> Gallium Arsenide. IN-<br />

SPEC, London, 1996.<br />

[98] J. K. Furdyna <strong>and</strong> J. Kossut. Superlatt. <strong>and</strong> Microstruct., 2:89, 1986.<br />

[99] J. Blinowski <strong>and</strong> P. Kacman. Phys. Rev. B, 67:121204, 2003.<br />

[100] P. Kacman <strong>and</strong> I. Kuryliszyn-Kudelska. Lecture Notes in Physics, in press.<br />

[101] D. E. Bliss, W. Walukiewicz, J. W. Ager, E. E. Haller, K. T. Chan, <strong>and</strong> S. Tanigawa.<br />

J. Appl. Phys., 71:1699(R), 1992.<br />

[102] R. M. A. Azzam <strong>and</strong> N. M. Bashara. Ellipsometry <strong>and</strong> polarized light. North–<br />

Holl<strong>and</strong> P. C., New York, 1977.


LIST OF FIGURES<br />

2.1 The schematic view <strong>of</strong> the Tracor X-ray Spectrace 5000. . . . . . . . . . 11<br />

2.2 The experimental setup for magneto<strong>transport</strong> measurements. . . . . . . . 13<br />

2.3 The experimental setup for magneto<strong>transport</strong> measurements (configuration<br />

with 7001 Keithley Scanner <strong>and</strong> 7065 Hall Card). . . . . . . . . . . . 14<br />

2.4 The configuration with two power suppliers (Oxford Instruments <strong>and</strong><br />

Lake Shore power supplier) connected in parallel. This configuration allowed<br />

to reverse fluently the direction <strong>of</strong> magnetic field. . . . . . . . . . . 14<br />

2.5 Experimental setup for AC susceptibility/DC magnetization measurements<br />

- 7229 LakeShore Susceptometer/Magnetometer system. . . . . . 17<br />

2.6 The cross-sectional view <strong>of</strong> the coil assembly. The two sensing coils are<br />

connected in opposition in order to cancel the voltages induced by the AC<br />

field itself or voltages induced by unwanted external sources. . . . . . . . 18<br />

2.7 The schematic view <strong>of</strong> experimental setup for magnetooptical Kerr effect<br />

measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

3.1 Chemical composition distribution along the crystal growth direction for<br />

the crystal <strong>of</strong> PbSnMnEuTe. . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2 RHEED oscillations observed during the growth <strong>of</strong> a Ga 1−x Mn x As film<br />

with x = 0.062. The first 7 periods correspond to LT-GaAs. The "jump"<br />

in the signal occurs at the point when the Mn shutter has been opened <strong>and</strong><br />

the rate <strong>of</strong> oscillations increased. . . . . . . . . . . . . . . . . . . . . . . 24<br />

4.1 The b<strong>and</strong> structure model <strong>of</strong> Pb 1−x−y Mn x Sn y Te mixed crystals . . . . . . 28<br />

4.2 Curie-Weiss temperature (Θ) versus free carrier concentration in<br />

Pb 1−x−y Mn x Sn y Te. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

4.3 Inverse AC susceptibility versus temperature measured for<br />

Pb 1−x−y Mn x Eu y Te samples (793 − 2: x=0.010, y=0; 793 − 4: x=0.010,<br />

y=0.001; 793 − 10: x=0.005, y=0.004; 793 − 12: x=0.007, y=0.003;<br />

793 − 14: x=0.005, y=0.005). The solid lines correspond to Curie -Weiss<br />

law fits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

4.4 The high temperature inverse AC susceptibility measured for several<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples (809 − 2: x=0.031, y=0.003, z=0.850;<br />

809 − 12: x=0.022, y=0.003, z=0.760; 809 − 32: x=0.026, y=0.014,<br />

z=0.69; 809 − 34: x=0.027, y=0.017, z=0.680).The solid lines correspond<br />

to Curie -Weiss law fits. . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.5 The high temperature inverse AC susceptibility measured for several<br />

Sn 1−x−y Mn x Eu y Te samples (841 − 18: x=0.131, y=0.013; 841 − 14:<br />

x=0.116, y=0.011; 842 − 20: x=0.091, y=0.009; 848 − 4: x=0.061,<br />

y=0.0115; 848 − 16: x=0.050, y=0.011; 848 − 24: x=0.051, y=0.019). . . . 38


List <strong>of</strong> Figures 117<br />

4.6 The low temperature behaviour <strong>of</strong> real part <strong>of</strong> susceptibility for several<br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples (809 − 12: x=0.022, y=0.003, z=0.760;<br />

809 − 36: x=0.025, y=0.013, z=0.690; 809 − 30: x=0.024, y=0.010,<br />

z=0.710 ;809 − 32: x=0.026, y=0.014, z=0.690; 809 − 34: x=0.027,<br />

y=0.017, z=0.680). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.7 The low temperature behaviour <strong>of</strong> real part <strong>of</strong> susceptibility for several<br />

Sn 1−x−y Mn x Eu y Te samples (841 − 14: x=0.121, y=0.011; 842 − 20:<br />

x=0.090, y=0.009; 841 − 18: x=0.128, y=0.014; 848 − 24: x=0.055,<br />

y=0.0175; 848 − 16: x=0.054, y=0.011; 848 − 4: x=0.058, y=0.011). . . . 40<br />

4.8 Magnetization measured at various temperatures <strong>and</strong> magnetic fields up to<br />

9T for Pb 1−x−y−z Mn x Eu y Sn z Te 809 − 10 sample with x=0.030, y=0.003,<br />

z=0.780 <strong>and</strong> hole concentration p=6 10 20 cm −3 . . . . . . . . . . . . . . 41<br />

4.9 The low temperature behaviour <strong>of</strong> both real Re(χ) <strong>and</strong> imaginary Im(χ)<br />

component <strong>of</strong> susceptibility for two samples <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te:<br />

809 − 2 (x=0.031, y=0.003, p=1·10 21 cm −3 ) <strong>and</strong> 809 − 12 (x=0.022,<br />

y=0.003, p=4·10 20 cm −3 ). The typical <strong>ferromagnetic</strong> characteristics is<br />

observed for 809 − 12 sample <strong>and</strong> spin glass behaviour for the sample with<br />

higher free hole concentration. . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4.10 The frequency dependence <strong>of</strong> real component <strong>of</strong> susceptibility Re(χ)<br />

for the sample <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 2: x=0.031, y=0.003,<br />

p=1·10 21 cm −3 . The shift <strong>of</strong> the freezing temperature T f towards higher<br />

temperatures with the frequency increase is clearly visible. . . . . . . . . 43<br />

4.11 The frequency dependence <strong>of</strong> imaginary part <strong>of</strong> susceptibility Im(χ)<br />

for the sample <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te: 809 − 2 (x=0.031, y=0.003,<br />

p=1·10 21 cm −3 ). The maximum <strong>of</strong> the observed cusp shifts towards<br />

higher temperatures with the frequency increase. . . . . . . . . . . . . . 44<br />

4.12 <strong>Magnetic</strong> phase diagram for Pb 1−x−y Mn x Sn y Te [27] <strong>and</strong><br />

Pb 1−x−y−z Mn x Eu y Sn z Te samples. Red triangles correspond to PbMnEuSnTe<br />

ferromagnets, red circles to PbMnEuSnTe spin glasses, black<br />

circles to PbMnSnTe ferromagnets, green circles to PbMnSnTe spin<br />

glasses, blue circles to reentrant spin glasses. Lines present model<br />

calculations <strong>of</strong> the phase boundary (see Ref. [27]): solid line presents<br />

geometric model, dashed <strong>and</strong> dot dashed lines correspond to Sherrington-<br />

Kirkpatrick model, dot line correspond to Sherrington-Southern model . 45<br />

4.13 Curie temperature calculated for Pb 1−x−y−z Mn x Eu y Sn z Te mixed crystals<br />

as a function <strong>of</strong> Eu content y for various values <strong>of</strong> Sn concentration 0.6<br />

≤ z ≤ 1 <strong>and</strong> Mn concentration x=0.02 . . . . . . . . . . . . . . . . . . 45<br />

4.14 Zero field cooled magnetization measured as a function <strong>of</strong> temperature for<br />

Sn 1−x−y Mn x Eu y Te sample (842 − 8 with x=0.068, y=0.003, p=1.6·10 21<br />

cm −3 ) at ambient <strong>and</strong> equal to 11.2 kbar pressure. . . . . . . . . . . . . . 46<br />

4.15 Zero field cooled magnetization measured as a function <strong>of</strong> temperature<br />

for Sn 1−x Mn x Te sample with x=0.10 at ambient <strong>and</strong> equal to 10.5 kbar<br />

pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

4.16 Zero field cooled magnetization measured as a function <strong>of</strong> magnetic field<br />

at low temperature T =5K for Sn 1−x−y Mn x Eu y Te sample (842 − 8 with<br />

x=0.068, y=0.003, p=1.6·10 21 cm −3 ) at ambient <strong>and</strong> equal to 11.2 kbar<br />

pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47


List <strong>of</strong> Figures 118<br />

4.17 Zero field cooled magnetization measured as a function <strong>of</strong> magnetic field<br />

at low temperature T =4.2K for Sn 1−x Mn x Te sample with x=0.10 at ambient<br />

<strong>and</strong> equal to 10.5 kbar pressure. . . . . . . . . . . . . . . . . . . . 47<br />

5.1 The zero-field resistivity <strong>of</strong> the as-grown GaMnAs epilayers in the wide<br />

range <strong>of</strong> Mn concentration 0.01≤ x ≤0.093. . . . . . . . . . . . . . . . . 52<br />

5.2 The temperature dependence <strong>of</strong> the zero-field resistivity (a hump structure<br />

slightly above Curie temperature is visible) <strong>and</strong> magnetization versus<br />

temperature measured by a SQUID magnetometer in small magnetic field<br />

(B=0.01T) parallel to the sample surface. A good agreement between<br />

<strong>transport</strong> <strong>and</strong> magnetic data is visible. . . . . . . . . . . . . . . . . . . . 53<br />

5.3 Curie temperature T ρ estimated from the the zero-field resistivity measurements<br />

versus temperatures <strong>of</strong> annealing for GaMnAs samples with<br />

different Mn concentration 0.01≤ x ≤0.093. . . . . . . . . . . . . . . . . 54<br />

5.4 Temperature dependence <strong>of</strong> the zero-field resistivity <strong>of</strong> GaMnAs sample<br />

with high Mn concentration (x=0.086) annealed at various temperatures. . 55<br />

5.5 The temperature dependence <strong>of</strong> zero-field resistivity for as-grown <strong>and</strong><br />

annealed at various temperatures sample with low Mn concentration<br />

x=0.027. Inset shows the Curie temperature estimated from the resistivity<br />

measurements versus annealing temperature. . . . . . . . . . . . . . . . . 55<br />

5.6 Conductivity versus annealing temperature for the samples with high Mn<br />

concentration x=0.086 <strong>and</strong> with low Mn concentration x=0.027. . . . . . 57<br />

5.7 The Hall voltage versus magnetic field for two samples <strong>of</strong> Ga 1−x Mn x As<br />

with x=0.086, as-grown <strong>and</strong> annealed at the optimal conditions, measured<br />

at T=4.2K. Note that dominant contribution to the Hall voltage comes<br />

from the AHE term; the Hall voltage reflects the M(B) behaviour. . . . . . 59<br />

5.8 Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0,<br />

for the epilayer with high Mn content x=0.07 measured at various temperatures<br />

4.2K, 40K, 110K, 150K <strong>and</strong> 200K for as-grown sample annealed<br />

at the optimal conditions (T a =289 0 C) <strong>and</strong> after annealing at higher temperature<br />

(T a =350 0 C). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

5.9 Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0,<br />

for the epilayer with high Mn content x=0.086: as-grown, annealed at the<br />

optimal conditions (T a =280 0 C) <strong>and</strong> after annealing at higher temperature<br />

(T a =350 0 C). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

5.10 Magnetoresistivity (R-R 0 )/R 0 , where R 0 is the value <strong>of</strong> resistivity at B=0,<br />

for the epilayer with low Mn content x=0.027: as-grown <strong>and</strong> annealed at<br />

the optimal conditions (T a =300 0 C). . . . . . . . . . . . . . . . . . . . . 62<br />

5.11 Magnetoresistivity R for the sample with x=0.01 in low magnetic fields at<br />

T=20K. The magnetic field B was applied perpendicular to the film. Note<br />

that hysteretic behavior is visible. The arrows <strong>and</strong> numbers indicate the<br />

history <strong>and</strong> direction <strong>of</strong> applied magnetic field. . . . . . . . . . . . . . . 63<br />

5.12 Low magnetic field magnetoresistivity for the Ga 1−x Mn x As sample with<br />

x=0.01 measured at various temperatures. The magnetic field B was applied<br />

perpendicular to the film. The arrows <strong>and</strong> numbers indicate the<br />

history <strong>and</strong> direction <strong>of</strong> applied magnetic field. . . . . . . . . . . . . . . 64


List <strong>of</strong> Figures 119<br />

5.13 Magnetization versus temperature M(T) measured in small magnetic field<br />

B = 10Gs for the as-grown <strong>and</strong> annealed at 289 0 C GaMnAs sample with<br />

x=0.07. It is clearly visible that annealing procedure at the optimal conditions<br />

leads to the increase <strong>of</strong> saturation magnetization. . . . . . . . . . 65<br />

5.14 Magnetization versus temperature M(T) measured in small magnetic field<br />

B = 10Gs for the as-grown <strong>and</strong> annealed at 289 0 C GaMnAs sample with<br />

x=0.07. It is clearly visible that annealing procedure at the optimal conditions<br />

leads to the increase <strong>of</strong> saturation magnetization. . . . . . . . . . 66<br />

5.15 The hysteresis loop M(B) for the as-grown <strong>and</strong> annealed at the optimal<br />

conditions sample with x=0.07 measured at T=5K. The decrease <strong>of</strong> coercive<br />

field is visible after optimal annealing. . . . . . . . . . . . . . . . . 67<br />

5.16 Magnetization versus temperature measured by SQUID magnetometer at<br />

small magnetic field B=10Gs for three samples with x=0.062 (10727D<br />

epilayer): as-grown, annealed at the temperature 289 0 C <strong>and</strong> 350 0 C. . . . . 68<br />

5.17 The hysteresis loops M(B) for the sample with x=0.062 as-grown, annealed<br />

at the temperature equal to 289 0 C <strong>and</strong> higher temperature 350 0 C. . 68<br />

5.18 The principle <strong>of</strong> magnetooptical Kerr effect method. . . . . . . . . . . . . 69<br />

5.19 Three configurations for magnetooptical Kerr effect measurements: longitudinal,<br />

transverse, polar. . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

5.20 Kerr rotation angle θ K versus magnetic field up to 6T for the as grown<br />

epilayer with high Mn content x=0.086 <strong>and</strong> a Curie temperature T C =88K. 72<br />

5.21 The magnetic field dependence <strong>of</strong> Kerr rotation angle θ K at different temperatures<br />

for the epilayer with x=0.086 annealed at the optimal conditions<br />

T a = 289 0 C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

5.22 The results <strong>of</strong> MOKE obtained at T=5K for λ=632.8nm for all investigated<br />

samples as grown <strong>and</strong> annealed 10727E epilayer with x=0.086 <strong>and</strong><br />

for epilayers with lower Mn concentration: 11127A (x=0.048) as well as<br />

10529A (x=0.014). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.23 The results <strong>of</strong> MOKE obtained at T=5K for λ=540.5nm for all investigated<br />

samples as grown <strong>and</strong> annealed 10727E epilayer with x=0.086 <strong>and</strong><br />

for epilayers with lower Mn concentration: 11127A (x=0.048) as well as<br />

10529A (x=0.014). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

5.24 The comparison <strong>of</strong> θ K (B) behaviour for two used wavelengths (two wavelengths<br />

<strong>of</strong> incident light λ=632.8nm <strong>and</strong> λ=540.5nm) for as grown epilayer<br />

with x=0.086. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

5.25 The comparison <strong>of</strong> θ K (B) behaviour for two used wavelengths (two wavelengths<br />

<strong>of</strong> incident light λ=632.8nm <strong>and</strong> λ=540.5nm) for as grown epilayer<br />

with x=0.014. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.26 SQUID magnetization M versus magnetic field at 5K. The magnetic field<br />

was applied perpendicular to the sample surface. . . . . . . . . . . . . . . 76<br />

5.27 The comparison <strong>of</strong> the two data sets: Kerr rotation angle with the SQUID<br />

magnetization M(B) collected at T=5K. . . . . . . . . . . . . . . . . . . 76<br />

5.28 The results <strong>of</strong> MOKE obtained at T=5K for λ=632.8nm in the range <strong>of</strong><br />

high magnetic fields for all investigated samples as grown <strong>and</strong> annealed<br />

10727E epilayer with x=0.086 <strong>and</strong> for epilayers with lower Mn concentration:<br />

11127A (x=0.048) as well as 10529A (x=0.014) . . . . . . . . . . 77


List <strong>of</strong> Figures 120<br />

5.29 The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> inverse magnetic<br />

field measured at T=5K for two epilayers: 10529A (x=0.014) <strong>and</strong><br />

11127A (x=0.048). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

5.30 The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> inverse magnetic<br />

field measured at T=5K for as-grown as well as annealed sample <strong>of</strong><br />

10727E epilayer with x=0.086. . . . . . . . . . . . . . . . . . . . . . . . 80<br />

5.31 The results <strong>of</strong> Fourier transform analysis <strong>of</strong> Kerr rotation angle curves<br />

measured at T=5K for two epilayers: 10529A (x=0.014) <strong>and</strong> 11127A<br />

(x=0.048) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.32 The results <strong>of</strong> Fourier transform analysis <strong>of</strong> Kerr rotation angle curves<br />

measured at T=5K for as-grown as well as annealed sample <strong>of</strong> 10727E<br />

epilayer with x=0.086. . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.33 The oscillatory terms <strong>of</strong> Kerr rotation angle as a function <strong>of</strong> magnetic<br />

field B measured at T=5K for as-grown as well as annealed sample <strong>of</strong><br />

10727E epilayer with x=0.086. The shift <strong>of</strong> oscillating terms towards<br />

higher magnetic field after thermal annealing is visible. . . . . . . . . . . 83<br />

5.34 Schematic view <strong>of</strong> the channeling <strong>of</strong> ions directed at an angle Ψ to a<br />

close-packed row <strong>of</strong> atoms in a crystal. . . . . . . . . . . . . . . . . . . 84<br />

5.35 The normalized yield χ h <strong>of</strong> ions that are backscattered from host atoms<br />

(the RBS yield) shows a strong dip at Ψ=0. If 50% <strong>of</strong> solute atoms are<br />

displaced into the channel, the normalized yield χ s <strong>of</strong> ions backscattered<br />

from the solute atoms is approximately half the r<strong>and</strong>om yield; i.e., χ s =0.5<br />

at Ψ=0 (broken curve). If displaced solute atoms are located near the<br />

center <strong>of</strong> the channel, a peak in yield may occur (dotted line). . . . . . . . 85<br />

5.36 Schematic <strong>of</strong> the tetrahedral interstitial positions for a zinc-blende lattice<br />

along the various axial directions. . . . . . . . . . . . . . . . . . . . . . 86<br />

5.37 The PIXE <strong>and</strong> RBS angular scans about the 〈100〉 axes for as-grown sample<br />

with x=0.092. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

5.38 The PIXE <strong>and</strong> RBS angular scans about the 〈110〉 axes for as-grown sample<br />

with x=0.092. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.39 The PIXE <strong>and</strong> RBS angular scans about the 〈111〉 axes for as-grown sample<br />

with x=0.092. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.40 The fraction <strong>of</strong> nonr<strong>and</strong>om Mn f nr , calculated by comparing normalized<br />

yields χ Mn <strong>and</strong> χ GaAs for the 〈100〉, 〈110〉 <strong>and</strong> 〈111〉 projections. . . . . . 89<br />

5.41 The ω/2Θ scan for the symmetric (004) Bragg reflection for as-grown<br />

samples with x=0.027, 0.062, <strong>and</strong> 0.086. . . . . . . . . . . . . . . . . . . 91<br />

5.42 The ω/2Θ scan for the symmetric (004) Bragg reflection for as-grown <strong>and</strong><br />

annealed sample with x=0.086. . . . . . . . . . . . . . . . . . . . . . . . 91<br />

5.43 Comparison <strong>of</strong> the measured ω/2Θ scan for the symmetric (004) Bragg<br />

reflection (dashed curve) with simulation results (solid curve) for the annealed<br />

sample with x= 0.086. . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

5.44 Reciprocal space maps <strong>of</strong> the symmetric (004) reflection for the as-grown<br />

<strong>and</strong> annealed sample with x=0.086. (Q x <strong>and</strong> Q y represent reciprocal space<br />

vectors (Q x is in the direction parallel to the surface, Q y is in the direction<br />

perpendicular to the surface), both given in λ/2d units, λ=0.15406<br />

nm, d denotes the interplanar spacing). . . . . . . . . . . . . . . . . . . . 94


List <strong>of</strong> Figures 121<br />

5.45 Reciprocal space maps <strong>of</strong> the asymmetric (224) reflection for the asgrown<br />

<strong>and</strong> annealed sample with x=0.086. (Q x <strong>and</strong> Q y represent reciprocal<br />

space vectors (Q x is in the direction parallel to the surface, Q y<br />

is in the direction perpendicular to the surface), both given in λ/2d units,<br />

λ=0.15406 nm, d denotes the interplanar spacing). . . . . . . . . . . . . . 95<br />

5.46 Relaxed lattice parameter <strong>of</strong> Ga 1−x Mn x As plotted as a function <strong>of</strong> the Mn<br />

concentration x for as-grown <strong>and</strong> annealed samples. . . . . . . . . . . . . 97<br />

7.1 The parameters that define the ellipse <strong>of</strong> polarization in its plane: 1) the<br />

azimuth θ, that defines orientation <strong>of</strong> the ellipse in its plane (− π ≤ θ ≤<br />

2<br />

π<br />

2 a 4<br />

π<br />

4 (a2 +b 2 ) 1 2 ); 5) the absolute phase δ defined by<br />

the initial electric field <strong>and</strong> auxiliary (dashed) circle (-π ≤ δ ≤ π), it<br />

determines the angle between the initial position <strong>of</strong> the electric vector at<br />

t=0 <strong>and</strong> major axis <strong>of</strong> the ellipse. . . . . . . . . . . . . . . . . . . . . . . 104<br />

7.2 1 Energy b<strong>and</strong>s for a simple <strong>semiconductor</strong> for B=0. . . . . . . . . . . . 105<br />

7.3 Energy b<strong>and</strong>s for a simple <strong>semiconductor</strong> for B>0. . . . . . . . . . . . . 107<br />

7.4 The intrerb<strong>and</strong> transitions between L<strong>and</strong>au levels (n=0,1,2,3,4) <strong>of</strong> two<br />

simple parabolic b<strong>and</strong>s ( the left side <strong>of</strong> the picture represents simple energy<br />

b<strong>and</strong>s for B=0). The arrows indicate the conditions for resonance<br />

transitions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109


LIST OF TABLES<br />

2.1 Specifications <strong>of</strong> 7229 LakeShore Susceptometer/Magnetometer system. 15<br />

3.1 The chemical composition <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples determined<br />

by means <strong>of</strong> X-ray dispersive fluorescence analysis technique. . . 22<br />

3.2 The lattice constant a 0 <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples determined<br />

by the st<strong>and</strong>ard powder X-ray measurements <strong>and</strong> the values <strong>of</strong> the lattice<br />

constant <strong>of</strong> Pb 1−x−y Mn x Sn y Te a [11] with similar content <strong>of</strong> Mn <strong>and</strong> Sn<br />

as for the samples investigated in the thesis. . . . . . . . . . . . . . . . . 22<br />

3.3 The parameters <strong>of</strong> the LT MBE growth <strong>of</strong> GaMnAs samples - substrate<br />

temperature T S <strong>and</strong> temperature <strong>of</strong> the Mn effusion cell T Mn , thickness<br />

<strong>of</strong> the GaMnAs layers d GaMnAs determined from RHEED oscillations <strong>and</strong><br />

Mn composition <strong>of</strong> investigated samples x (determined from RHEED oscillations,<br />

X-ray diffraction <strong>and</strong> high resolution X-ray diffraction measurements<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

4.1 The results <strong>of</strong> <strong>transport</strong> characterization <strong>of</strong> Pb 1−x−y−z Mn x Eu y Sn z Te samples<br />

- hole concentration p [10 21 cm −3 ], conductivity σ[(Ωcm) −1 ], mobility<br />

µ [cm 2 /Vs]) measured at the room <strong>and</strong> liquid nitrogen temperature. . . 33<br />

4.2 . The results <strong>of</strong> magnetic measurements for IV-VI mixed crystals. . . . . 35<br />

5.1 Curie temperature T ρ estimated from the zero-field resistivity measurements<br />

for as-grown (a.g.) <strong>and</strong> annealed at different temperatures T a for 1<br />

hour in nitrogen atmosphere GaMnAs samples with different Mn composition<br />

(0.1≤x≤0.093) <strong>and</strong> various layer thickness d. . . . . . . . . . . . . 54<br />

5.2 The measured values <strong>of</strong> perpendicular to the layer plane lattice<br />

parameter(a ⊥ ), in-plane lattice parameter (a ‖ ), the calculated values <strong>of</strong><br />

the relaxed mismatch (a relax -a s )/a s , <strong>and</strong> thickness <strong>of</strong> the GaMnAs epilayers<br />

determined from both RHEED oscilations <strong>and</strong> XRD measurements<br />

before <strong>and</strong> after annealing. . . . . . . . . . . . . . . . . . . . . . . . . . 92

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!