175
University of South Florida Scholar Commons Graduate eses and Dissertations Graduate School January 2013 Magnetism in Complex Oxides Probed by Magnetocaloric Effect and Transverse Susceptibility Nicholas Steven Bingham University of South Florida, [email protected] Follow this and additional works at: hp://scholarcommons.usf.edu/etd Part of the Condensed Maer Physics Commons is Dissertation is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in Graduate eses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact [email protected]. Scholar Commons Citation Bingham, Nicholas Steven, "Magnetism in Complex Oxides Probed by Magnetocaloric Effect and Transverse Susceptibility" (2013). Graduate eses and Dissertations. hp://scholarcommons.usf.edu/etd/4440

Magnetism in Complex Oxides Probed by Magnetocaloric

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Magnetism in Complex Oxides Probed by Magnetocaloric

University of South FloridaScholar Commons

Graduate Theses and Dissertations Graduate School

January 2013

Magnetism in Complex Oxides Probed byMagnetocaloric Effect and TransverseSusceptibilityNicholas Steven BinghamUniversity of South Florida, [email protected]

Follow this and additional works at: http://scholarcommons.usf.edu/etd

Part of the Condensed Matter Physics Commons

This Dissertation is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion inGraduate Theses and Dissertations by an authorized administrator of Scholar Commons. For more information, please [email protected].

Scholar Commons CitationBingham, Nicholas Steven, "Magnetism in Complex Oxides Probed by Magnetocaloric Effect and Transverse Susceptibility" (2013).Graduate Theses and Dissertations.http://scholarcommons.usf.edu/etd/4440

Page 2: Magnetism in Complex Oxides Probed by Magnetocaloric

Magnetism in Complex Oxides Probed by Magnetocaloric Effect and Transverse

Susceptibility

by

Nicholas S. Bingham

A dissertation submitted in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

Department of Physics

College of Arts and Sciences

University of South Florida

Co-Major Professor: Hariharan Srikanth Ph.D.

Co-Major Professor: Manh-Huong Phan Ph.D.

Casey Miller Ph.D.

Sarath Witanachchi Ph.D.

Gerald Woods Ph.D.

Michael Osofsky Ph.D.

Date of Approval

April 8, 2013

Keywords: Manganites, Cobaltites, Frustrated Magnets, Phase Separation, Magnetism

Copyright © 2013, Nicholas S. Bingham

Page 3: Magnetism in Complex Oxides Probed by Magnetocaloric

DEDICATION

I dedicate this dissertation to all of my friends and family who have helped and

supported me throughout this process. In particular, my wife Gina Bingham, my siblings

Cynda, Matt, Callie, Sarah, Kayla, Erin and Carl, as well as, their spouses and wonderful

children. I also want to specially recognize my parents Dr. Paul Bingham and Julie

Popick, my step-parents Dr. Edward Popick and Cindy Bingham, my father-in-law David

McIntyre and very special thanks to my late mother-in-law Anita McIntyre. I would not

have been able to accomplish this without the support from all of them.

Page 4: Magnetism in Complex Oxides Probed by Magnetocaloric

ACKNOWLEDGMENTS

I would like to extend my utmost gratitude to all of the individuals who have

helped me throughout my research to make this dissertation possible. Of course, I would

like to thank my adviser Professor Hariharan Srikanth for his guidance, constant support,

and patience through the writing of this dissertation and all of my Ph.D. work in general.

Professor Srikanth pushed me to ask questions and taught me how to present my research

in a professional manner. I also want to give special thanks to Dr. Manh-Huong Phan,

whose exuberance for science, physics and education was contagious, not just for me but

for everyone he talks to. I will always be grateful to Dr. Phan, for all of his support,

encouragement and guidance throughout all of the research I have been part of at USF. I

will always be grateful to Dr. Alberto Pique, Dr. Michael Osofsky and Dr. Huengsoo

Kim at the Naval Research Laboratory (NRL). They supported me through my Industrial

Practicum, as well as several summers thereafter. I want to thank Professor Victorino

Franco for all of his help with my research as well as his friendship.

I would like to thank my committee members, Dr. Casey Miller, Dr. Sarath

Witanachchi, Dr. Gerald Woods and Dr. Michael Osofsky for stimulating discussions and

general guidance throughout this dissertation. I would like to also thank our collaborators,

in particular, Dr. Christopher Leighton and Dr. Sang-Wook Cheong for providing

excellent quality samples that greatly helped me throughout my research. I want to give

very special thanks to my lovely wife Gina Bingham, without her constant support I

Page 5: Magnetism in Complex Oxides Probed by Magnetocaloric

would have never been able to accomplish any of my research. Also, special thanks to my

former lab mate Dr. Anurag Chaturvedi, who has always been there for me as friend and

colleague. I would like to thank all of my current and former lab-mates as well, in

particular Kristen, Paige, Corey, Monet and Natalie for all of the collaboration and fun

throughout my time at USF.

Page 6: Magnetism in Complex Oxides Probed by Magnetocaloric

i

TABLE OF CONTENTS

LIST OF FIGURES………………………………………………………………iv

ABSTRACT…………………….…………………………………………….......ix

CHAPTER 1 INTRODUCTION .............................................................................1

1.1 Overview and Motivation ..................................................................... 1

1.2 Objectives of the Dissertation ............................................................... 2

1.3 Outline of the Dissertation .................................................................... 3

CHAPTER 2. FUNDAMENTAL ASPECTS OF MANGANITES ........................6

2.1 Crystal Structure ................................................................................... 7

2.2 Crystal field splitting and the Jahn-Teller effect................................... 9

2.3 Magnetic Interactions.......................................................................... 11

2.4 Conclusions ......................................................................................... 16

References ................................................................................................. 16

CHAPTER 3. EXPERIMENTAL METHODS .....................................................19

3.1 Magnetocaloric Effect ......................................................................... 19

3.1.1 What is the Magnetocaloric Effect? ....................................19

3.1.2 Theoretical Aspects of MCE ...............................................21

3.2 Magnetocaloric Effect as a Fundamental probe…………..….………29

3.2.1 Order of Transitions and Critical Phenomena…………….

............................................................................................................. 25

3.3 Transverse Susceptibility .................................................................... 29

References ................................................................................................. 33

CHAPTER 4 IMPACT OF REDUCED DIMENSIONALITY ON THE

MAGNETIC AND MAGNETOCALORIC RESPONSE OF

La0.7Ca0.3MnO3 ..........................................................................................36

4.1 Introduction ......................................................................................... 36

Page 7: Magnetism in Complex Oxides Probed by Magnetocaloric

ii

4.2 Experiment .......................................................................................... 38

4.3 Results and Discussion ....................................................................... 40

4.4 Conclusions ......................................................................................... 45

References ................................................................................................. 45

CHAPTER 5 INFLUENCE OF Sr DOPING ON THE MAGNETIC

TRANSITIONS AND CRITICAL BEHAVIOR OF

La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2, AND 0.25) SINGLE

CRYSTALS ...............................................................................................48

5.1 Introduction ......................................................................................... 49

5.2 Experiment .......................................................................................... 50

5.3 Results and Discussion ....................................................................... 51

5.4 Conclusions ......................................................................................... 62

References ................................................................................................. 62

CHAPTER 6 MAGNETIC TRANSITIONS, MAGNETOCALORIC

EFFECT, MAGNETIC ANISOTROPY, CRITICAL

EXPONENTS AND THEIR CORRELATIONS IN Pr0.5Sr0.5MnO3.........68

6.1 Introduction ......................................................................................... 69

6.2 Experiment .......................................................................................... 72

6.3 Results and Discussion ....................................................................... 72

6.3.1 Influence of first- and second-order magnetic phase

transitions on the magnetocaloric effect and refrigerant capacity of

Pr0.5Sr0.5MnO3 ................................................................................72

6.3.2 Magnetic Anisotropy and Magnetization Dynamics in

Pr0.5Sr0.5MnO3 ................................................................................86

6.4 Conclusions ......................................................................................... 91

References: ................................................................................................ 92

CHAPTER 7 PROBING MULTIPLE MAGNETIC TRANSITIONS AND

PHASE COEXISTENCE IN La5/8−xPrxCa3/8MnO3 (x = 0.275)

SINGLE CRYSTALS ................................................................................98

7.1 Introduction ......................................................................................... 99

7.2 Results and Discussion ..................................................................... 100

7.2.1 Phase Coexistence and Magnetocaloric Effect ...................100

7.2.2 Transverse Susceptibility ................................................... 109

7.3 Conclusions ....................................................................................... 112

References: .............................................................................................. 112

Page 8: Magnetism in Complex Oxides Probed by Magnetocaloric

iii

CHAPTER 8 MAGNETOCALORIC EFFECT AND TRANSVERSE

SUSCEPTIBILITY OF Pr1-xSrxCoO3 (x =0.3-0.5): IMPACT OF

THE MAGNETOCRYSTALLINE ANISOTROPY-DRIVEN

PHASE TRANSITION ............................................................................116

8.1 Introduction ....................................................................................... 117

8.2 Results and Discussion ..................................................................... 118

8.2.1 Anomalous magnetism and Magnetocaloric effect in Pr1-

xSrxCoO3 (0.3 ≤ x ≤ 0.5) ...............................................................118

8.2.2 Transverse susceptibility as a probe of the coupled

structural/magnetocrystalline anisotropy transition in Pr1-xSrxCoO3

(x = 0.5) ........................................................................................125

8.3 Conclusions ....................................................................................... 129

References: .............................................................................................. 130

CHAPTER 9 A COMPLEX MAGNETIC PHASE DIAGRAM AND

MAGNETOCALORIC EFFECT IN Ca3Co2O6 SINGLE

CRYSTALS .............................................................................................133

9.1 Introduction ....................................................................................... 133

9.2 Results and Discussion ..................................................................... 135

9.3 Conclusions ....................................................................................... 142

References: .............................................................................................. 142

CHAPTER 10 CONCLUSIONS AND OUTLOOK ...........................................145

10.1 Conclusions ..................................................................................... 145

10.2 Outlook ........................................................................................... 148

References: .............................................................................................. 150

APPENDICES .....................................................................................................151

APPENDIX A LIST OF PUBLICATIONS ........................................... 152

APPENDIX B LIST OF CONFERENCE PRESENTATION................ 156

Page 9: Magnetism in Complex Oxides Probed by Magnetocaloric

iv

LIST OF FIGURES

Figure 2.1: Idealized cubic perovskites phase: Mn3+

/4+

(green) occupy the center of the

cube, oxygen ions (red) form octahedra around the Mn3+

/4+

ions. And the trivalent rare

earth or divalent alkali earth (blue) forms the corners of the cube. .................................... 7

Figure 2.2: The Pnma unit cell of La1-xCaxMnO3 (taken from [8]) showing general

distortions of the standard cubic lattice. The ions are represented by black (manganese),

grey (La or Ca) and white (oxygen) spheres. ...................................................................... 8

Figure 2.3: Crystal field splits degenerate 3d orbitals into eg and t2g levels, then crystal

distortion due to the Jahn-Teller effect, further splits the degenerate eg and t2g levels

(taken from [10]). .............................................................................................................. 10

Figure 2.4: A schematic representation of the double exchange mechanism showing the

simultaneous transfer of electrons between Mn3+

to O2-

and from the O2-

to Mn4+

ions

taken from [14]. ................................................................................................................ 12

Figure 2.5: Schematic representation showing the arrangement of spins and orbitals in

superexchange taken from [14]. ........................................................................................ 14

Figure 2.6: CE -type charge ordering along ab-plane in La0.5Sr0.5MnO4 taken from

[http://folk.uio.no/ravi/activity/ordering/chargeordering.html] ........................................ 15

Figure 2.7: Various forms of orbital ordering in the manganites taken from

[http://folk.uio.no/ravi/activity/ordering/orbitalordering.html] ........................................ 16

Figure 3.1: Schematic of a working magnetic refrigerator. Image credit Tegus et. al.

Nature 415 (2002). ............................................................................................................ 20

Figure 3.2: The method for calculating the RC from the −SM(T) curve using Eq. (3.13)

for two types of transitions in Pr0.5Sr0.5MnO3. .................................................................. 24

Figure 3.3: (a)Schematic diagram of the transverse susceptibility circuit, (b) schematic

depiction of transverse susceptibility probe , (c) Quantum Design PPMS. ...................... 31

Figure 3.4: Transverse and parallel susceptibility (T and P respectively) for single-

domain magnetic particles. Image from reference [20]. ................................................... 32

Figure 4.1: Phase diagram of La1-xCaxMnO3, showing the subtle balance between

chemical doping and magnetic properties (taken from [2]). ............................................. 37

Page 10: Magnetism in Complex Oxides Probed by Magnetocaloric

v

Figure 4.2: Schematic representation of a pulsed laser deposition system. Image credit

Andor Technology. ........................................................................................................... 39

Figure 4.3: Temperature dependence of magnetization recorded on cooling in a field of

500 Oe and normalized to 25 K value. Lines are guide to the eye. Inset: First derivative of

magnetization. ................................................................................................................... 41

Figure 4.4: (a) Comparison of temperature-dependent entropy change in bulk and thin-

film La0.7Ca0.3MnO3 samples under an applied field change of 5T. (b) The refrigerant

capacity as a function of applied magnetic field. .............................................................. 42

Figure 4.5: H/M vs M2 for bulk and thin-film La0.7Ca0.3MnO3 ....................................... 43

Figure 4.6: Universal curve calculations as described in the text for the polycrystalline

bulk (a) and thin-film (b) forms of La0.7Ca0.3MnO3. ......................................................... 44

Figure 5.1: Temperature dependence of magnetization taken at 5 kOe. Inset shows the

dependence of the Curie temperature (TC) on the Sr-doped content. The boundary line

between the orthorhombic (Pbnm) and rhombohedral (R3c) phases is taken at x = 0.15.

From [44]. ......................................................................................................................... 52

Figure 5.2: The H/M vs. M2 plots for representative temperatures around the TC for the

La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1 and 0.2) samples. From [44]. ................................ 53

Figure 5.3: Modified Arrott plot isotherms with 1K temperature interval for the

La0.7Ca0.3-xSrxMnO3 (x=0.1, 0.2 and 0.25) samples. From [44] ........................................ 55

Figure 5.4: Temperature dependence of spontaneous Ms (square) and inverse initial

susceptibility 0-1

(circles) for the x=0.2 sample; solid lines are fitting curves to Eqs.

(3.15) and (3.16), respectively. From [44]. ....................................................................... 56

Figure 5.5: The linearity of the M(T = TC) versus H

=H

curves validates the

values of the critical exponents. From [44]. ..................................................................... 57

Figure 5.6: Normalized isotherms of La0.7Ca0.3−xSrxMnO3 (x = 0.1 and 0.2) samples

below and above Curie temperature (TC) using the values of and determined from K-F

method. From [44]. ........................................................................................................... 59

Figure 6.1: Temperature dependence of ZFC and FC magnetizations taken at a field of

0.05 T. ............................................................................................................................... 73

Figure 6.2: Temperature dependence of magnetization taken at different magnetic fields

up to 5T. ............................................................................................................................ 74

Figure 6.3: Isothermal magnetization curves taken at different fixed temperatures

between 65 and 300 K for the Pr0.5Sr0.5MnO3 manganite: (a) around TC and (b) around

TCO. .................................................................................................................................... 75

Page 11: Magnetism in Complex Oxides Probed by Magnetocaloric

vi

Figure 6.4: Temperature dependence of magnetic entropy change (−SM) at different

applied fields up to 5 T. .................................................................................................... 76

Figure 6.5: Magnetic field dependence of RC for the cases around TC and TCO (without

and with subtracted hysteretic losses). The inset shows the magnetic field dependence of

magnetization taken at 150 K (below TCO) and at 180 K (below TC). .............................. 77

Figure 6.6: Determination of the critical exponents n (panel a) and Δ (panel b) after

fitting of δs and εr, respectively, as a power law of the reduced magnetic field H/Hf. .... 80

Figure 6.7: Dimensionless renormalized temperature t* or field h* dependence of the

dimensionless temperature renormalized magnetization mt* (a), of the dimensionless

field renormalized magnetization mh* (b in logarithmic scales) ...................................... 81

Figure 6.8: Inverse of the dimensionless isothermal susceptibility jh* ............................ 82

Figure 6.9: Temperature (a) and field (b) dependence of the magnetization, and

temperature dependence of SM (c). The dark sold lines in (a) and (b) and the solid lines

in (c) are fits to the data via the ANEOS .......................................................................... 84

Figure 6.10: (a) Temperature dependence above TC of the inverse of the experimental

isothermal initial susceptibility 0 (solid symbol *), along with the ones offered by Eq.

(6) (open symbol ○) and Eq. (15) (lines). (b) Temperature dependence above TC of the

critical exponent ( )ef T (solid symbol) and the value of ·=1.440 presented in the sample

when the FM clusters and their interactions have disappeared above 300 K (dashed line).

........................................................................................................................................... 86

Figure 6.11: An example of unipolar transverse susceptibility scan of Pr0.5Sr0.5MnO3 (a).

3-D Unipolar scans of transverse susceptibility as a function of magnetic field and

temperature (b). ................................................................................................................. 87

Figure 6.12: Temperature dependence of effective anisotropy field (HK), switching field

(HS), and peak height of transverse susceptibility curves ([T/T]max) ........................... 90

Figure 6.13: Unipolar transverse susceptibility and first magnetization with respect to

applied magnetic field at T=140K. ................................................................................... 91

Figure 7.1: Zero-field-cooled and field cooled M(T) with 10mT applied field, measured

on warming [24]. ............................................................................................................. 101

Figure 7.2: The M(H) curves for some selected temperatures. The arrows indicate the

way in which the virgin, return, and second magnetization curves were measured [24].

......................................................................................................................................... 103

Figure 7.3: Temperature dependence of magnetic entropy change (−∆SM) for LPCMO

for the magnetic field change of 1.5 T and 6 T, respectively [24] .................................. 104

Page 12: Magnetism in Complex Oxides Probed by Magnetocaloric

vii

Figure 7.4: (a) Magnetic field dependence of maximum magnetic-entropy change

([−SM]max) for LPCMO at 75 K; (b) the magnetic hysteresis loop M(H) measured at 75K

[24]. ................................................................................................................................. 106

Figure 7.5: (a) Magnetic field dependence of maximum magnetic-entropy change

([−SM]max) for LPCMO at 205 K; (b) the magnetic hysteresis loop M(H) measured at

205 K [24]. ...................................................................................................................... 108

Figure 7.6: (a-d): Bipolar TS scans below TC (a) and above TC (b-d). .......................... 110

Figure 7.7: New phase diagrams for LPCMO developed from TS vs HDC (a) Positive and

negative switching field as a function of temperature. (b) Maximum change in TS at

HDC=0 as a function of temperature. ............................................................................... 111

Figure 8.1: Temperature dependence of the magnetization of Pr1−xSrxCoO3 (x=0.3, 0.35,

0.4, and 0.5) compounds when a magnetic field 0H=5 T is applied. Inset: Temperature

dependence of the magnetization at low field (0H=1 mT) and intermediate field

(0H=0.1T) in the x=0.5 sample. .................................................................................... 119

Figure 8.2: Field dependence, from 5 K to 320 K in 5 K increments, of the magnetization

of the polycrystalline Pr0.5Sr0.5CoO3 compounds. The magnetization curve is marked

(open symbol) at the Curie temperature of the sample TC(x=0.5)=230 K. ..................... 120

Figure 8.3: Temperature dependence of the magnetic entropy change for 0H=5 T of the

Pr1−xSrxCoO3 (x = 0.3, 0.35, 0.4, and 0.5) compounds. Solid arrows indicate the

temperatures (TA) of the second phase transition that occurs at low temperature Inset:

Reduced temperature dependence of the magnetic entropy change near TC .................. 121

Figure 8.4: Field dependence of the maximum magnetic entropy change (panel a) and

the refrigerant capacity (panel c) in the studied polycrystalline Pr1−xSrxCoO3 (x=0.3, 0.35,

0.4, and 0.5) compounds. Dimensionless field dependence of the dimensionless

maximum magnetic entropy change s (panel b), and dimensionless refrigerant capacity

rc (panel d). The non-collapse into two master curves indicates that the exponents n1 and

n2 are composition dependent. ........................................................................................ 123

Figure 8.5: Bipolar transverse susceptibility scans of Pr0.5Sr0.5CoO3 as a function of

applied field for 20K (a), 95K (b), 110K (c), and 225K (d). On 8.5(a) the arrows indicate

the sequence of measurement and the anisotropy (Hk), crossover (Hcr), and switching (HS)

peaks are labeled [22]. .................................................................................................... 125

Figure 8.6: Unipolar transverse susceptibility scans for several different temperature

plotted on two plots depicting the two different ferromagnetic phases ((a) is FM1 and (b)

is FM2). The signal intensity appears in arbitrary units as soon of the curves have been

shifted upward or downward for clarity [22]. ................................................................. 126

Page 13: Magnetism in Complex Oxides Probed by Magnetocaloric

viii

Figure 8.7: Temperature dependence of the peaks positions in the transverse

susceptibility measurement. (a) Anisotropy field (+HK) (b) Switching field (HS) (c)

Crossover field (Hcr) [22]... ............................................................................................. 128

Figure 9.1: Temperature dependence of (a) dc-magnetization at an applied field of 100

Oe (b) ac susceptibility with small ac magnetic field ~10 Oe at a variety of frequencies.

Also, magnetic field dependence of magnetization at 25K (c) and 5K (d). .................... 136

Figure 9.2: Isothermal magnetization vs applied field, for a temperature range of 120K-

5K with a temperature interval of 5K, and magnetic field from 0-7T. ........................... 137

Figure 9.3: Change in magnetic entropy as a function of temperature, calculated using

the thermodynamic Maxwell relation (Eqn. 3.9). ........................................................... 138

Figure 9.4: Change in magnetic entropy as a function of applied field at various constant

temperatures, (a) 30K, (b) 20K, (c) 15K and (d) 10K. ................................................... 140

Figure 9.5: (a) First derivative of the field dependent change in entropy. (b) Magnetic

phase diagram derived from the field and temperature dependent change in magnetic

entropy. ........................................................................................................................... 141

Page 14: Magnetism in Complex Oxides Probed by Magnetocaloric

9

ABSTRACT

Magnetic oxides exhibit rich complexity in their fundamental physical properties

determined by the intricate interplay between structural, electronic and magnetic degrees

of freedom. The common themes that are often present in these systems are the phase

coexistence, strong magnetostructural coupling, and possible spin frustration induced by

lattice geometry. While a complete understanding of the ground state magnetic properties

and cooperative phenomena in this class of compounds is key to manipulating their

functionality for applications, it remains among the most challenging problems facing

condensed-matter physics today. To address these outstanding issues, it is essential to

employ experimental methods that allow for detailed investigations of the temperature

and magnetic field response of the different phases.

In this PhD dissertation, I will demonstrate the relatively unconventional

experimental methods of magnetocaloric effect (MCE) and radio-frequency transverse

susceptibility (TS) as powerful probes of multiple magnetic transitions, glassy

phenomena, and ground state magnetic properties in a large class of complex magnetic

oxides, including La0.7Ca0.3-xSrxMnO3 (x = 0, 0.05, 0.1, 0.2 and 0.25), Pr0.5Sr0.5MnO3, Pr1-

xSrxCoO3 (x = 0.3, 0.35, 0.4 and 0.5), La5/8−xPrxCa3/8MnO3 (x = 0.275 and 0.375), and

Ca3Co2O6.

First, the influences of strain and grain boundaries, via chemical substitution and

reduced dimensionality, were studied via MCE in La0.7Ca0.3-xSrxMnO3. Polycrystalline,

single crystalline, and thin-film La0.7Ca0.3-xSrxMnO3 samples show a paramagnetic to

Page 15: Magnetism in Complex Oxides Probed by Magnetocaloric

10

ferromagnetic transition at a wide variety of temperatures as well as an observed change

in the fundamental nature of the transition (i.e. first-order magnetic transition to second

order magnetic transition) that is dependent on the chemical concentration and

dimensionality.

Systematic TS and MCE experiments on Pr0.5Sr0.5MnO3 and Pr0.5Sr0.5CoO3 have

uncovered the different nature of low-temperature magnetic phases and demonstrate the

importance of coupled structural/magnetocrystalline anisotropy in these half-doped

perovskite systems. These findings point to the existence of a distinct class of phenomena

in transition-metal oxide materials due to the unique interplay between structure and

magnetic anisotropy, and provide evidence for the interplay of spin and orbital order as

the origin of intrinsic phase separation in manganites.

While Pr0.5Sr0.5MnO3 provides important insights into the influence of first- and

second-order transitions on the MCE and refrigerant capacity (RC) in a single material,

giving a good guidance on the development of magnetocaloric materials for active

magnetic refrigeration, Pr1-xSrxCoO3 provides an excellent system for determining the

structural entropy change and its contribution to the MCE in magnetocaloric materials.

We have demonstrated that the structural entropy contributes significantly to the total

entropy change and the structurally coupled magnetocrystalline anisotropy plays a crucial

role in tailoring the magnetocaloric properties for active magnetic refrigeration

technology.

In the case of La5/8−xPrxCa3/8MnO3, whose bulk form is comprised of micron-sized

regions of ferromagnetic (FM), paramagnetic (PM), and charge-ordered (CO) phases, TS

and MCE experiments have evidenced the dominance of low-temperature FM and high-

Page 16: Magnetism in Complex Oxides Probed by Magnetocaloric

11

temperature CO phases. The “dynamic” strain liquid state is strongly dependent on

magnetic field, while the “frozen” strain-glass state is almost magnetic field independent.

The sharp changes in the magnetization, electrical resistivity, and magnetic entropy just

below the Curie temperature occur via the growth of FM domains already present in the

material, even in zero magnetic field. The subtle balance of coexisting phases and kinetic

arrest are also probed by MCE and TS experiments, leading to a new and more

comprehensive magnetic phase diagram.

A geometrically frustrated spin chain compound Ca3Co2O6 provides an interesting

case study for understanding the cooperative phenomena of low-dimensional magnetism

and topological magnetic frustration in a single material. Our MCE studies have yielded

new insights into the nature of switching between multi-states and competing interactions

within spin chains and between them, leading to a more comprehensive magnetic phase

diagram.

Page 17: Magnetism in Complex Oxides Probed by Magnetocaloric

1

CHAPTER 1.

INTRODUCTION

1.1 Overview and Motivation

Materials that exhibit a coupling among magnetic, structural and electronic degrees

of freedom have been the center of focus for many years now from, an applications point

of view, as well as for a fundamental understanding. Research of complex oxide

materials exploded in 1986 with the discovery of the high critical temperature (high-TC)

cuprate superconductors [1]. Since then a whole host of oxide compounds have been

discovered with a seemingly endless number of interesting phenomena. These

phenomena range from all types of magnetism, i.e. ferromagnetism (FM), ferrimagnetism

(FIM), antiferromagnetism (AFM), paramagnetism (PM), canted-AFM etc. [2-10].

Electronic properties such as ferroelectricity, charge/orbital ordering, metallic, insulating,

semiconducting, superconductivity, spin-state transitions etc. all occur in a wide variety

of structures, which can cause localized lattice distortions leading to phase-separation and

geometrical magnetic frustration among other phenomena [3]. All of these properties are

greatly influenced by external fields (such as magnetic, electric, temperature and external

pressure), as well as internal forces induced by chemical doping [2, 5-8, 10].

Interestingly, most of the properties listed above can exist in a single material when

there is a strong coupling amongst all of the degrees of freedom, leading to complex

Page 18: Magnetism in Complex Oxides Probed by Magnetocaloric

2

phase diagrams. All of the aforementioned properties in oxides have been studied in great

detail; however, there is still not a generalized theory on the mechanism causing such

complex behavior. For example, high-TC superconductivity and colossal

magnetoresistance have been under debate for more than 20 years, now.

The properties of complex oxides can be realized for a wide variety of potential

applications, ranging from spintronic devices, sensors, MOSFETS (metal-oxide-

semiconductor field-effect transistors), magnetic refrigeration, etc. However wide the

applications are for this class of materials, one could imagine a seemingly endless

number of applications, if the nature of the materials was fully understood.

1.2 Objectives of the Dissertation

The overall objectives of this dissertation are to explore the fundamental nature of

magnetic phase transitions in complex correlated electron oxides with the magnetocaloric

effect (MCE) and transverse susceptibility (TS) probes.

Specific objectives are:

1. To implement MCE and TS as fundamental probes for magnetic, electrical and

structural properties of complex oxide materials synthesized in the Functional

Materials Laboratory at the University of South Florida or acquired from our

collaborators.

2. To systematically investigate, by chemical doping or size reduction, the influence

of first/second- order magnetic phase transitions on MCE.

3. To investigate complex phenomena related to the coupling of the crystalline

structure and magnetocrystalline anisotropy.

Page 19: Magnetism in Complex Oxides Probed by Magnetocaloric

3

4. To use MCE and TS probes to study materials that exhibit multiple magnetic and

electronic phenomena simultaneously, known as phase-separation.

5. To investigate geometrically-induced magnetic frustration and related

phenomena, using MCE, in Ca3Co2O6.

1.3 Outline of the Dissertation

The present dissertation aims to provide a comprehensive understanding of MCE

and TS effects on complex magnetic oxide materials, as well as provide a detailed

description of the fundamental nature of magnetic phase transitions in these materials.

For these reasons, this dissertation will be presented in the following chapters:

Chapter 1 gives an overall overview of the motivation of this Ph.D. research

work.

Chapter 2 discusses the fundamental properties, currently known, in doped R1-

xMxMnO3 (where R is a trivalent rare-earth, M is a divalent alkaline earth metal)

manganite compounds.

Chapter 3 gives an overview of the experimental methods used throughout this

dissertation, namely, MCE and TS.

Chapter 4 shows the influence of reduced dimensions on the magnetocaloric

properties of La0.7Ca0.3MnO3 bulk polycrystalline and thin-film samples.

Chapter 5 provides a detailed description of the effects of Sr doping in the

La0.7Ca0.3-xSrxMnO3 manganite on the magnetic phase transitions, as well as the critical

phenomena.

Page 20: Magnetism in Complex Oxides Probed by Magnetocaloric

4

Chapter 6 discusses the influence of a first/second- order magnetic transitions on

MCE, as well as the influence of a structural transition on the magnetic properties in

Pr0.5Sr0.5MnO3, probed by TS.

Chapter 7 presents a systematic study on the subtle balance between multiple

phases coexisting in the same temperature regime in phase-separated La5/8−xPrxCa3/8MnO3

(x = 0.275 and 0.375) single crystals.

Chapter 8 demonstrates very strong coupling between structure and

magnetocrystalline anisotropy and its impact on the MCE in Pr1-xSrxCoO3 (x = 0.3, 0.35,

0.4, and 0.5).

Chapter 9 demonstrates the overall usefulness of the MCE as a fundamental

probe, when a new phase diagram is presented in the very complicated, geometrically

frustrated, Ca3Co2O6 single-crystal.

Chapter 10 summarizes all of the important results presented throughout the

dissertation, as well as various plans for future implementation of the MCE and TS

probes.

References

[1] J.G. Bednorz, K.A. Muller, Possible High-Tc Superconductivity in the Ba-La-Cu-O

System, Z Phys B Con Mat, 64 (1986) 189-193.

[2] E. Dagotto, Open questions in CMR manganites, relevance of clustered states and

analogies with other compounds including the cuprates, New J Phys, 7 (2005).

[3] E. Dagotto, Complexity in strongly correlated electronic systems, Science, 309 (2005)

257-262.

Page 21: Magnetism in Complex Oxides Probed by Magnetocaloric

5

[4] M.H. Phan, N.A. Frey, M. Angst, J. de Groot, B.C. Sales, D.G. Mandrus, H. Srikanth,

Complex magnetic phases in LuFe2O4, Solid State Commun, 150 (2010) 341-345.

[5] M.H. Phan, S.C. Yu, Review of the magnetocaloric effect in manganite materials, J

Magn Magn Mater, 308 (2007) 325-340.

[6] C.N.R. Rao, B. Raveau, Colossal magnetoresistance, charge ordering and related

properties of manganese oxides, World Scientific, Singapore ; River Edge, N.J., 1998.

[7] P.A. Sharma, S. El-Khatib, I. Mihut, J.B. Betts, A. Migliori, S.B. Kim, S. Guha, S.W.

Cheong, Phase-segregated glass formation linked to freezing of structural interface

motion, Phys Rev B, 78 (2008).

[8] P.A. Sharma, S.B. Kim, T.Y. Koo, S. Guha, S.W. Cheong, Reentrant charge ordering

transition in the manganites as experimental evidence for a strain glass, Phys Rev B, 71

(2005).

[9] Y. Tokura, N. Nagaosa, Orbital physics in transition-metal oxides, Science, 288

(2000) 462-468.

[10] Y. Tokura, Y. Tomioka, Colossal magnetoresistive manganites, J Magn Magn

Mater, 200 (1999) 1-23.

Page 22: Magnetism in Complex Oxides Probed by Magnetocaloric

6

CHAPTER 2.

FUNDAMENTAL ASPECTS OF MANGANITES

The discovery of colossal magnetoresistance (CMR) in doped manganites with

the general formula R1−xMxMnO3 (R = La, Pr, Nd, Sm, and M = Sr, Ca, Ba, and Pb) has

stimulated intense research into their physical properties [1]. The relationship between

the ferromagnetism and conductivity (e.g. the relationship between the metal–insulator

(MI) transition and the ferromagnetic-paramagnetic (FM-PM) transition) in several CMR

materials has continued to generate interest and reveal new insights, primarily due to the

complexity of the systems [2-7].

It has been experimentally shown that while the parent compound RMnO3 is an

insulating antiferromagnet, substitution of the trivalent R3+

ion by a divalent M2+

ion

leads to coexistence of Mn3+

and Mn4+

ions and, at sufficiently high doping levels (x), the

material becomes a conducting ferromagnet [1]. The metallic ferromagnetic state in

doped manganites was widely interpreted using the double-exchange (DE) mechanism

[8]. According to this model, the transfer of an itinerant eg electron between the

neighboring Mn sites (local t2g spins) via the O2−

ion results in a ferromagnetic interaction

due to the on-site Hund’s rule coupling. This model is effective for certain types of

manganites, but, it leaves a lot to be desired. In particular, the understanding of the

complex nature of phase transitions, phase coexistence and separation, and the

magnetostructural coupling phenomena in several mixed-valent manganite systems has

Page 23: Magnetism in Complex Oxides Probed by Magnetocaloric

7

remained elusive. By using the magnetocaloric effect (MCE) and transverse

susceptibility (TS) measurements, we hope to gain deeper physical insights into the

ground state magnetic properties and magnetic field-induced phenomena in these

complex materials.

2.1 Crystal Structure

R1−xMxMnO3 belongs to the general perovskite family, ABO3. The Mn ions

occupy the B-site at the center of the unit cell, while oxygen ions coordinate around the

Mn ion, forming MnO6 octahedra. The R3+

and M2+

ions are then distributed randomly

over the A-sites in the crystal. The high temperature (~1000 K) cubic phase can be seen

in figure 2.1.

Figure 2.1: Idealized cubic perovskites phase: Mn3+

/4+

(green) occupy the center of the

cube, oxygen ions (red) form octahedra around the Mn3+

/4+

ions. And the trivalent rare

earth or divalent alkali earth (blue) forms the corners of the cube.

At lower temperatures, the MnO6

octahedra become distorted, thereby reducing

the symmetry of the structure. This distortion from the idealized cubic structure can be

simply parameterized via the Goldschmidt tolerance factor (t):

Page 24: Magnetism in Complex Oxides Probed by Magnetocaloric

8

⟨ ⟩

√ ( ),

(2.1)

where ⟨ ⟩ (1.21 Ǻ) and (Mn3+ = 0.58 Ǻ, Mn

4+=0.53 Ǻ) [9], are the

average radii of the A-site cation, oxygen, and manganese ions, respectively. For the

idealized cubic case presented in Figure 2.1, t = 1, however, stable values range from 0.8

< t < 1.

Cubic: t = 1

Rhombohedral: 0.96 ≤ t <1

Orthorhombic: t < 0.96

Figure 2.2: The Pnma unit cell of La1-xCaxMnO3 (taken from [8]) showing general

distortions of the standard cubic lattice. The ions are represented by black (manganese),

grey (La or Ca) and white (oxygen) spheres.

Generally, the size of the A-site cations differs greatly from the idealized cubic

model (a linear variation from 1.32Ǻ to 1.25Ǻ for x = 0 and x = 1, respectively), this will

lead the MnO6 octahedra to rotate or distort to improve packing. This distortion in the

Page 25: Magnetism in Complex Oxides Probed by Magnetocaloric

9

MnO6 leads this class of perovskites to have a true orthorhombic Pnma unit cell, as

shown in figure 2.2.

Electrical conduction and magnetic ground states are greatly influenced by the

distortion of the MnO6 octahedra. A conduction band is formed in the manganites via the

overlap of Mn 3d and O 2p orbitals. Therefore, by distorting the MnO6 octahedra, there

will be lengthening or bending of the Mn-O-Mn bond, thereby reducing orbital overlap

and hopping amplitude of the itinerant eg electrons. Further evidence for the importance

of chemical doping will be discussed throughout this thesis.

2.2 Crystal field splitting and the Jahn-Teller effect

In the idealized cubic perovskite structure, the five-fold degeneracy of the Mn 3d

orbitals is split into two levels (t2g triplet and eg doublet) by the octahedral crystal field.

The electron state will be defined from Hund’s rules:

1. Maximize spin angular momentum |S|

2. Maximize orbital angular momentum |L|

3. If an orbital is less than half-filled, the total angular momentum (J) is

J=|L-S|. If the orbital is more than half-filled, J=|L+S|. If the orbital is

half-filled then L=0 via rule 1 and the Pauli Exclusion Principle.

Therefore the three 3d electrons in Mn4+

are strongly spin aligned in the lower t2g

level, forming a “core” spin S=3/2. Meanwhile, Mn3+

has an extra electron, which, due to

Pauli Exclusion Principle and strong Coulomb repulsion, resides in the eg level. Due to

Hund’s coupling the eg electron is spin coupled to the “core” S=3/2 electrons.

Page 26: Magnetism in Complex Oxides Probed by Magnetocaloric

10

Figure 2.3: Crystal field splits degenerate 3d orbitals into eg and t2g levels, then

crystal distortion due to the Jahn-Teller effect, further splits the degenerate eg and t2g

levels (taken from [10]).

The Jahn-Teller (JT) theorem [10] states that degenerate orbital ground state

levels are generally unstable, which requires decreased symmetry to lift the degeneracy.

Therefore, the MnO6 octahedra spontaneously distorts, further splitting the eg and t2g

energy levels; this distortion is known as the JT effect. A schematic is shown in figure

2.3. The JT effect is only energetically favorable if either of the t2g

or eg

states is partially

occupied. Therefore, Mn3+

, which has an electronic configuration of 3d4, leads to a single

electron in the eg

level, which leads to a very strong deformation in the MnO6 octahedra

in order to lift the degeneracy.

Electrical conduction and magnetic properties of these manganite materials are

primarily attributed to the Mn-O bonds. In the MnO6 octahedral environment, the Mn t2g

triplet orbitals (3dxy, 3dyz and 3dzx) have very little overlap with the oxygen 2p orbitals

Page 27: Magnetism in Complex Oxides Probed by Magnetocaloric

11

and are strongly localized. The eg orbitals (3d3z2

-r2 and 3dx

2-y

2) on the other hand, are

farther reaching and are oriented towards the oxygen 2p orbitals. The probability of

overlap between the Mn eg and O 2p orbitals is large enough for electron hopping, thus

forming a method of conduction. By shortening the distance between Mn-O, there will be

an increased probability of orbital overlap, thus a larger occurrence of electron hopping,

and in turn enhancement of ferromagnetic properties. The overall length and angle of the

Mn-O-Mn bond depends strongly on the radii of the A-site cations. The A-site cations

will determine the ratio of Mn3+

to Mn4+

ions in the system. For example, a doped

manganite of the form R(1-x)MxMnO3 will exhibit a Mn4+

/ Mn3+

ratio as:

(2.2)

The ideal ratio for a ferromagnetic-metal ground state has been shown to be x=1/3

[1], however, by changing the ratio, one can observe competition between various

magnetic ground states, as well as charge and orbital ordering, which will be discussed

throughout this dissertation. An eg electron can hop between nearest-neighbor Mn ions,

causing JT lattice distortion. The strong coupling between the eg electron and the lattice

distortions is known as a polaron. In the paramagnetic state, polarons are free to move

about the lattice, thus creating a polaronic liquid [11].

2.3 Magnetic Interactions

The A-site (A = R, M) doping has been shown to control an effective one electron

bandwidth (W), which primarily governs the magnetic and magnetotransport properties

of a material [1-7]. Manganites are generally separated into three classes, namely large,

intermediate and low bandwidth manganites. Larger W, leads to a larger probability of

Page 28: Magnetism in Complex Oxides Probed by Magnetocaloric

12

electron hopping, hence large conductivity, and a more stabilized ferromagnetic (FM)

phase. As W is reduced, FM state tends be less prominent and there is a larger probability

of phase separation. W can be tuned by applying external pressure, internal strain due to

chemical doping, etc., thus creating a wide variety of magnetic and structural ground

states in which to investigate.

2.3.1 Double Exchange

Double exchange (DE), first coined by Zener [12], involves the simultaneous

transfer of one electron from the Mn3+

eg orbital to an oxygen 2p orbital, and from that

oxygen 2p orbital to a neighboring Mn4+

eg orbital, as shown in figure 2.4.

Figure 2.4: A schematic representation of the double exchange mechanism showing

the simultaneous transfer of electrons between Mn3+

to O2-

and from the O2-

to Mn4+

ions taken from [14].

The hopping of the itinerate electrons via DE leads to the main conduction channel

in the manganites. The two electrons involved in the DE process must have the same spin

i.e. ferromagnetic coupling, due to Hund’s coupling and the Pauli Exclusion Principle.

The strength of the DE mechanism is defined by the charge transfer integral [13]:

( ) (2.3)

where is the relative angle between local spins. As a magnetic field is applied to the

material, it will force the local t2g spins to align thus reducing spin scattering (i.e.

Page 29: Magnetism in Complex Oxides Probed by Magnetocaloric

13

resistivity decreases) and enhancing the ferromagnetic phase. The DE theory has been

shown to accurately describe the properties of the metallic ferromagnetic state in doped

manganites with relatively large W, such as La0.7Sr0.3MnO3 [13]. However, it alone

cannot explain the features of the MI transition and CMR observed in manganites with

narrow W such as La0.7Ca0.3MnO3 [3],

where other effects such as collective JT

distortions and antiferromagnetic (AFM) interactions coexist and strongly compete with

the ferromagnetic phase.

2.3.2 Superexchange

In general, exchange interactions between magnetic ions are short-range

interactions, occurring only when electrons’ spatial wave functions overlap. The range of

these interactions can be slightly extended by the transfer of electrons that do not

contribute to the overall magnetic behavior; this is called superexchange (SE). In the SE

model, magnetic interactions between adjacent ions are mediated by an intermediate non-

magnetic ion. This is a common interaction in insulating oxides, such as the manganites.

If partially-filled and fully-filled orbitals on nearest-neighboring ions point towards each

other, then an electron can be shared between the two ions. For the case of the

manganites, if a vacancy in the eg orbital and the fully occupied O 2p orbital point

towards each other, an electron from the O 2p will be shared between the two ions, as in

the schematic shown in figure 2.5.

SE interactions can produce ferromagnetic or antiferromagnetic states, depending

on the occupancy of Mn orbitals. There are two possible options for SE to occur in

manganites, namely Mn3+

-O2-

-Mn3+

(figure 2.5(a)) and Mn4+

-O2-

-Mn4+

(figure 2.5(b)).

For the Mn3+

(figure 2.5(a)) case, there is an extra electron in the eg state which is spin-

Page 30: Magnetism in Complex Oxides Probed by Magnetocaloric

14

aligned to the t2g electrons via Hund’s rules, so the “shared” spin with the oxygen will be

anti-aligned to the eg electron due to the Pauli Exclusion Principle. However, for the

Mn4+

case (figure 2.5(b)), the t2g electrons will align parallel to the O 2p electron. Unlike

the DE mechanism discussed previously, SE will always result in an insulating state.

Figure 2.5: Schematic representation showing the arrangement of spins and orbitals

in superexchange taken from [14].

2.3.3 Charge and Orbital Ordering

In A-site doped manganites, the substitution of a trivalent rare-earth ion with a

divalent alkaline earth metal introduces excess holes (or electrons) into the material. At

relatively high temperatures (>300 K) the excess holes (or electrons) are largely

distributed randomly throughout the crystal, however, as the material is cooled, the

excess holes (or electrons) may arrange in a periodic lattice due to repulsive Coulomb

interactions between them; this ordered state is known as charge order (CO). The pattern

of the ordering can be drastically influenced by the choice of doping, as well as the

doping concentration. The CO state is often accompanied by slight localized lattice

distortions, thereby making it directly observable, although this observation can be rather

difficult. Generally the CO state can be characterized by large changes in the transport

and magnetic properties due to localized charges disrupting the conduction channel

Page 31: Magnetism in Complex Oxides Probed by Magnetocaloric

15

leading to the CO state being associated with the AFM (or PM) insulating magnetic state.

Interestingly, regions in the material that are not associated with CO can be ferromagnetic

and metallic in nature, therefore CO manganites can have alternating regions of FM

metallic and AFM insulating regions, which is known as phase separation.

Figure 2.6: CE -type charge ordering along ab-plane in La0.5Sr0.5MnO4 taken from

[http://folk.uio.no/ravi/activity/ordering/chargeordering.html]

Figure 2.6 shows a schematic of CE-type charge order of the half-doped

La0.5Sr0.5MnO4. In this figure, the Mn3+

ions automatically arrange themselves in a

checkerboard type configuration. Another interesting ordering process in the manganites

comes from the lack of spherical symmetry in the Mn 3d orbitals, where the symmetry of

the local environment will determine the most favorable orbital ground-state, therefore

affecting the orientation of the orbitals throughout the crystal. The crystal-field splitting

and JT effect discussed earlier can lead to orbitals aligning into periodic patterns; this

effect is known as orbital ordering (O-O). The O-O can also arrange itself in several

ways; figure 2.7 shows a few orientations, signifying the complex nature of this behavior.

Page 32: Magnetism in Complex Oxides Probed by Magnetocaloric

16

Figure 2.7: Various forms of orbital ordering in the manganites taken from

[http://folk.uio.no/ravi/activity/ordering/orbitalordering.html]

2.4 Conclusions

Manganites exhibit a wide range of magnetic, electronic and structural properties,

all of which depend heavily on the chemical composition. Small changes in the A-site, B-

site or oxygen stoichiometry can lead to extraordinarily altered physical properties,

ranging from a simple shift in magnetic transition temperature to a completely different

structure. This class of materials gives scientists a relatively endless supply of complex

phenomena to investigate.

References

[1] Y. Tokura, Y. Tomioka, Colossal magnetoresistive manganites, J Magn Magn Mater,

200 (1999) 1-23.

[2] P.G. Radaelli, D.E. Cox, M. Marezio, S.W. Cheong, P.E. Schiffer, A.P. Ramirez,

Simultaneous Structural, Magnetic, and Electronic-Transitions in La1-xCaxMnO3 with x =

0.25 and 0.50, Phys Rev Lett, 75 (1995) 4488-4491.

[3] Y. Lyanda-Geller, et. al., Charge transport in manganites: Hopping conduction, the

anomalous Hall effect, and universal scaling, Phys Rev B, 63 (2001).

Page 33: Magnetism in Complex Oxides Probed by Magnetocaloric

17

[4] E. Dagotto, Open questions in CMR manganites, relevance of clustered states and

analogies with other compounds including the cuprates, New J Phys, 7 (2005).

[5] H.Y. Hwang, S.W. Cheong, P.G. Radaelli, M. Marezio, B. Batlogg, Lattice effects on

the magnetoresistance in doped LaMnO3, Phys Rev Lett, 75 (1995) 914-917.

[6] V.B. Shenoy, C.N.R. Rao, Electronic phase separation and other novel phenomena

and properties exhibited by mixed-valent rare-earth manganites and related materials,

Philos T R Soc A, 366 (2008) 63-82.

[7] J. Tao, D. Niebieskikwiat, M. Varela, W. Luo, M.A. Schofield, Y. Zhu, M.B.

Salamon, J.M. Zuo, S.T. Pantelides, S.J. Pennycook, Direct Imaging of Nanoscale Phase

Separation in La0.55Ca0.45MnO3: Relationship to Colossal Magnetoresistance, Phys Rev

Lett, 103 (2009).

[8] W.E. Pickett, D.J. Singh, Electronic structure and half-metallic transport in the La1-

xCaxMnO3 system, Phys Rev B, 53 (1996) 1146-1160.

[9] R.D. Shannon, Revised Effective Ionic-Radii and Systematic Studies of Interatomic

Distances in Halides and Chalcogenides, Acta Crystallogr A, 32 (1976) 751-767.

[10] H.A. Jahn, E. Teller, Stability of polyatomic molecules in degenerate electronic

states. I. Orbital degeneracy, Proc R Soc Lon Ser-A, 161 (1937) 220-235.

[11] M. Ziese, C. Srinitiwarawong, Polaronic effects on the resistivity of manganite thin

films, Phys Rev B, 58 (1998) 11519-11525.

[12] C. Zener, Interaction between the D-Shells in the Transition Metals .2.

Ferromagnetic Compounds of Manganese with Perovskite Structure, Phys Rev, 82 (1951)

403-405.

Page 34: Magnetism in Complex Oxides Probed by Magnetocaloric

18

[13] P.W. Anderson, H. Hasegawa, Considerations on Double Exchange, Phys Rev, 100

(1955) 675-681.

[14] James Christopher Chapman (2005) Phase Coexistence in Manganites (Doctoral

Dissertation)

Page 35: Magnetism in Complex Oxides Probed by Magnetocaloric

19

CHAPTER 3.

EXPERIMENTAL METHODS

3.1 Magnetocaloric Effect

3.1.1 What is the Magnetocaloric Effect?

The magnetocaloric effect (MCE) describes the isothermal change in entropy (or

adiabatic change in temperature) of a magnetic material through the application and

removal of an external magnetic field. MCE is best known for its contribution to industry

via magnetic refrigeration (MR)[1].

Figure 3.1 is a schematic representation of an MR; the general operation works as

a magnetic material, initially demagnetized, is at a temperature T. The material is then

adiabatically magnetized, thus decreasing the magnetic entropy of the material, which in

turn increases the entropy of the lattice, therefore leading to an increase in the material’s

temperature (T+T). Then, the small increase in temperature (T) is removed, bringing

the temperature of the material back to its original state (T). Finally, the material is

demagnetized, therefore increasing the magnetic entropy, which decreases the lattice

entropy, and the material’s temperature. Vapor-compression refrigerators are the most

widely used thus far; they are based on the compression and expansion of greenhouse

gases, such as chlorofluorocarbons (CFCs) and hydrochlorofluorocarbons (HCFCs). This

refrigeration process is not very efficient, with an efficiency of just 5–10% of the ideal

Page 36: Magnetism in Complex Oxides Probed by Magnetocaloric

20

Carnot cycle as opposed 30–60% for MR. Refrigeration accounts for 25% of residential

and 15% of commercial power consumption; therefore, new technology is needed to

reduce these numbers.

Figure 3.1: Schematic of a working magnetic refrigerator. Image credit Tegus et. al.

Nature 415 (2002).

The concept of MCE itself is very old with magnetic cooling for producing ultra-low

temperatures dating back to the 1920’s. However, recently, complex oxides have shown

promise in MR due to large MCE and tunable phase transitions.

There has been a lot of work on MCE-based MR as an alternative to conventional

gas compression (CGC) methods. The MR technology has several advantages over the

CGC technology, such as:

Magnetic refrigerators can be more compactly built when using solid

substances as working materials.

MR does not use ozone-depleting or global-warming gases, and therefore is

considered an environmentally friendly cooling technology.

Page 37: Magnetism in Complex Oxides Probed by Magnetocaloric

21

MR has found wide applications in energy-intensive industrial and

commercial refrigerators such as large-scale air conditioners, heat pumps,

supermarket refrigeration units, waste separation, chemical processing, gas

liquefaction, liquor distilling, sugar refining, grain drying, and so forth.

Even though MCE is known, primarily, as an application-based tool, it also can be

utilized as a particularly good probe for studying fundamentals of magnetic, structural

and electronic phase transitions in magnetic materials. For example, doped manganites

with a general formula of R1−xMxMnO3 (R=La, Pr, Nd, etc., and M=Sr, Ca, Ba, etc.)

exhibit a rich variety of phenomena such as colossal magnetoresistance [2] and large

MCE [3]. From the nature of the measurement, MCE can probe into the coupling of

magnetic- and temperature-induced transitions, which is of utmost importance in these

types of materials. Manganites and the MCE measurement process will be discussed in

further detail in subsequent sections.

3.1.2 Theoretical Aspects of MCE

The thermodynamics of a magnetic material in an applied magnetic field (H), at a

temperature (T) and pressure (p) can be completely described via the Gibbs free energy:

(3.1)

where U is the internal energy, S is the total entropy, V is the volume and M the

magnetization of the material. V, M and S are given by taking the first derivative of G:

( ) (

)

(3.2)

( ) (

)

(3.3)

( ) (

)

(3.4)

Page 38: Magnetism in Complex Oxides Probed by Magnetocaloric

22

The specific heat can then be described as the second derivative of G:

( ) (

)

(3.5)

By the definition of phase transitions, if the first derivative of the Gibbs free energy is

discontinuous, the transition is of the first order. Second order transitions occur when the

second derivative of the Gibbs free energy is discontinuous.

The total change in entropy of a magnetic material can be described by:

( ) ( ) ( ) ( ) (3.6)

where Sl,SM, Se are the lattice, magnetic and electronic entropy, respectively. When a

magnetic material is subjected to a sufficiently high magnetic field, the magnetic

moments of the atoms become reoriented, therefore decreasing the magnetic entropy. If

the magnetic field is applied adiabatically, the system must make up for the decrease in

SM by increasing Sl, thus the temperature of the material rises. By running this process

in reverse, we can achieve a decrease in the sample temperature. This warming and

cooling in response to the application and removal of an external magnetic field is called

the MCE. Since entropy is a state function, the total differential can be expressed as:

(

) (

)

(

)

(3.7)

Isobaric (dp=0) and isothermal (dT=0) measurements will lead to a measurement of the

magnetic entropy change alone. The change of dS of a magnetic material upon the

application of an H is related to M with respect to T through the thermodynamic Maxwell

relation:

(

) (

)

(3.8)

Page 39: Magnetism in Complex Oxides Probed by Magnetocaloric

23

The magnetic entropy change, ΔSM (T,H), is then calculated by:

( ) ( ) ( ) ∫ (

)

(3.9)

It is interesting to note that since ΔSM (T,H) depends directly on the derivative of the

M(T), this measurement is inherently more sensitive for studying phase transitions than

standard magnetometry. For magnetization measurements made at discrete field and

temperature intervals, ΔSM (T,H) can be approximately calculated by the following

expression:

( ) ∑

( ) ( )

(3.10)

Alternatively, ΔSM (T,H) can be obtained from calorimetric measurements of the field

dependence of the heat capacity and subsequent integration:

( ) ∫

( ) ( )

(3.11)

where C(T,H) and C(T,0) are the values of the heat capacity measured in a field, H, and

in zero field (H = 0), respectively. Therefore, the adiabatic temperature change (ΔTad) can

be evaluated by integrating Eq. (3.11) over the magnetic field, which is given by:

(

)

(3.12)

The refrigerant capacity (RC), is defined as the heat transferred from the cold end (T1) to

the hot end (T2) of an ideal thermodynamic refrigeration process. The RC of a

magnetocaloric material can be, in simple cases, evaluated by considering the magnitude

of ΔSM and its full-width at half maximum (δTFWHM):

( ) ∫ ( )

(3.13)

Page 40: Magnetism in Complex Oxides Probed by Magnetocaloric

24

Figure 3.2 shows the method of calculating RC from the -SM curve. From this

figure the RC corresponds to the hatched area under the curve. Immediately following

from equations (3.8–3.11), materials whose total entropy is strongly influenced by a

magnetic field and whose magnetization vary rapidly with temperature are expected to

exhibit an enhanced MCE, i.e. around a magnetic phase transition temperature such as TC

or TN in a conventional ferromagnet or antiferromagnet respectively.

Figure 3.2: The method for calculating the RC from the −SM(T) curve using Eq. (3.13)

for two types of transitions in Pr0.5Sr0.5MnO3.

However, according to the Ehrenfest classification of phase transitions [4], since a

first-order magnetic transition (FOMT) is accompanied by a discontinuity in the first

derivative of the free energy (SM or M), care must be used when calculating SM. For

this kind of phase transition, the ΔSM values obtained from Eq. (3.9) are often much

higher than the ones obtained from pure magnetic contribution alone [5, 6].

3.2 MCE as a Fundamental Research Tool

From equation (3.9) we note that SM(T) is directly related to the first derivative

of magnetization with respect to temperature (M/T), so the MCE is expected to be

Page 41: Magnetism in Complex Oxides Probed by Magnetocaloric

25

inherently more sensitive for probing magnetic transitions than conventional

magnetization and resistivity measurements. A very small change in M can give rise to a

more pronounced effect in SM(T) than in M(T) or resistivity measurements. More

importantly, the sign of SM, which is determined by the slope change of the dM/dT

curve, can allow probing the magnetic transitions further to better understand the nature

of the different phases in a material with a rich and complex H-T magnetic phase diagram

[10-13]. Following the convention in MCE analysis, the value of -SM is positive for

materials exhibiting an FM transition, because of the fully magnetically ordered

configuration with the application of an external magnetic field [9, 10]. Meanwhile,

negative values of -SM are found in AFM ordering systems due to orientational disorder

of the magnetic sublattice anti-parallel to the applied magnetic field [11, 12]. An

excellent example of this is the Pr0.5Sr0.5MnO3 system as shown in Figure 3.2, where the

material undergoes a transition from the paramagnetic to ferromagnetic state and a

second transition from the ferromagnetic to antiferromagnetic phase. Recently, von

Ranke et al. [13] have theoretically investigated the implications of positive and negative

MCE in antiferromagnetic and ferromagnetic arrangements. In this dissertation, I will

demonstrate the usefulness of the MCE method for probing the nature of phase

transitions, phase coexistence, magnetic ground states, and the subtle balance of

competing phases in phase separated manganite systems such as Pr0.5Sr0.5MnO3, La5/8-

xPrxCa3/8MnO3 (x = 0.275 and 0.375) and half-doped cobaltites Pr0.5Sr0.5CoO3.

3.2.1 Order of Transitions and Critical Phenomena

Understanding the nature of phase transitions and critical magnetic behaviors near

these transitions is essential to access the underlying origins of the magnetic field induced

Page 42: Magnetism in Complex Oxides Probed by Magnetocaloric

26

phenomena such as colossal magnetoresistance (CMR) and the magnetocaloric effect

(MCE). Below, we show the methods used for the analysis of second-order magnetic

transitions and critical exponents of the material systems studied in this PhD research

work.

It has been shown that for materials that exhibit a second-order magnetic

transition (SOMT), the Kouvel–Fisher (K–F) method [14] can be used to precisely

determine the critical exponents of these samples. This method consists of an iterative

procedure which starts by constructing the Arrott–Noakes (A–N) plot (i.e. the plot of M2.5

vs. (H/M)0.75

). From it, the values for M0(T) are computed from the intercepts of various

isothermal magnetization vs. field curves on the ordinate of the plot (for temperatures

below TC). The intercept on the abscissa (for temperatures above TC), allows calculating

0(T). Once the M0(T) and 0(T) curves have been constructed, two additional parameter

data sets, X(T) and Y(T), may be determined:

( )

(

)

(3.15)

( ) (

)

(3.16)

In the critical region, both X(T) and Y(T) should be linear, with slopes that give

the values of the critical exponents, and intercepts of the temperature axis, which

correspond to TC. The values of the critical exponents are refined by using an iterative

method: once Eqs. (3.15) and (3.16) produce the values of the critical exponents, a

generalized A-N plot (M1/

vs. (H/M)1/

) is constructed and used to calculate new M0(T)

and 0(T) curves, which are subsequently input into Eqs. (3.15) and (3.16), resulting in

Page 43: Magnetism in Complex Oxides Probed by Magnetocaloric

27

newer values for and . The procedure terminates when the desired convergence of the

parameters is achieved. TC is obtained as the intercept on the abscissa of both X and Y

lines.

According to Widom’s scaling hypothesis [15], for SOMT materials, the

spontaneous magnetization M(,0) below the TC, the critical magnetization M(0,H), the

isothermal initial susceptibility ( ,0),T and the maximum magnetic entropy change

0,MS H possess the following power-law dependences:

0( ,0) ( ) ( 1, ) ;0 0M M M

(3.17)

1/(0, ) ( ) 0 0, ) ;( 1CM H M H M H

(3.18)

0

0

0

;(1,0) 0( ,0) ( )

( 1,0) ; 0T

(3.19)

(0, ) (0,1)M M

nS H S H (3.20)

where /C CT T T is the reduced temperature, M(1,0), M(0,1), 0(1,0) , 0( 1,0) ,

and Δ (0,1)MS are the critical amplitudes, and , , , and n are the critical exponents.

The Widom scaling hypothesis also allows for the determining of the equations of

state of a magnetic system; they can be expressed as [16]:

1/ 1/

( , )( )t

M Hm t

H H

(3.21)

( )h

M Hff m h

(3.22)

Page 44: Magnetism in Complex Oxides Probed by Magnetocaloric

28

where 1/( , ) /tm M H H is the renormalized temperature dependence of the

magnetization, 1//t H is the renormalized temperature, is the temperature scaling

function, /hm M

is the renormalized field dependence of the magnetization,

/h H

is the renormalized field, f+ (for T > TC) and f (for T < TC) are the field

scaling functions, and is the so-called gap exponent. According to Eq.

(3.21), if appropriate values of the critical exponents (, , and ) and TC are used to plot

mt versus t, all experimental data points will collapse onto one universal curve, . On the

other hand, according to Eq. (3.22), all experimental data of the plot of mh versus h will

collapse onto two universal curves: f+ (for T > TC) and f (for T < TC). This is an

important criterion to validate the reliability of the procedure used to obtain the critical

exponents. In addition, from Eq. (3.19) the plot of the renormalized field dependence of

the isothermal susceptibility, ( , ) /Thj H

versus h, will collapse onto two

universal curves ( )hj g h : one above TC (g+) and another below TC (g), and from Eq.

(3.20). The plot of the renormalized temperature dependence of the magnetic entropy

change, Δ ( , ) /M

n

ts S H H versus t, will collapse onto one universal curve st=·(t)

[17-20].

Recently, it has been shown theoretically [19] and experimentally [20] that the

latter universal behavior can be obtained without knowledge of the critical exponents,

and that the magnetic field dependence of ΔSM ·can be represented as [18, 21]:

,Δ , Δ ,1

T H

M M

nS T H S T H (3.23)

Page 45: Magnetism in Complex Oxides Probed by Magnetocaloric

29

where the amplitude, ,1MS T , is T-dependent, and the exponent n generally depends

on temperature and field, taking the asymptotic values n = 1 and n = 2 when the values of

T are quite far below and above TC, respectively. At T = TC (or at the temperature that

ΔSM attains its maximum value Δ M

pkS , T = Tpk), the exponent n is field-independent and

can be expressed in terms of the other critical exponents, as:

( ) ( )

( )

(3.24)

Since only two critical exponents (n and ) are independent, it is then possible to

completely define the critical exponents using a non-iterative method [22]. The exponent

n can be obtained from the fitting of the Δ M

pkS values as a power law of the magnetic

field using Eq. (3.20), while the exponent , according to Eq. (3.21), can be obtained

from the fitting of the reference temperatures (selected as that corresponding to a certain

fraction of ΔSM in the ΔSM vs. T representation) as a power law of the field 1/r H

or, alternatively, according to Eq. (3.22), as the fitting of the reference fields (selected as

that corresponding to a certain fraction of ΔSM in the ΔSM vs. H representation) as a

power law of the temperature rH

.

3.3 Transverse Susceptibility

Resonant based measurements are advantageous when it comes to probing

relatively minor changes in the physical properties of a material. An easy way to perform

resonant measurements on magnetic materials is based on an LC tank circuit, where a

material is placed inside of the capacitor or the inductor as is the case in this study.

Therefore, any change in material properties will induce a change in the capacitance (C)

Page 46: Magnetism in Complex Oxides Probed by Magnetocaloric

30

or inductance (L), which in turn changes the resonant frequency of the circuit. Thus,

measurement of the frequency shift results in a direct correlation in the electronic,

dielectric, or magnetic response of the material [23]. Here we introduce the radio

frequency (RF) transverse susceptibility as a very powerful probe of magnetic anisotropy,

switching fields and spin dynamics in strongly correlated electron magnetic systems such

as complex oxides, as discussed throughout this dissertation.

The transverse susceptibility measurement is an LC tank circuit that is driven at a

constant resonance by supplying the circuit with external power (driven by a tunnel diode

oscillator (TDO)) to compensate for dissipation. This arrangement makes a self-resonant

circuit as the power supplied by the TDO maintains continuous oscillation of the LC tank

circuit operating at a frequency given by the expression [24].

LC

1 (3.25)

Inserting a sample into the inductive coil will produce a small change in the coil

inductance ΔL. If ΔL/L <<1, one can differentiate equation (3.25) and obtain the

expression:

L

L

2

(3.26)

Since ΔL is the related to material properties. For magnetic materials, this is

proportional to the real part (µ′) of the complex permeability.

µ = µ′-iµ″ (3.27)

In this setup, the sample is encapsulated in a gel cap, which can be placed inside

the inductive coil. The entire coil is inserted into the sample chamber of a Physical

Property Measurement System (PPMS) by Quantum Design (Fig. 3.3). The advantage of

Page 47: Magnetism in Complex Oxides Probed by Magnetocaloric

31

inserting the probe into the PPMS is that we gain control over a wide range of

temperatures (1.8K – 350K) and magnetic fields (±7 T). Since the LC circuit is driven at

resonance there is a small oscillating RF field (HRF) produced by the inductor, which is

oriented perpendicularly to the DC field (HDC), produced by the PPMS, therefore giving

us transverse geometry.

Figure 3.3: (a)Schematic diagram of the transverse susceptibility circuit, (b) Schematic

depiction of transverse susceptibility probe , (c) Quantum Design PPMS.

In this configuration, the change in inductance is determined by the change in

transverse permeability, µT, of the sample. The TS ratio can be written as:

ΔχT/χT (%) =

sat

T

sat

TT H

100)( (3.28)

where χTsat

is the TS at the saturating field, Hsat. Since this is a measure of the overall

change in TS, there is no dependence on the geometrical parameters, therefore proving to

be useful for many systems [24].

The theory of reversible TS was first studied theoretically by Aharoni et al. in

1957 [25]. In his work, the expression for TS was analytically derived for a single-

domain particle based on the Stoner-Wolfarth model. When ΔχT/χT is plotted as a

function of HDC, singularities are observed at the anisotropy fields (±HK), as seen in

(a) (b) (c)

Page 48: Magnetism in Complex Oxides Probed by Magnetocaloric

32

figure 3.4. The effective anisotropy constant (K) can be extracted from this through the

relation:

HK = 2K/MS (3.29)

where MS is the saturation magnetization.

Figure 3.4: Transverse and parallel susceptibility (T and P respectively) for single-

domain magnetic particles. Image from reference [20].

Many known coupling phenomena are characterized by an overall increase in

effective anisotropy of the system, with exchange coupling being the best known

example. In these systems, an increase in coercivity is almost always observed [26]. We

have shown how TS can also be used to complement traditional static magnetic

measurements for systems whose effective anisotropy has increased due to the presence

of a phase transition. In the examples outlined in subsequent sections, we explain how the

TS curves, for complicated systems, can probe regions in complex phase diagrams that

general magnetometry cannot.

Page 49: Magnetism in Complex Oxides Probed by Magnetocaloric

33

References

[1] A.M.Tishin and Y.I. Spichkin, The Magnetocaloric Effect and its Applications,

Institute of Physics Publishing, Bristol and Philadelphia, 2003.

[2] Y. Tokura, Y. Tomioka, Colossal magnetoresistive manganites, J Magn Magn Mater,

200 (1999) 1-23.

[3] M.-H. Phan, S.-C. Yu, Review of the magnetocaloric effect in manganite materials, J

Magn Magn Mater, 308 (2007) 325-340.

[4] J.M. Yeomans, Statistical Mechanics of Phase Transitions, Claredon Press, Oxford,

1992.

[5] L. Tocado, E. Palacios, R. Burriel, Entropy determinations and magnetocaloric

parameters in systems with first-order transitions: Study of MnAs, J Appl Phys, 105

(2009).

[6] N.A.de Oliveira and P.J.von Ranke, Physics Reports, 489 (2010).

[7] A. Giguere, M. Foldeaki, B.R. Gopal, R. Chahine, T.K. Bose, A. Frydman, J.A.

Barclay, Direct measurement of the "giant" adiabatic temperature change in Gd5Si2Ge2,

Phys Rev Lett, 83 (1999) 2262-2265.

[8] M. Balli, D. Fruchart, D. Gignoux, R. Zach, The "colossal" magnetocaloric effect in

Mn1-xFexAs: What are we really measuring?, Appl Phys Lett, 95 (2009).

[9] M.H. Phan, S.C. Yu, Review of the magnetocaloric effect in manganite materials, J

Magn Magn Mater, 308 (2007) 325-340.

[10] M.H. Phan, G.T. Woods, A. Chaturvedi, S. Stefanoski, G.S. Nolas, H. Srikanth,

Long-range ferromagnetism and giant magnetocaloric effect in type VIII Eu8Ga16Ge30

clathrates, Appl Phys Lett, 93 (2008).

Page 50: Magnetism in Complex Oxides Probed by Magnetocaloric

34

[11] A. Biswas, et. al., Observation of large low field magnetoresistance and large

magnetocaloric effects in polycrystalline Pr(0.65)(Ca(0.7)Sr(0.3))(0.35)MnO(3), Appl Phys Lett,

92 (2008).

[12] P. Sande, et.al., Large magnetocaloric effect in manganites with charge order, Appl

Phys Lett, 79 (2001) 2040-2042.

[13] P.J. von Ranke, N.A. de Oliveira, B.P. Alho, E.J.R. Plaza, V.S.R. de Sousa, L.

Caron, M.S. Reis, Understanding the inverse magnetocaloric effect in antiferro- and

ferrimagnetic arrangements, J Phys-Condens Mat, 21 (2009).

[14] J.S. Kouvel, M.E. Fisher, Detailed magnetic behavior of nickel near its curie point,

Physical Review a-General Physics, 136 (1964) 1626-&.

[15] B. Widom, Equation of state in neighborhood of critical point, J Chem Phys, 43

(1965) 3898-&.

[16] R.B. Griffiths, Thermodynamic Functions for Fluids and Ferromagnets near the

Critical Point, Phys Rev, 158 (1967) 176-187.

[17] V. Franco, J.S. Blazquez, A. Conde, Field dependence of the magnetocaloric effect

in materials with a second order phase transition: A master curve for the magnetic

entropy change, Appl Phys Lett, 89 (2006).

[18] V. Franco, A. Conde, M.D. Kuz'min, J.M. Romero-Enrique, The magnetocaloric

effect in materials with a second order phase transition: Are T-C and T-peak necessarily

coincident?, J Appl Phys, 105 (2009).

[19] V. Franco, A. Conde, J.M. Romero-Enrique, J.S. Blazquez, A universal curve for the

magnetocaloric effect: an analysis based on scaling relations, J Phys-Condens Mat, 20

(2008).

Page 51: Magnetism in Complex Oxides Probed by Magnetocaloric

35

[20] V. Franco, A. Conde, Scaling laws for the magnetocaloric effect in second order

phase transitions: From physics to applications for the characterization of materials, Int J

Refrig, 33 (2010) 465-473.

[21] T.D. Shen, R.B. Schwarz, J.Y. Coulter, J.D. Thompson, Magnetocaloric effect in

bulk amorphous Pd40Ni22.5Fe17.5P20 alloy, J Appl Phys, 91 (2002) 5240-5245.

[22] V. Franco, A. Conde, V. Provenzano, R.D. Shull, Scaling analysis of the

magnetocaloric effect in Gd5Si2Ge1.9X0.1 (X=Al, Cu, Ga, Mn, Fe, Co), J Magn Magn

Mater, 322 (2010) 218-223.

[23] H. Srikanth, J. Wiggins, H. Rees, Radio-frequency impedance measurements using a

tunnel-diode oscillator technique, Rev Sci Instrum, 70 (1999) 3097-3101.

[24] L. Spinu, C.J. O'Connor, H. Srikanth, Radio frequency probe studies of magnetic

nanostructures, Ieee T Magn, 37 (2001) 2188-2193.

[25] E.H.F. A. Aharoni, S. Shtrikman and D. Treves, Bulletin of the Research Council of

Israel, 6A (1957).

[26] J. Nogues, I.K. Schuller, Exchange bias, J Magn Magn Mater, 192 (1999) 203-232.

Page 52: Magnetism in Complex Oxides Probed by Magnetocaloric

36

CHAPTER 4.

IMPACT OF REDUCED DIMENSIONALITY ON THE MAGNETIC AND

MAGNETOCALORIC RESPONSE OF La0.7Ca0.3MnO3

This chapter presents a systematic investigation of dimensionality effects and

grain boundaries on the magnetic and magnetocaloric properties of La0.7Ca0.3MnO3 by

contrasting the behavior of poly-crystalline and pulsed laser-deposited thin-film forms of

the system. The paramagnetic to ferromagnetic transition and saturation magnetization

were found to broaden and shift to lower temperatures in the thin-film sample. These

factors serve to reduce the magnitude of the magnetocaloric response in the thin-film

samples. However, a large broadening of the magnetic entropy change peak in the thin-

film sample leads to enhanced refrigerant capacity (RC) when compared to the bulk

sample. Universal curves, based on re-scaled entropy change curves, tend toward collapse

with reduced dimensionality, indicating a crossover from a first- to second-order

magnetic transition in the thin-film.

4.1 Introduction

La1-xCaxMnO3 (LCMO) is a canonical example of the mixed-valent perovskite

compounds, exhibiting a rich phase diagram (figure 4.1), which has fueled a great deal of

research into these systems over the past two decades [1]. The average radius of the A-

site ion can lead to various magnetic phenomena, such as canted-antiferromagnetism

(CAF), charge-ordering (CO), ferromagnetic metallic/insulator (FM/FI) and

Page 53: Magnetism in Complex Oxides Probed by Magnetocaloric

37

antiferromagnetism (AFM). In the doping range 0.25 < x < 0.33, bulk LCMO exhibits a

paramagnetic to ferromagnetic transition between 230 K and 260 K, concurrent with the

metal to insulator (MI) transition.

Figure 4.1: Phase diagram of La1-xCaxMnO3, showing the subtle balance between

chemical doping and magnetic properties (taken from [2]).

This critical temperature and doping range is characterized by colossal magnetoresistance

(CMR) and a large magnetocaloric effect (MCE), prompting many studies focused on

understanding the underlying transport and magnetic properties in thin-films [3-8] and

bulk [9, 10] samples of LCMO.

The properties of thin-films as compared to their bulk counterparts vary widely

depending on thickness, substrate, details of deposition, post annealing, oxygen content,

etc. However the general trend upon thickness reduction is a suppression of the Curie

temperature (TC) and saturation magnetization (MS), and an enhancement of resistivity

and magnetoresistance (MR) [7]. In very thin coherently-strained films, disagreement

Page 54: Magnetism in Complex Oxides Probed by Magnetocaloric

38

exists as to whether these effects can be attributed primarily to strain or finite size effects,

[5, 8] due to magnetically dead layers on the surface of the film, as well as at the interface

between the substrate and film. However, in moderately thick films (>100 nm) these

features are likely disorder-induced, as strain relaxation can give rise to extrinsic defects

such as dislocations, stacking faults, cationic vacancies, and grain boundaries.

In this chapter, I will examine the re-structuring of the first-order ferromagnetic

LCMO compound in a thin-film sample and compare the results to a bulk polycrystalline

sample of the same composition. The magnetic and magnetocaloric properties of the

samples are characterized, and it is found that the greater deviation from the bulk

behavior occurs in the thin-film. Universal scaling, based on magnetic entropy change

curves, is applied to understand the impact of reduced dimensionality on the nature of the

ferromagnetic transition in LCMO.

4.2 Experiment

A frequently used method for producing manganite thin-films is pulsed laser

deposition (PLD). A PLD system generally consists of a target holder, a heated substrate,

a vacuum chamber (base pressure 10-6

Torr) and a laser, as seen in the schematic in figure

4.2.

A pulsed laser beam is focused onto a stoichiometric target through a quartz window. The

immense temperature generated from the laser onto the target makes the target molten,

leading to evaporation and ionization of the target. The collection of particles ejected

from the surface, called the plume, is then transported to the substrate. PLD has a few

disadvantages, primarily particulates (up to ~10 m) formed in the plume can be

deposited on to the substrate and the plume is highly directional, therefore, non-uniform

Page 55: Magnetism in Complex Oxides Probed by Magnetocaloric

39

film thickness may occur. Rastering the laser across the target helps to ensure uniformity

of the film, by minimizing the development of large droplets on the substrate’s surface.

The deposition is performed off-axis.

Figure 4.2: Schematic representation of a pulsed laser deposition system. Image credit

Andor Technology.

Substrate temperature, oxygen partial pressure, laser energy density and laser

repetition rate greatly affect the quality of the thin-film. Samples of LCMO in various

forms were prepared at the Naval Research Laboratory (NRL). LCMO thin-films (~150

nm thickness) were deposited on MgO (100) substrates, from a commercially purchased

(Kurt J.Lesker Company) stoichiometric polycrystalline target, using a KrF excimer laser

(Lambda Physik LPX 305, = 248nm, FWHM = 30 ns) that was operated at 5Hz and

focused through a lens with a 50-cm focal length onto a rotating target at a 45o angle of

incidence. The energy density of the laser beam at the target surface was maintained at ~2

J/cm2. The target-to-substrate distance was 5.5 cm.

Films were deposited at a substrate temperature of 777 oC in an oxygen pressure

of 300 mTorr followed by in-situ annealing at 600 oC for 30 min in an oxygen-rich

Page 56: Magnetism in Complex Oxides Probed by Magnetocaloric

40

background. To minimize the effects of variation in stoichiometry and purity, a portion of

the same target was used as the bulk polycrystalline reference sample, while conducting

magnetic measurements. Profilometry confirmed the thickness of the film to be 150 nm.

Magnetic measurements were carried out from 25 K – 350 K under fields up to 5 T. The

diamagnetic contribution from the MgO substrate was corrected by using a linear fitting

and subtraction.

4.3 Results and Discussion

Figure 4.3 shows the temperature dependence of magnetization in a

polycrystalline bulk, and a thin-film of LCMO. Two effects are immediately obvious

from inspection of the results: first, the sharp paramagnetic to ferromagnetic (PM-FM)

transition that occurs in the bulk material is considerably broadened in the thin-film.

Secondly, the TC – estimated here from the minima in the derivative of magnetization

with temperature – shifts to a lower temperature (~235 K) in the thin-film. The

broadening phenomenon can be attributed to the distribution of TC owing to the local

variation of strain near grain boundaries and defects in the thin-film [3, 6]. In a

percolative system, the distribution in transition temperature is Gaussian with a width that

can be qualitatively linked to the disorder present in the system [6]. A Gaussian fit to the

dM/dT curves in the inset of Fig. 4.3 yields distribution widths of Г=5 K and 35 K for the

polycrystalline and thin-film samples, respectively. The large distribution is indicative of

considerable variation in the local structure of the film. The reduction in the average

value of TC has been observed before in both nanocrystalline and thin-film forms of

LCMO, and is usually attributed to finite size effects and disorder [7, 11]. To evaluate the

magnetic entropy change in the system, isothermal magnetization vs. field (M (H)) curves

Page 57: Magnetism in Complex Oxides Probed by Magnetocaloric

41

were recorded around the transition temperature in each sample. The inset of Fig. 4.4

compares the M (H) isotherms at 120 K. The 5 T value of magnetization reached 3.5 µB

and 2.4 µB in the polycrystalline and thin-film samples, respectively.

50 100 150 200 250 300 3500.0

0.2

0.4

0.6

0.8

1.0

150 200 250 300

dM

/dT

T (K)

Poly-crystalline

Thin-Film

M/M

(25

K)

T (K)

Figure 4.3: Temperature dependence of magnetization recorded on cooling in a field of

500 Oe and normalized to 25 K value. Lines are guide to the eye. Inset: First derivative of

magnetization.

The expected spin-only magnetic moment is determined by the Ca2+

-doping-dependent

ratio of Mn4+

(S = 3/2) to high-spin Mn3+

(S = 2), which predicts a value of 3.7 µB in

LCMO. Defects and oxygen off-stoichiometry could reasonably account for the ~5%

discrepancy between the polycrystalline sample and the ideal value.

The entropy change in the system is obtained by integrating between successive

M (H) isotherms according to the thermodynamic Maxwell relation. Under an applied

field change of 5 T, the absolute value of the magnetic entropy change (| |) reaches a

peak value near the TC of 7.7 J/kg K in the polycrystalline sample. This value is in good

agreement with Ref. [12], in which a sol-gel method combined with high temperature

sintering was used to prepare the sample, in contrast to Ref. [10] in which a modified

Page 58: Magnetism in Complex Oxides Probed by Magnetocaloric

42

solid state reaction process gave rise to unusually large values of magnetic entropy

change (9.9 J/kg K for µ0ΔH = 5T).

50 100 150 200 250 3000

2

4

6

8

0 1 2 3 4 5

0

50

100

150

200

250(b)

-

S (

J/k

g K

)

T (K)

0H = 5T

(a) Poly-crystalline

Thin-Film

RC

(J/k

g)

0H (T)

Figure 4.4: (a) Comparison of temperature-dependent entropy change in bulk and thin-

film La0.7Ca0.3MnO3 samples under an applied field change of 5T. (b) The refrigerant

capacity as a function of applied magnetic field.

The decline in the peak | | from the bulk sample to the thin-film sample

(approximately one-third of the bulk value) is consistent with the fall-off in the

magnitude of magnetic moment and the broadening of the FM-PM transition that reduces

⁄ .

Figure 4.4 (b) shows the field-dependence of the refrigerant capacity (RC). The

RC, given by ∫ ( )

, where T1 and T2 are the temperatures defining the full

width at half maximum, is a measure of the heat exchanged between the hot and cold

ends of an ideal refrigeration cycle. The RC is considered to be an important figure of

merit in the evaluation of a magnetocaloric material. It has been observed on a number of

occasions that the reduced maximum value of | | that often accompanies broad

magnetic entropy change peaks can be compensated for by the increased width, resulting

Page 59: Magnetism in Complex Oxides Probed by Magnetocaloric

43

in an enhanced RC over sharper transitions [13]. It can be seen that this scenario holds

true in the present case, where the RC reaches its largest values in the thin-film, for which

the breadth of the transition overcomes the deleterious effects of the drop in .

0 2500 5000 75000.00

0.05

0.10

0.15

0 1000 2000 3000 4000 50000.0

0.1

0.2

0.3

0.4

Thin film

(b)

0H

/M (

T g

/em

u)

M2 (emu/g)

2

(a)

Bulk

300 K

120 K

300 K

0H

/M (

T g

/em

u)

M2 (emu/g)

2

50 K

Figure 4.5: H/M vs. M2 for bulk and thin-film La0.7Ca0.3MnO3

In general, the order of the magnetic phase transition is determined via the

Banerjee criterion, which using the Landau-Lifshitz theory of first-order transitions [14],

a negative slope in the H/M vs. M2 plot is of first-order, otherwise the transition is

second-order (fig 4.5). The Banerjee criterion has been implemented for a wide variety of

materials, however, controversial reports have emerged for DyCo2 [15]. The first-order

nature of the transition in DyCo2 is not obvious using the Banerjee criterion, due to the

small size of the discontinuity in the first derivative of the free energy. This results in

only positive slopes of the H/M vs. M2 isotherms, leading to the false classification of a

second-order transition. A new criterion for determining the order of a transition has

recently been proposed based on a re-scaling of entropy change curves [16].

Universal behavior manifested in the collapse of ΔSM (T) curves after a scaling

procedure has been established for materials undergoing an SOMT, such as that

described in Chapter 3 of this dissertation.

Page 60: Magnetism in Complex Oxides Probed by Magnetocaloric

44

-12 -8 -4 0 40.01

0.1

1

(b)

Thin-Film

S

'

S

'

Polycrystalline

(a)

-4 -2 0 2 4

0.1

1

Figure 4.6: Universal curve calculations as described in the text for the polycrystalline

bulk (a), and thin-film (b) forms of La0.7Ca0.3MnO3.

However, the scaling assumptions that underlie this behavior break down when applied to

an FOMT, and the expected collapse of the modified ( ) curves fails. As a

consequence, whether or not collapse is achieved, the universal curve method can be

applied as a method of distinguishing first and second order magnetic transitions. Figure

4.6 compares the universal curve constructions for each sample by plotting the

normalized entropy change (ΔS') against a reduced temperature (θ), where ΔS'= ΔSM /

Δ is the re-scaled entropy change and θ is the temperature variable defined by:

{ ( ) ( )⁄

( ) ( )⁄

.

Here the reference temperatures and are chosen such that ∆ ( )

∆ ( ) ⁄ . In Fig. 4.6 (a), the divergence of the curves is clear in the

polycrystalline compound, particularly above the TC. In the thin-film there is a clear

collapse of the curves below TC, which is consistent with an SOMT. We note that there is

not perfect agreement of the data above TC, however a check of the Arrott plot

constructions confirms the second order nature of the transition in the thin-film. This

Page 61: Magnetism in Complex Oxides Probed by Magnetocaloric

45

suggests that T<TC is the essential region for collapse when applying this criterion. The

presence of quenched disorder is known to force fluctuation-driven FOMTs to become

continuous [17], and this appears to be the case in the thin-film sample.

4.4 Conclusions

The magnetic and magnetocaloric properties of bulk polycrystalline and thin-film

samples of the CMR manganite LCMO were investigated to observe the effects of

reduced dimensionality in the system. Broadened transitions, along with reduced Curie

temperature, magnetic moment, and magnetic entropy change were observed in the

nanocrystalline and thin-film samples. Even though the thin-film exhibited vastly

different properties, there was a minor increase in the RC. The FOMT in bulk LCMO is

converted to a SOMT in the thin-film. In the next chapter the role of Sr-doping on the A-

site of La0.7Ca0.3-xSrxMnO3 will be discussed.

References

[1] Y. Tokura, Critical features of colossal magnetoresistive manganites, Rep Prog Phys,

69 (2006) 797-851.

[2] S.-W.C and H.Y. Hwang, Colossal Magnetoresistive Oxides, in: Y. Tokura (Ed.),

Gordon & Breach, New York, 1997.

[3] Y.A. Soh, G. Aeppli, N.D. Mathur, M.G. Blamire, Mesoscale magnetism at the grain

boundaries in colossal magnetoresistive films, Phys Rev B, 63 (2001) 020402.

[4] P.K. Siwach, et. al., Influence of strain relaxation on magnetotransport properties of

epitaxial La0.7Ca0.3MnO3 films, J Phys-Condens Mat, 18 (2006) 9783-9794.

[5] J. Dvorak, et. al., Are strain-induced effects truly strain induced? A comprehensive

study of strained LCMO thin films, J Appl Phys, 97 (2005) 10C102.

Page 62: Magnetism in Complex Oxides Probed by Magnetocaloric

46

[6] M. Egilmez, K.H. Chow, J. Jung, Percolative model of the effect of disorder on the

resistive peak broadening in La(2/3)Ca(1/3)MnO(3) near the metal-insulator transition, Appl

Phys Lett, 92 (2008) 162515.

[7] M. Ziese, H.C. Semmelhack, K.H. Han, S.P. Sena, H.J. Blythe, Thickness dependent

magnetic and magnetotransport properties of strain-relaxed La0.7Ca0.3MnO3 films, J Appl

Phys, 91 (2002) 9930-9936.

[8] A. de Andres, J. Rubio, G. Castro, S. Taboada, J.L. Martinez, J.M. Colino, Structural

and magnetic properties of ultrathin epitaxial La0.7Ca0.3MnO3 manganite films: Strain

versus finite size effects, Appl Phys Lett, 83 (2003) 713-715.

[9] D. Kim, B. Revaz, B.L. Zink, F. Hellman, J.J. Rhyne, J.F. Mitchell, Tricritical point

and the doping dependence of the order of the ferromagnetic phase transition of La1-

xCaxMnO3, Phys Rev Lett, 89 (2002) 227202.

[10] A.N. Ulyanov, et.al., Metamagnetic transition and extremely large low field

magnetocaloric effect in La(0.7)Ca(0.3)MnO(3) manganite, J Appl Phys, 103 (2008) 07B328.

[11] T. Sarkar, A.K. Raychaudhuri, A.K. Bera, S.M. Yusuf, Effect of size reduction on

the ferromagnetism of the manganite La1-xCaxMnO3 (x=0.33), New J Phys, (2010)

123026.

[12] W. Tang, W.J. Lu, X. Luo, B.S. Wang, X.B. Zhu, W.H. Song, Z.R. Yang, Y.P. Sun,

Particle size effects on La0.7Ca0.3MnO3: size-induced changes of magnetic phase

transition order and magnetocaloric study, J Magn Magn Mater, 322 (2010) 2360-2368.

[13] N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija, C. Leighton, Magnetocaloric

effect and refrigerant capacity in charge-ordered manganites, J Appl Phys, 106 (2009)

023909.

Page 63: Magnetism in Complex Oxides Probed by Magnetocaloric

47

[14] L.D. Landau, E.M. Lifshits, L.P. Pitaevskii, Statistical physics, Pergamon Press,

Oxford ; New York, 1980.

[15] M. Parra-Borderias, F. Bartolome, J. Herrero-Albillos, L.M. Garcia, Detailed

discrimination of the order of magnetic transitions and magnetocaloric effect in pure and

pseudobinary Co Laves phases, J Alloy Compd, 481 (2009) 48-56.

[16] C.M. Bonilla, F. Bartolome, L.M. Garcia, M. Parra-Borderias, J. Herrero-Albillos,

V. Franco, A new criterion to distinguish the order of magnetic transitions by means of

magnetic measurements, J Appl Phys, 107 (2010) 09E131.

[17] S. Rossler, U.K. Rossler, K. Nenkov, D. Eckert, S.M. Yusuf, K. Dorr, K.H. Muller,

Rounding of a first-order magnetic phase transition in Ga-doped La0.67Ca0.33MnO3, Phys

Rev B, 70 (2004) 104417.

Page 64: Magnetism in Complex Oxides Probed by Magnetocaloric

48

CHAPTER 5.

INFLUENCE OF Sr DOPING ON THE MAGNETIC TRANSITIONS AND

CRITICAL BEHAVIOR OF La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2, AND 0.25)

SINGLE CRYSTALS

This chapter presents a comprehensive study of the ferromagnetic phase

transitions and critical exponent trends near these transitions in La0.7Ca0.3-xSrxMnO3 (x =

0, 0.05, 0.1, 0.2 and 0.25) single crystals. Based on the H/M vs. M2 analyses and using

Banerjee criterion, we demonstrate a transition from the discontinuous FOMT to the

continuous SOMT at x∼0.1. The critical analyses, based on the magnetic data using the

Kouvel–Fisher method, affirm that x∼0.1 is a tri-critical point that separates the FOMT

for x < 0.1 from the SOMT for x > 0.1. Above the tri-critical point (i.e. x ≥ 0.2), the

system exhibits a SOMT with the critical exponents ( = 0.36±0.01, = 1.22±0.01)

belonging to the Heisenberg universality class ( = 0.365±0.003, = 1.336±0.004) with

short-range exchange interactions. This indicates that the magnetic interaction in these

manganites is of a short-range nature. Our results and analyses reveal that while the DE

mechanism and formation of ferromagnetic clusters can account for the canonical MR

and metal-like conductivity in La0.7Ca0.3−xSrxMnO3 with x = 0.2 and 0.25, other effects

such as cooperative Jahn–Teller (JT) distortions and antiferromagnetic (AFM) coupling

are important additions for understanding the nature of the ferromagnetic transition, the

MI transition and CMR in La0.7Ca0.3−xSrxMnO3 with x = 0, 0.05 and 0.1. Our studies

Page 65: Magnetism in Complex Oxides Probed by Magnetocaloric

49

provide physical insights into the relationship between the ferromagnetism and

conductivity in doped manganites.

5.1 Introduction

In perovskite manganites, the choice of A-site (A = rare-earth, alkaline earth)

dopants has been shown to control an effective one-electron bandwidth (W), which in

turn governs the magnetic and magneto-transport properties of the materials [1-7]. The

metallic FM state in doped manganites has been widely interpreted in the context of the

DE mechanism; as a magnetic field is applied to the material, the local t2g spins are

forced to align, thus reducing spin scattering (i.e. resistivity decreases). The transfer

integral that dictates the strength of the DE interaction depends on the angle between

these local spins, thus linking ferromagnetism and metallicity. The DE theory provides a

good description of the metallic FM state in large bandwidth manganites such as

La0.7Sr0.3MnO3 [8]. However, it is not sufficient to explain the features of the MI

transition and CMR observed in La0.7Ca0.3MnO3 [3] and other narrow bandwidth

manganites in which collective JT distortions and AFM interactions coexist and compete

with the ferromagnetic phase. Indeed, experimental studies have revealed that the

occurrence of the MI transition and CMR in La0.7Ca0.3MnO3 results mainly from the

combination of the DE interaction between Mn3+

and Mn4+

ions and a strong JT effect [3,

9]. This system has been found to undergo a discontinuous FOMT at TC [10, 11], and the

first-order nature of the transition is associated with a strong electron-phonon coupling

and/or an intrinsic inhomogeneity in the material that gives rise to competing coexisting

ground states [12, 13]. Meanwhile, the La0.7Sr0.3MnO3 system (which is free from JT

lattice distortions) has been found to undergo a continuous SOMT and exhibit a canonical

Page 66: Magnetism in Complex Oxides Probed by Magnetocaloric

50

MR behavior.[3, 14] From a crystallographic standpoint, we note that La0.7Ca0.3MnO3

crystallizes in an orthorhombic (Pbnm) structure, whereas La0.7Sr0.3MnO3 possesses a

rhombohedral (R 3c) structure [15-17].

One way to stabilize intermediate phases bridging the two systems is co-

substitution of both Sr and Ca in the lattice. The substitution of large Sr2+

ions for smaller

Ca2+

ions in La0.7Ca0.3−xSrxMnO3 (0≤x≤0.3) manganites leads to a structural change from

the orthorhombic to rhombohedral structure which consequently induces a change in the

physical properties of the materials.[3, 15, 17] Notably, Tomioka et al.[3] reported that

with increasing Sr content, the sharp MI transition and CMR behavior for La0.7Ca0.3MnO3

(x = 0) were transformed to the metallic state and exhibited canonical MR behavior for

La0.7Sr0.3MnO3 (x = 0.3). However, the physical origin of the observed phenomena is still

not fully understood. Particularly, the mechanism of a metal-like conductivity in the PM

region in La0.7Ca0.3−xSrxMnO3 compounds with x > 0.1 remains an open question [3]. In

addition, it is unclear how the magnetic interactions are renormalized near the PM–FM

transition range and what universality class governs the PM–FM transitions in these

systems.

To address these important issues, we have conducted a comprehensive study of

the FM phase transitions and critical exponents near these transitions in La0.7Ca0.3-

xSrxMnO3 (x = 0, 0.05, 0.1, 0.2 and 0.25) single crystals. The samples were provided by

Professor Nan Hwi Hur group at Sogang University, South Korea.

5.2 Experiment

Single crystals of La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2, 0.25) were prepared by

the floating-zone method using an infrared radiation convergence-type image furnace that

Page 67: Magnetism in Complex Oxides Probed by Magnetocaloric

51

consist of four mirrors and halogen lamps; details of the growth conditions can be found

elsewhere.[18] The starting polycrystalline ceramic feed rods were obtained from the

solid-state reaction of a stoichiometric mixture of La2O3, CaCO3, SrCO3 and MnCO3

were ground, pelletized, and calcined at 1000 °C for 20 h. Sintering was carried out in air

at 1300 °C for 80 h with intermediate regrinding. The feed rod was treated at 1300 °C for

20 h in air. The crystal growth apparatus (Crystal Systems Inc.) is an infrared radiation

convergence type image furnace that consists of four mirrors and four halogen lamps.

The input power for the halogen lamp is 1000 W. The temperature of the image furnace

was measured by detecting blackbody radiation of a graphite rod with a pyrometer. X-ray

diffraction (XRD) data and electron-probe microanalysis confirmed the quality of the

crystals. The XRD analyses indicated that the crystal structure is orthorhombic for x = 0,

0.05 and 0.1 compositions and is rhombohedral for x = 0.2 and 0.25 compositions. These

results are fully consistent with those reported in previous works. [15-17]

5.3 Results and Discussion

Figure 5.1 shows the temperature dependence of magnetization taken at an applied

field of 5 kOe for the La0.7Ca0.3−xSrxMnO3 (x=0, 0.05, 0.1 and 0.25) samples. It is

observed in Figure. 5.1 that all of the samples undergo a PM–FM transition and this

transition broadens gradually with increasing Sr doping. The TC of each sample, defined

by the minimum in dM/dT, are plotted as a function of Sr-doping, as shown in the inset of

Figure. 5.1. In connection with the crystal structure of the samples, one can see clearly in

the inset of Figure. 5.1 that with increasing Sr doping, the TC increases at a rate faster in

the orthorhombic phase (x = 0, 0.05 and 0.1) than in the rhombohedral phase (x = 0.2 and

0.25). The boundary line between these two crystalline phases is taken at x = 0.15, which

Page 68: Magnetism in Complex Oxides Probed by Magnetocaloric

52

corresponds to a tolerance factor t = 0.92 as determined from previous studies [3, 17, 19].

It has been noted in doped manganites that cooperative JT distortions are present in the

orthorhombic phase but are not allowed due to the higher symmetry of the MnO6

octahedra in the rhombohedral phase [20-22].

0 50 100 150 200 250 300 350 4000.0

0.2

0.4

0.6

0.8

1.0

0.0 0.1 0.2 0.3210

240

270

300

330

360

M

/MS

T (K)

x=0

x=0.05

x=0.10

x=0.25

R3c

Pbnm

TC (

K)

x

Figure 5.1: Temperature dependence of magnetization taken at 5 kOe. Inset shows the

dependence of the Curie temperature (TC) on the Sr-doped content. The boundary line

between the orthorhombic (Pbnm) and rhombohedral (R3c) phases is taken at x = 0.15.

From [44]

This leads to a general expectation in the present case that the JT effect is significant in

La0.7Ca0.3−xSrxMnO3 with x = 0, 0.05 and 0.1 but is negligible in La0.7Ca0.3−xSrxMnO3 with

x = 0.2 and 0.25. Since JT distortions decrease and the Mn–O–Mn bond angle increases

(and W increases) with Sr doping, the DE interaction is strengthened and the metallic FM

state is stabilized in the samples with larger Sr doping [3, 15]. This clearly explains the

increase of TC with increasing Sr doping (see Figure. 5.1). This also explains a faster

increase in the rate of TC with Sr addition in the orthorhombic phase than in the

Page 69: Magnetism in Complex Oxides Probed by Magnetocaloric

53

rhombohedral phase (see the inset of Figure. 5.1). To determine the order of the magnetic

phase transition in the La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2 and 0.25) samples, we

have analyzed H/M vs. M2 curves (which were constructed from the isothermal M(H)

data) using Banerjee criterion [23], the results of which are presented in Figure. 5.2.

0 1000 2000 3000 4000 5000 6000

200

400

600

800

0 500 1000 1500 2000 2500 3000 3500

100

200

300

400

0 500 1000 1500 2000

100

200

300

400

500

600

0 1000 2000 3000 4000 5000 6000 7000

500

1000

1500

2000

La0.7

Ca0.3-x

SrxMnO

3

(a) x = 0

224K220K 214K 208K

228K

232K

236K

H/M

(O

e.g

/em

u)

M2 (emu/g)

2

(c) x = 0.1

280K283K286K

289K292K

295K

298K

301K

H/M

(O

e.g

/em

u)

M2 (emu/g)

2

(d) x = 0.2

320K323K

326K329K

332K

335K

338K

341K

H/M

(O

e.g

/em

u)

M2 (emu/g)

2

(b) x = 0.05

230K240K250K

260K270K

280K

290K

300K

H/M

(O

e.g

/em

u)

M2 (emu/g)

2

Figure 5.2: The H/M vs. M2 plots for representative temperatures around the TC for the

La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1 and 0.2) samples. From [44]

According to this criterion, the magnetic transition is of the second-order if all of

the H/M vs. M2 curves have a uniformly positive slope [24]. On the other hand, if some of

the H/M vs. M2 curves shows a negative slope at some point, the transition is of first-

order [24, 25]. It can be observed in Figure. 5.2 that the H/M vs. M2 curves of the x = 0

Page 70: Magnetism in Complex Oxides Probed by Magnetocaloric

54

and 0.05 samples show a negative slope at T > TC, indicating a FOMT for these samples.

However, the H/M vs. M2 curves of the x = 0.1, 0.2 and 0.25 samples have only positive

slopes, implying that these samples belong to the class of SOMT materials. Nevertheless,

a closer examination of the H/M vs. M2 curves at low magnetic fields for x = 0.1 reveals

that this sample is not a purely SOMT material and some degree of FOMT may be still

present in the material. It has been noted, in chapter 4, that a precise determination of the

type of magnetic transition using the Banerjee criterion becomes difficult when the

magnetic transition is a mixture of FOMT and SOMT [26]. Our study shows that the Sr

doping in La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2 and 0.25) suppresses the FOMT but

favors a SOMT and a FOMT to SOMT transition occurs at x∼0.1. A similar trend has

been reported on polycrystalline La2/3(Ca1−xSrx)1/3MnO3 (x = 0, 0.05, 0.15 and 0.3)

manganites [22-24].

Figure. 5.3 shows the A–N plots of the La0.7Ca0.3−xSrxMnO3 (x = 0.1, 0.2 and

0.25) samples with optimized critical exponents ( and ) obtained from the K–F method.

Figure. 5.4 shows the K–F plot for one representative sample (x = 0.2). The best fits yield

the values of TC = 289 K, = 0.26 ± 0.01 and = 1.06 ± 0.02 for the x = 0.1 sample; TC =

326 K, = 0.36 ± 0.01 and = 1.22 ± 0.01 for the x = 0.2 sample; TC = 344 K, = 0.42 ±

0.02 and = 1.14 ± 0.05 for the x = 0.25 sample. Using the Widom scaling relationship

[27], + =, the critical exponent () is determined to be 5.1 ± 0.2, 4.4 ± 0.2 and 3.7 ±

0.2 for x = 0.1, 0.2 and 0.25 compositions, respectively. This relationship has been tested

by plotting M(T = TC) versus H/()

=H1/

and checking the linearity of the curve as

shown in Figure. 5.5. The reliability of the obtained exponents and Curie temperatures

can also be ascertained by checking the scaling of the magnetization curves.

Page 71: Magnetism in Complex Oxides Probed by Magnetocaloric

55

50 100 150 200 250 3000

1

2

3

4

5

6

50 100 1500

1

2

3

4

50 100 150 2000

2

4

6

8

= 1.222

300K

M1/ (

10

6 (

em

u/g

)1/)

(H/M)1/

(Oe g/emu)1/

(a)

x = 0.1

280K

= 0.260

= 1.064

x = 0.25

340K

= 0.363

x = 0.2

(b)M

1/ (

10

4 (

em

u/g

)1/)

(H/M)1/

(Oe g/emu)1/

320K

= 0.417

= 1.142

355K

341K(c)

M1/ (

10

3 (

em

u/g

)1/)

(H/M)1/

(Oe g/emu)1/

Figure 5.3: Modified Arrott plot isotherms with 1K temperature interval for the

La0.7Ca0.3-xSrxMnO3 (x=0.1, 0.2 and 0.25) samples. From [44].

According to Eq. (3.21), if the appropriate values for the critical exponents and for the TC

are used, the plot of M/H1/

versus ε/H1/

should correspond to a universal curve onto

which all experimental data points collapse. Two different constructions have been used

in this work, both based on the scaling equation of state discussed in chapter 3.2.1.

Page 72: Magnetism in Complex Oxides Probed by Magnetocaloric

56

315 318 321 324 327 330 3330

5

10

15

20

25

30

35

0

2

4

6Kouvel-Fisher method

=1.222

TC=326.793

=0.363

TC=326.710

Y(K

)=M

S(d

MS/d

T)-1

T (K)

X(K

)=

0

-1(d

0

-1/d

T)-1

Figure 5.4: Temperature dependence of spontaneous Ms (square) and inverse initial

susceptibility 0-1

(circles) for the x=0.2 sample; solid lines are fitting curves to Eqs.

(3.15) and (3.16), respectively. From [44].

Alternatively, Eq. (3.22) indicates that M/|ε| versus H/|ε|

should result in two

universal curves, one for ε > 0 (T > TC) and the other for ε < 0 (T < TC). For a more

convenient visualization of the results, this plot is usually represented in logarithmic

scale. However, this second representation tends to cover the small deviations of the

experimental data with respect to the universal curves caused by an inappropriate choice

of the parameters. Using the values of and TC obtained from the K–F method, the

scaled data are plotted in Figure. 5.6 for the x = 0.1 and 0.2 samples, respectively. In the

case of scaling using Eq. (3.22), it can be observed that all of the experimental points fall

on two curves, one for T < TC and the other for T > TC. This clearly indicates that the

obtained values of and TC for these samples are reliable and in agreement with the

scaling hypothesis. A less perfect overlap of the data points has been observed for the x =

0.1 sample in comparison with the x = 0.2 sample (see Figure. 5.6a and b). This agrees

Page 73: Magnetism in Complex Oxides Probed by Magnetocaloric

57

with our previous observation (Figure. 5.2c) and argument that this sample is not a purely

SOMT material and some degree of FOMT is still present in it.

0.7 0.8 0.9 1.0

25

30

35

x = 0.2M

(T

)

H1/

(T0.23

)

0.7 0.8 0.9 1.0

30

35

40

45

x = 0.1

H1/

(T0.20

)

La0.7

Ca0.3-x

SrxMnO

3

Figure 5.5: The linearity of the M(T = TC) versus H

=H

curves validates the

value of the critical exponents. From [44].

It has been argued that in homogeneously magnetic systems the universality class

of the magnetic phase transition should depend on the range of the exchange interaction,

J(r) [28]. If J(r) decays with distance (r) at a rate faster than r−5

then the Heisenberg

exponents ( = 0.365 ± 0.003, = 1.336 ± 0.004) are valid for a 3D isotropic ferromagnet

[29]. However, if J(r) decays at a rate slower than r−4.5

then the mean-field exponents (

= 0.5, =1) are valid. According to the DE theory, the effective FM interaction is driven

by the kinetics of the electrons which favor extended states. Therefore, one could expect

the critical exponents in the DE model to be described within the framework of the mean-

field theory [30, 31]. However, computational studies have demonstrated that the critical

exponents in the DE model are consistent with those expected for the 3D Heisenberg

model [32]. It has also been theoretically shown that PM–FM transitions may become

discontinuous, depending on doping level and competition with superexchange AFM

Page 74: Magnetism in Complex Oxides Probed by Magnetocaloric

58

interactions [33]. The values of the critical exponents of La0.7Ca0.3−xSrxMnO3 (x = 0.1, 0.2

and 0.25) samples are in agreement with those predicted by the latter models. [32, 33] It

is clear that the critical exponents of the x = 0.1 sample ( = 0.26 ± 0.01, = 1.06 ± 0.02)

match well with those derived from the tricritical mean-field model ( = 0.25, = 1). The

existence of this tricritical point (x∼0.1) clearly sets a boundary between FOMT (x < 0.1)

and SOMT (x > 0.1) within the range of compositions under study (0≤ x ≤0.25). A similar

case has also been reported on La1−xCaxMnO3 (0.2≤ x ≤0.5) polycrystalline manganites,

where the tricritical point is found at x∼0.4 [29].

To further confirm that the PM–FM transition becomes a conventional SOMT

above the tricritical point, the critical exponents of the samples with x > 0.1 are expected

to match those predicted for the 3D Heisenberg model [34]. The values of the critical

exponents of the x = 0.2 sample ( = 0.36 ± 0.01, = 1.22 ± 0.01) are close to those

expected for the 3D Heisenberg model ( = 0.365 ± 0.003, = 1.336 ± 0.004). For the x =

0.25 sample, the value of = 0.42 ± 0.02 is relatively larger than that expected for the 3D

Heisenberg model ( = 0.365 ± 0.003) but is consistent with that of La0.7Sr0.3MnO3 ( =

0.45±0.02) reported previously by Ziese[35] and Taran et al.[36] using the Arrott-Noakes

method and by Lofland et al. [14], using microwave absorption methods.

The critical exponent of the x = 0.25 sample ( = 1.14 ± 0.05) is smaller than that

of the x = 0.2 sample ( = 1.22±0.01). It has been argued that for a true SOMT the critical

exponents are independent of the microscopic details of a system due to the divergence of

correlation length and correlation time close to a transition point and hence their values

are almost the same for a transition that may occur in different physical systems [37].

However, one must note secondary effects on the PM–FM transition (hence on the

Page 75: Magnetism in Complex Oxides Probed by Magnetocaloric

59

critical exponents) of a system due to magnetic anisotropies or dipolar long-range

couplings of FM clusters [38-41]. It has been pointed out that a Mn-ion triplet containing

an Mn3+–Mn

4+–Mn

3+ cluster has significant binding energy of about half the binding

energy of the bulk [42].

-0.5 0.0 0.5 1.0 1.5

2

4

6

8

10

105

106

107

108

101

102

-2 -1 0 1 2 3

1

2

3

4

5

6

105

106

107

108

109

101

102

x = 0.1

MH

-1/ (

em

u g

-1 O

e-0

.20)

H-1/(+)

(10-4 Oe

-0.75)

(a)

-M

(e

mu

g-1)

-(+)

H (Oe)

x = 0.1

(b)

MH

-1/ (

em

u g

-1 O

e-0

.23)

H-1/(+)

(10-4 Oe

-0.63)

x = 0.2

(c)

-M

(e

mu

g-1)

-(+)

H (Oe)

x = 0.2

(d)

Figure 5.6: Normalized isotherms of La0.7Ca0.3−xSrxMnO3 (x = 0.1 and 0.2) samples

below and above Curie temperature (TC) using the values of and determined from K-F

method. From [44].

The large spin moments of these FM clusters are expected to enhance the dipole–dipole

interaction in the case of the 3D Heisenberg model thus resulting in larger values of the

critical exponents than those predicted by this model [36, 42]. Recently, magnetization

and electron paramagnetic resonance (EPR) studies have revealed that FM clusters

persisting, even in the paramagnetic region, have significant influence on the magnetic

Page 76: Magnetism in Complex Oxides Probed by Magnetocaloric

60

order parameters (i.e. the critical exponents) in doped manganites, such as Pr0.5Sr0.5MnO3

which will be discussed in great detail in the following chapter.

In the case of La0.7Ca0.3−xSrxMnO3 (x = 0.1, 0.2 and 0.25) samples, the x = 0.1

composition is in the crossover region of FOMT→SOMT. A crossover in the MI

transition is also observed in this composition (for x < 0.1 the MI transition is well

defined, but is largely suppressed and instead a metal-like conductivity is observed in the

PM region for x > 0.1) [3]. This concurrence clearly demonstrates a coherent correlation

between the magnetism and conductivity in the doped manganites. Furthermore, we note

the nonlinearity of the modified Arrott plots (Figure. 5.3), signaling the presence of FM

clusters in the La0.7Ca0.3−xSrxMnO3 samples. Therefore, it can be proposed that it is the

formation of FM clusters that leads to a percolation mechanism for conduction and

metallic behaviors observed at T > TC in the PM region for the x = 0.2 and 0.25 samples.

A recent study has shown that the PM–FM transition of a (Sm0.7Nd0.3)0.52Sr0.48MnO3

manganite transforms from a FOMT to a SOMT when subjected to an external pressure

[43]. It has been argued that the presence of external pressure at a certain level can

suppress the polaronic state, increase the bandwidth of the system, and as a result the

FOMT converts to SOMT [43]. In the case of La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2

and 0.25) manganites, we argue that the substitution of large Sr2+

ions (<rA>∼1.31Ǻ ) for

smaller Ca2+

ions (<rA>∼1.18Ǻ ) introduces chemical (internal) pressure without

affecting the valency of the Mn ions (the ratio of Mn3+

/Mn4+

is unchanged i.e. there is no

change of electronic density) [16, 17]. It has been shown that the chemical pressure

modifies local structural parameters such as the Mn–O bond distance and Mn–O–Mn

bond angle, which directly influence the probability of electron hopping between Mn ions

Page 77: Magnetism in Complex Oxides Probed by Magnetocaloric

61

[1, 3, 16, 17]. Therefore, the change of internal pressure with Sr addition is expected to

increase W which consequently increases the TC and converts the FOMT into SOMT at a

threshold pressure (which corresponds to a Sr-doping level, x = 0.1). Above the critical

point (x > 0.1), the system shows a SOMT with the critical exponents belonging to the

Heisenberg universality class. The change in nature of the PM–FM transition with

chemical (internal) pressure (Sr doping) clearly points to a strong coupling between the

magnetic order parameter and lattice strain in these doped manganites [43].

Furthermore, the theory predicts that the PM–FM transition of La1−xMxMnO3 (M=

Ca, Sr) becomes continuous or discontinuous, depending upon the change in the relative

strength of the DE interaction (i.e. the FM interaction) and AFM coupling [33]. In the

present study, we note that the ground state of La0.7Ca0.3MnO3 (x = 0) is FM, but the

AFM phase is also present and competing with the FM phase [12, 13]. The substitution of

Sr for Ca in La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2 and 0.25) suppresses the AFM

tendency and enhances the FM phase. As a result, the PM–FM transition changes from

discontinuous to continuous at x∼0.1. In addition, we recall that strong JT lattice

distortions are possible in the x = 0, 0.05 and 0.1 samples that crystallize in an

orthorhombic structure, whereas this effect is negligible in the x = 0.2 and 0.25 samples

possessing a rhombohedral structure. These important observations, coupled with the

magnetic, magnetotransport and critical exponent analyses, clearly suggest that in

addition to the DE interaction, cooperative JT effects and AFM coupling are important

ingredients for assessing the nature of the FM transition and the MI transition including

CMR in La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05 and 0.1) manganites. While the DE model is

sufficient to explain the canonical MR behavior in La0.7Ca0.3−xSrxMnO3 (x = 0.2 and 0.25)

Page 78: Magnetism in Complex Oxides Probed by Magnetocaloric

62

samples where AFM interactions are weak and JT lattice distortions are negligible, the

formation of FM clusters likely leads to a percolation mechanism for the conduction and

metallic behavior observed in the PM region for these samples.

5.4 Conclusions

The FM phase transitions and critical behavior of La0.7Ca0.3−xSrxMnO3 (x = 0,

0.05, 0.1, 0.2 and 0.25) single crystals have been studied systematically. Using the

Banerjee criterion and Kouvel–Fisher method, we show that x∼0.1 is a tricritical point

that separates a FOMT for x < 0.1 from a SOMT for x > 0.1. Above the tricritical point,

the system exhibits a SOMT with critical exponents belonging to the Heisenberg

universality class with short-range exchange interactions. The change of the PM–FM

transition with chemical (internal) pressure introduced by substitution of larger Sr ions

for smaller Ca ions points to the strong coupling between the magnetic order and

structural parameters in these doped manganites.

References

[1] Y. Tokura, Y. Tomioka, Colossal magnetoresistive manganites, J Magn Magn Mater,

200 (1999) 1-23.

[2] P.G. Radaelli, D.E. Cox, M. Marezio, S.W. Cheong, P.E. Schiffer, A.P. Ramirez,

Simultaneous structural, magnetic, and electronic-transitions in La1-xCaxMno3 with

x=0.25 and 0.50, Phys Rev Lett, 75 (1995) 4488-4491.

[3] Y. Lyanda-Geller, et. al., Charge transport in manganites: Hopping conduction, the

anomalous Hall effect, and universal scaling, Phys Rev B, 63 (2001).

[4] H.Y. Hwang, S.W. Cheong, P.G. Radaelli, M. Marezio, B. Batlogg, Lattice effects on

the magnetoresistance in doped LaMno3, Phys Rev Lett, 75 (1995) 914-917.

Page 79: Magnetism in Complex Oxides Probed by Magnetocaloric

63

[5] E. Dagotto, Open questions in CMR manganites, relevance of clustered states and

analogies with other compounds including the cuprates, New J Phys, 7 (2005).

[6] V.B. Shenoy, C.N.R. Rao, Electronic phase separation and other novel phenomena

and properties exhibited by mixed-valent rare-earth manganites and related materials,

Philos T R Soc A, 366 (2008) 63-82.

[7] J. Tao, D. Niebieskikwiat, M. Varela, W. Luo, M.A. Schofield, Y. Zhu, M.B.

Salamon, J.M. Zuo, S.T. Pantelides, S.J. Pennycook, Direct Imaging of Nanoscale Phase

Separation in La0.55Ca0.45MnO3: Relationship to Colossal Magnetoresistance, Phys Rev

Lett, 103 (2009).

[8] Y. Motome, N. Furukawa, Disorder effect on spin excitation in double-exchange

systems, Phys Rev B, 71 (2005).

[9] P. Dai, J.D. Zhang, H.A. Mook, S.H. Liou, P.A. Dowben, E.W. Plummer,

Experimental evidence for the dynamic Jahn-Teller effect in La0.65Ca0.35MnO3, Phys Rev

B, 54 (1996) R3694-R3697.

[10] C.S. Hong, W.S. Kim, N.H. Hur, Transport and magnetic properties in the

ferromagnetic regime of La1-xCaxMnO3, Phys Rev B, 63 (2001).

[11] A.N. Ulyanov, J.S. Kim, Y.M. Kang, D.G. Yoo, S.I. Yoo, Oxygen deficiency as a

driving force for metamagnetism and large low field magnetocaloric effect in La0.7Ca0.3-

xSrxMnO3-delta manganites, J Appl Phys, 104 (2008).

[12] A.S. Alexandrov, A.M. Bratkovsky, Carrier density collapse and colossal

magnetoresistance in doped manganites, Phys Rev Lett, 82 (1999) 141-144.

[13] J.A. Verges, V. Martin-Mayor, L. Brey, Lattice-spin mechanism in colossal

magnetoresistive manganites, Phys Rev Lett, 88 (2002).

Page 80: Magnetism in Complex Oxides Probed by Magnetocaloric

64

[14] S.E. Lofland, V. Ray, P.H. Kim, S.M. Bhagat, M.A. Manheimer, S.D. Tyagi,

Magnetic phase transition in La0.7Sr0.3MnO3: Microwave absorption studies, Phys Rev B,

55 (1997) 2749-2751.

[15] P.G. Radaelli, M. Marezio, H.Y. Hwang, S.W. Cheong, B. Batlogg, Charge

localization by static and dynamic distortions of the MnO6 octahedra in perovskite

manganites, Phys Rev B, 54 (1996) 8992-8995.

[16] A.N. Ulyanov, G.V. Gusakov, V.A. Borodin, N.Y. Starostyuk, A.B. Mukhin,

Ferromagnetic ordering temperature and internal pressure in manganites La0.7Ca0.3-

xSrxMnO3, Solid State Commun, 118 (2001) 103-105.

[17] A.N. Ulyanov, I.S. Maksimov, E.B. Nyeanchi, Y.V. Medvedev, S.C. Yu, N.Y.

Starostyuk, B. Sundqvist, Structure and pressure effect on the properties of La0.7Ca0.3-

xSrxMnO3 manganites, J Appl Phys, 91 (2002) 7739-7741.

[18] C.S. Hong, W.S. Kim, E.O. Chi, K.W. Lee, N.H. Hur, Colossal magnetoresistance in

La0.7Ca0.3MnO3-delta: Comparative study of single-crystal and polycrystalline material,

Chem Mater, 12 (2000) 3509-3515.

[19] P.G. Radaelli, G. Iannone, M. Marezio, H.Y. Hwang, S.W. Cheong, J.D. Jorgensen,

D.N. Argyriou, Structural effects on the magnetic and transport properties of perovskite

A(1-x)A’(x)MnO3 (x=0.25, 0.30), Phys Rev B, 56 (1997) 8265-8276.

[20] C.H. Booth, F. Bridges, G.H. Kwei, J.M. Lawrence, A.L. Cornelius, J.J. Neumeier,

Direct relationship between magnetism and MnO6 distortions in La1-xCaxMnO3, Phys

Rev Lett, 80 (1998) 853-856.

[21] J.B. Goodenough, Electronic structure of CMR manganites (invited), J Appl Phys,

81 (1997) 5330-5335.

Page 81: Magnetism in Complex Oxides Probed by Magnetocaloric

65

[22] J. Mira, J. Rivas, L.E. Hueso, F. Rivadulla, M.A.L. Quintela, Drop of

magnetocaloric effect related to the change from first- to second-order magnetic phase

transition in La2/3(Ca1-xSrx)(1/3)MnO3, J Appl Phys, 91 (2002) 8903-8905.

[23] S.K. Banerjee, On a generalised approach to 1st and 2nd order magnetic transitions,

Physics Letters, 12 (1964) 16-17.

[24] J. Mira, J. Rivas, F. Rivadulla, C. Vazquez-Vazquez, M.A. Lopez-Quintela, Change

from first- to second-order magnetic phase transition in La2/3(Ca,Sr)(1/3)MnO3

perovskites, Phys Rev B, 60 (1999) 2998-3001.

[25] M.H. Phan, S.C. Yu, N.H. Hur, Y.H. Jeong, Large magnetocaloric effect in a

La0.7Ca0.3MnO3 single crystal, J Appl Phys, 96 (2004) 1154-1158.

[26] V. Franco, A. Conde, V. Provenzano, R.D. Shull, Scaling analysis of the

magnetocaloric effect in Gd5Si2Ge1.9X0.1 (X=Al, Cu, Ga, Mn, Fe, Co), J Magn Magn

Mater, 322 (2010) 218-223.

[27] B. Widom, Equation of state in neighborhood of critical point, J Chem Phys, 43

(1965) 3898-&.

[28] M.E. Fisher, B.G. Nickel, S. Ma, Critical exponents for long-range interactions,

Phys Rev Lett, 29 (1972) 917-&.

[29] D. Kim, et. al., Tricritical point and the doping dependence of the order of the

ferromagnetic phase transition of La1-xCaxMnO3, Phys Rev Lett, 89 (2002).

[30] K. Kubo, N. Ohata, Quantum-theory of double exchange .1, J Phys Soc Jpn, 33

(1972) 21.

[31] C.V. Mohan, et. al., Critical behaviour near the ferromagnetic-paramagnetic phase

transition in La0.8Sr0.2MnO3, J Magn Magn Mater, 183 (1998) 348-355.

Page 82: Magnetism in Complex Oxides Probed by Magnetocaloric

66

[32] Y. Motome, N. Furukawa, A Monte Carlo method for fermion systems coupled with

classical degrees of freedom, J Phys Soc Jpn, 68 (1999) 3853-3858.

[33] J.L. Alonso, L.A. Fernandez, F. Guinea, V. Laliena, V. Martin-Mayor,

Discontinuous transitions in double-exchange materials, Phys Rev B, 63 (2001).

[34] J. Yang, Y. Lee, Y. Li, Critical behavior of the electron-doped manganite

La0.9Te0.1MnO3, Phys Rev B, 76 (2007).

[35] M. Ziese, Critical scaling and percolation in manganite films, J Phys-Condens Mat,

13 (2001) 2919-2934.

[36] S. Taran, B.K. Chaudhuri, S. Chatterjee, H.D. Yang, S. Neeleshwar, Y.Y. Chen,

Critical exponents of the La0.7Sr0.3MnO3, La0.7Ca0.3MnO3, and Pr0.7Ca0.3MnO3 systems

showing correlation between transport and magnetic properties, J Appl Phys, 98 (2005).

[37] H.E. Stanley, Introduction to phase transitions and critical phenomena, Oxford

University Press, New York,, 1971.

[38] H. Huhtinen, et. al., Unconventional critical behavior of magnetic susceptibility as a

consequence of phase separation and cluster formation in La0.7Ca0.3MnO3 thin films, J

Appl Phys, 91 (2002) 7944-7946.

[39] M. Kar, A. Perumal, S. Ravi, Critical behavior studies in La1-xAgxMnO3 double-

exchange ferromagnet, Physica Status Solidi B-Basic Solid State Physics, 243 (2006)

1908-1913.

[40] T.L. Phan, S.G. Min, S.C. Yu, S.K. Oh, Critical exponents of La(0.9)Pb(0.1)MnO(3)

perovskite, J Magn Magn Mater, 304 (2006) E778-E780.

[41] M. Seeger, S.N. Kaul, H. Kronmuller, R. Reisser, Asymptotic critical-behavior of

Ni, Phys Rev B, 51 (1995) 12585-12594.

Page 83: Magnetism in Complex Oxides Probed by Magnetocaloric

67

[42] G.A. Gehring, D.J. Coombes, The theory of small polarons in manganite, J Magn

Magn Mater, 177 (1998) 873-874.

[43] P. Sarkar, S. Arumugam, P. Mandal, A. Murugeswari, R. Thiyagarajan, S.E. Muthu,

D.M. Radheep, C. Ganguli, K. Matsubayshi, Y. Uwatoko, Pressure Induced Critical

Behavior of Ferromagnetic Phase Transition in Sm-Nd-Sr Manganites, Phys Rev Lett,

103 (2009).

[44] Portions of this chapter are reprinted from Journal of Alloys and Comounds, 508,

M.H. Phan, V. Franco, N.S. Bingham, H. Srikanth, N.H. Hur, S.C. Yu, Tricritical point

and critical exponents of La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05, 0.1, 0.2,

0.25) single crystals, Copyright (2010), with permission from Elsevier

Page 84: Magnetism in Complex Oxides Probed by Magnetocaloric

68

CHAPTER 6.

MAGNETIC TRANSITIONS, MAGNETOCALORIC EFFECT, MAGNETIC

ANISOTROPY, CRITICAL EXPONENTS AND THEIR CORRELATIONS IN

Pr0.5Sr0.5MnO3

In the previous chapter, we discussed the influence on the nature of the

paramagnetic (PM) to ferromagnetic (FM) transition on Sr doping in the La0.7Ca0.3-

xSrxMnO3 system. However, by changing the doping concentration from the ideal (in

terms of stabilized FM-metallic nature) x=1/3, to x=1/2 (where there are an equal number

of Mn3+

and Mn4+

ions) there are new phases that arise in addition to FM ordering,

namely antiferromagnetic (AFM) order, charge/orbital ordering (CO/OO), and in some

cases phase-separation i.e. competition among all aforementioned states.

In this chapter, we present systematic studies of the influence of a first-order

magnetic transition (FOMT) and a second-order magnetic transition (SOMT) on the

MCE and RC of charge-ordered (CO) Pr0.5Sr0.5MnO3 (PSMO). Our results reveal that

while the FOMT at TCO induces a larger MCE, it is restricted to a narrow temperature

range resulting in a smaller RC. The SOMT at TC induces a smaller MCE but with a

distribution over a broader temperature range, thus resulting in a larger RC. In addition,

hysteretic losses associated with the FOMT are very large below TCO and, therefore,

detrimental to the RC, whereas these effects are very small or negligible below TC due to

the nature of the SOMT.

Page 85: Magnetism in Complex Oxides Probed by Magnetocaloric

69

In addition to the existence of deleterious magnetic and thermal effects on RC, the

Maxwell relation used to calculate SM is suspect due to the discontinuous nature of a

first-order transition. Because PSMO exhibits two fundamentally different transitions,

well-separated in temperature, we were able to deconvolute the two transitions, therefore

allowing the critical analyses of the SOMT in order to understand the type of magnetic

order and interactions of the FM phase in PSMO. The results obtained reveal the

existence of a short-range ferromagnetic order at T < TC and the presence of

ferromagnetic clusters in the paramagnetic region (T > TC).

To probe the magnetic anisotropy and its correlations with the MCE and critical

exponents in this material, transverse susceptibility measurements were performed. These

experiments reveal an abrupt change in magnetic anisotropy field (HK) at the FM to AFM

transition (TCO ~150 K), which is associated with a simultaneous structural phase

transition from tetragonal to orthorhombic symmetry. This provides evidence of strong

coupling between the magnetism and the lattice in Pr0.5Sr0.5MnO3. The data also clearly

indicate the presence of HK in both the PM and AFM states, pointing to the importance of

the d(x2-y

2)-type orbital order and the orbital-order-induced intrinsic phase separation in

the A-type AFM. The polycrystalline sample was synthesized using standard solid-state

reaction method and provided by Professor Christopher Leighton at the University of

Minnesota.

6.1 Introduction

As we discussed in Chapter 3, the refrigerant capacity (RC) is considered to be

the most important factor for assessing the usefulness of a magnetic refrigerant material

[1-3]. The RC depends not only on the magnitude of ∆SM, but also on the temperature

Page 86: Magnetism in Complex Oxides Probed by Magnetocaloric

70

dependence of ∆SM (e.g. the full width at half maximum of the ∆SM(T) peak) [1, 3]. A

good magnetic refrigerant material with large RC requires both a large magnitude of ∆SM

as well as a broad ∆SM(T) curve. Most previous studies on charge-ordered manganites [4-

9] were focused mainly on exploring large MCE (large magnitudes of ∆SM) around TCO

and did not consider the issues of refrigerant capacity and hysteretic losses in detail.

Thus, from fundamental and practical perspectives, it is essential to understand the

influence of the nature of magnetic phase transitions on both the MCE and RC in these

materials.

We also note that although the use of the Maxwell relation has been validated

over the years for determining SM of SOMT materials [1, 10], its applicability to FOMT

materials still remains under debate [11-18]. Gschneidner, Jr. et al.[11] calculated the

SM of Gd5Ge2Si2 from M(H) measurements using the Maxwell relation and showed its

excellent agreement with that calculated from the heat capacity (C) data. However,

Giguere et al. [19] later argued that the use of the Maxwell relation overestimated the

value of SM for this FOMT compound and suggested an alternative approach using the

Clausius-Clapeyron equation. In a recent study, Liu et al. [20] have shown that the use of

the Maxwell relation yields huge “spike” peaks of SM in the FOMT compounds MnAs,

Mn1-xFexAs, and La0.7Pr0.3Fe11.5Si1.5, in which both PM and FM phases coexist in the

vicinity of TC. Those authors proposed an alternative “geometric” solution for removing

these spurious SM peaks. De Oliveira and von Ranke have reformulated the Maxwell

relation showing its limitation for the calculation of SM in FOMT materials. Having

theoretically studied the effect of magnetic irreversibility on estimating the SM from

M(H) measurements, J.S. Amaral and V.S. Amaral argue that the spurious SM peaks

Page 87: Magnetism in Complex Oxides Probed by Magnetocaloric

71

arise mainly from the presence of magnetic irreversibility or a mixed-phase regime that is

not considered in the Maxwell relation [21]. In this case, the authors suggest that if the

Landau theory or the mean field model is applied to experimental (nonequilirium) data

then the equilibrium M(H) curves can be estimated and the true value of SM can be

consequently obtained [21, 22]. Very recently, Cui et al. [23] have proposed a simple

modification to the Maxwell relation by taking into account the mass variations in FM

and PM phases on temperature in the two-phase region for MnAsCx, (Mn,Al)As, and

Mn0.994Fe0.004As compounds, which in effect eliminates the unphysical SM peaks. While

previous studies were focused mainly on magnetocaloric materials with a first-order PM-

FM transition [13, 19-24], an interesting question emerges regarding the validity of using

the Maxwell relation for calculation of SM in magnetocaloric materials with a first-order

FM-AFM transition.

Furthermore, it has been noted that there exists a static d(x2-y

2)-type orbital order,

which mediates the FM coupling within the OO planes while favoring the AFM coupling

perpendicular to the OO planes, thus resulting in an overall A-type AFM spin state in

PSMO [25-29]. This implies that OO-induced magnetic anisotropy could play an

important role in the magnetism of the material. Indeed, torque magnetization and

electron paramagnetic resonance (EPR) studies have recently revealed strong OO-

induced magnetic anisotropy in the FM state, and the signature of magnetic anisotropy

detected in the PM and AFM state suggests FM clusters persisting within these states

[28]. Despite the light this study could shed on the phase separation problem, a clear

understanding of the magnetic anisotropy and its correlations with the magnetic entropy

and critical exponents in this material is lacking.

Page 88: Magnetism in Complex Oxides Probed by Magnetocaloric

72

In order to address the outstanding issues we have conducted a systematic study

of the magnetism, MCE and TS of PSMO. Our results clear-up some of the discrepancies

on the nature of first- and second-order phase transitions and their influences on the MCE

and highlight the important role of magnetic anisotropy and its correlations with the MCE

and critical exponents in mixed phase manganite systems like PSMO.

6.2 Experiment

Polycrystalline samples of Pr0.5Sr0.5MnO3 were fabricated from Pr2O3, SrCO3, and

MnO using standard solid-state reaction methods. The starting powders were thoroughly

ground and then reacted in air for 10 days at 1500 °C with several intermediate grindings.

The reacted powders were then cold pressed into disks of 1.5 mm thickness and sintered

in air for 1 day at 1500 °C. The final sintered samples were slow cooled over a period of

40 h to room temperature. Temperature-dependent X-ray diffraction (XRD)

measurements were performed on the sample over a wide temperature range of 100 to

300 K. The XRD patterns reveal a structural transition from the tetragonal (I4/mcm) to

orthorhombic (Fmmm) symmetry. This result is consistent with what was reported

previously for the same composition [30, 31].

6.3 Results and Discussion

6.3.1 Influence of first- and second-order magnetic phase transitions on the

magnetocaloric effect and refrigerant capacity of Pr0.5Sr0.5MnO3

Figure 6.1 shows the temperature dependence of magnetization (M(T)) taken at a

field of 0.05T. It can be observed that the PSMO system undergoes a SOMT from the PM

Page 89: Magnetism in Complex Oxides Probed by Magnetocaloric

73

to FM state at TC ~250 K followed by a FOMT from the FM charge-disordered to AFM

CO state at TCO ~152 K.

0 50 100 150 200 250 3000

20

40

60

PMFM

M

(e

mu

/g)

T (K)

ZFC

FC

AFM

Pr0.5

Sr0.5

MnO3

Figure 6.1: Temperature dependence of zero-field cooled (ZFC) and field-cooled (FC)

magnetizations taken at a field of 0.05 T.

To investigate the effect of magnetic field on the SOMT and FOMT, M(T) curves

were measured in different magnetic fields (µ0H = 1 T, 2 T, 3 T, 4 T, and 5 T), the results

of which are displayed in Figure 6.2. It can be observed that the SOMT at TC continues to

broaden as the magnetic field is increased, whereas the FOMT at TCO remains reasonably

sharp even at fields of up to 5 T due to the strong coupling between the magnetism and

the lattice in the vicinity of the TCO [32]. Due to the sharp change in the M(T) we expect

to observe a very large MCE in the FOMT region.

Page 90: Magnetism in Complex Oxides Probed by Magnetocaloric

74

Figure 6.2: Temperature dependence of magnetization taken at different magnetic fields

up to 5T [49].

In order to evaluate the MCE, the isothermal magnetization curves of the sample

were measured with a field step of 0.05 T in a range of 0-5 T and a temperature interval

of 3 K in a range of temperatures around the TC and around the TCO. Such families of

M(H) curves are shown in Figure 6.3(a) and (b), respectively. As expected from Figure

6.1, there is a more drastic change of the magnetization around the TCO (see Figure

6.3(b)) than around the TC (see Figure 6.3(a)), indicating a larger magnetic entropy

change in the vicinity of the TCO. It is worth mentioning here that around the TCO the

sample shows S-shape magnetization, which is typical for metamagnetic materials[33]. It

is believed that the metamagnetism arises mainly from the coexistence of the competing

AFM/CO and FM phases and the collapse of the AFM/CO state that occurs in the

presence of an applied magnetic field [32].

Page 91: Magnetism in Complex Oxides Probed by Magnetocaloric

75

0 1 2 3 4 5 60

10

20

30

40

50

60

70(b)

T=65K

T=300K

T=165K

0H (T)

M (

em

u/g

)

M (

em

u/g

)

0H (T)

T=165K(a)

0 1 2 3 4 5 60

10

20

30

40

50

60

70

Figure 6.3: Isothermal magnetization curves taken at different fixed temperatures

between 65 and 300 K for the Pr0.5Sr0.5MnO3 manganite: (a) around TC and (b) around

TCO [49].

Figure 6.4 shows the temperature dependence of SM calculated using the

Maxwell relation at different magnetic fields ranging from 0.15 T to 5 T. It can be

observed that the Pr0.5Sr0.5MnO3 system exhibits large magnetic entropy changes around

the TC and around the TCO. As expected from the M(T) and M(H) data, the SM around

the TCO is much larger than that around the TC. For µ0ΔH = 5 T, the magnitude of ∆SM (-

7.5 J/kg K) at the TCO is about twice that (3.2 J/kg K) at the TC. We note from Figure 6.4

that the large ∆SM peak around the TCO is concentrated in a narrow temperature range,

whereas the ∆SM peak around the TC is spread over a broader temperature range raising

the possibility of an enhanced RC.

We have calculated the RC for both cases around the TC and around the TCO using

the method described in Chapter 3.2, and plotted the results as a function of magnetic

field in Figure 6.5. It can be observed that in both cases the RC increases with the applied

magnetic field. An important fact to be emphasized here is that the magnitude of RC is

larger around the TC than around the TCO for µ0H < 3.7 T, which belongs to the magnetic

Page 92: Magnetism in Complex Oxides Probed by Magnetocaloric

76

field range useful for practical use of magnetic refrigerators. As shown previously in Ref.

[1], in the same refrigeration cycle, a material with higher RC is preferred, because it

would support transport of a greater amount of heat in a real refrigeration cycle.

Figure 6.4: Temperature dependence of magnetic entropy change (−SM) at different

applied fields up to 5 T [49].

Therefore, for the case of Pr0.5Sr0.5MnO3, a larger value of RC around the TC indicates

that magnetic refrigeration in the vicinity of the TC is more effective than that around the

TCO.

Furthermore, we recall that hysteretic losses (magnetic and thermal hysteresis) are

often involved in FOMT [33]. Because these hysteretic losses are the costs in energy to

make one cycle of the magnetic field, they must be considered when calculating the

usefulness of a magnetic refrigerant material being subjected to cyclic fields [2].

Page 93: Magnetism in Complex Oxides Probed by Magnetocaloric

77

1 2 3 4 5 6 7

50

100

150

200

250

300

RC

(J/k

g)

0H (T)

TC

TCO

(hysteresis)

TCO

(subtracted hysteresis)

Figure 6.5: Magnetic field dependence of RC for the cases around TC and TCO (without

and with subtracted hysteretic losses). The inset shows the magnetic field dependence of

magnetization taken at 150 K (below TCO) and at 180 K (below TC).

To evaluate possible hysteretic losses involved in the magnetic phase transitions

in Pr0.5Sr0.5MnO3, we measured the M(H) loops at temperatures around the TC and the

TCO. The inset of Fig 6.5 shows, for example, the M(H) curves measured at 180 K (below

TC) and at 150 K (below TCO). It can be seen that the hysteretic losses (the area mapped

out by the increasing and decreasing field curves) are very large below the TCO, whereas

they are very small or negligible below the TC. To be precise, we have subtracted the

average hysteretic losses from the RC values calculated without considering hysteretic

losses. For comparison, the RC obtained after subtracting the average hysteretic losses for

the case around the TCO are also included in Figure 6.5. In the range of magnetic fields

investigated, the superiority of the RC around the TC is even more pronounced after

Page 94: Magnetism in Complex Oxides Probed by Magnetocaloric

78

accounting for hysteresis around the TCO. This clearly indicates that the hysteretic losses

involved in a FOMT significantly reduce the RC and are, therefore, undesirable for use in

a real magnetic refrigeration cycle.

An important fact that emerges from the present study is that the comparison of

MCE among magnetocaloric materials [34] by considering the magnitude of SM only is

inadequate. Instead, a proper estimation should be made with the use of RC, with

attention paid to the fact that magnetic hysteresis losses must be estimated and subtracted

from the RC calculation. In addition, this comparison should be made only in the same

temperature range. From a device standpoint, it is believed that not only FOMT materials

but also SOMT materials are promising candidates for active magnetic refrigeration

applications. Some SOMT materials with zero hysteretic losses are even more

advantageous. An example of this is Gd – the best magnetic refrigerant candidate

material to date for practical use in sub-room-temperature magnetic refrigerators [35].

Considering the fact that SOMT materials with a large magnetic entropy change over a

broad temperature range usually possess large refrigerant capacity [35, 36], it would be

worthwhile to search for enhanced RC in materials that undergo multiple magnetic phase

transitions [9, 37, 38]. From this perspective, the MCE research in magnetic

nanocomposites with a SOMT may be of great interest [37].

We now turn our focus to the complete description of the SOMT near TC in

PSMO, including the critical exponents and equation of state, based on analyses of the

family of SM curves. As previously noted in Chapter 3, only two critical exponents (n

and ) of the SOMT are independent. We have used the non-iterative procedure to obtain

the other related critical exponents (, , and ) using the following relationships:

Page 95: Magnetism in Complex Oxides Probed by Magnetocaloric

79

1 Δ Δ; 1 2 ; 1/ 1/ Δ 1n n n (6.1)

In order to obtain the critical exponent n as outlined in Eq. 3.20, the experimental

Δ pk

MS values normalized to their corresponding maximum values,

s=Δ ( ) / Δ ( )M M

pk pkfS H S H , have been fitted as a power law of the magnetic field H in the

whole experimental magnetic field range. The fitting yields a value of the critical

exponent n = 0.670, as presented in Figure 6.6(a). Alternatively, the reference

temperature ( ) /r r C CT T T , where r CT T is chosen such that

( ) ( )⁄ , is also expected to show a power law dependence on H

with the exponent equal to Δ. This fitting, as shown in Figure 6.6(b), yields a value of the

gap exponent 1.835 (1/ 0.545 ). In this fitting procedure magnetic fields higher

than 1.4 T have been used, to ensure technical saturation.

In order to validate the reliability of the critical exponents obtained according to

this procedure, the temperature and field dependence of the normalized magnetization

and isothermal susceptibility (mt, mh and jh respectively) has been plotted versus the

renormalized temperature, t, and field, h, respectively.

Figure 6.7 shows the dimensionless mt* vs. t*, Ln(mh*) vs. Ln(h*) and Figure 6.8

shows Ln(jh*)-1

vs. Ln(h*) representations, where mt*, mh*, jh*, t*, and h* are the

corresponding dimensionless magnitudes obtained according to the relationship x*=x/xo,

where x=mt, mh, jh, t, h, and xo is in the units of x. It can be observed that the experimental

data fall on the same curve, in the mt* vs. t* representation (Figure 6.7(a)), and on the

two curves f+ (for T>TC) and f (T<TC) in the Ln(mh*) vs. Ln(h*) representation (Figure

6.7(b)).

Page 96: Magnetism in Complex Oxides Probed by Magnetocaloric

80

0.25

0.50

0.75

1.000.00 0.25 0.50 0.75 1.00

0.50 0.75 1.000.50

0.75

1.00

(H/Hf)

0.670

s

a.

b.

r

(H/Hf)

0.545

Figure 6.6: Determination of the critical exponents n (panel a) and Δ (panel b) after

fitting of δs and εr, respectively, as a power law of the reduced magnetic field H/Hf.

This fact indicates that the obtained values of the critical exponents, = 0.394 and =

4.651, of PSMO are reliable. These values are close to those reported previously by A. K.

Pramanik and A. Banerjee for the same composition [29].

By combining Eqs. (3.21) and (3.22), it is possible to find relationships between

the critical amplitudes M(0,1) and M(1,0), and the values of the temperature and field

scaling functions for ·(t=0) and f·(h=0), respectively:

(0,1) (0)M , (6.2)

(0) 0; (0) ( 1,0)f f M , (6.3)

Page 97: Magnetism in Complex Oxides Probed by Magnetocaloric

81

as has been presented in Figs. 6.7(a) and 6.7(b), 0,1 32.264M emu g1

T1/

and

( 1,0) 98.962M emu g1

, respectively.

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.40

10

20

30

40

50

60

70

-12 -8 -4 0 4 80

50

100

150

200

250

mt*

t*

M(0,1)

=(t*)

(a)

mt*=(t*)

0 1 2 3 4 5 6 7 8 9 100

1

2

3

4

5

6

-4

-2

0

2

4

6

-5 0 5 10

f+

f-

T=TC

T>TC

T<TC

Ln

mh*

Ln h*

Ln(M(-1,0))

f=(h*)1/

(b)

mh*=f

-(h*)

mh*=f

+(h*)

Figure 6.7: Dimensionless renormalized temperature t* and field h* dependence of

the dimensionless temperature renormalized magnetization mt* (a), of the

dimensionless field renormalized magnetization mh* (b in logarithmic scales)

Additionally, taking into account the value of the critical exponent = 1.440, the

representation Ln(jh*)-1

vs. Ln(h*) has been plotted in Figure 6.8 showing that the

experimental data fall on two curves g+ (for T>TC) and g (T<TC), which gives further

evidence of the validity of the fitting procedure used to obtain the critical exponents.

Similarly, by combining the equation jh=g(h) and Eq. (3.19) it is possible to obtain the

relationships between the critical amplitudes 0(1,0) and 0( 1,0) , and the values of the

field scaling functions for 0g h :

(0) (1,0); (0) ( 1,0)o og g , (6.4)

as has been presented in Figure 6.8, 0 1,0 1.716 T emu1

g. In Figs. 6.7(b) and 6.8 it

can be seen that the asymptotic (or critical) values of mh* and jh* evolve as power-law of

Page 98: Magnetism in Complex Oxides Probed by Magnetocaloric

82

field with critical exponents 1/ (according to Eq. (3.18)) and

, respectively. To

achieve as much information about the FM state as possible we need to separate the two

transitions.

-2 0 2 4 6 8 10 12

0

1

2

3

4

5

6

7

g=(h*)- /

g-

g+

Ln(o(1,0))

Ln

(j h

*)-1

Ln h*

T<TC

T>TC

T=TC

Figure 6.8: Inverse of the dimensionless isothermal susceptibility jh*.

In order to be able to de-convolute the field- and temperature-dependent ,SOM T H

curves of the SOMT, subsequently used to obtain their deconvoluted Δ ,SO

MS T H

curves, the Arrott–Noakes equation of state (ANEOS) has been used [39]:

1

1/

/

( )Ca T T bH

MM

,

(6.5)

where the parameters a and b are related to the critical amplitudes of the isothermal initial

susceptibility ( 0(1,0) ), the spontaneous magnetization ( ( 1,0)M ), and critical

magnetization (M(0,1)), along with the critical exponents (, , and ) and Curie

Page 99: Magnetism in Complex Oxides Probed by Magnetocaloric

83

temperature (TC). These relationships can be obtained using Eqs. (3.17), (3.18), (3.19)

and (6.5), according to:

1/(1,0)o

C

aT

, (6.6)

/(0,1)b M , (6.7)

1/( 1,0)

C

M ba

T

,

(6.8)

Recently, it has been shown that the ANEOS is able to reproduce the universal

curve σ for amorphous alloys [18], and also allows predicting of the shape of the ΔSM

curves [40] when the parameters a and b have been obtained from nonlinear fit of the

experimental data, and the critical exponents β and γ from the Kouvel–Fisher iterative

method. Using Eq. (6.5) with the directly obtained parameters a = 6.055 10-3

emu1/

g1/

K1

T1/

, and b = 1.343 10-5

emu/(·)

g/(·)

T1/

predicted by the universal behaviors, the

critical exponents and , and the TC value, the thermomagnetic response of the

deconvoluted SOMT in the magnetic field range up to a value of 0   5 TfH in 10 mT

increments, and in the temperature range from 30 K to 400 K in 0.1 K increments, has

been simulated. To compare with the experimental data, the corresponding simulated data

have been presented with solid lines in Figure 6.9 (a-c). Good agreement between the

experimental and simulated data has been found, which indicates that the ANEOS can be

used to represent the thermomagnetic response of manganites in the vicinity of their TC.

Figure 6.9(c) shows the deconvoluted magnetocaloric response of the SOMT,

Δ ,SO

MS T H , when the ANEOS is used [18, 40].

Page 100: Magnetism in Complex Oxides Probed by Magnetocaloric

84

25 75 125 175 225 275 3250

20

40

60

80

Tf

CO

M (

em

u g

-1)

T (K)

TC

(a)

0 1 2 3 4 5 60

20

40

60

80(b)

T=

30

0 K

M

(e

mu

g-1)

0H (T)

TC=

24

7 K

T=

17

3 K

30 80 130 180 230 280 330 380

0

1

2

3

-S

SO

M (

J K

-1 k

g-1)

T (K)

oH (T)

1

2

3

4

5

Lines: Thoer. data Eq. 15

Symbols: Exp. data

(c)

Figure 6.9: Temperature (a), and field (b), dependence of the magnetization, and

temperature dependence of SM (c). The dark sold lines in (a) and (b) and the solid lines

in (c) are fits to the data via the ANEOS.

We have found good agreement between the experimental and simulated data in the

vicinity of TC, but a noticeable discrepancy for T << TC. This discrepancy likely arises

from the coexistence of AFM and FM phases that are present in the material below the

Page 101: Magnetism in Complex Oxides Probed by Magnetocaloric

85

TC. It is also worth noting that for small field changes ( 0 1 TH ) the experimentally

calculated value of Δ ,SO

MS T H is considerably smaller than the theoretically calculated

Δ ,SO

MS T H , but becomes equal for large field changes ( 0 5 TH ). This can be

attributed to the presence of ferromagnetic clusters [28, 41, 42] and their influence on the

magnetic entropy change and critical exponents near the SOMT in PSMO. We propose

that for small applied magnetic fields, the magnetic field response of magnetization is

weak in materials exhibiting FM clusters thus resulting in a small change in magnetic

entropy, while for large applied magnetic fields, all of the magnetic moments are aligned

with the magnetic field resulting in the overall change in magnetic entropy that is close to

the theoretically predicted value. This explains the small values of Δ ,SO

MS T H reported

for magnetocaloric materials with FM clusters present [43].

The experimentally measured 0 (solid symbol ), along with the predictions

offered by Eq. (2.19) (open symbol ○) or by Eq. (3.4) as 0 Ca T T

(lines), are

presented in Figure 6.10(a). The discrepancy between the experimental and theoretical

data is due to the upward deviation of 0( )T above TC,

This deviation once again suggests the existence of FM correlations/clusters in this

compound above TC. This observation has consequences for the critical behavior of the

system and consequently the equation of state. The effects of this FM interaction above

TC can be taken into account either if a temperature dependence of the parameter a(T) of

the ANEOS is introduced, or if an effective critical exponent, ef, is defined. Figure

6.10(b) presents the temperature dependence of the critical exponent ( )ef T obtained

numerically, demonstrating its tendency towards a constant value of ·=1.440, as

Page 102: Magnetism in Complex Oxides Probed by Magnetocaloric

86

temperature increases as a result of the disappearance of the FM clusters and their

interactions.

0.00

0.05

0.10

0.15

0.20

250 260 270 280 290 300

1.20

1.25

1.30

1.35

1.40

1.45

(b)

-1 o (

T e

mu

-1 g

)

(a)

T>TC (K)

Figure 6.10: (a) Temperature dependence above TC of the inverse of the experimental

isothermal initial susceptibility 0 (solid symbol *), along with the ones offered by Eq.

(6) (open symbol ○) and Eq. (15) (lines). (b) Temperature dependence above TC of the

critical exponent ( )ef T (solid symbol) and the value of ·=1.440 presented in the sample

when the FM clusters and their interactions have disappeared above 300 K (dashed line).

6.3.2 Magnetic Anisotropy and Magnetization Dynamics in Pr0.5Sr0.5MnO3

In order in confirm these results we have also studied the temperature dependence

of effective magnetic anisotropy field (HK, a measure of ferromagnetic correlations) by

Page 103: Magnetism in Complex Oxides Probed by Magnetocaloric

87

using the radio-frequency transverse susceptibility (TS). The HK has been found to persist

in the temperature range TC < T < 300 K, suggesting the existence of FM correlations or

FM clusters in the PM regime and their disappearance above 300 K, which is fully

consistent with the critical analysis and discussions presented above.

As discussed in the introduction, we expect the magnetic anisotropy to provide

considerable additional information on the magnetic interactions in PSMO. Figure 6.11

(a) shows an example of the TS profile of PSMO taken at 160 K. The change in T with

DC magnetic field (Hdc) is expressed from equation (3.28). The plot represents a unique

uni-polar field sweep from positive (+5 kOe) to negative (-5 kOe) fields. As expected,

two symmetric, broad peaks are seen in the scans at the anisotropy fields (HK), while the

sharp peak corresponds to the switching field (HS). Figure 6.11(b) displays a 3D plot of

the magnetic field and temperature dependence of the change in TS for unipolar field

sweeps from positive (+5 kOe) to negative (-5 kOe) fields. It is observed that there are

remarkable changes in peak location (HK) and peak height ([∆/]max) with temperature

(T) at the PM-FM and FM-AFM transition regions.

-4 -3 -2 -1 0 1 2 3 40.2

0.4

0.6

0.8

positive to negative

-HK

HS

HK

(

%)

H (kOe)

T = 160 K

Pr0.5

Sr0.5

MnO3

(a)

Figure 6.11: An example of unipolar transverse susceptibility scan of Pr0.5Sr0.5MnO3 (a).

3-D Unipolar scans of transverse susceptibility as a function of magnetic field and

temperature (b).

-6-4

-2

0

2

4

6

0.0

0.5

1.0

1.5

2.0

2.5

050

100150

200250

300

TC

/

(

%)

T (K)

H (kOe)

TN

(b)

Page 104: Magnetism in Complex Oxides Probed by Magnetocaloric

88

To better illustrate these, the temperature dependence of HK, HS, and [∆/]max

are depicted in Figure 6.12. As one can see clearly in Figure 6.12(a), there is a strong

change in HK in the vicinity of the FM-AFM transition and finite values of HK are

detected over a wide temperature range, even above the TC (in the PM region) and below

the TCO (in the AFM region). To understand this dependence, we recall that the crystal

structure of PSMO favors a Jahn-Teller (JT) distortion that likely induces the d(x2-y

2)-

type orbital order even in the PM state. It has been shown that in the PM region, the

dynamic spin fluctuations are anisotropic due to the polarization of the d(x2-y

2)-type

orbital [27, 28]. Below TC, the PSMO system enters the FM state with an enhanced lattice

distortion, resulting in the change of the polarization of the orbital state. Therefore, we

attribute the persistence of magnetic anisotropy in the PM range (Figure 6.12(a)) to the

anisotropic spin fluctuations resulting from the polarization of the d(x2-y

2)-type orbital.

This is fully consistent with the previous observation based on torque magnetization and

EPR [42], suggesting that the existence of FM clusters in the PM region originates from

the d(x2-y

2)-type orbital order. It is likely that the existence of FM clusters in the vicinity

of the PM-FM transition (the role of magnetic anisotropy) governs the critical exponents

of Pr0.5Sr0.5MnO3 [31]. In this case, the change in shape and size of FM clusters above

and below the TC is expected to reduce the effective dimensionality of spin interaction

(D<3). This may provide some important clues for understanding the unusual critical

behaviors reported in a large class of doped manganites [44-46].

With further decrease in temperature from the TC, the HK first increases sharply at

190 K, reaches a maximum at ~150 K (TCO), decreases sharply at ~ 135 K, and finally

decreases gradually at lower temperatures. Based on previous neutron scattering studies

Page 105: Magnetism in Complex Oxides Probed by Magnetocaloric

89

[27, 28], the AFM spin correlations are known to develop between orbital-ordered planes

in the FM state and progressively increase below 190 K yielding a weak parasitic A-type

AFM order before the system undergoes a first-order transition at TCO to the pure AFM

phase [27, 28].Therefore, we can attribute the sharp increase in HK at ~190 K to the

progressive development of the AFM phase that may result in the formation of highly

anisotropic FM clusters, while the sharp decrease in HK at ~135 K arises mainly from the

strong decrease of the volume fraction of the FM phase. We recall that a structural

transition from tetragonal to orthorhombic symmetry occurs at the TCO in PSMO [31].

The strong variation of HK coupled with the structural change at the TN clearly suggests a

strong coupling between the magnetism and the lattice in this material. The strong

coupling between the structure and magnetic anisotropy has also been studied in

Pr0.5Sr0.5CoO3, which will be discussed in chapter 8.

As shown in Figure 6.12(a), HK increases in the FM region but decreases in the

AFM region with lowering temperature. The temperature dependence of HS (Figure

6.12(b)) also reveals a remarkable variation around the TCO, where the structural

transition has been documented [31]. This observation once again points to the strong

coupling between the magnetism and the lattice in PSMO. The HS is found to increase

with decreasing temperature in both FM and AFM regions.

As one can see clearly in Figure 6.12(c), [∆]max is small and almost constant

in the AFM range (T < TCO). This is consistent with the small permeability of the AFM

phase. It is worth noting from Figure 6.12(c) that there is a sharp increase in [∆]max

at the TCO, which is associated with the increase in magnetic permeability (µ). In the

temperature range of TCO ≤ T ≤ TC, [∆]max increases rapidly with increasing

Page 106: Magnetism in Complex Oxides Probed by Magnetocaloric

90

temperature, as a result of the strong increase in ∆µ. At T ~ TC, a drop in [∆]max is

observed, which is associated with the PM-FM transition. In the PM range, [∆]max is

expected to be zero [29, 47].

0 50 100 150 200 250 3000

300

600

900

1200

0 50 100 150 200 250 3000

1

2

3

0 50 100 150 200 250 3000

100

200

300

400

500

(c)

PM

FM

Hk(O

e)

T(K)

AFM

(a)

PM

FM

AFM

[/

] m

ax (

%)

T (K)

(b)

PM

FM

AFM

Hs (

Oe)

T (K)

Figure 6.12: Temperature dependence of effective anisotropy field (HK), switching field

(HS), and peak height of transverse susceptibility curves ([T/T]max)

However, the non-zero value of [∆]max for Pr0.5Sr0.5MnO3 is retained until T =

300 K (see Figure 6.12(c)). This fully agrees with the temperature dependence of HK and

HS, supporting our argument that FM clusters persist at temperatures above the TC in

PSMO. Finally, we show that the TS is also useful for probing metamagnetic transitions

in PSMO. As one can see clearly in Figure 6.13, a composite plot of the TS and M(H)

data taken at 140 K, the PSMO system is converted from the AFM to the FM state as the

applied magnetic field is increased.

Page 107: Magnetism in Complex Oxides Probed by Magnetocaloric

91

-40 -30 -20 -10 0 10 20 30 400.0

0.2

0.4

0.6

/

(

%)

H (kOe)

T = 140 K

0

20

40

60

80

M (

em

u/g

)

-HC1

-HC2

HC2

HC1

Figure 6.13: Unipolar transverse susceptibility and first magnetization with respect to

applied magnetic field at T=140K.

Here, HC1 corresponds to a critical field at which the AFM state begins

transforming into the FM state and HC2 corresponds to the critical field at which the AFM

state fully converts into the FM one.The clear coincidence between the [∆] and

M(H) data demonstrates the ability of this technique for probing magnetic phase

conversion in metamagnetic systems [48].

6.4 Conclusions

We have demonstrated the usefulness of MCE and TS in studying complex phase

transitions and the magnetic anisotropy in half-doped manganites such as Pr0.5Sr0.5MnO3.

First, we have studied the influence of first- and second-order magnetic phase transitions

on the magnetocaloric effect and refrigerant capacity of charge-ordered Pr0.5Sr0.5MnO3. It

is shown that the first-order magnetic transition at TCO induces a larger MCE but confines

Page 108: Magnetism in Complex Oxides Probed by Magnetocaloric

92

the peak in a narrower temperature limiting the value of RC, while the second-order

magnetic transition at TC induces a smaller maximum MCE over a broader temperature

range resulting in larger RC. Hysteretic losses accompanying the first-order magnetic

transition are very large below TCO and therefore detrimental to the RC, whereas they are

very small or negligible below TC due to the nature of the SOMT. The Maxwell relation,

and non-iterative critical analysis methods have been used to characterize the MCE and

the nature of phase transitions in phase-separated manganites of Pr0.5Sr0.5MnO3. We show

that around the second-order PM-FM transition the SM can be precisely determined from

magnetization measurements using the Maxwell relation, but around the first-order FM-

AFM transition the values of SM are overestimated by the Maxwell relation in the

magnetic field range where a conversion between the AFM and FM phases occurs. The

presence of AFM and FM phase coexistence as well as FM clusters has a significant

impact on the MCE and critical exponents near the SOMT in phase-separated materials.

With TS it is evidenced that there is an abrupt change in magnetic anisotropy at the FM-

AFM transition, which is associated with the structural phase transition that occurs at the

same temperature. This is clear indication of the strong coupling between the magnetism

and the lattice in Pr0.5Sr0.5MnO3.

References:

[1] V.K. Pecharsky, K.A. Gschneidner, Some common misconceptions concerning

magnetic refrigerant materials, J Appl Phys, 90 (2001) 4614-4622.

[2] V. Provenzano, A.J. Shapiro, R.D. Shull, Reduction of hysteresis losses in the

magnetic refrigerant Gd5Ge2Si2 by the addition of iron, Nature, 429 (2004) 853-857.

Page 109: Magnetism in Complex Oxides Probed by Magnetocaloric

93

[3] V.K. Sharma, M.K. Chattopadhyay, S.B. Roy, Large inverse magnetocaloric effect in

Ni50Mn34In16, J Phys D Appl Phys, 40 (2007) 1869-1873.

[4] P. Sande, L.E. Hueso, D.R. Miguens, J. Rivas, F. Rivadulla, M.A. Lopez-Quintela,

Large magnetocaloric effect in manganites with charge order, Appl Phys Lett, 79 (2001)

2040-2042.

[5] P. Chen, Y.W. Du, G. Ni, Low-field magnetocaloric effect in Pr0.5Sr0.5MnO3,

Europhys Lett, 52 (2000) 589-593.

[6] M.S. Reis, V.S. Amaral, J.P. Araujo, P.B. Tavares, A.M. Gomes, I.S. Oliveira,

Magnetic entropy change of Pr1-xCaxMnO3 manganites (0.2 <= x <= 0.95), Phys Rev B,

71 (2005).

[7] R. Venkatesh, M. Pattabiraman, K. Sethupathi, G. Rangarajan, S.N. Jammalamadaka,

Enhanced magnetocaloric effect in single crystalline Nd0.5Sr0.5MnO3, J Appl Phys, 101

(2007).

[8] S. Karmakar, et. al., Magnetocaloric effect in charge ordered Nd(0.5)Ca(0.5)MnO(3)

manganite, J Appl Phys, 103 (2008).

[9] A. Biswas, T. Samanta, S. Banerjee, I. Das, Magnetocaloric properties of

nanocrystalline Pr(0.65)(Ca(0.6)Sr(0.4))(0.35)MnO(3), J Appl Phys, 103 (2008).

[10] A.M. Tishin and Y.I. Spichkin, The Magnetocaloric Effect and its Applications,

Institute of Physics Publishing, Bristol and Philadelphia, 2003.

[11] K.A. Gschneidner, V.K. Pecharsky, Magnetocaloric materials, Annu Rev Mater Sci,

30 (2000) 387-429.

[12] L. Tocado, et.al., Entropy determinations and magnetocaloric parameters in systems

with first-order transitions: Study of MnAs, J Appl Phys, 105 (2009).

Page 110: Magnetism in Complex Oxides Probed by Magnetocaloric

94

[13] N.A.de Oliveria P.J von Ranke, Physics Reports, 489 (2010).

[14] B. Widom, J Chem Phys, 43 (1965).

[15] Griffith.Rb, Thermodynamic Functions for Fluids and Ferromagnets near Critical

Point, Phys Rev, 158 (1967) 176.

[16] V. Franco, J.S. Blazquez, A. Conde, Field dependence of the magnetocaloric effect

in materials with a second order phase transition: A master curve for the magnetic

entropy change, Appl Phys Lett, 89 (2006).

[17] V. Franco, A. Conde, M.D. Kuz'min, J.M. Romero-Enrique, The magnetocaloric

effect in materials with a second order phase transition: Are T-C and T-peak necessarily

coincident?, J Appl Phys, 105 (2009).

[18] V. Franco, A. Conde, J.M. Romero-Enrique, J.S. Blazquez, A universal curve for the

magnetocaloric effect: an analysis based on scaling relations, J Phys-Condens Mat, 20

(2008).

[19] A. Giguere, M. Foldeaki, B.R. Gopal, R. Chahine, T.K. Bose, A. Frydman, J.A.

Barclay, Direct measurement of the "giant" adiabatic temperature change in Gd5Si2Ge2,

Phys Rev Lett, 83 (1999) 2262-2265.

[20] G.J. Liu, J.R. Sun, J. Shen, B. Gao, H.W. Zhang, F.X. Hu, B.G. Shen, Determination

of the entropy changes in the compounds with a first-order magnetic transition, Appl

Phys Lett, 90 (2007).

[21] J.S. Amaral, V.S. Amaral, The effect of magnetic irreversibility on estimating the

magnetocaloric effect from magnetization measurements, Appl Phys Lett, 94 (2009).

[22] J.S. Amaral, V.S. Amaral, On estimating the magnetocaloric effect from

magnetization measurements, J Magn Magn Mater, 322 (2010) 1552-1557.

Page 111: Magnetism in Complex Oxides Probed by Magnetocaloric

95

[23] W.B. Cui, W. Liu, Z.D. Zhang, The origin of large overestimation of the magnetic

entropy changes calculated directly by Maxwell relation, Appl Phys Lett, 96 (2010).

[24] V.K. Pecharsky, K.A. Gschneidner, Giant magnetocaloric effect in Gd5(Si2Ge2),

Phys Rev Lett, 78 (1997) 4494-4497.

[25] R. Kajimoto, H. Yoshizawa, Y. Tomioka, Y. Tokura, Stripe-type charge ordering in

the metallic A-type antiferromagnet Pr0.5Sr0.5MnO3, Phys Rev B, 66 (2002).

[26] H. Kawano-Furukawa, R. Kajimoto, H. Yoshizawa, Y. Tomioka, H. Kuwahara, Y.

Tokura, Orbital order and a canted phase in the paramagnetic and ferromagnetic states of

50% hole-doped colossal magnetoresistance manganites, Phys Rev B, 67 (2003).

[27] R. Mahendiran, M.R. Ibarra, C. Marquina, B. Garcia-Landa, L. Morellon, A.

Maignan, B. Raveau, A. Arulraj, C.N.R. Rao, Giant anisotropic magnetostriction in

Pr0.5Sr0.5MnO3, Appl Phys Lett, 82 (2003) 242-244.

[28] S.X. Cao, B.J. Kang, J.C. Zhang, S.J. Yuan, Orbital-order-induced magnetic

anisotropy and intrinsic phase separation of the Pr0.50+XSr0.50-XMnO3 single crystals, Appl

Phys Lett, 88 (2006).

[29] A.K. Pramanik, A. Banerjee, Critical behavior at paramagnetic to ferromagnetic

phase transition in Pr0.5Sr0.5MnO3: A bulk magnetization study, Phys Rev B, 79 (2009).

[30] K. Knizek, et. al., Structure and Magnetic-Properties of Pr1-xSrxMnO3 Perovskites, J

Solid State Chem, 100 (1992) 292-300.

[31] F. Damay, et. al., Structural transitions in the manganite Pr0.5Sr0.5MnO3, J Magn

Magn Mater, 184 (1998) 71-82.

[32] Y. Tomioka, et. al., Collapse of a Charge-Ordered State under a Magnetic-Field in

Pr1/2Sr1/2MnO3, Phys Rev Lett, 74 (1995) 5108-5111.

Page 112: Magnetism in Complex Oxides Probed by Magnetocaloric

96

[33] M.H. Phan, G.T. Woods, A. Chaturvedi, S. Stefanoski, G.S. Nolas, H. Srikanth,

Long-range ferromagnetism and giant magnetocaloric effect in type VIII Eu8Ga16Ge30

clathrates, Appl Phys Lett, 93 (2008).

[34] M.H. Phan, S.C. Yu, Review of the magnetocaloric effect in manganite materials, J

Magn Magn Mater, 308 (2007) 325-340.

[35] A.N. Ulyanov, J.S. Kim, Y.M. Kang, D.G. Yoo, S.I. Yoo, Oxygen deficiency as a

driving force for metamagnetism and large low field magnetocaloric effect in La0.7Ca0.3-

xSrxMnO3-delta manganites, J Appl Phys, 104 (2008).

[36] O. Tegus, E. Bruck, L. Zhang, Dagula, K.H.J. Buschow, F.R. de Boer, Magnetic-

phase transitions and magnetocaloric effects, Physica B-Condensed Matter, 319 (2002)

174-192.

[37] V.K. Pecharsky, K.A. Gschneidner, Magnetocaloric effect and magnetic

refrigeration, J Magn Magn Mater, 200 (1999) 44-56.

[38] P. Kumar, et. al., Magnetocaloric and magnetotransport properties of R2Ni2Sn

compounds (R = Ce, Nd, Sm, Gd, and Tb), Phys Rev B, 77 (2008).

[39] A. Arrott, J.E. Noakes, Approximate Equation of State for Nickel near Its Critical

Temperature, Phys Rev Lett, 19 (1967) 786-&.

[40] V. Franco, et. al., Magnetocaloric response of FeCrB amorphous alloys: Predicting

the magnetic entropy change from the Arrott-Noakes equation of state, J Appl Phys, 104

(2008).

[41] J.M. DeTeresa, M.R. Ibarra, P.A. Algarabel, C. Ritter, C. Marquina, J. Blasco, J.

Garcia, A. delMoral, Z. Arnold, Evidence for magnetic polarons in the magnetoresistive

perovskites, Nature, 386 (1997) 256-259.

Page 113: Magnetism in Complex Oxides Probed by Magnetocaloric

97

[42] A.K. Pramanik, A. Banerjee, Phase separation and the effect of quenched disorder in

Pr(0.5)Sr(0.5)MnO(3), J Phys-Condens Mat, 20 (2008).

[43] A. Rebello, V.B. Naik, R. Mahendiran, Large reversible magnetocaloric effect in

La0.7-xPrxCa0.3MnO3, J Appl Phys, 110 (2011).

[44] N.A. Frey, S. Srinath, H. Srikanth, M. Varela, S. Pennycook, G.X. Miao, A. Gupta,

Magnetic anisotropy in epitaxial CrO2 and CrO2/Cr2O3 bilayer thin films, Phys Rev B, 74

(2006).

[45] P. Poddar, J.L. Wilson, H. Srikanth, D.F. Farrell, S.A. Majetich, In-plane and out-of-

plane transverse susceptibility in close-packed arrays of monodisperse Fe nanoparticles,

Phys Rev B, 68 (2003).

[46] G.T. Woods, P. Poddar, H. Srikanth, Y.M. Mukovskii, Observation of charge

ordering and the ferromagnetic phase transition in single crystal LSMO using rf

transverse susceptibility, J Appl Phys, 97 (2005).

[47] R.S. Freitas, L. Ghivelder, P. Levy, F. Parisi, Magnetization studies of phase

separation in La0.5Ca0.5MnO3, Phys Rev B, 65 (2002).

[48] H. Srikanth, J. Wiggins, H. Rees, Radio-frequency impedance measurements using a

tunnel-diode oscillator technique, Rev Sci Instrum, 70 (1999) 3097-3101.

[49] N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija, and C. Leighton,

Magnetocaloric effect and refrigerant capacity in charge-ordered manganites Journal of

Applied Physics 106, 023909 (2009)

Page 114: Magnetism in Complex Oxides Probed by Magnetocaloric

98

CHAPTER 7.

PROBING MULTIPLE MAGNETIC TRANSITIONS AND PHASE

COEXISTENCE IN La5/8−xPrxCa3/8MnO3 (x = 0.275) SINGLE CRYSTALS

In the previous chapters, we have demonstrated the versatility of MCE and TS as

useful tools for studying phase transitions of various origins, however, these probes can

also be utilized to study interesting phenomena such as phase separation. In this chapter,

we will show how the multiple magnetic transitions, phase coexistence, and kinetic arrest

of microscale phase-separated La5/8−xPrxCa3/8MnO3 (LPCMO) manganites are probed by

the MCE and TS techniques. Bulk LPCMO is comprised of micron-sized regions of

ferromagnetic (FM), paramagnetic (PM), and charge-ordered (CO) phases. TS and MCE

experiments have evidenced the dominance of low-temperature FM and high-temperature

CO phases. The “dynamic” strain liquid state is strongly dependent on magnetic field,

while the “frozen” strain-glass state is almost magnetic field independent. In combination

with magnetic, magneto-transport, and magnetic force microscopy (MFM)

measurements, MCE studies provide solid evidence that the sharp change in the

magnetization, electrical resistivity, and magnetic entropy just below the Curie

temperature (TC)occur via the growth of FM domains already present in the material even

in zero magnetic field. The subtle balance of coexisting phases and kinetic arrest are also

probed by MCE and TS experiments, leading to a new and more comprehensive magnetic

phase diagram.

Page 115: Magnetism in Complex Oxides Probed by Magnetocaloric

99

7.1 Introduction

La5/8−xPrxCa3/8MnO3 (LPCMO) is a mixture of La5/8Ca3/8MnO3 (x = 0) and

Pr5/8Ca3/8MnO3 (x = 5/8) exhibiting low-temperature ferromagnetic metallic (FMM) and

charge-ordered insulating (COI) ground states, respectively. In this system, the

substitution of smaller Pr ions for larger La ions reduces bandwidth (W), thus leading to

micrometer scale phase separation with multiple phases coexisting in the material [1, 2].

It has been argued that in the presence of quenched disorder induced by the ions of

different radii, the similarity of the free energies leads to the coexistence of the

competing FMM and COI phases [3, 4]. If this is the case, the phase separation should be

static because the phase boundaries are pinned by the disorder sites. However,

experimental studies have revealed that these phase boundaries are not fully pinned in

LPCMO [2, 5-10], and hence the inherent coexistence of the FMM and COI phases at the

micron length scale is inconsistent with the notion of a charge-segregation type of phase

separation [11-13]. It has been suggested that the different crystal structures of the FMM

and COI phases generate long-range strain interactions leading to an intrinsic variation in

elastic energy landscape, which in turn leads to phase separation (PS) in the strain-liquid

and strain-glass regimes [5, 9, 10, 14]. It has also been suggested that phase separated

regions strongly interact with each other via martensitic accommodation strain, which

leads to a cooperative freezing of the combined charge/spin/ strain degrees of freedom [2,

8]. As a result, LPCMO undergoes a transition from the strain-liquid state to the strain

glass state. Since the strain-liquid state shows large fluctuations in resistivity [8, 9] and

can easily be transformed into a metallic state by applying an external field [5, 10], it is

considered a “dynamic” PS. In contrast, effect of electric field is negligible on the frozen

Page 116: Magnetism in Complex Oxides Probed by Magnetocaloric

100

strain-glass state thus classifying it as a “static” PS [5, 10]. Despite a number of previous

works [8, 14, 15], the effect of magnetic field on the strain-liquid and strain-glass states

has not been studied in detail. Another emerging issue, still under debate, is the

underlying origin of the sharp increase in the magnetization below TC in the strain-liquid

region. Two different mechanisms have been proposed for interpreting this observation

[1, 5, 6, 8, 10, 14]. In the first scenario, it is proposed that with lowering temperature, the

COI state is spontaneously destabilized to the FMM phase giving rise to a coexistence of

these two phases, and the sharp increase of the magnetization below TC is due to the

melting of the COI state [8, 14]. This is similar to the case of charge-ordered

R0.5Sr0.5MnO3 (R=La, Pr, and Nd) manganites [11, 16, 17]. In the second scenario, it has

been suggested that the increase in magnetization occurs via the growth of FMM domain

regions that are already present in the material even in zero magnetic field [1, 5, 6, 10].

To shed light on these important unresolved issues, it is essential to employ two

experimental methods that allow detailed investigations of the temperature and magnetic

field responses of the different phases. In this study, we use our ideally suited MCE and

TS experiments for this purpose. Our studies were performed on La5/8−xPrxCa3/8MnO3 (x

= 0.275) single crystals, which were grown in an optical floating-zone furnace and

provided by Professor Sang-Wook Cheong at Rutgers University [7, 18].

7.2 Results and Discussion

7.2.1 Phase Coexistence and Magnetocaloric Effect

Figure 7.1 shows the zero-field-cooled (ZFC) and field-cooled (FC)

magnetization curves taken at 10 mT applied field with the data recorded while warming

up. It is observed that the LPCMO system undergoes multiple magnetic transitions. A

Page 117: Magnetism in Complex Oxides Probed by Magnetocaloric

101

peak at TCO ~ 205 K is due to the COI transition [8], and a shoulder at a lower

temperature of about 175 K arises from antiferromagnetic (AFM) ordering [14]. As T is

further decreased, the magnetization sharply increases and an FMM transition is observed

at TC ~ 90 K. A drop in magnetization is observed at Tg ~ 30 K, below which the system

enters the frozen strain-glass state from the dynamic strain-liquid state [2, 8, 14, 19]. It

has been shown that Tg is actually the re-entrant COI transition temperature [2, 8].

0 50 100 150 200 250 3000.0

0.3

0.6

0.9

Tg

TC

TN

TCO

ZFC

FC

0H = 10 mT

M (

em

u/g

)

T (K)

Figure 7.1: Zero-field-cooled and field cooled M(T) with 10mT applied field, measured

on warming [24].

The MCE measurement process for LPCMO is slightly different than the previous

samples; it appears that large magnetic irreversibility occurs just below TCO, due to the

coexistence of the FMM and COI phases. Therefore, we measured the M(H) curves using

the following measurement protocol to eliminate any influence from the hysteresis.

Before conducting any M(H) measurements at temperatures below TCO, the sample was

cooled in zero magnetic field from above TCO. At each temperature, the magnetization

was measured as the magnetic field was continuously swept from 0 to 6 T (labeled the

Page 118: Magnetism in Complex Oxides Probed by Magnetocaloric

102

virgin M(H) curve); then from 6 T to 0 (labeled the return M(H) curve); and finally from

0 to 6 T (labeled the second M(H) curve). The M(H) data were taken first at 300 K and

subsequently at lower temperatures following the same measurement protocol.

Figure 7.2 shows the M(H) data at selected temperatures with the “virgin,”

“return,” and “second” curves labeled. These data clearly indicate large field hysteresis in

LPCMO. It is worth noting that for T = 75 K, a reversible magnetization is observed as

the applied magnetic field is cycled (i.e., the second M(H) curves coincide with the virgin

M(H) curves). However, for T < 75 K the second M(H) curves do not coincide with the

virgin M(H) curves but coincide with the return M(H) curves below 65 K. Identical

curves are then observed for subsequent field cycles. To capture these intriguing features

from the perspective of MCE, we have calculated the magnetic entropy change from the

M(H) curves taken after completing one cycle of applied field (i.e., the second M(H)

curves). This helps avoid the hysteresis problem (discussed in section 3) when calculating

the magnetic entropy change of LPCMO using equation (1.3).

Figure 7.3 shows the temperature dependence of −SM for the magnetic field

change of 1.5 T and 6 T, respectively. As expected, −SM curves exhibit peaks around

TCO, TC, and Tg. The positive values of −SM around TC and the negative values of −SM

around TCO and Tg (at low applied fields, 0H < 2 T) are consistent with the LPCMO

undergoing the FM, CO and re-entrant CO transitions, respectively.

It is noted in Fig. 7.3 that the −SM has the largest variation at T ~ 75 K (in the

dynamic PS state), while it is comparatively very small at T < Tg (in the frozen PS state).

This indicates that the strain-liquid state is strongly affected by an applied magnetic field,

whereas the strain-glass state is relatively magnetic field independent.

Page 119: Magnetism in Complex Oxides Probed by Magnetocaloric

103

0 1 2 3 4 50

20

40

60

80

100

0 1 2 3 4 50

20

40

60

80

100

0 1 2 3 4 50

20

40

60

80

100

0 1 2 3 4 50

20

40

60

80

100

(b) T = 70 K

2nd

(a) T = 75 K

M (

em

u/g

)

0H (T)

0T -> 5T (virgin)

5T -> 0T (return)

0T -> 5T (second)

virgin

2nd

(d) T = 25 K

2nd

virgin

M (

em

u/g

)

0H (T)

2nd

virgin(c) T = 65 K

M (

em

u/g

)

0H (T)

2nd

virgin

M (

em

u/g

)

0H (T)

Figure 7.2: The M(H) curves for some selected temperatures. The arrows indicate the

way in which the virgin, return, and second magnetization curves were measured [24].

In the strain-liquid region, the large variation of −SM (Fig. 7.3) is attributed to the

suppression of dynamic fluctuations (dynamic phase separation) in magnetic fields. The

large variation of −SM in the dynamic PS region (Fig. 7.3) can also be correlated with

the strong increase in the magnetization below TC (Fig. 7.1). The contribution to the

−SM results, in LPCMO, from the low-field magnetization change in the ferromagnetic

phase and the high-field magnetization change related to the fact that the field-induced

metamagnetic transition takes place in the AFM phase.

To clarify, if the strong increase in the magnetization below TC is attributed to the

destabilization of the COI phase [8, 14] or due to the enhancement of pre-existing FMM

domains in the material [1, 5, 6, 10], we plot in Fig. 7.4 the magnetic field dependencies

Page 120: Magnetism in Complex Oxides Probed by Magnetocaloric

104

of the maximum magnetic-entropy change (−SMmax

) and the magnetization (M) at 75

K.It is observed in Fig. 7.4(a) that the −SMmax

increases rapidly and quite linearly with

increasing H up to 2.6 T and then remains almost constant for H > 2.6 T. This

dependence of −SMmax

(H) can be correlated with the M(H) dependence. We note that at

75 K the COI and FMM phases coexist and both of them are magnetic field dependent.

0 50 100 150 200 250 300

0

3

6

9

(-)

(+)T

N

TC

TCO

Tg

6.0T

1.5T

-S

M (

J/k

g K

)

T (K)

Figure 7.3: Temperature dependence of magnetic entropy change (−∆SM) for LPCMO

for the magnetic field change of 1.5 T and 6 T, respectively [24].

The change in the FMM phase can be achieved at a lower magnetic field while a

higher magnetic field is needed to change the COI phase. As extracted from the M(H)

curve (Fig. 7.4(b)), HS1 = 1.5 T is a critical magnetic field at which the COI phase starts

to convert into the FMM phase while HS2 = 2.6 T is a critical magnetic field at which the

COI phase converts fully into the FMM phase. Therefore it can be concluded that for H <

HS1 the −SMmax

results solely from the variation of the magnetization in the FMM phase,

Page 121: Magnetism in Complex Oxides Probed by Magnetocaloric

105

since the applied magnetic field has a negligible effect on the COI phase. For HS1 < H <

HS2, however, the COI phase converts into the FMM phase, thus also contributing to

−SMmax

. For H > HS2, the constancy of −SMmax

can be attributed to the complete

conversion of the COI phase into the FMM phase. According to this, it is quite natural to

infer, at first glance, that the sharp increase of the magnetization below TC for LPCMO is

due to the destabilization of the COI phase [8, 14].

However, we note that a critical magnetic field needed to fully convert COI into

FMM is often very high for charge-ordered manganites (for example, at T~75 K, HS2 = 12

T for La1−xCaxMnO3 (x = 0.5) (Ref. [1]) and HS2 = 8–17 T for Pr1−xCaxMnO3 (0.3 < x <

0.5) (Ref. [15])). For the case of LPCMO, the volume fraction of the COI phase at 75 K

is large (69%) determined from the M(H) curve in Fig. 7.4(b) (using the same method

employed in Ref. [6]) and the application of a magnetic field of ~2.6 T is unlikely to be

strong enough to convert COI fully into FMM. This can also be reconciled with the fact

that at 75 K the −SMmax

resulting from the variation of the magnetization in the FM

phase (~6.49 J/kg K) is about twice larger than that resulting from the COI→FMM

conversion (~2.44 J/kg K).

These findings clearly suggest that the sharp increase in the magnetization below

TC in LPCMO cannot be due to the destabilization of the COI phase [8, 14], but instead

can be attributed to the enhancement of the pre-existing FMM domain regions [1, 5-7,

10]. The hysteresis appearing below HS2 ~ 2.6 T is the result of the coexistence of the

COI and FMM phases, whereas the application of higher fields (H > 2.6 T) completely

suppresses the COI phase and as a result the FMM phase with no hysteresis is observed at

these magnetic fields.

Page 122: Magnetism in Complex Oxides Probed by Magnetocaloric

106

0 1 2 3 4 52

4

6

8

10

0 1 2 3 4 50

20

40

60

80

100

(III)(II)

FM

CO + FM

T = 75 K

Phase coexistence

[-S

M] m

ax (

J/k

g K

)

0H (T)

Hs2

= 2.6 T

CO + FM

Phase coexistence

(I)

(a)

(b)

Hs1

= 1.5 T

Hs2

= 2.6 T

T = 75 K

FMCO + FM

CO + FM

M (

em

u/g

)

0H (T)

Figure 7.4: (a) Magnetic field dependence of maximum magnetic-entropy change

([−SM]max) for LPCMO at 75 K; (b) the magnetic hysteresis loop M(H) measured at 75K

[24].

Finally, we note that the subtle balance between the competing COI and FMM

phases in LPCMO is readily affected by applied magnetic field, and study of such a

balance can be of great importance in elucidating the physical origin of magnetic/electric

field-induced “colossal” effects [5, 9, 10]. Here, we show that the change in magnitude

Page 123: Magnetism in Complex Oxides Probed by Magnetocaloric

107

and sign of -SM can be an indicator of the intricate balance between the COI and FMM

phases as seen in the change in sign of -SMmax

at 205 K (~TCO) as the applied magnetic

field is increased. As shown in Fig. 7.5(a), the -SMmax

is negative and first increases in

magnitude with increasing H up to HC1~2.2 T, and then decreases and reaches zero at HC2

~3.1 T. For H > HC2, it is positive and increases gradually with increasing H up to HC3 ~

3.9 T and finally increases rapidly for H > HC3. Here HC1 = 2.2 T is a critical magnetic

field at which the COI phase starts to convert into the FMM phase, HC2 = 3.1 T is a

critical magnetic field at which the half of the COI phase converts into the FMM phase,

and HC3 = 3.9 T is a critical magnetic field at which the COI phase converts fully into the

FMM phase. The magnetic field dependence of -SMmax

can be interpreted as follows. For

H < HC1, the applied magnetic field is not strong enough to convert the COI phase into

the FMM phase, so the negative -SMmax

and its increase with H result from the

contribution of the COI phase. However, for HC1 < H < HC2, the positive contribution to -

SMmax

from the FMM phase becomes significant because the COI phase is partially

converted into the FMM phase. Since the contribution from the FMM phase is opposite

to that from the COI phase, the sum of the two components lead to a decrease in

magnitude of the negative -SM with H in the range HC1 < H < HC2.

In other words, both the COI and FMM phases coexist but the COI phase is

dominant over the FMM phase, since the sign of -SM is negative. At H=HC2, the positive

and negative contributions to -SMmax

from the COI and FMM phases are equal or

compensated and so -SMmax

crosses zero. For HC2 < H < HC3, the COI phase is largely

converted into the FMM phase which now dominates over the COI phase leading to a

positive -SM. For H > HC3, the COI phase is fully converted into the FMM phase leading

Page 124: Magnetism in Complex Oxides Probed by Magnetocaloric

108

to a rapid increase in magnitude of positive -SM. The values of HC1, HC2, and HC3

coincide with the critical magnetic fields determined from the M(H) curve (see Fig.

7.5(b)).

Figure 7.5: (a) Magnetic field dependence of maximum magnetic-entropy change

([−SM]max) for LPCMO at 205 K; (b) the magnetic hysteresis loop M(H), measured at

205 K [24].

The hysteresis seen in the M(H) curve between HC1 and HC3 (Figs. 7.2(b) and 7.5(b)) is

fully consistent with the coexistence of the COI and FMM phases as already revealed by

the MCE data (Fig. 7.5(a)). These results provide an important understanding of the

0 1 2 3 4 5 6

-0.4

0.0

0.4

0.8

1.2

+

FM

+ FM

CO

HC3

= 3.9 T

HC1

= 2.2 T

CO

FM

[-S

M] m

ax (

J/kg

K)

0H (T)

T = 205 K

HC2

= 3.1 T

(+)

(-)

CO

(a)

-6 -4 -2 0 2 4 6-100

-50

0

50

100

Hc3

= 3.9 T

Hc2

= 3.1 T

M (

em

u/g

)

0H (T)

T = 205 K

Hc1

= 2.2 T

(b)

Page 125: Magnetism in Complex Oxides Probed by Magnetocaloric

109

physical origin of the magnetic/electric field induced “colossal” effects, including

colossal magnetoresistance and large magnetocaloric effects in mixed-phased manganites

[9, 20-22].

7.2.2 Transverse Susceptibility

To understand the magnetization dynamics of LPCMO we have also performed

TS measurements in these LPCMO (x=0.275) single crystals. Figure 7.6 (a-c) shows the

change in T as a function of applied DC field (HDC) below the TCO. As seen from the -

SM data (Fig. 7.2) there is no appreciable dependence of -SM on external field at

temperatures below Tg, due to the system being in the static PS regime. However, when

probing near (and above) TC (Fig. 7.6 (b-f)), in the dynamic PS region, we see the

emergence of a sharp drop in the TS data, associated with the conversion of COI to

FMM, as seen from M(H) and –SM(T) (Figs 7.4, 7.5). At T < TC, the shape of TS spectra

remains almost the same with no field hysteresis as the HDC is swept between +5T and -

5T and vice versa, indicating a magnetic field-assisted kinetic arrest phenomenon as also

reported previously for Gd5Ge2 alloys [23]. It is worth noting that unlike the case of

Gd5Ge2, this kinetic arrest occurs just below TC in the LPCMO system. This finding

points to the important fact that the spin dynamics of the LPCMO system are frozen out

by the applied field even in the dynamic region, as the magnetic energy dominates over

the thermal and strain energies. As the temperature is increased further above TC, the drop

in TS occurs at higher values of HDC, signifying that LPCMO is in an increasingly stable

COI state above TC. It is noted that a largest field hysteresis (the area enclosed by the

increasing (blue) and decreasing (red) TS curves) is observed around 70 K, which is

associated with the perspective of the strongest phase separation that occurs in this

Page 126: Magnetism in Complex Oxides Probed by Magnetocaloric

110

temperature range. The field hysteresis is largely suppressed with increasing temperature,

which is consistent with the fact that the COI phase becomes dominant at the expense of

the FMM phase at high temperatures.

Having considered the variation in the magnetization dynamics of the LPCMO

system after the DC magnetic field is cycled, we plot in Fig. 7.7(a) the temperature

dependence of the switching fields (HS+ and HS

-) associated with the conversion between

the COI and FMM phases on increasing and decreasing DC fields, leading to a more

accurate H-T phase diagram in the dynamic and static PS regions at temperatures below

TCO.

Figure 7.6: (a-d): Bipolar TS scans below TC (a) and above TC (b-d).

It can be seen that below TC, after the material saturated, the COI phase is fully

suppressed and its dynamics are kinetically arrested. As a result, the material now

behaves as a soft ferromagnet irrespective of variation in applied field. As temperature

-60 -40 -20 0 20 40 60

0

1

2

3

T

T (

%)

H (kOe)

decreasing

increasingT = 150 K

-60 -40 -20 0 20 40 60

0

2

4

T

T (

%)

H (kOe)

decreasing

increasing

T = 90 K

(c)

-60 -40 -20 0 20 40 60

0

2

4

T

T (

%)

H (kOe)

decreasing

increasing

T = 80 K

(b)

-60 -40 -20 0 20 40 60

0.0

0.7

1.4

2.1

T (

%)

H (kOe)

decreasing

increasing

T = 60 K

(a)

(

c)

(

f)

Page 127: Magnetism in Complex Oxides Probed by Magnetocaloric

111

is increased further, the emergence of two competing phases becomes more evident,

where the drastic drop in TT is mainly due to the growth of the FMM domains inside

the COI ones.

Figure 7.7: New phase diagrams for LPCMO developed from TS vs. HDC (a) Positive

and negative switching field as a function of temperature. (b) Maximum change in TS at

HDC=0 as a function of temperature.

Probing even deeper insights into the temperature dependence of TS, we plot the

[TT]max at zero applied DC field (HDC = 0) (Fig. 7.7 (b)) and in so doing, realize a

new temperature dependent phase diagram for LPCMO. At high temperature [TT]max

~ 0, due to the small permeability in the PM region. On decreasing the temperature a

small increase in [TT]max is observed around TCO, signifying the beginning of the

dynamic PS region, where we begin to see FMM clusters forming in the COI region. The

curve dramatically increases to a maximum at TC, followed by a large drop off at Tg. The

dramatic increase below TCO is a signature of FMM domains nucleating in the sample,

followed by the large drop-off associated with the freezing of the PS region. The

strongest variation in [TT]max at T ~ TC reflects the fact that the most robust phase

separation occurs at this temperature. An important consequence that emerges from these

new TS phase diagrams is the absence of peaks associated with effective anisotropy

0 30 60 90 120 150 1800

1

2

3 H

+

S(T)

H-

S(T)

0H

(T

)

T (K)

CO

CO +

FM

FM

(a)

0 50 100 150 200 250 3000

1

2

3

4

5

FS

FM + CO

/

T] m

ax (

%)

T(K)

PM

(b)

Page 128: Magnetism in Complex Oxides Probed by Magnetocaloric

112

fields (HK), which are often observed in conventional FM systems such as Pr0.5Sr0.5MnO3

(Chapter 6) and Pr0.5Sr0.5CoO3 (Chapter 8), indicating that the FMM phase is not

stabilized in the bulk form of LPCMO. In other words, the phase separation in the

LPCMO system is dynamic and the phase boundaries are not pinned by the disorder sites.

7.3 ConclusionsIn summary, systematic magnetocaloric measurements on

La5/8−xPrxCa3/8MnO3 single crystals have revealed important insights into the complex

multiple-phase transitions. The system is ferromagnetic at low temperature and becomes

charge ordered at high temperature. The dynamic strain-liquid phase is strongly affected

by an applied magnetic field, whereas the frozen strain-glass phase is nearly magnetic

field independent. The origin of the large MCE in the strain-liquid region arises from the

suppression of dynamic fluctuations in magnetic fields. The MCE data clarify that the

sharp increase in the magnetization below TC may not be due to the destabilization of the

COI phase to the FMM phase, but favors the idea of the growth of pre-existing FMM

domain regions. TS experiments provide evidence for the instability of the FMM phase

and the unusual magnetic field-assisted kinetic arrest of the magnetization in this system.

Overall, MCE and TS have proven to be excellent probes of the magnetic transitions and

ground-state magnetic properties of phase-separated systems like LPCMO.

References:

[1] M. Kim, H. Barath, S.L. Cooper, P. Abbamonte, E. Fradkin, M. Ruebhausen, C.L.

Zhang, S.W. Cheong, Raman scattering studies of the temperature- and field-induced

melting of charge order in La(x)Pr(y)Ca(1-x-y)MnO(3), Phys Rev B, 77 (2008).

Page 129: Magnetism in Complex Oxides Probed by Magnetocaloric

113

[2] P.A. Sharma, S. El-Khatib, I. Mihut, J.B. Betts, A. Migliori, S.B. Kim, S. Guha, S.W.

Cheong, Phase-segregated glass formation linked to freezing of structural interface

motion, Phys Rev B, 78 (2008).

[3] E. Dagotto, Complexity in strongly correlated electronic systems, Science, 309 (2005)

257-262.

[4] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki, E. Dagotto, Giant cluster coexistence in

doped manganites and other compounds, Phys Rev Lett, 84 (2000) 5568-5571.

[5] T. Dhakal, J. Tosado, A. Biswas, Effect of strain and electric field on the electronic

soft matter in manganite thin films, Phys Rev B, 75 (2007).

[6] V. Kiryukhin, B.G. Kim, V. Podzorov, S.W. Cheong, T.Y. Koo, J.P. Hill, I. Moon,

Y.H. Jeong, Multiphase segregation and metal-insulator transition in single crystal La5/8-

yPryCa3/8MnO3, Phys Rev B, 63 (2001).

[7] H.J. Lee, et. al., Optical evidence of multiphase coexistence in single crystalline

(La,Pr,Ca)MnO3, Phys Rev B, 65 (2002).

[8] P.A. Sharma, S.B. Kim, T.Y. Koo, S. Guha, S.W. Cheong, Reentrant charge ordering

transition in the manganites as experimental evidence for a strain glass, Phys Rev B, 71

(2005).

[9] M. Uehara, S. Mori, C.H. Chen, S.W. Cheong, Percolative phase separation underlies

colossal magnetoresistance in mixed-valent manganites, Nature, 399 (1999) 560-563.

[10] S.H. Yun, et. al., Giant positive magnetoresistance in ultrathin films of mixed phase

manganites, J Appl Phys, 103 (2008).

[11] Y. Tokura, Y. Tomioka, Colossal magnetoresistive manganites, J Magn Magn

Mater, 200 (1999) 1-23.

Page 130: Magnetism in Complex Oxides Probed by Magnetocaloric

114

[12] C.N.R. Rao, B. Raveau, Colossal magnetoresistance, charge ordering and related

properties of manganese oxides, World Scientific, Singapore ; River Edge, N.J., 1998.

[13] M. Uehara, S.W. Cheong, Relaxation between charge order and ferromagnetism in

manganites: Indication of structural phase separation, Europhys Lett, 52 (2000) 674-680.

[14] L. Ghivelder, F. Parisi, Dynamic phase separation in La5/8-yPryCa3/8MnO3, Phys Rev

B, 71 (2005).

[15] I.G. Deac, S.V. Diaz, B.G. Kim, S.W. Cheong, P. Schiffer, Magnetic relaxation in

La0.250Pr0.375Ca0.375MnO3 with varying phase separation, Phys Rev B, 65 (2002).

[16] H. Kuwahara, et. al., A first-order phase-transition induced by a magnetic-field,

Science, 270 (1995) 961-963.

[17] Y. Tomioka, et.al., Collapse of a charge-ordered state under a magnetic-field in

Pr1/2Sr1/2MnO3, Phys Rev Lett, 74 (1995) 5108-5111.

[18] K.H. Kim, M. Uehara, C. Hess, P.A. Sharma, S.W. Cheong, Thermal and electronic

transport properties and two-phase mixtures in La5/8-xPrxCa3/8MnO3, Phys Rev Lett, 84

(2000) 2961-2964.

[19] W. Wu, C. Israel, N. Hur, S. Park, S.-W. Cheong, A. De Lozanne, Magnetic imaging

of a supercooling glass transition in a weakly disordered ferromagnet, Nat Mater, 5

(2006) 881-886.

[20] L. Ghivelder, et. al., Abrupt field-induced transition triggered by magnetocaloric

effect in phase-separated manganites, Phys Rev B, 69 (2004).

[21] A.L.L. Sharma, P.A. Sharma, S.K. McCall, S.B. Kim, S.W. Cheong, Enhanced

magnetic refrigeration capacity in phase separated manganites, Appl Phys Lett, 95

(2009).

Page 131: Magnetism in Complex Oxides Probed by Magnetocaloric

115

[22] J. Tosado, T. Dhakal, A. Biswas, Colossal piezoresistance in phase separated

manganites, J Phys-Condens Mat, 21 (2009).

[23] M.K. Chattopadhyay, M.A. Manekar, A.O. Pecharsky, V.K. Pecharsky, K.A.

Gschneidner, J. Moore, G.K. Perkins, Y.V. Bugoslavsky, S.B. Roy, P. Chaddah, L.F.

Cohen, Metastable magnetic response across the antiferromagnetic to ferromagnetic

transition in Gd5Ge4, Phys Rev B, 70 (2004).

[24] M. H. Phan, M. B. Morales, N. S. Bingham, and H. Srikanth, C. L. Zhang and S. W.

Cheong, Phase coexistence and magnetocaloric effect in La5/8-yPryCa3/8MnO3

(y=0.275). Physical Review B, 2010. 81(9).

Page 132: Magnetism in Complex Oxides Probed by Magnetocaloric

116

CHAPTER 8.

MAGNETOCALORIC EFFECT AND TRANSVERSE SUSCEPTIBILITY OF Pr1-

xSrxCoO3 (x =0.3-0.5): IMPACT OF THE MAGNETOCRYSTALLINE

ANISOTROPY-DRIVEN PHASE TRANSITION

In this chapter we demonstrate that the TS can be used as a powerful probe of the

structurally coupled magnetocrystalline anisotropy in complex oxides like Pr0.5Sr0.5CoO3,

which undergoes a PM-FM phase transition at TC ~235 K followed by a structurally

coupled magnetocrystalline anisotropy transition at TA ~120 K. Our findings point to the

existence of a distinct class of phenomena in correlated materials due to the unique

interplay between structure and magnetic anisotropy. Since the structural change at the TA

in Pr0.5Sr0.5CoO3 is not associated with any magnetic transition, Pr0.5Sr0.5CoO3 provides

an excellent system for determining the structural entropy change and its contribution to

the MCE in magnetocaloric materials. Having systematically studied the influence of the

structurally coupled magnetocrystalline anisotropy transition on the MCE in Pr1-xSrxCoO3

(x = 0.3, 0.35, 0.4, and 0.5) compounds, we have demonstrated, for the first time, that the

structural entropy contributes significantly to the total entropy change and the structurally

coupled magnetocrystalline anisotropy plays a crucial role in tailoring the magnetocaloric

properties for active magnetic refrigeration technology [1]. Our study has shed light on

one of the most challenging issues in the research field of magnetocaloric materials.

Page 133: Magnetism in Complex Oxides Probed by Magnetocaloric

117

8.1 Introduction

Although relatively less studied than the manganites, cobaltites of the formula R1-

xMxCoO3 (R= Lanthanide, M = Alkaline-Earth) present interesting characteristics,

perhaps the most well-known example being the spin-state transition in LaCoO3 [2-4].

The presence of Co on the perovskite B-site leads to an additional spin-state degree of

freedom due to similar magnitudes of the crystal field and Hund’s rule exchange

energies. This, along with the significantly larger magnetocrystalline anisotropy, makes

the study of cobaltites intriguing, both for fundamental understanding as well as device

applications, for which manipulation of the anisotropy is desirable.

Half-doped Pr1-xSrxCoO3 (x=0.5) is known to exhibit particularly unusual

magnetic behavior that is not consistent with the phase behavior often seen in manganites

and other complex oxide systems [5-9]. This system undergoes a PM-FM phase transition

at TC ~235 K and a magnetic anomaly in the field-cooled magnetization versus

temperature profiles is observed at TA ~120 K. In order to understand this anomaly,

systematic studies were recently undertaken to rule out the phase transitions that are most

routinely associated with perovskites such as charge ordering, antiferromagnetic

ordering, ferrimagnetism, or spin-flip transitions [10].

It was conclusively shown that all of the observed behavior can be explained by a

ferromagnetic to ferromagnetic (FM-FM) transition resulting from a structural change

that drives a transition in the magnetocrystalline anisotropy. Because this FM-FM

transition is not seen in transport measurements, and traditional magnetometry

measurements provided minimal information on the nature of the magnetocrystalline

anisotropy, we show below that the TS measurement technique is extremely well-suited

Page 134: Magnetism in Complex Oxides Probed by Magnetocaloric

118

to explore this particular structure-driven magnetocrystalline anisotropy transition.

Furthermore, it has been noted that the largest MCEs are observed in materials exhibiting

a FOMT coupled with a crystal structure change [9].

Since the magnetic and structural changes are often coupled with each other, it is

challenging to decouple the structural entropy contribution from the magnetic entropy

contribution to the total MCE. Since the structural change at the TA in Pr0.5Sr0.5CoO3 is

not associated with any magnetic transition, Pr0.5Sr0.5CoO3 provides an excellent system

for determining the structural entropy change and its contribution to the MCE in

magnetocaloric materials. Our systematic study of the influence of the structurally

coupled magnetocrystalline anisotropy transition on the MCE in Pr1-xSrxCoO3 (x = 0.3,

0.35, 0.4, and 0.5) compounds has addressed this outstanding issue. The polycrystalline

samples were supplied by Professor Christopher Leighton from the University of

Minnesota.

8.2 Results and Discussion

8.2.1 Anomalous magnetism and Magnetocaloric effect in Pr1-xSrxCoO3 (0.3 ≤ x ≤

0.5)

Figure 8.1 shows the M(T) curves for PSCO with various substitution values

recorded while cooling under a high field (0H=5T). While cooling, there exists a sharp

increase in M signifying a second order PM to FM transition at TC ~190K, 200K, 220K,

230K or x=0.3, 0.35, 0.4, 0.5 respectively. However, for smaller cooling fields, (inset of

Fig. 8.1), there is an appearance of a second transition. While cooling we see first a large

increase in the magnetization at the TC, followed by a decrease (increase) in

magnetization at lower temperature (TA~120K) under cooling fields of less than (greater

Page 135: Magnetism in Complex Oxides Probed by Magnetocaloric

119

than) 750 Oe [3]. At cooling fields in which the magnetization is saturated, no anomaly is

observed in the M(T). Note that the decrease (increase) in magnetization upon cooling in

low (high) field is manifest as a gradual change in curvature starting at around TA and

persists well into the low temperature regime.

0 50 100 150 200 250 3000.0

0.5

1.0

1.5

M

(

B/a

t.C

o)

T (K)

x=0.3

x=0.35

x=0.4

x=0.35

0H=5 T

0 50 100 150 200 2500.00

0.05

0.10

0.15

0H=1 mT

TA=120 K

0.0

0.5

1.0

0H=0.1 T

x=0.5

Figure 8.1: Temperature dependence of the magnetization of Pr1−xSrxCoO3 (x=0.3, 0.35,

0.4, and 0.5) compounds when a magnetic field 0H=5 T is applied. Inset: Temperature

dependence of the magnetization at low field (0H=1 mT) and intermediate field

(0H=0.1T) in the x=0.5 sample.

The ~120 K transition has been attributed to a strong coupling between

structural/magnetocrystalline anisotropy [11]. Because of the structural transition that

occurs at TA there is a large change in the direction of the easy axis of magnetization. It

has been shown from previous work [11] that the sample exhibits no change in overall

crystal symmetry, or even any non-negligible change in unit cell volume, however, there

is a drastic change in the lattice parameters (a~+1.15% and b~-1.10%) [4, 12] near the

transition at TA. The structural transition is believed to be due to Pr-O hybridization [13].

When Lopez-Morales et.al. examined PrBa2Cu3O7-, the only rare-earth 1:2:3 system that

Page 136: Magnetism in Complex Oxides Probed by Magnetocaloric

120

is insulating rather than metallic or superconducting, they found that the unusual

tendency of Pr 4f–O 2p hybridization is the cause of the nonsuperconducting state. The

lack of change in crystal symmetry in PSCO is drastically different from what is observed

in materials with similar properties [14-16]. Also, since the two transitions are well

separated in T, PSCO provides a rare opportunity to study the effect of a structural

transition on the MCE.

0 1 2 3 4 5

0.0

0.5

1.0

1.5

2.0

T=

32

0 K

TC=

23

0 K

M (

B/a

t.C

o)

0H (T)

T=

5 Kx=0.5

Figure 8.2: Field dependence, from 5 K to 320 K in 5 K increments, of the magnetization

of the polycrystalline Pr0.5Sr0.5CoO3 compounds. The magnetization curve is marked

(open symbol) at the Curie temperature of the sample TC (x = 0.5) =230 K.

Figure 8.2 shows the isothermal magnetization versus applied field (M(H)) as a

function of temperature for half-doped PSCO, used for calculating the SM. Since, the

equations used integrate between successive M(H) curves, the large gaps between

successive curve are associated with peaks in SM. The open symbols represent M(H) at

TC, interestingly, above TC, there appears to be a minor ferromagnetic signal attributed to

ferromagnetic clusters.

Figure 8.3, shows the calculated change in magnetic entropy for x = 0.3-0.5. The

first peak at high temperature at Tpk~190K, 200K, 220K, 230K for x = 0.30, x = 0.35, x =

Page 137: Magnetism in Complex Oxides Probed by Magnetocaloric

121

0.40, x = 0.50, respectively, is associated with the PM to FM transition. However, for x =

0.4 and 0.5, we see a rather large change in entropy at lower temperatures TA~70K and

120K for x = 0.4 and 0.5 respectively, which is associated with the

structural/magnetocrystalline anisotropy transition. Interestingly, the samples that exhibit

the second peak also show an enhanced and broadened MCE at TC, suggesting that there

is a clear advantage to coupling between a magnetic and structural transition from the

perspective of achieving the maximum RC.

0 50 100 150 200 250 3000.0

0.5

1.0

1.5

2.0

2.5

-100 -50 0 50 100

0.5

1.0

1.5

2.0

2.5

TA=120 K

-S

M (

J K

-1 k

g-1)

T (K)

x=0.3

x=0.35

x=0.4

x=0.5

0H=5 T

TA=70 K

T-TC (K)

Figure 8.3: Temperature dependence of the magnetic entropy change for 0H=5 T of the

Pr1−xSrxCoO3 (x = 0.3, 0.35, 0.4, and 0.5) compounds. Solid arrows indicate the

temperatures (TA) of the second phase transition that occurs at low temperature Inset:

Reduced temperature dependence of the magnetic entropy change near TC.

From Fig. 8.3 we can also see that, depending on the doping concentration, the main peak

can be shifted to lower (higher) temperatures, which is quite useful for industrial

applications of magnetic refrigeration. For the peak at TC the ΔSM taken at 5T remains

almost unchanged for x = 0.3 and 0.35 (1.41 J/kg K) but increases for x = 0.4 (1.67 J/kg

K) and for x = 0.5 (2.2 J/kg K). For the peak at TA, it increases from 0.29 J/kg K for x =

0.4 to 0.79 J/kg K for x = 0.5.

Page 138: Magnetism in Complex Oxides Probed by Magnetocaloric

122

According to the experimental results and theoretical validation [17-19], it can be

assumed that in a biphasic system the field dependence of ΔSM and RC follow power

laws of the field;

1( , )( , ) ( )M

T HnS T H a T H , (8.1)

( ) ( ) , (8.2)

where the exponent n1 depends, in general, on temperature and field, and its asymptotic

values are 1 and 2 when the values of temperature are quite far below and above TC,

respectively. The exponent n1 takes minimum values around the Curie temperatures of

the existing phases and an extreme value in the temperature range defined by the critical

temperatures of the constituent phases [20]. In a first approximation, this exponent can be

considered field independent at the temperature of TC and Tpk, and can be expressed in

terms of the critical exponents as [17] 1 1( ) ( ) 1 (1 ) / ( )C pkn T n T . On the

other hand, the exponent n2 that controls the field dependence of the RC is related to the

critical exponents by 2 1 / ( )n . Note that the exponents n1 and n2 are related

via the expression 2 1 1/ ( )n n .

Taking into in account that only two critical exponents are independent, all of

them can be calculated from n1 and n2. For instance, the critical exponents and can be

obtained according to

2 2

2 1 2 1

1 2;

n n

n n n n

. (8.3)

The field dependence of M

pkS and RC has been presented in Figs. 8.4a,c for the

whole studied compositional range, indicating two different, well-defined behaviors for

the x=0.3 and x=0.35 single phase systems (approximately the same M

pkS and RC

Page 139: Magnetism in Complex Oxides Probed by Magnetocaloric

123

values), and for the x=0.4 and x=0.5 biphasic systems (for which those values are

different). As can be seen also in the inset of Fig. 8.3 for the single-phase systems

(approximately the same ΔSM(T) values) and for the biphasic systems (approximately the

same ΔSM(T-TC) values in the temperature range between the critical temperatures of the

two constituent phases).

0.0

0.5

1.0

1.5

2.0

2.50.00 0.25 0.50 0.75 1.00

0.25

0.50

0.75

1.00

0 1 2 3 4 5

0

25

50

75

100

0.00 0.25 0.50 0.75 1.000.00

0.25

0.50

0.75

1.00

-S

pk

M (

J K

-1 k

g-1)

x=0.3

x=0.35

x=0.4

x=0.5

a.

b.

s

h0.76

c.

RC

(J k

g-1)

0H (T)

d.rc

h1.33

Figure 8.4: Field dependence of the maximum magnetic entropy change (a) and the

refrigerant capacity (c) in the studied polycrystalline Pr1−xSrxCoO3 (x=0.3, 0.35, 0.4, and

0.5) compounds. Dimensionless field dependence of the dimensionless maximum

magnetic entropy change s (b), and dimensionless refrigerant capacity, rc (d). The non-

collapse into two master curves indicates that the exponents n1 and n2 are composition

dependent.

In order to compare the field evolution of the experimental data presented in Fig.

8.4a-c, it would be necessary to eliminate the factors of Eqs. (8.1) and (8.2) that depend

Page 140: Magnetism in Complex Oxides Probed by Magnetocaloric

124

only on the composition and on the previously defined temperature span T (in this work

T=TFWHM for the whole composition range). By normalizing these expressions with the

values corresponding to the maximum applied field, dimensionless relationships can be

written for the different studied compositions [21].

1

1

1

2

2

2

max max

max max

( , ) ( )

( , ) ( )

( , , ) ( , )

( , , ) ( , )

pk

M

pk

M

AREA

AREA

nn

n

nn

n

S H x a x Hs h

S H x a x H

RC T H x b T x Hrc h

RC T H x b T x H

, (8.4)

where max/h H H . When the latter two power laws are plotted for the different

compositions (Figs. 8.4b,d), it can be seen that in general the exponents n1(x) and n2(x)

are composition dependent.

While Fig. 8.4(b) shows that the exponent n1 takes the values n1(x=0.3)=

n1(x=0.35)= n1(x=0.4)=0.76 and n1(x=0.5)=0.58, the Fig. 8.4(d) indicates that the

exponent n2 varies slightly in the full series taking the values n2(x=0.3)=1.33,

n2(x=0.35)=1.31, n2(x=0.4)=1.24 and n2(x=0.5)=1.23. According to Eq. (8.3), this means

that the critical exponent takes a value = 0.54 0.05 close to that in the mean field

theory (MFT=0.5) for x=0.3, 0.35 and 0.4, and a value = 0.358 0.001 close to that in

the Heisenberg model (H = 0.365 0.003) for x=0.5. These results reveal that the

magnetic interaction in the x=0.5 cobaltite is of short-range type. The smaller values of

the uncertainly for x=0.5 is due to the higher resolution of the field used for this

composition. On the other hand, the critical exponent takes increasing values when the

doping increases (x = 0.3) = 1.17 0.04, (x = 0.35) = 1.25 0.04, (x = 0.4) = 1.58

0.04, and decreasing value for the half doped cobaltite ( x = 0.5) = 1.178 0.003, very

similar to that of x=0.3.

Page 141: Magnetism in Complex Oxides Probed by Magnetocaloric

125

8.2.2 Transverse susceptibility as a probe of the coupled

structural/magnetocrystalline anisotropy transition in Pr1-xSrxCoO3 (x = 0.5)

TS measurements were also performed on the Pr0.5Sr0.5CoO3 sample at a number

of temperatures to examine the temperature dependence of the anisotropic features across

TA. Figure 8.5 shows bipolar TS scans of Pr0.5Sr0.5CoO3 taken at four representative

temperatures: 20 K (8.5a), 95 K (8.5b), 110 K (8.5c), and 225 K (8.5d).

Figure 8.5: Bipolar transverse susceptibility scans of Pr0.5Sr0.5CoO3 as a function of

applied field for 20K (a), 95K (b), 110K (c), and 225K (d). On 8.5(a) the arrows indicate

the sequence of measurement; the anisotropy (Hk), crossover (Hcr), and switching (HS)

peaks are labeled [22].

The broader, high-field peaks seen on either side of μ0H=0, closest to saturation,

are the anisotropy peaks indicating the anisotropy fields, ±HK. The second peak observed

upon decreasing the field after positive saturation corresponds to the “crossover field”,

Hcr, which is defined as the field which separates the lower FC magnetization state from

the higher FC magnetization state for any given temperature occurring below TA, as

Page 142: Magnetism in Complex Oxides Probed by Magnetocaloric

126

described above. The third peak observed is the prominent switching peak, HS.Figure

8.6a shows the temperature evolution of the TS curves for phase FM1 and Fig. 8.6b

shows the temperature evolution of the TS curves for FM2. Unlike in Fig. 8.5, the

magnitude of the TS signal appears in arbitrary units so that all of the curves could be fit

onto either graph in a manner that still clearly shows the important features. The degree

to which the two phases differ in appearance is remarkable. Whereas FM1 has a very

well-defined +HK peak for all temperatures up to the transition, and displays the

crossover field peak, the FM2 curve is largely dominated by the intense switching peak.

The anisotropy peak appears much broader. We note here that while the TS experiments

reveal clear differences in anisotropy features between FM1 and FM2, this picture

remains slightly ambiguous in the M(T) and M(H) data.

Figure 8.6: Unipolar transverse susceptibility scans for several different temperatures

plotted on two plots depicting the two different ferromagnetic phases ((a) is FM1 and (b)

is FM2). The signal intensity appears in arbitrary units as some of the curves have been

shifted upward or downward for clarity [22].

Page 143: Magnetism in Complex Oxides Probed by Magnetocaloric

127

To better illustrate the difference in anisotropy features between the ferromagnetic

states FM1 ( ) and FM2 ( ), we have superimposed the TS curves for

each phase onto two separate plots. The relative appearance of the curves for FM1 and

FM2 is akin to the comparison of two different materials entirely, rather than the

comparison of two structural phases of the same material. This, once again, indicates that

the TS technique is more suitable for studying anisotropy-driven transitions.

Figure 8.7 (a) shows the anisotropy field (+HK) as a function of temperature

where it is conclusively demonstrated that at higher temperatures, Pr0.5Sr0.5CoO3 has a

higher magnetocrystalline anisotropy phase (the FM2 phase) than at lower temperatures

(FM1 phase). For lower temperatures (T < TA,) the anisotropy field decreases with

increasing temperature, which is typical of most magnetic systems as the thermal energy

begins to compete with the anisotropy energy of the system. The structural transition at

120 K then appears as a dramatic increase in the HK to values even higher than those seen

at the lowest temperatures (μ0HK ≈ 184 mT at 120 K versus ≈125 mT at 10 K). The sharp

change in HK at TA is a direct consequence of the coupled structural/magnetocrystalline

anisotropy transition. After reaching this maximum, HK then slowly decreases again until

TC, where it goes to zero. The decrease of HK with temperature for T < TA and for TA < T

< TC is fully consistent with the perspective that the Pr0.5Sr0.5CoO3 system undergoes a

transition from one FM state to another.

The switching field is tracked as a function of temperature in Fig. 8.7 (b). Its

shape closely follows that reported in reference [10] for both the coercivity and fraction

of irreversible magnetization as measured by the first order reversal curve method [2].

This is not surprising, as all three properties are direct consequences of irreversible

Page 144: Magnetism in Complex Oxides Probed by Magnetocaloric

128

hysteretic processes. At low temperatures, magnetization decreases rapidly until the

approach to TA where it experiences an uptick and a cusp at 120 K and then decreases

again until TC. While empirical data suggested this field should be around 75 mT, the

temperature dependence of this peak reveals that this crossover field is different for

different regions of the M(T) plots. Around the structural transition, the crossover field is

indeed measured by TS to be 75 mT.

Figure 8.7 (c) shows the evolution of the peak position associated with the

crossover field (Hcr) as a function of temperature. However, at lower temperatures, the

change in shape of the magnetization curves appears to occur at much lower fields,

around 20 mT, which then increases rapidly with temperature up to TA. It has already

been discussed that the presence of this peak lends insight to the crossover behavior

between the two types of anomalous M(T) curves.

Figure 8.7: Temperature dependence of the peaks positions in the transverse

susceptibility measurement. (a) Anisotropy field (+HK), (b) Switching field (HS), (c)

Crossover field (Hcr) [22].

Page 145: Magnetism in Complex Oxides Probed by Magnetocaloric

129

It has been noted that due to poor magnetic coupling between the grains at low

temperatures (T << TA), the initial susceptibility is smaller in the FM1 region than in the

FM2 regions [16]. In addition, the magnetization has been found to increase with

increasing temperature in the FM1 region below the crossover field. Therefore, the

increase of Hcr with temperature at T < TA revealed in the TS profile is as expected,

consistently pointing to thermally activated improvement in intergranular coupling in this

temperature range.

8.3 Conclusions

We have used the transverse susceptibility and magnetocaloric measurement

techniques to examine the anisotropic magnetic properties of Pr0.5Sr0.5CoO3, specifically

the structure-driven magnetocrystalline anisotropy transition at 120 K. By using these

techniques, we were able to show that the FM-FM phase transition is clearly manifested

in the evolution of the anisotropy and switching peaks with temperature. The well-

documented unusual M(T) behavior, dependent upon cooling field, is present in the TS as

well in the form of a sharp peak at the crossover field which disappears above TA. We

showed (figure 8.6) that the rotation of the easy axis can also be deduced by comparing

the signal intensity from two different measurement orientations where a crossover

behavior is observed. We also observed a relatively rare structural transition, which is not

coupled with a magnetic transition, using the magnetocaloric effect. Collectively these

findings show that transverse susceptibility and the magnetocaloric effect are very useful

tools for lending insight into the unusual magnetic behavior of doped perovskites.

Page 146: Magnetism in Complex Oxides Probed by Magnetocaloric

130

References:

[1] O. Tegus, E. Bruck, K.H.J. Buschow, F.R. de Boer, Transition-metal-based magnetic

refrigerants for room-temperature applications, Nature, 415 (2002) 150-152.

[2] R.F. Klie, J.C. Zheng, Y. Zhu, M. Varela, J. Wu, C. Leighton, Direct measurement of

the low-temperature spin-state transition in LaCoO(3), Phys Rev Lett, 99 (2007).

[3] D.P. Kozlenko, N.O. Golosova, Z. Jirak, L.S. Dubrovinsky, B.N. Savenko, M.G.

Tucker, Y. Le Godec, V.P. Glazkov, Temperature- and pressure-driven spin-state

transitions in LaCoO3, Phys Rev B, 75 (2007).

[4] A. Podlesnyak, S. Streule, J. Mesot, M. Medarde, E. Pomjakushina, K. Conder, A.

Tanaka, M.W. Haverkort, D.I. Khomskii, Spin-state transition in LaCoO3: Direct neutron

spectroscopic evidence of excited magnetic states, Phys Rev Lett, 97 (2006).

[5] S. Hirahara, Y. Nakai, K. Miyoshi, K. Fujiwara, J. Takeuchi, Magnetic properties,

thermal expansion and magneto striction of Pr0.5Sr0.5CoO3 single crystal, J Magn Magn

Mater, 310 (2007) 1866-1867.

[6] R. Mahendiran, P. Schiffer, Double magnetic transition in Pr0.5Sr0.5CoO3, Phys Rev B,

68 (2003).

[7] M. Patra, S. Majumdar, S. Giri, Cluster-glass-like state and exchange bias effect in

spontaneously phase separated, Pr0.7Sr0.3CoO3, J Appl Phys, 107 (2010).

[8] M. Uchida, R. Mahendiran, Y. Tomioka, Y. Matsui, K. Ishizuka, Y. Tokura, Changes

of magnetic domain structure induced by temperature-variation and electron-beam

irradiation in Pr0.5Sr0.5CoO3, Appl Phys Lett, 86 (2005).

[9] K. Yoshii, A. Nakamura, H. Abe, M. Mizumaki, T. Muro, Magnetism and transport of

Ln(0.5)Sr(0.5)CoO(3) (Ln = Pr, Nd, Sm, Eu and Gd), J Magn Magn Mater, 239 (2002) 85-87.

Page 147: Magnetism in Complex Oxides Probed by Magnetocaloric

131

[10] C. Leighton, et. al., Coupled structural/magnetocrystalline anisotropy transitions in

the doped perovskite cobaltite Pr1-xSrxCoO3, Phys Rev B, 79 (2009).

[11] Y. Ji, C.L. Chien, Y. Tomioka, Y. Tokura, Measurement of spin polarization of

single crystals of La0.7Sr0.3MnO3 and La0.6Sr0.4MnO3, Phys Rev B, 66 (2002).

[12] V.B. Shenoy, C.N.R. Rao, Electronic phase separation and other novel phenomena

and properties exhibited by mixed-valent rare-earth manganites and related materials,

Philos T R Soc A, 366 (2008) 63-82.

[13] M.E. Lopezmorales, D. Riosjara, J. Taguena, R. Escudero, S. Laplaca, A. Bezinge,

V.Y. Lee, E.M. Engler, P.M. Grant, Role of Oxygen in PrBa2Cu3O7-Y - Effect on

Structural and Physical-Properties, Phys Rev B, 41 (1990) 6655-6667.

[14] V.K. Pecharsky, K.A. Gschneidner, Giant magnetocaloric effect in Gd5(Si2Ge2),

Phys Rev Lett, 78 (1997) 4494-4497.

[15] V.K. Pecharsky, K.A. Gschneidner, Y. Mudryk, D. Paudyal, Making the most of the

magnetic and lattice entropy changes, J Magn Magn Mater, 321 (2009) 3541-3547.

[16] A.M. Pereira, A.M. dos Santos, C. Magen, J.B. Sousa, P.A. Algarabel, Y. Ren, C.

Ritter, L. Morellon, M.R. Ibarra, J.P. Araujo, Understanding the role played by Fe on the

tuning of magnetocaloric effect in Tb5Si2Ge2, Appl Phys Lett, 98 (2011).

[17] V. Franco, A. Conde, Scaling laws for the magnetocaloric effect in second order

phase transitions: From physics to applications for the characterization of materials, Int J

Refrig, 33 (2010) 465-473.

[18] V. Franco, A. Conde, M.D. Kuz'min, J.M. Romero-Enrique, The magnetocaloric

effect in materials with a second order phase transition: Are T-C and T-peak necessarily

coincident?, J Appl Phys, 105 (2009).

Page 148: Magnetism in Complex Oxides Probed by Magnetocaloric

132

[19] V. Franco, A. Conde, J.M. Romero-Enrique, J.S. Blazquez, A universal curve for the

magnetocaloric effect: an analysis based on scaling relations, J Phys-Condens Mat, 20

(2008).

[20] R. Caballero-Flores, V. Franco, A. Conde, Q.Y. Dong, H.W. Zhang, Study of the

field dependence of the magnetocaloric effect in Nd1.25Fe11Ti: A multiphase magnetic

system, J Magn Magn Mater, 322 (2010) 804-807.

[21] R. Caballero-Flores, V. Franco, A. Conde, K.E. Knipling, M.A. Willard, Influence

of Co and Ni addition on the magnetocaloric effect in Fe88-2xCoxNixZr7B4Cu1 soft

magnetic amorphous alloys, Appl Phys Lett, 96 (2010).

[22] N. A. F. Huls, N. S. Bingham, M. H. Phan, H. Srikanth, D. D. Stauffer, and C.

Leighton, Transverse susceptibility as a probe of the magnetocrystalline anisotropy-

driven phase transition in Pr0.5Sr0.5CoO3. Physical Review B, 2011. 83(2).

Page 149: Magnetism in Complex Oxides Probed by Magnetocaloric

133

CHAPTER 9.

A COMPLEX MAGNETIC PHASE DIAGRAM AND MAGNETOCALORIC

EFFECT IN Ca3Co2O6 SINGLE CRYSTALS

In the previous chapters we have shown the usefulness of MCE for probing the

magnetic ground states, phase coexistence, and field-induced kinetic arrest phenomena in

phase-separated manganites. In this chapter, we demonstrate that it is also very useful for

probing the complex ground state magnetism of Ca3Co2O6 (CCO), which exhibits spin

frustration and intrinsic low dimensionality due to the formation of 1D spin chains. Our

MCE experiments have provided new insights into the nature of switching between

multi-states and competing interactions within spin chains, and between them, leading to

a more comprehensive magnetic phase diagram.

9.1 Introduction

Ca3Co2O6 (CCO) has received considerable interest over the past decades due to the

complex interplay among the magnetic, electronic and structural properties. The refined

neutron and X-Ray powder diffraction data have shown that the material crystallizes in

the rhombohedral space group R c[1]. CCO consists of face sharing CoO6 trigonal

prisms and CoO6 octahedra chains running along the c-axis with six nearest neighbor

chains, forming a triangular lattice in the ab plane and all separated by Ca atoms.

Intrachain Co-Co separation is 2.59Ǻ, while interchain Co-Co separation is roughly twice

Page 150: Magnetism in Complex Oxides Probed by Magnetocaloric

134

that, giving rise to large anisotropy and a quasi 1D structure. The Co ions in CCO are

found in the 3+ oxidation state, with alternating low-spin (LS) and high-spin (HS) states

for the octahedral and trigonal configurations, respectively [2]. Due to the nature of the

geometry, and the strong anisotropy of Co, CCO has long been considered to be an Ising-

like material, where each chain can be represented by a single spin in a 2-D lattice.

However, recent theoretical predictions [3] and experimental results [4] show that this is

an oversimplified description; therefore a full 3-D picture is needed.

The origin of a broad feature above TN (centered around 100K) revealed by

calorimetric measurements [5] is currently under debate, and has been attributed to 1-D

magnetic ordering along the chains [6] or a spin-state transition, which is common for

this class of cobaltite materials [7]. Neutron diffraction studies have revealed long-range

magnetic ordering at TN ~ 25K; this is generally accepted as very strong FM intrachain

coupling, and a slightly weaker AFM interchain coupling. The 2-D Ising model discussed

above gives rise to a partially disordered antiferromagnet (PDA), where 2/3 of the chains

are coupled antiferromagnetically, and 1/3 are incoherent [8]. However, in the

intermediate temperature range (12K < T < 25K), resonant X-ray scattering [9]

discovered incommensuration in the magnetic order along the chain, implying a long

wavelength spin-density wave (SDW). Thermal conductivity measurements [10] have

shown an exchange-mediated heat transfer, which supports helical exchange pathways

rather than Ising-type behavior. Below T ~ 12K, there is an increase in the number of

“steps” in the field-driven magnetization process, as well as an increase in magnetic

hysteresis. These steps had led many to believe that the system exhibits quantum

tunneling of magnetization (QTM), similar to molecular magnets [11], however the idea

Page 151: Magnetism in Complex Oxides Probed by Magnetocaloric

135

of QTM has been challenged, and the behavior is suggested to correspond more closely

to that of superparamagnetic clusters [12], or a host of metastable states. Below TFS ~ 8K,

CCO exhibits extremely slow spin dynamics, large magnetic hysteresis, and a very

pronounced frequency dependence in AC susceptibility measurements [13], which can all

be described as a frozen spin state.

Recent theoretical and experimental evidence suggests that below 25K, nearest

and next nearest neighbor AFM interactions stabilize the AFM state in the form of a

longitudinal amplitude-modulated SDW propagating along the chains [4]. As temperature

is decreased, the SDW becomes more unstable, leading to a large volume fraction of

short-range ordering giving rise to an overall highly disordered state [14].

It is clear that the complex magnetic behavior described above provides a unique

opportunity to study a variety of interactions with numerous probes. Much of the research

on CCO has been focused on probing the microstates either experimentally or

theoretically. This study focuses on the macroscopic details of the field- and temperature-

dependence of the change in magnetic entropy probed by the MCE. CCO single crystals

were grown by the floating zone method and provided by Prof. Sang-Wook Cheong’s

group at Rutgers University.

9.2 Results and Discussion

Figure 9.1 (a) shows the DC M(T) curve acquired under a field of H = 100 Oe.

The small deviation from M = 0 beginning below T ~ 70K is attributed to the onset of 1-

D magnetic ordering along the chains. Then at TC ~ 25K, there is a large jump in M

associated with FM intrachain coupling. At this ordering temperature, it is important to

note that the coupling between neighboring chains is AFM in nature, thus the large jump

Page 152: Magnetism in Complex Oxides Probed by Magnetocaloric

136

in M(T) leads to the understanding that the intrachain FM coupling dominates in strength.

At even lower temperatures, the system enters a complex glassy regime.

Figure 9.1: Temperature dependence of (a) DC-magnetization at an applied field of 100

Oe (b) AC susceptibility with small AC magnetic field ~10 Oe at a variety of frequencies.

Also, magnetic field dependence of magnetization at (c) 25K and (d) 5K.

Magnetization dynamics were probed (Figure 9.1(b)) using AC magnetization

measurements in a small (HAC ~ 10 Oe) AC- magnetic field at various frequencies. The

strong frequency dependence can be described via an Arrehnius-like regime at higher

temperatures (T > 10K) and a quantum regime at lower temperatures (T <10 K) [15].

Figure 9.1 (c,d) shows M(H) curves at T = 25K and 5K; dM/dT is plotted in the inset. For

T < 10K, there is an increase in the number of steps in the M(H) curves, leading to strong

magnetization dynamics and coupling phenomena.

(a) (b)

(c) (d)

Page 153: Magnetism in Complex Oxides Probed by Magnetocaloric

137

Figure 9.2 shows the isothermal M(H) curves used to calculate the MCE. Due to

the measurement process, H and T need to be stabilized in order to collect the data. For

CCO, this results in a smoother step in the magnetization due to the slow field sweep rate

[15]. For T > TN, there is a linear increase in M with increase in H, however, an approach

to saturation is evident in the curves.

0 2 4 6 8

0

20

40

60

80

100

T=5K

M (

em

u/g

)

0H (T)

T=120K

Figure 9.2: Isothermal magnetization vs. applied field for a temperature range of 120K-

5K with a temperature interval of 5K, and magnetic field from 0-7T.

This leads to the belief that the large bump in the heat capacity [5] at high

temperature is due to short-range FM order along the chains. Below TN, steps emerge in

the magnetization. A large plateau occurs at low field when M = MS/3 (MS is the

saturation magnetization), as the geometrically frustrated matrix enters the ferrimagnetic

(FIM) up-up-down (UUD) state in which 2/3 of the chains are aligned ferromagnetically

and 1/3 are aligned antiferromagnetically. There is another step 0H = 3.6T, above which

the system becomes fully ferromagnetically aligned. For T = 5K, more steps in M occur

Page 154: Magnetism in Complex Oxides Probed by Magnetocaloric

138

at a regular interval [16] (0H = 0T, 1.2T, 2.4T and 3.6T). This is due to the stabilization

of micro-clusters of disorder within the chain [14].

Figure 9.3 shows the calculated temperature dependence of SM. The most

striking feature is the very large peak (–SMmax

~ 6.5 J/kgK) at 30K. This peak has the

signature of short-range FM ordering, which is believed to be correlated to the large

maximum seen in the heat capacity measurements [5].

0 20 40 60 80 100 120

-2

0

2

4

6

0H = 7 T

-S

M (

J / K

g K

)

T ( K )

0H = 0 T

Figure 9.3: Change in magnetic entropy as a function of temperature, calculated using

the thermodynamic Maxwell relation (Eqn. 3.9).

At TN, there is a small peak associated with FM intrachain alignment and AFM interchain

alignment. Due to the nature of the transition, –SM < 0 is expected, however, since –SM

> 0, this highlights the dominance of the intrachain FM coupling over the interchain

AFM coupling. For T ≤ 10K, –SM < 0, where the thermal fluctuations have stabilized the

AFM phase.

Page 155: Magnetism in Complex Oxides Probed by Magnetocaloric

139

The MCE measurements become even more enlightening when looking into the

field dependence of SM. Figure 9.4 (a-d) shows –SM(H) for 30K, 20K, 15K and 10K

respectively. At T = 30K (Figure 9.4 (a)), –SM(H) shows a gradual increase with H,

similar to that of short-range FM ordering, leading to the belief that there are magnetic

interactions occurring above the long-range ordering temperature. These interactions are

believed to be the growth of FM clusters along the chains, and can extend up to

temperatures as high as 100K. The fact that the large peak is not observed until T = 30K

could be due to thermal fluctuations that cannot be overcome by the magnetic fields

available in our laboratory (up to ±7T). As the temperature is decreased below TN, a very

interesting feature emerges. Figure 9.4 (b), shows –SM(H) at 20K. As H is increased

there is an increase in –SM, as would be expected for FM ordering throughout CCO, i.e.

the UUD model discussed previously. Interestingly, –SM exhibits a maximum at 0H =

2T, then starts decreasing for fields in the range 2T < 0H < 3.6T. In this region, the FM

coupling has been stabilized, and the AFM coupling between nearest-neighbor (NN) and

next-nearest neighbor (NNN) chains is strong enough to slightly distort the FM ordering

present in each chain; this can be thought of as a disorder induced by order. A similar

feature has been observed in the helimagnet, Dy [17].

Finally, for large enough fields the disorder can be overcome and there is full FM

order throughout the sample. At T = 15K, (Figure 9.4 (c)) –SM(H) exhibits similar

features to those observed at higher temperatures, yet the dip in –SM at 0H = 2T

reaches negative values, in contrast to the T=20 K –SM(H) curve. The T = 15 K behavior

can be attributed to the system cooling below the “cross-over” temperature (T ~ 18K),

where the AFM coupling between NN and NNN chains increases, therefore further

Page 156: Magnetism in Complex Oxides Probed by Magnetocaloric

140

stabilizing the AFM phase and increasing its negative contribution to –SM. At T = 10K

(Figure 9.4 (d)), the SDW is destabilized giving rise to a large volume fraction of short-

range order FM clusters, aligned either parallel or anti-parallel to the applied magnetic

field.

0 2 4 6 8

0

2

4

6

-S

M (

J/k

g K

)

-S

M (

J/k

g K

)

0H (T)

0H (T)

0H (T)

-S

M (

J/k

g K

)

-S

M (

J/k

g K

)

0H (T)

(a) T = 30 K

0 2 4 6 8

0

2

4

(b) T = 20 K

0 2 4 6 8-0.5

0.0

0.5

1.0

1.5

2.0

2.5

HC3

HC2

HC1

(c) T = 15 K

0 2 4 6 8

-1.2

-0.8

-0.4

0.0

0.4

(d) T = 10 K

Figure 9.4: Change in magnetic entropy as a function of applied field at various constant

temperatures, (a) 30K, (b) 20K, (c) 15K and (d) 10K.

Therefore, on average CCO is in an AFM-like state leading to a purely negative –SM(H).

The slight upturn in –SM near oH = 3T is related to a small volume fraction of FM

correlations ordering with the field. However, by ordering these regions, there is, once

more, an increase in disorder brought about by distortion induced by neighboring chains.

Figure 9.5, represents a phase diagram constructed from the MCE data, where HC1

(shown in Figure 9.4 (c)) is related to the maximum (minimum) in –SM before the onset

Page 157: Magnetism in Complex Oxides Probed by Magnetocaloric

141

of field-induced disorder. HC2 represents the field at which the maximum disorder is

induced in the chains.

0 2 4 6

(a)

dS

/dH

0H (T)

S

D

W

FIM

Disordered

FM

T = 20 K

6 8 10 12 14 16 18 20 22 240

1

2

3

4

5

6

7

8 (b)

SDW

Disorder Due to Order

AFM

FM

FIM

0H

(T

)

T (K)

HC3

HC2

HC1

HSDW

FIM

Figure 9.5: (a) First derivative of the field-dependent change in entropy. (b)

Magnetic phase diagram derived from the field- and temperature-dependent

change in magnetic entropy.

HC3 indicates the field at which the volume fraction of FM ordering starts to

become the dominant phase. From fig 9.5, as T is decreases below 25K it takes a

significantly larger value of H in order to achieve a large volume fraction of FM ordering

Page 158: Magnetism in Complex Oxides Probed by Magnetocaloric

142

in CCO (in fact, for T = 5K –SM never crosses over zero, therefore we extrapolated

linearly), leading to the paradoxical conclusion that CCO becomes more disordered as T

decreases. HC2 remains almost constant for all T, and is associated with the large plateau

in M present at all T. For HC1, a crossover temperature exists; for T ≥ 15K a relatively

small field will globally align the material in the UUD FIM state described above.

However, for T < 15K, the system is in the short range order-SDW state which gives rise

to AFM-like behavior.

9.3 Conclusions

MCE data was taken for a wide range of temperatures and applied magnetic fields

for Ca3Co2O6. The MCE data seem to confirm the spin-density wave description that has

been proposed recently. The MCE data has also confirmed that CCO becomes more

disordered with a decrease in temperature and also exhibits field-driven and order-

induced disorder. These observations are consistent with previously studied helimagnets.

A new and comprehensive magnetic phase diagram is constructed for the first time from

MCE experiments.

References:

[1] H. Fjellvag, et. al., Crystal structure and possible charge ordering in one-dimensional

Ca3Co2O6, J Solid State Chem, 124 (1996) 190-194.

[2] S. Aasland, H. Fjellvag, B. Hauback, Magnetic properties of the one-dimensional

Ca3Co2O6, Solid State Commun, 101 (1997) 187-192.

[3] Y. Kamiya, C.D. Batista, Formation of Magnetic Microphases in Ca3Co2O6, Phys Rev

Lett, 109 (2012).

Page 159: Magnetism in Complex Oxides Probed by Magnetocaloric

143

[4] S. Agrestini, C.L. Fleck, L.C. Chapon, C. Mazzoli, A. Bombardi, M.R. Lees, O.A.

Petrenko, Slow Magnetic Order-Order Transition in the Spin Chain Antiferromagnet

Ca3Co2O6, Phys Rev Lett, 106 (2011).

[5] V. Hardy, S. Lambert, M.R. Lees, D.M. Paul, Specific heat and magnetization study

on single crystals of the frustrated quasi-one-dimensional oxide Ca3Co2O6, Phys Rev B,

68 (2003).

[6] L.J. Dejongh, A.R. Miedema, Experiments on Simple Magnetic Model Systems, Adv

Phys, 23 (1974) 1-260.

[7] J.B. Goodenough, An Interpretation of the Magnetic Properties of the Perovskite-

Type Mixed Crystals La1-xSrxCoO3-Lambda, J Phys Chem Solids, 6 (1958) 287-297.

[8] H. Kageyama, K. Yoshimura, K. Kosuge, H. Mitamura, T. Goto, Field-induced

magnetic transitions in the one-dimensional compound Ca3Co2O6, J Phys Soc Jpn, 66

(1997) 1607-1610.

[9] S. Agrestini, C. Mazzoli, A. Bombardi, M.R. Lees, Incommensurate magnetic ground

state revealed by resonant x-ray scattering in the frustrated spin system Ca3Co2O6, Phys

Rev B, 77 (2008).

[10] J.G. Cheng, J.S. Zhou, J.B. Goodenough, Thermal conductivity, electron transport,

and magnetic properties of single-crystal Ca3Co2O6, Phys Rev B, 79 (2009).

[11] D. Gatteschi, R. Sessoli, Quantum tunneling of magnetization and related

phenomena in molecular materials, Angew Chem Int Edit, 42 (2003) 268-297.

[12] E.V. Sampathkumaran, N. Fujiwara, S. Rayaprol, P.K. Madhu, Y. Uwatoko,

Magnetic behavior of Co ions in the exotic spin-chain compound Ca3Co2O6 from Co-59

NMR studies, Phys Rev B, 70 (2004).

Page 160: Magnetism in Complex Oxides Probed by Magnetocaloric

144

[13] A. Maignan, V. Hardy, S. Hebert, M. Drillon, M.R. Lees, O. Petrenko, D.M. Paul,

D. Khomskii, Quantum tunneling of the magnetization in the Ising chain compound

Ca3Co2O6, J Mater Chem, 14 (2004) 1231-1234.

[14] T. Moyoshi, K. Motoya, Incommensurate Magnetic Structure and Its Long-Time

Variation in a Geometrically Frustrated Magnet Ca3Co2O6, J Phys Soc Jpn, 80 (2011).

[15] V. Hardy, M.R. Lees, O.A. Petrenko, D.M. Paul, D. Flahaut, S. Hebert, A. Maignan,

Temperature and time dependence of the field-driven magnetization steps in Ca3Co2O6

single crystals, Phys Rev B, 70 (2004).

[16] Y.B. Kudasov, Magnetic structure and phase diagram in a spin-chain system:

Ca3Co2O6, Epl-Europhys Lett, 78 (2007).

[17] M. Foldeaki, R. Chahine, T.K. Bose, Magnetic Measurements - a Powerful Tool in

Magnetic Refrigerator Design, J Appl Phys, 77 (1995) 3528-3537.

Page 161: Magnetism in Complex Oxides Probed by Magnetocaloric

145

CHAPTER 10.

CONCLUSIONS AND OUTLOOK

10.1 Conclusions

Throughout this dissertation I have demonstrated the effectiveness of the

magnetocaloric effect (MCE) and transverse susceptibility (TS) as probes used for

fundamental research rather than the standard application-based probes seen widely

throughout the community. First, the basic properties of manganite materials were

discussed, where the strongly coupled magnetic, electric and structural degrees of

freedom depend strongly on local strain caused by chemical doping. Manganites,

although very similar in structure, can be “tuned” drastically by varying the A-site radius

and applying external fields (i.e. pressure, magnetic and electric). This makes manganites

ideally suited for this kind of comprehensive study. Next, the magnetic and

magnetocaloric properties of bulk polycrystalline and thin-film samples of the Collosal

Magenetoresistive (CMR) manganite La0.7Ca0.3MnO3 were investigated to observe the

effects of reduced dimensionality in the system. Broadened transitions along with

reduced Curie temperature, magnetic moment, and magnetic entropy change were

observed in the thin-film sample. The film exhibited enhanced refrigerant capacity over

the polycrystalline sample, due to the change of the nature of the phase transition being

converted from a first-order to second-order paramagnetic to ferromagnetic transition.

Page 162: Magnetism in Complex Oxides Probed by Magnetocaloric

146

Then, a systematic study on the effects of A-site cation doping on the

ferromagnetic phase transitions and critical behavior of La0.7Ca0.3−xSrxMnO3 (x = 0, 0.05,

0.1, 0.2 and 0.25) single crystals was presented. Using the Banerjee criterion and

Kouvel–Fisher method, it is shown that x∼0.1 is a tricritical point that separates the first-

order magnetic transition for x < 0.1 from a second-order magnetic transition for x > 0.1.

Above the tricritical point, the system exhibits a second-order magnetic transition with

the critical exponents belonging to the Heisenberg universality class with short-range

exchange interactions. This indicates that short-range magnetic interactions dominate in

these systems. It is shown that while the Double-Exchange (DE) mechanism and

formation of ferromagnetic clusters can account for the canonical MR and metal-like

conducting behavior in La0.7Ca0.3−xSrxMnO3 with x = 0.2 and 0.25. Other effects, such as

cooperative Jahn–Teller distortions and antiferromagnetic coupling are important

additions for understanding of the relationship between the PM–FM transition and the MI

transition, including CMR in La0.7Ca0.3−xSrxMnO3 with x = 0, 0.05 and 0.1. The change of

the PM–FM transition with chemical (internal) pressure introduced by substitution of

larger Sr ions for smaller Ca ions points to the strong coupling between the magnetic

order and structural parameters in these doped manganites.

The usefulness of MCE and TS for studying complex phase transitions and the

magnetic anisotropy along with its variation with temperature in half-doped manganites

such as Pr0.5Sr0.5MnO3 was demonstrated. The influence of first- and second-order

magnetic phase transitions on the MCE and RC of charge-ordered Pr0.5Sr0.5MnO3 was

displayed first. It is shown that the first-order magnetic transition at TCO induces a larger

MCE, but concentrates the MCE in a narrower temperature range, resulting in smaller

Page 163: Magnetism in Complex Oxides Probed by Magnetocaloric

147

RC. However, the second-order magnetic transition at TC induces a smaller MCE, but

spreads the MCE over a broader temperature range, resulting in larger RC. Hysteretic

losses accompanying the first-order magnetic transition are very large below TCO and

therefore detrimental to the RC, whereas they are very small or negligible below TC, due

to the nature of the second-order magnetic transition. With TS, it is evidenced that there

is an abrupt change in magnetic anisotropy at the FM-AFM transition, which is

associated with the structural phase transition that occurs at the same temperature. This is

a clear indication of the strong correlations between magnetic and structural properties n

Pr0.5Sr0.5MnO3.

Systematic study of magnetocaloric measurements on La5/8−xPrxCa3/8MnO3 (x =

0.275) single crystals have revealed further insights into the complex multiple-phase

transitions. The system is FM at low temperature and becomes charge ordered at high

temperature. The dynamic strain-liquid phase is strongly affected by an applied magnetic

field, whereas the frozen strain-glass phase is nearly magnetic field independent. The

origin of the large MCE in the strain-liquid region arises from the suppression of dynamic

fluctuations in magnetic fields. The MCE data clarify that the sharp increase in the

magnetization below TC may not be due to the destabilization of the COI phase to the

FMM phase, but favors the idea of the growth of pre-existing FMM domain regions.

Overall, MCE and TS have proven to be excellent probes of the magnetic transitions and

ground-state magnetic properties of mixed-phase systems.

TS and MCE measurement techniques were used to examine the anisotropic

magnetic properties of Pr0.5Sr0.5CoO3, specifically the structure-driven magnetocrystalline

anisotropy transition at 120 K. By using these techniques, the FM-FM phase transition is

Page 164: Magnetism in Complex Oxides Probed by Magnetocaloric

148

clearly manifested in the evolution of the anisotropy and switching peaks with

temperature. The well-documented unusual M(T) behavior, dependent upon cooling field,

is present in the TS, as well, in the form of a sharp peak at the crossover field which

disappears above TA. The rotation of the easy axis can also be deduced by comparing the

signal intensity from two different measurement orientations where a crossover behavior

is observed. Also observed was a relatively rare structural transition, which is not coupled

with a magnetic transition, using the MCE. Collectively these findings show that TS and

the MCE are very useful tools for lending insight into the unusual magnetic behavior of

doped perovskites.

Finally, magnetic and magnetocaloric data were taken for a wide range of

temperatures and applied magnetic fields for Ca3Co2O6 (CCO). The data seem to confirm

the spin-density wave description that has been proposed recently. MCE data has also

confirmed that CCO becomes more disordered with a decrease in temperature, and also

exhibits a field-driven, order-induced disorder. A new phase diagram was produced from

this data, further proving the usefulness of the MCE as a very powerful fundamental

probe.

10.2 Outlook

The CMR manganites are sensitive to all types of perturbations. In particular, it

has been shown in bulk that the internal (through the average size of the A-site cation) or

external pressure (via hydrostatic pressure) can strongly influence the magnetotransport

properties. Thus, strains affect the properties of manganite thin films, and, in

consequence, one needs to correctly understand the effects in order to obtain the desired

properties.

Page 165: Magnetism in Complex Oxides Probed by Magnetocaloric

149

Van Tendeloo et al. [1] have studied the evolution of the microstructure as

function of the thickness in La0.7Sr0.3MnO3 films grown on LaAlO3. Close to the

interface, both the film and the substrate are elastically strained in opposite directions in

such a way that the interface is perfectly coherent. In the thicker films, the stress is partly

relieved after annealing by the formation of misfit dislocations. In general, a strain is

observed due to the epitaxial growth in very thin films i.e. lattice parameters adopt those

of the cubic lattice. However, when films reach a critical thickness (generally around

~100nm) the strain becomes relaxed and the film takes on the properties of its bulk

counterpart.

Another way to achieve exotic behavior in these materials is via the growth of

artificial superlattices and multilayer films. The interfaces of thin-film manganite

heterostructures are well documented sites for fundamentally altered magnetism.

Superlattices of FM and AFM layers can lead to an overall enhancement in magnetization

through an induced FM ordering extending into the AFM layer [2], while ferromagnetism

has also been observed at the interface of two AFM manganites [3]. Intriguingly, induced

magnetism can also occur in paramagnetic (PM) layers at FM/PM and AFM/PM

interfaces [4, 5].

La0.7Sr0.3MnO3 (La0.7Ca0.3MnO3) films grown on BaTiO3 (BTO) have recently [6]

been shown to exhibit large jumps in temperature-dependent magnetization due to strain

from first-order structural phase transitions, where BTO changes from rhombohedral to

orthorhombic at 200 K, from orthorhombic to tetragonal at 300 K and from tetragonal to

cubic at the ferroelectric Curie temperature TFEC ~ 400 K). However, since BTO is also a

piezoelectric material, the application of a relatively small electric field can also change

Page 166: Magnetism in Complex Oxides Probed by Magnetocaloric

150

the substrate’s lattice parameters, allowing for the control of strain at temperatures below

TFEC.

Thus, by depositing complex oxide thin-films onto variable substrates, or in

composite multi-layered systems, the physical properties of the films will in turn be

“tunable,” thereby making these materials functional for a wide variety of applications,

and creating a vast playground for fundamental research.

References:

[1] F.S. Razavi, G. Gross, H.U. Habermeier, O. Lebedev, S. Amelinckx, G. Van

Tendeloo, A. Vigliante, Epitaxial strain induced metal insulator transition in

La0.9Sr0.1MnO3 and La0.88Sr0.1MnO3 thin films, Appl Phys Lett, 76 (2000) 155-157.

[2] A. Hoffmann, S.J. May, S.G.E. te Velthuis, S. Park, M.R. Fitzsimmons, G. Campillo,

M.E. Gomez, Magnetic depth profile of a modulation-doped La1-xCaxMnO3 exchange-

biased system, Phys Rev B, 80 (2009).

[3] T.S. Santos, B.J. Kirby, S. Kumar, S.J. May, J.A. Borchers, B.B. Maranville, J.

Zarestky, S.G.E.T. Velthuis, J. van den Brink, A. Bhattacharya, Delta Doping of

Ferromagnetism in Antiferromagnetic Manganite Superlattices, Phys Rev Lett, 107

(2011).

[4] D. Niebieskikwiat, et. al., Nanoscale magnetic structure of

ferromagnet/antiferromagnet manganite multilayers, Phys Rev Lett, 99 (2007).

[5] M. Gibert, P. Zubko, R. Scherwitzl, J. Iniguez, J.M. Triscone, Exchange bias in

LaNiO3-LaMnO3 superlattices, Nat Mater, 11 (2012) 195-198.

[6] X. Moya, et. al., Giant and reversible extrinsic magnetocaloric effects in

La0.7Ca0.3MnO3 films due to strain, Nat Mater, 12 (2013) 52-5

Page 167: Magnetism in Complex Oxides Probed by Magnetocaloric

151

APPENDICES

Page 168: Magnetism in Complex Oxides Probed by Magnetocaloric

152

APPENDIX A LIST OF PUBLICATIONS

1. N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija, and C. Leighton,

Magnetocaloric effect and refrigerant capacity in charge-ordered manganites

Journal of Applied Physics 106, 023909 (2009)

2. N.S. Bingham, P. Lampen, M.H. Phan, T.D. Hoang, H.D. Chinh, C.L. Zhang,

S.W. Cheong, and H. Srikanth, Impact of nanostructuring on the magnetic and

magnetocaloric properties of microscale phase-separated La5/8-yPryCa3/8MnO3

manganites, Physical Review B 86, 064420 (2012)

3. N.S. Bingham, H. Wang, F. Qin, H.X. Peng, J. F. Sun, V. Franco, H. Srikanth,

and M.H. Phan, Excellent magnetocaloric properties of melt-extracted Gd-based

amorphous microwires, Applied Physics Letters 101, 102407 (2012)

4. N.S. Bingham, P. Lampen, T.L. Phan, M.H. Phan, S.C. Yu, and H. Srikanth,

Magnetocaloric effect and refrigerant capacity in Sm1-xSrxMnO3 manganites,

Journal of Applied Physics 111, 07D705 (2012)

5. M.H. Phan, M.B. Morales, N.S. Bingham, H. Srikanth, C.L. Zhang, and S.W.

Cheong, Phase Coexistence and Magnetocaloric Effect in La5/8-yPryCa3/8MnO3 (y

= 0.275), Physical Review B 81, 094413 (2010)

6. N.A. Frey, N.S. Bingham, M.H. Phan, H. Srikanth, D. D. Stauffer, and C.

Leighton, Transverse Susceptibility as a Probe of the Magnetocrystalline

Anisotropy – Driven Phase Transition in Pr0.5Sr0.5CoO3, Physical Review B 83,

024406 (2011).

Page 169: Magnetism in Complex Oxides Probed by Magnetocaloric

153

7. M.H. Phan, S. Chandra, N.S. Bingham, H. Srikanth, C.L. Zhang, S.W. Cheong,

T. D. Hoang, and H. D. Chinh, Collapse of charge ordering and enhancement of

magnetocaloric effect in nanocrystalline La0.35Pr0.275Ca0.375MnO3, Applied

Physics Letters 97, 242506 (2010)

8. P. Lampen, N.S. Bingham, M.H. Phan, H. Kim, M. Osofsky, A. Piqué, T.L.

Phan, S.C. Yu, and H. Srikanth, Impact of reduced dimensionality on the

magnetic and magnetocaloric response of La0.7Ca0.3MnO3, Applied Physics

Letters 102, 062414 (2013)

9. M.H. Phan, V. Franco, N.S. Bingham, H. Srikanth, N.H Hur, and S.C. Yu,

Tricritical point and critical exponents of La0.7Ca0.3-xSrxMnO3 (x = 0, 0.05, 0.1,

0.2, 0.25) single crystals, Journal o Alloys and Compounds 2, 508 (2010).

10. F. Qin, H. Wang, H.X. Peng, N.S. Bingham, D.W. Xing, J. F. Sun, V. Franco, H.

Srikanth, and M.H. Phan, Mechanical and magnetocaloric properties of Gd-based

amorphous microwires fabricated by melt-extraction technique, Acta Materialia

61, 1284 (2013)

11. N.H. Hong, C.-K. Park, A. T. Raghavender, O. Ciftja, N.S. Bingham, M.H.

Phan, and H. Srikanth, Room temperature ferromagnetism in monoclinic Mn-

doped ZrO2 thin films, Journal of Applied Physics 111, 07C302 (2012)

12. D. Mukherjee, N.S. Bingham, M.H. Phan, H. Srikanth, P. Mukherjee, and S.

Witanachchi, Ziz-zag Interface and strain-influenced ferromagnetism in epitaxial

Mn3O4/La0.7Sr0.3MnO3 thin films grown on SrTiO3 (100) substrates, Journal of

Applied Physics 111, 07D730 (2012)

Page 170: Magnetism in Complex Oxides Probed by Magnetocaloric

154

13. D. Mukherjee, N.S. Bingham, M. Hordagoda, M.H. Phan, H. Srikanth, P.

Mukherjee, and S. Witanachchi, Influence of microstructure and interfacial strain

on the magnetic properties of epitaxial Mn3O4/La0.7Sr0.3MnO3 layered-composite

thin films, Journal of Applied Physics 112, 083910 (2012)

14. D. Mukherjee, T. Dhakal, N.S. Bingham, M.H. Phan, H. Srikanth, P. Mukherjee,

and S. Witanachchi Role of crystal orientation on the magnetic properties of

CoFe2O4 thin films grown on Si (100) and Al2O3 (0001) substrates using pulsed

laser deposition, Physica B: Condensed Matter 406, 2663 (2011)

15. C. Miller, D. Williams, N.S. Bingham, and H. Srikanth, Magnetocaloric Effect

in Gd/W Thin Film Heterostructures, Journal of Applied Physics 107, 09A903

(2010)

16. D. Mukherjee, R. Hyde, M. Hordagoda, N.S. Bingham, H. Srikanth, S.

Witanachchi, and P. Mukherjee, Challenges in the stoichiometric growth of

polycrystalline and epitaxial PbZr0.52Ti0.48O3/La0.7Sr0.3MnO3 multiferroic

heterostructures using pulsed laser deposition, Journal of Applied Physics 112,

064101 (2012)

17. N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija, and C. Leighton , Magnetic

anisotropy and spin-lattice coupling in Pr0.5Sr0.5MnO3, Physical Review B, 2013

(under consideration)

18. R. Caballero-Flores, N.S. Bingham, M.H. Phan, M.A. Torija, C. Leighton, V.

Franco, A. Conde, and H. Srikanth, Magnetocaloric effect and critical behavior in

Pr0.5Sr0.5MnO3: An analysis of the validity of the Maxwell relation and the nature

of phase transitions, Physical Review B, 2013 (under review)

Page 171: Magnetism in Complex Oxides Probed by Magnetocaloric

155

19. N.S. Bingham, P. Lampen, M.H. Phan, S.W. Cheong, and H. Srikanth, A

complex magnetic diagram and magnetocaloric effect in Ca3Co2O6, Physical

Review B 2013 (rapid communication, under consideration)

20. N.S. Bingham, R. Caballero-Flores, M.H. Phan, M.A. Torija, C. Leighton, V.

Franco, A. Conde, and H. Srikanth, Magnetocaloric effect cross a coupled

structural/magnetocrystalline anisotropy transition in Pr1-xSrxMnO3 (x = 0.3, 0.35,

0.4, and 0.5) cobaltites, Applied Physics Letters, 2013 (under consideration)

21. A.R. Gorges, N.S. Bingham, M.K. DeAngelo, M.S. Hamilton, and J.L. Roberts,

Light Assisted Collisional Loss in 85/87 Rb Ultracold Optical Trap, Physical

Review A 78, 033420 (2008)

22. W.M. Tiernan, N.S. Bingham, J.C. Combs, Magnetization and Resistance of

Melt-Textured Growth YBCO Near TC and at Low magnetic Fields, AIP Conf.

Proc. Vol. 850, pp. 459-460 (2006)

Page 172: Magnetism in Complex Oxides Probed by Magnetocaloric

156

APPENDIX B LIST OF CONFERENCE PRESENTATIONS

1. N.S. Bingham, M.H. Phan, H. Srikanth, and S.W. Cheong, Complex magnetism

in geometrically frustrated spin-chain Ca3Co2O6 probed by transverse

susceptibility and magnetocaloric effect, 12th INTERMAG/MMM Joint

conference, Chicago, January13th -18th

, 2013

2. D. Mukherjee, M. Hordagoda, R. Hyde, N.S. Bingham, H. Srikanth, P.

Mukherjee, S. Witanachchi, Epitaxial Growth of Multiferroic Heterostructures of

Magnetic and Ferroelectric Oxides using the Dual-laser Ablation Technique,

American Vacuum Society (AVS) 59, Tampa, FL, Oct. 28th

- Nov 2nd

, 2012

3. D. Mukherjee, M. Hordagoda, R. Hyde, N.S. Bingham, H. Srikanth, P.

Mukherjee, S. Witanachchi, Role of Dual-laser Ablation in Controlling Mn Oxide

Precipitation during the Epitaxial Growth of Mn Doped ZnO Thin Films with

Higher Doping Concentrations, American Vacuum Society (AVS) 59, Tampa,

FL, Oct. 28th

- Nov 2nd

, 2012

4. N.S. Bingham, R. Caballeo-Flores, M.H. Phan, M.A. Torija, C. Leighton, and H.

Srikanth, Magnetocaloric effect and critical behavior in Pr0.5Sr0.5MnO3: An

analysis on the validity of Maxwell relation and nature of phase transitions.

INTERMAG 2012, Vancouver B.C., May 7 – 11, 2012.

5. N.S. Bingham, T.L. Phan, M.H. Phan, S.C. Yu and H. Srikanth, Phase

coexistence and magnetocaloric effect in Sm1-xSrxMnO3 (x = 0.42, 0.44, 0.46)

manganites, 56th

annual Magnetism and Magnetic Materials (MMM) conference,

Scottsdale, Az., Oct. 30 - Nov 4, 2011

Page 173: Magnetism in Complex Oxides Probed by Magnetocaloric

157

6. N.S. Bingham, M.H. Phan, H. Srikanth, C.L. Zhang, and S.W. Cheong,

Multiphase transitions and complex phase diagram in mixed phase

(La,Pr,Ca)MnO3 manganites, 56th

annual Magnetism and Magnetic Materials

(MMM) conference, Scottsdale, Az., Oct. 30 - Nov 4, 2011

7. P. Lampen, N.S. Bingham, M.H. Phan, H. Srikanth, C.L. Zhang, S. W. Cheong,

T.D. Hoang, and H.D. Chinh, Impact of nanostructuring on the magnetic and

magnetocaloric properties of La0.25Pr0.375Ca0.375MnO3, 56th

annual Magnetism and

Magnetic Materials (MMM) conference, Scottsdale, Az., Oct. 30 - Nov 4, 2011

8. D. Mukherjee, N.S. Bingham, M.H. Phan, H. Srikanth, P. Mukherjee and S.

Witanachchi, Ziz-zag interface and strain-influenced ferromagnetism in epitaxial

Mn3O4/La0.7Sr0.3MnO3 thin films grown on MgO (100) and SrTiO3 (100)

substrates , 56th

annual Magnetism and Magnetic Materials (MMM) conference,

Scottsdale, Az., Oct. 30 - Nov 4, 2011

9. N.S. Bingham, P. Lampen1 M.H. Phan, H. Srikanth, C.L. Zhang, S.W. Cheong,

T. H. Hoang, and H. D. Chinh, Impact of nanostructuring on the magnetic and

magnetocaloric properties of phase separated LaPrCaMnO3 manganites, Presented

at 1st Centennial of Superconductivity: Trends on Nanoscale Superconductivity

and Magnetism International Workshop, Cali, Colombia, June 29th

- July 1st, 2011

10. N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija and C. Leighton,

Magnetocaloric effect across the coupled structural magnetocrystalline anisotropy

transition in Pr1-xSrxCoO3 (x = 0.3-0.5), Presented at APS March Meeting, Dallas,

Texas, March 21 - 25, 2011

Page 174: Magnetism in Complex Oxides Probed by Magnetocaloric

158

11. M.H. Phan, N.S. Bingham, H. Srikanth, C.L. Zhang, and S.W. Cheong, Probing

multiple magnetic transitions and phase coexistence in mixed phase manganites,

Presented at APS March Meeting, Dallas, Texas, March 21 - 25, 2011

12. P. Lampen, N.S. Bingham, M.H. Phan, H. Srikanth, T.D. Hoang, and H.D.

Chinh, Influence of particle size on the magnetic and magnetocaloric properties of

nanocrystalline La2/8Pr3/8Ca3/8MnO3, Presented at APS March Meeting, Dallas,

Texas, March 21 - 25, 2011

13. D.V. Williams, C. Bauer, N.S. Bingham, H. Srikanth and C.W. Miller, Impact of

post-deposition annealing on the magnetic entropy change in Gd thin films. 55th

annual Magnetism and Magnetic Materials (MMM) conference, Nov. 14-18, 2010

Atlanta GA. (Winner of the best poster award)

14. N. S. Bingham, M.H. Phan, M. A. Torija, C. Leighton, and H. Srikanth,

Influence of the coupled structural/magnetocrystalline anisotropy transition on

magnetic entropy change in Pr1-xSrxCoO3, 55th

annual Magnetism and Magnetic

Materials (MMM) conference, Atlanta, GA, Nov. 14-18, 2010

15. N.S. Bingham, M.H. Phan, H. Srikanth, M. A. Torija, C. Leighton,

Magnetocaloric Effect and Refrigerant Capacity in Charge-Ordered

Pr0.5Sr0.5MnO3, APS March Meeting, Portland Oregon, March 15—19, 2010

16. N. Laurita, S. Chandra, N.S. Bingham, M.H. Phan, H. Srikanth, T. H. Hoang, H.

D. Chinh, T.Z. Ward, J. Shen, Phase coexistence and collapse of charge ordering

in low-dimensional (La,Pr)CaMnO3, APS March Meeting, Portland Oregon,

March 15-19, 2010

Page 175: Magnetism in Complex Oxides Probed by Magnetocaloric

159

17. M.H. Phan, N.S. Bingham, H. Srikanth, M. Torija and C. Leighton,

Magnetocaloric effect and transverse susceptibility in Pr-Sr-Mn-O, 2009

INTERMAG Conference, Sacramento CA, May 4-8, 2009

18. M.H. Phan, N.S. Bingham, N. A. Frey, M. A. Torija, C. Leighton and H.

Srikanth, Probing magnetic anisotropy and phase transitions in PSMO using RF

transverse susceptibility, 11th

joint MMM-INTERMAG conference, Washington

DC, January 18-22, 2010

19. M.H. Phan, M. B. Morales, N.S. Bingham, S. Chandra, C.L. Zhang, S.-W.

Cheong, T.D. Hoang, H. D. Chinh, and H. Srikanth, Collapse of charge-order and

enhanced magnetocaloric effect in nanostructured (La,Pr,Sr)CaMnO3, 11th

joint

MMM-INTERMAG conference, Washington DC, January 18-22, 2010

20. N.S. Bingham, M.H. Phan, H. Srikanth, M.A. Torija, and C. Leighton, A

Comparative Study of the Influence of First and Second Order Transitions on the

Magnetocaloric Effect and Refrigerant Capacity in Half-doped Manganites, 2nd

Annual IEEE Summer school, Nanjing, China, Sept. 2009, Poster

21. W.M. Tiernan, N.S. Bingham, J.C. Combs, Magnetization and Resistance of

Melt-Textured Growth YBCO Near TC and at Low magnetic Fields, 24th

International Conference on Low Temperature Physics, Orlando, FL, Aug. 2005

22. W.M. Tiernan, N.S. Bingham, J.C. Combs, Magnetization and Resistance of

Melt-Textured Growth YBCO Near TC and at Low magnetic Fields, Colorado

Research Symposium, Grand Junction, Colorado, Aug. 2003