131
ABSTRACT PROPERTIES OF TRANSITION METAL OXIDES FOR GAS SEPARATION AND OXYGEN STORAGE APPLICATIONS Steven Remsen, Ph.D. Department of Physics Northern Illinois University, 2011 Bogdan Dabrowski, Director Ceramic oxide materials have been heavily researched for selective oxygen separation, oxygen storage, chemical energy conversion, and related energy applications. They exhibit a range of properties, such as large reversible changes in oxygen content and enhanced electronic and oxygen ion conductivity that are temperature and oxygen atmosphere dependent, which are ideal for such applications. This work explores the application-related properties of new hexagonal (R = Dy, Y and M = Mn) and perovskite (R = La, Sr and M = Mn, Fe, Co) RMO 3+δ materials with reversible oxygen storage/release capacities and mixed oxygen ion electronic conductivity, respectively. These materials were achieved by special synthesis techniques guided by the temperature and oxygen content dependence of the Goldschmidt tolerance factor. Hexagonal P6 3 cm Dy 1-x Y x MnO 3+δ materials were found to have reversible oxygen storage/release capacities in oxygen atmospheres comparable to best-known materials (~2000 μmol-O/g) while operating at considerably lower temperatures. Thermogravimetric measurements in oxygen atmosphere and neutron and x-ray diffraction identified two new hexagonal phases with δ ≈ 0.25 and 0.40. These large uptakes of oxygen at 200 – 300°C were observed

ABSTRACT PROPERTIES OF TRANSITION METAL … of transition metal oxides for gas separation and oxygen storage applications . ... 4.2 synthesis and stability

Embed Size (px)

Citation preview

ABSTRACT

PROPERTIES OF TRANSITION METAL OXIDES FOR GAS SEPARATION AND OXYGEN STORAGE APPLICATIONS

Steven Remsen, Ph.D. Department of Physics

Northern Illinois University, 2011 Bogdan Dabrowski, Director

Ceramic oxide materials have been heavily researched for selective oxygen

separation, oxygen storage, chemical energy conversion, and related energy

applications. They exhibit a range of properties, such as large reversible changes in

oxygen content and enhanced electronic and oxygen ion conductivity that are

temperature and oxygen atmosphere dependent, which are ideal for such applications.

This work explores the application-related properties of new hexagonal (R = Dy, Y

and M = Mn) and perovskite (R = La, Sr and M = Mn, Fe, Co) RMO3+δ materials with

reversible oxygen storage/release capacities and mixed oxygen ion electronic

conductivity, respectively. These materials were achieved by special synthesis

techniques guided by the temperature and oxygen content dependence of the

Goldschmidt tolerance factor. Hexagonal P63cm Dy1-xYxMnO3+δ materials were found

to have reversible oxygen storage/release capacities in oxygen atmospheres

comparable to best-known materials (~2000 μmol-O/g) while operating at

considerably lower temperatures. Thermogravimetric measurements in oxygen

atmosphere and neutron and x-ray diffraction identified two new hexagonal phases

with δ ≈ 0.25 and 0.40. These large uptakes of oxygen at 200 – 300°C were observed

to completely release when materials were heated to 275 – 375°C or exposed to lower

oxygen partial-pressures. Oxygen non-stoichiometry was also found to have

considerable impact on the structural, thermal/chemical expansion, transport, and

magnetic properties of Dy1-xYxMnO3+δ materials. Several perovskite RMO3+δ materials

were discovered that displayed properties indicating their superior bulk oxygen ion

and electrical conductivity compared to other commonly used mixed oxygen ion

electronic conducting materials. Thermogravimetric, conductivity, and

electrochemical impedance measurements of La1-xSrxFe1-yCoyO3+δ, La.2Sr.8MnO3+δ,

and SrFe.7Mn.3O3+δ samples suggested these compounds to have large fractional

oxygen ion vacancies, high electric conductivity, and exceptionally low activation

energies of oxygen ion conductivity. These materials were also found to have

considerably high oxygen storage/release capacities (~3000 μmol-O/g) for hydrogen-

oxygen atmosphere cycling. Several La1-xSrxFe1-yCoyO3+δ samples showed the

remarkable ability to reabsorb oxygen at room temperature, making them excellent

candidates for application.

NORTHERN ILLINOIS UNIVERSITY DE KALB, ILLINOIS

MAY 2011

PROPERTIES OF TRANSITION METAL OXIDES FOR GAS SEPARATION

AND OXYGEN STORAGE APPLICATIONS

BY

STEVEN REMSEN ©2010 Steven Remsen

A DISSERATION SUBMITTED TO THE GRADUATE SCHOOL

IN PARTIAL FULFILMENT OF THE REQUIREMENTS

FOR THE DEGREE

DOCTOR OF PHILOSOPHY

DEPARTMENT OF PHYSICS

Dissertation Director: Bogdan Dabrowski

All rights reserved

INFORMATION TO ALL USERSThe quality of this reproduction is dependent on the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscriptand there are missing pages, these will be noted. Also, if material had to be removed,

a note will indicate the deletion.

All rights reserved. This edition of the work is protected againstunauthorized copying under Title 17, United States Code.

ProQuest LLC.789 East Eisenhower Parkway

P.O. Box 1346Ann Arbor, MI 48106 - 1346

UMI 3457804

Copyright 2011 by ProQuest LLC.

UMI Number: 3457804

ACKNOWLEDGEMENTS

I would like to thank, first and foremost, my advisor, Dr. Bogdan Dabrowski,

whose guidance, patience, and mentorship made this project possible and helped

transform a young man into a scientist. I also gained invaluable experience and

assistance throughout this project from Dr. Stanislaw Kolesnik, Dr. Omar Chmaissem,

Dr. Leopoldo Suescun, Dr. Konrad Świerczak, Dr. Andrzej Szewezyk, Dr. David

Carter, Dr. Brian Ingram, Dr. Terry Cruse, and James Mais. I am also indebted to my

fellow graduate students, Stephen Boona, Michael Himes, Dr. Sevda Avci, Manasa

Majjiga, Tim Maxwell, Donald Johnson, Seyed Ahmad Sabok-Sayr, Ben Stillwell,

Simon Murphy, and Nathan Styx, for their laboratory support, teamwork, and

friendship. I would like to thank Amber Remsen for her proofreading of this

dissertation. Finally, I would like to acknowledge the National Science Foundation for

funding this work (NSF-DMR 0706610).

To my father, mother, and wife, Amber,

whose constant love, support, and inspiration made this work possible.

TABLE OF CONTENTS

Page

LIST OF TABLES .................................................................................................... vi

LIST OF FIGURES .................................................................................................. vii

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

1.1 Air Separation Methods, Past to Present ............................................................ 1

1.2 Elevated-Temperature Ceramic Materials for Air Separation and Oxygen Storage ................................................................................................................... 3

1.3 Other Applications of Ceramic Oxygen Sorbents and Mixed Electronic Ionic Conductors ............................................................................................................. 8

1.3.1 Chemical Looping Combustion .................................................................. 8

1.3.2 Solid Oxide Fuel Cells .............................................................................. 10

1.3.3 Oxygen Sensors ........................................................................................ 12

1.3.4 Waste Heat Air Separation for High-Temperature Systems ...................... 13

CHAPTER 2: EXPERIMENTAL METHODS ......................................................... 15

CHAPTER 3: STUDY OF HEXAGONAL Dy1-xYxMnO3+δ (-0.2 ≤ δ ≤ 0.4) MATERIALS FOR OSC APPLICATIONS ............................................................. 18

3.1 Introduction .................................................................................................... 18

3.2 Synthesis and Stability .................................................................................... 23

3.3 Oxygen Storage Measurements ....................................................................... 29

3.4 Crystal Structure ............................................................................................. 39

3.5 Thermal and Chemical Expansion ................................................................... 46

3.6 Transport Properties ........................................................................................ 52

v

Page

3.7 Magnetic Measurements of DyMnO3+δ (x = 0) ................................................ 59

3.8 Conclusions .................................................................................................... 62

CHAPTER 4: STUDY OF PEROVSKITE La1-xSrxFe1-yCoyO3+δ (x ≥ 0.5), La0.2Sr0.8MnO3+δ, AND SrFe0.7Mn0.3O3+δ MATERIALS FOR MIEC AND OXYGEN CARRIER APPLICATIONS .................................................................................... 65

4.1 Introduction .................................................................................................... 65

4.2 Synthesis and Stability .................................................................................... 71

4.3 Oxygen Storage and Oxygen Content Behavior Measurements ....................... 74

4.4 Electrical Conductivity ................................................................................... 83

4.5 Total Ionic Conductivity ................................................................................. 89

4.6 Thermal and Chemical Expansion ................................................................... 99

4.7 Conclusions ...................................................................................................104

REFERENCES ...................................................................................................... 107

APPENDIX: STRUCTURAL PARAMETERS AND AGREEMENT FACTORS FOR Dy1-xYxMnO3+δ ...................................................................................................... 117

vi

LIST OF TABLES

Page

Table 3.1 Hexagonal-Perovskite transition δ-values: t(δTheo.) = 0.855, δobs. are observed values from TGA, and values of t are calculated with Shannon values……………………………………………………………………..26

Table 3.2 OSC (μmol-O/g) of Dy1-xYxMnO3+δ. …………………………………….30

Table 3.3 List of annealed DyMnO3+δ samples……………………………………...40

Table 3.4 Refinable positions of Dy1-xYxMnO3+δ…………….……………………..45

Table 3.5 Activation energies in their respective temperature ranges and coefficients of determination of linear fits based on the phonon assisted conduction……………………………………………………….54

Table 3.6 Activation energies in their respective temperature ranges and coefficients of determination of linear fits based on semiconductor behavior………...54

Table 4.1 Oxygenation temperatures (TOx, RT = room temperature) of samples after reduction at 300-400°C in 42% H2 and observed OSC and calculated OSC……………………………………………………….80

Table 4.2 Activation energies for electronic conduction and observed calculated transition temperatures (*referenced from text)…………………………..87

Table 4.3 Activation energies for total ionic conduction (*referenced in text).….....95

Table 4.4 TEC (*10-6 K-1) and CE (*10-2 mol-1) values and effective TEC is calculated without separating CE (*referenced in text)…………………101

Table A.1 Atomic coordinates, lattice parameters, oxygen site occupancies, thermal factors and agreement factors of P63cm DyMnO3+δ (*Deqv.) from IPNS data………………………………………………………………..117

Table A.2 Atomic coordinates, lattice parameters, oxygen site occupancies, thermal factors and agreement factors of P63cm Dy1-xYxMnO3+δ from Echidna data……………………………………………………….118

vii

LIST OF FIGURES

Page

Figure 1.1 Original drawing of Linde’s cryogenic distillation apparatus for air separation from his US patent filed in 1895……………………………..1

Figure 1.2 Schematic drawing of thermal swing absorption (TSA) with ABO3 sorbent material…………………………………………………………..5

Figure 1.3 Schematic drawings of oxygen separation membranes with solid oxide electrolyte (left) and mixed ionic electronic conductor (right)…………..8

Figure 1.4 Schematic drawing of chemical looping combustion (CLC) with perovskite oxygen carrier……………………………………………….10

Figure 1.5 Schematic drawing of a solid oxide fuel cell (SOFC)………………….11

Figure 3.1 Schematic drawings of the crystal structures and 3d orbitals of orthorhombic perovskite Pnma (top), hexagonal P63cm (middle), and pyrochlore Fd3m (bottom) where red spheres, yellow polyhedrons, and blue spheres represent O2- anions, MnOn polyhedrons, and R3+ cations, respectively………………………….………………………………….19

Figure 3.2 XRD patterns examples of hexagonal DyMnO2.963 and YMnO3.004 after initial synthesis………………………………………………………….24

Figure 3.3 Phase diagram during synthesis of Dy1-xYxMnO3+δ…………………....25

Figure 3.4 TGA of annealing after initial synthesis in argon (curve 1) and subsequent reduction of DyMnO3+δ in 42%H2 to Dy2O3 and MnO…….29

Figure 3.5 TGA oxygen content vs. temperature for Dy1-xYxMnO3+δ with a heating (top) and a cooling (bottom) rate of 0.1°C/min in O2………………….31

Figure 3.6 TGA of percent weight vs. time for Dy1-xYxMnO3+δ………...…………31

Figure 3.7 Dy1-xYxMnO3+δ TSA relevant temperatures……………………………32

Figure 3.8 TGA reductions in 21% O2 of Dy1-xYxMnO3+δ to the stable 3.0 oxygen content…………………………………………………………………..33

Figure 3.9 TGA of Dy1-xYxMnO3+δ switching between Ar and O2 atmospheres at 330, 300, 280, 250, and 230°C for x = 0, 0.3, 0.5, 0.7, and 1, respectively……………………………………………………………..35

viii

Page Figure 3.10 Example of XRD patterns for DyMnO3+δ, with δ = -0.037, 0.0, 0.18,

0.21, 0.24, 0.35.........................................................................................41

Figure 3.11 XRD patterns of Dy1-xYxMnO3+δ (x = 0.5, δ = 0.37 and x = 0.3, δ = 0.40), which show an increased ratio of the Hex3/Hex2 phases after high-pressure annealings………………………………………………..43

Figure 3.12 XRD comparison for DyMnO3+δ samples: P63cm (δ = 0), nearly single phase Hex2 (δ = 0.24), and mixed phase of Hex2 and Hex3 (δ = 0.35)………………………………………………………………..43

Figure 3.13 Example of NPD pattern for DyMnO2.963. Crosses are observed data and the line below is the difference between experimental data and best fit calculated from the Rietveld refinement method………………….……45

Figure 3.14 Dilatometry measurement of perovskite DyMnO3.0 in 21% O2………...48

Figure 3.15 TGA measurement of hexagonal DyMnO3+δ in 21% O2.………………49

Figure 3.16 Dilatometry measurement of hexagonal DyMnO3+δ in 21% O2………..50

Figure 3.17 TEC values for P63cm and Hex2 versus average R-site ionic radius…...50

Figure 3.18 Chemical expansion values for Dy1-xYxMnO3+δ….……………………50

Figure 3.19 a: Conductivity measurements of Dy1-xYxMnO3+δ in air, b: phonon-assisted conduction model, σT ~ e-Ea/kT, of P63cm (high temperature ) and Hex2 (low-temperature) phases, c: semiconductor conduction model, σ ~ e-Ea/kT, of P63cm (high-temperature) and Hex2 (low-temperature) phases……………………………………………….…………………..53

Figure 3.20 Resistivity measurements of DyMnO3+δ in cycling O2 and Ar atmospheres at 330°C…………………………………………………...58

Figure 3.21 Resistivity measurements of YMnO3+δ in cycling O2 and Ar atmospheres at 230°C…………………………………………………………………58

Figure 3.22 Inverse of DC susceptibility measurements of DyMnO3+δ, δ = -0.037 and 0.00 curves overlapping on top and δ = 0.22 curve on bottom…………59

Figure 3.23 Temperature derivative of inverse DC susceptibility of DyMnO3+δ where TN was found to be slightly lower for non-stoichiometric sample; insert of δ = 0.22………………………………………………………………61

ix

Page Figure 3.24 Temperature derivative of inverse DC susceptibility of DyMnO3+δ where

transition temperature was found to be lower for non-stoichiometric sample; insert of δ =0.22………………………………….……………61

Figure 4.1 Representational drawing of the Brownmillerite crystal structure……..70

Figure 4.2 XRD patterns of the perovskite a) La0.3Sr0.7Fe0.5Co0.5O3, b) La0.2Sr0.8MnO3, c) SrFe0.7Mn0.3O3 structures after initial synthesis……………………………………………………….………..72

Figure 4.3 Example of TGA hydrogen reduction for La0.3Sr0.7Fe0.5Co0.5O3+δ (1°C/min, 42% H2/Ar) to determine oxygen content after initial synthesis………………………………………………………………...72

Figure 4.4 Fraction of oxygen vacancies versus temperature in 21% O2/Ar for La0.2Sr0.8MnO3+δ, SrFe0.7Mn0.3O3+δ, and La0.3Sr0.7Fe0.5Co0.5O3+δ determined by TGA (δ < 0)……………………………………………..76

Figure 4.5 Oxygen vacancy-ordered phases of SrMnO3+δ: a) monoclinic Sr7Mn7O19 (SrMnO2.714), b) tetragonal Sr5Mn5O13 (SrMnO2.6), and c) orthorhombic Sr2Mn2O5 (SrMnO2.5)…………………………………………………...77

Figure 4.6 Example of TGA hydrogen reduction for La0.2Sr0.8MnO3+δ (0.1°C/min, 42% H2/Ar), which shows enhanced stability at δ = -0.4 and δ = -0.5 of vacancy-ordered phases...………………………………………………77

Figure 4.7 Representational drawing of the polyhedral network with oxygen ordered vacancies in Sr8Fe4Co4O23 sample……………………………………...77

Figure 4.8 Example of TGA reduction for La0.3Sr0.7Fe0.6Co0.4O3+δ to Fe3+/Co3+ (δ = -0.55) with 42% H2/Ar at 350°C (1°C/min heating/cooling), followed by room-temperature oxygenation in O2…………………………………...79

Figure 4.9 Example of TGA reduction for SrFe0.7Mn0.3O3+δ to Brownmillerite phase (δ ≈ -0.5) with 42% H2/Ar at 400°C (0.5°C/min heating and fast cooling), followed by heating 1°C/min in O2 to 500°C…………………………..79

Figure 4.10 Example of TGA reduction for La0.2Sr0.8MnO3+δ to (La0.2Sr0.8)5Mn5O13(δ ≈ -0.4) with 42% H2/Ar at 350°C (0.5°C/min heating and fast cooling), followed by heating 1°C/min in O2 to 500°C…………………………..79

Figure 4.11 Conductivity measurements of La1-xSrxFe1-yCoyO3+δ, SrFe0.7Mn0.3O3+δ, and La0.2Sr0.8MnO3+δ………………………………………………...….83

x

Page Figure 4.12 Logarithm of the product of conductivity and temperature versus inverse

temperature of La1-xSrxFe1-yCoyO3+δ, SrFe0.7Mn0.3O3+δ, and La0.2Sr0.8MnO3+δ to find electrical activation energies………………....86

Figure 4.13 Diagrams of half (left) and full (right) test cells for EIS measurements………………………………………………...…………91

Figure 4.14 Circuit diagram of a Randles cell and Nyquist plot of its response.........94

Figure 4.15 Nyquist plots of La0.3Sr0.7Fe0.5Co0.5O3+δ half cell at various temperatures.............................................................................................95

Figure 4.16 Arrhenius plots for La0.2Sr0.8MnO3+δ, SrFe0.7Mn0.3O3+δ, and La0.3Sr0.7Fe0.5Co0.5O3+δ half cells and La0.3Sr0.7Fe0.5Co0.5O3+δ full cell…95

Figure 4.17 Arrhenius plots for La0.2Sr0.8MnO3+δ (LSM* Stillwell data)…………...96

Figure 4.18 Arrhenius plots for SrFe0.7Mn0.3O3+δ (SFM* Stillwell data)…………...97

Figure 4.19 Dilatometry measurements of La1-xSrxFe1-yCoyO3+δ, SrFe0.7Mn0.3O3+δ, La0.2Sr0.8MnO3+δ, and reference La0.8Sr0.2MnO3+δ……………………100

Figure 4.20 Dilatometry measurement of La0.4Sr0.6Fe0.5Co0.5O3+δ during oxygenation from reduced state (δ = -0.19) to better determine chemical expansion……………………………………………………………...102

CHAPTER 1: INTRODUCTION

1.1 Air Separation Methods, Past to Present

Methods to separate and enrich oxygen from the air were first developed by

both Carl Linde in Germany and William Hampson in England independently in 1895

by a method known as cryogenic distillation [1][2]. At a basic level, this method

primarily relies on the difference in boiling points of the primary components of air,

which are oxygen (-183.0°C), nitrogen (-195.8°C), and argon (-189.3°C). Thus, by

slowly cooling from each boiling point to the next, each component can be siphoned

off in liquid form and distilled one at a time (Figure 1.1). Soon after, primarily due to

the invention of oxy-acetylene welding and new methods for smelting steel from iron

2

ore, the demand for large quantities of high-purity oxygen greatly increased. By the

end of World War II, new methods and technologies were put in place that

significantly improved upon Linde’s and Hampson’s original methods and made

oxygen separation and purification commercially viable. Improvements to the original

Linde-Hampson machine still continue today [3][4] and cryogenic distillation has

remained the technology of choice to meet the world’s demand for the production of

large volumes of high-purity oxygen, which in 2001 exceeded 100 Mtons/yr and

continues to be one of the top-ten traded elemental commodities worldwide today

[3][5]. In the past twenty-five years, pressure swing adsorption (PSA) and vacuum-

pressure swing adsorption (VPSA) have also become cost-effective methods to

produce oxygen of smaller volumes with lower purities (90 - 95%) [6]. These,

typically smaller systems, function at high pressure from low to room temperatures by

using a two high-pressure chamber system, each with nitrogen and oxygen sorbent

pressure-sensitive materials. Materials for oxygen and nitrogen sorbents have typically

been various zeolite compounds and carbon molecular sieves, respectively [7];

however, there has been recent work with porous, metal-organic frameworks that may

have the potential to replace zeolite materials as oxygen sorbents in these systems [8].

The need for high-purity oxygen for various industrial production processes (e.g.,

steel, glass, plastics, etc.), medical applications, welding methods, and a rocket fuel

component [9] will certainly insure a strong continued future demand for high-purity

oxygen and the need for the development of new materials and improved methods for

air separation and oxygen storage.

3

1.2 Elevated-Temperature Ceramic Materials for Air Separation

and Oxygen Storage

Recently, elevated-temperature ceramic materials have been increasingly

researched by the Department of Energy (DOE), the Linde Group, Air Products and

Chemicals, Praxair, and many small start-up companies for air separation and oxygen

storage methods. Compared to cryogenic distillation, these materials have great

potential to both significantly lower capital and operation costs (projected 30%

reduction), while also operating with considerably less power consumption [10].

Ceramic materials for elevated-temperature air separation come in two primary

varieties: materials with reversible oxygen storage and release capacities (OSC, OSC

materials) and oxygen-ion conductors.

OSC materials can reversibly and selectively absorb, store, and release oxygen.

The mechanism for oxygen sorption in these materials is, most commonly, dependent

on the creation of oxygen ion vacancies or interstitial sites at high temperatures, due to

changes in stoichiometry or intrinsic defects such as Schottky or Frenkel defects.

These changes in oxygen content are usually accommodated by the material’s crystal

structure, thus no structural phase transitions occur and the change in oxygen content

is easily reversible. The rate of oxygen ion diffusion for OSC materials can be

approximated by the Arrhenius equation (where D0 is diffusion rate at

infinite temperature, EA is activation energy, R is the gas constant, and T is

temperature). For application, ideal OSC materials must have large values of OSC

4

(typically measured in moles of oxygen per weight of material) and their

absorption/desorption of oxygen must occur over a narrow temperature range at near

atmospheric conditions. Additional properties, such as oxygen partial-pressure

dependence of absorption/desorption, exothermic absorption and endothermic

reduction, and stability/recoverability in strong reducing conditions (e.g., CO and H2

atmospheres at high temperatures), are also desired and being researched for various

applications (see next section). A common method for air separation, which relies on

temperature-dependent oxygen absorption/desorption, is thermal swing absorption

(TSA). In this method, multiple beds of sorbent cycle in between two chambers that

are at different temperatures. This creates oxygen-rich and oxygen-deficient

atmospheres in each chamber (Figure 1.2). More recently, Lin et al. patented a method

in 2000 for perovskite materials, which combines TSA and PSA techniques in a

process named ceramic autothermal recovery (CAR) [11]. Again, multiple beds filled

with sorbent are cycled through two chambers with a temperature gradient; however,

in this method the chambers are also maintained at different oxygen partial-pressures.

Again, this creates two chambers that are oxygen rich and deficient. Sorbents designed

for CAR also have strong endothermic reduction and exothermic absorption; therefore,

the process operates autothermally, needing little or no heat added once operational.

5

Commercially, CeO2-ZrO2 compositions have been the recent, ceramic OSC

materials of choice for air separation, which function around 500°C and have OSCs of

~400 – 500 μmol-O/g in oxygen atmosphere [12] [13] or as high as 1500 μmol-O/g

with 20% H2 reversible reduction [14]. The OSC of these materials comes from the

ability of the CeO2 fluorite structure to accommodate a large number of oxygen ion

vacancies when doped and the ease of reducing the Ce4+ cation to Ce3+. Recent studies

with Ce1-xCrxO2 have further boosted the OSC of the fluorite structure as high as 2500

μmol-O/g in air and hydrogen atmospheres but require considerably higher reduction

temperatures (550 – 700°C) [15]. Currently, P63mc RBaCo4O7+δ (R = Y, Dy, Ho, Er,

Tm, Yb, and Lu) [16]–[18] and YBaCo4-xAlxO7+δ [19] materials have the best reported

OSC values at low temperatures, storage up to ~2700 μmol-O/g, and completely

desorb at ~400 – 425°C in O2. The ease of reversible phase transitions between the

hexagonal P63mc YBaCo4O7 and orthorhombic Pbc21 YBaCo4O8.1 (which is a mixture

of tetrahedrally and octahedrally coordinated cobalt) [20] is responsible for its oxygen

6

storage behavior. Transition between stable phases with large differences in oxygen

content, as seen here, is a new mechanism for OSC materials and has great potential

for storage and operation temperatures, as will be further discussed in Chapter 3 and 4.

Unlike OSC materials, which are a relatively new topic of study, oxygen ion

conductors have been studied over a hundred years. In 1899, Nernst reported O2- anion

conduction in ZrO2+9%Y2O3 (9YSZ) [21]. Oxygen ion conduction is directly related

to oxygen ion diffusion and can be approximated by the Arrhenius relation:

(where σ0 is a function of occupied ionic sites in the lattice); thus, bulk oxygen

ion conduction is mainly attributed by two material properties: the activation energy of

oxygen ion migration and the fraction of oxygen vacancies [22]. Typically, oxygen ion

conductors for air separation can be divided into two basic types: solid oxide

electrolytes, which have high O2- conductivity but are electrically insulating, and

mixed ionic electronic conductors (MIEC), which have both high electric and O2-

conductivity. If an oxygen ion conductor (of either type) is used to separate two

different partial-pressures of oxygen, a measurable potential difference arises, which is

governed by the Nernst potential:

(where Ɛ is electromotive

force, F is the Faraday constant, and PO2/ P’O2 is the ratio of partial-pressures of

oxygen). Assuming that lattice diffusion of oxygen determines the overall rate of O2-

permeation, the oxygen flux through such a membrane can be approximated by

σ

σ

σ σ

by Wagner’s derivation [23] (where L is the

thickness of the membrane). Thus, oxygen can be “pumped” by transporting O2- ions

7

from the cathode to the anode side of an electrolyte material by applying current,

which can be used to create an effective oxygen pump (Figure 1.3) [24]. This has been

demonstrated with various stabilized zirconia materials [25]. However, the difficulty

of finding suitable gas-permeable electrodes and the required electric power for

operation (due to high resistivity) are currently limiting factors for practical

application. MIEC membranes, however, can selectively migrate oxygen ions without

electrodes or applied current by using the difference in partial-pressure of oxygen as a

driving force. O2 ionizes on the high-pressure side of the membrane by picking up

electrons in the MIEC’s conduction band and then migrates to the low-pressure side,

where the oxygen ions release electrons back to the membrane to reform oxygen

molecules. The O2- flux is charge compensated by the simultaneous flow of electronic

charge carrier flux in the opposite direction (Figure 1.3). Substituted perovskites, such

as La1-xAxM1-yM’yO3+δ (A= Ca, Sr, Ba; M/M’ = Cr, Mn, Fe, Co, Ni, Ga, δ < 0) [26] –

[31], have been thoroughly researched for MIEC for such application (discussed

further in Chapter 4) but currently require high temperatures for operation (600-

900°C). It is important to note that for many MIEC (σ σ ), conductivities can

be effectively treated as independent of oxygen partial-pressure and the oxygen flux

can be approximated as σ

; thus, EA and the number of

oxygen vacancies are the only inherently important material properties for these MIEC

for oxygen separation membrane applications.

8

1.3 Other Applications of Ceramic Oxygen Sorbents

and Mixed Electronic Ionic Conductors

1.3.1 Chemical Looping Combustion

There has been a growing concern about the increasing levels of CO2

emissions from fossil fuel combustion and the resulting effects on the earth’s climate.

A growing majority of the scientific community agrees that carbon emissions must be

greatly reduced in the near future; however, our strong dependence on fossil fuels and

their relatively large abundance makes replacing these systems exceedingly difficult.

Currently, the leading technology for “clean coal” energy production is to remove CO2

9

from flue gas of existing fossil fuel power plants [32]. Selectively capturing CO2 from

the complex mixture of high-temperature flue gases requires large amounts of energy,

which would significantly reduce the net efficiency of existing systems. One solution

to this problem is to combust with high partial-pressures of oxygen instead of with

ambient air, which has the added benefit of significantly reducing NOx and SOx

emissions [33][34]. The resulting combustion products are primarily H2O and CO2,

which are much easier to separate and store. One of the major hurdles for such “oxy-

fuel” combustion, as cited by a recent DOE report [35], is to develop improved and

cost-effective air separation units. Ceramic OSC and MIEC are uniquely qualified to

fulfill this role by operating with reactor waste heat or by direct exposure to the

reaction chamber. Methods by using waste heat include redirection of thermal energy

to OSC materials to perform a TSA process, as described in the previous section.

Chemical looping combustion (CLC) is another solution to add oxygen directly to the

combustion process (oxygen sorbents are typically called oxygen carriers for this

application) [36][37]. The CLC process is a two-step cycling procedure, where first,

reduction of the oxygen carrier takes place at the reactor to oxide fossil fuels and then,

the reduced oxygen carrier is removed from the reactor for reoxygenation (Figure 1.4).

Thus, materials for CLC must not only have large values of OSC but also must have

stability in highly reducing conditions at high temperatures, which are recoverable in

air to stoichiometric oxygen content. Previous materials studied for oxygen carriers

have been perovskite La1-xSrxFe1-yCoyO3-δ (where x ≥ 0.5 and 0 ≥ y ≥ 1) and various

metal oxides (e.g., Fe and Mn) [38]–[40].

10

1.3.2 Solid Oxide Fuel Cells

The first functional solid oxide fuel cells (SOFC) were demonstrated with

current densities of ~1 mA/cm2 at 1000°C by Baur and Preis in 1937, which were

largely based on the Nernst glower from the early 1900’s [41]. In the 1960’s,

significant progress toward commercial production was done by Westinghouse in the

U.S and by Brown, Boveri, and Cie in Germany, which achieved current densities up

to 100 mA/cm2 [42]. SOFC are multilayer, ceramic electrochemical conversion

devices that generate electricity using gaseous fuels and an oxidant at high

temperatures (600 – 1000°C). SOFC consist of an interconnect and three basic layered

components: a cathode, an electrolyte, and an anode (Figure 1.5). Oxygen is reduced

11

at the cathode ( ) by free electrons from the anode and hydrogen is

oxidized to H+ at the anode ( ) from O2- anions from the

cathode.

The electrolyte, as discussed in the previous section, is a good oxygen ion

conductor but is electrically insulating; thus, O2- anions conduct through the

electrolyte from the cathode to the anode and electrons flow through the interconnect.

Compared to traditional energy conversion systems, SOFC are not limited by the

Carnot efficiency, have superior fuel adaptability and reliability, and produce very low

levels of CO2, NOx, and SOx emissions [43]. Additionally, the efficiencies of SOFC

do not drop with decreases in scale, which is a significant problem with steam turbines

and internal combustion engines, making them ideal for personal transportation and

Third-World energy production. Compared to other types of fuel cells, SOFC also

typically have less corrosive and cheaper catalyst materials. However, SOFC do have

12

several major drawbacks, which have limited their development for wide-scale

application: high operation temperature, electrode over-potential, chemical reaction

between layers, thermal expansion mismatch of components, and ideal SOFC

materials are typically brittle in nature.

MIEC are ideal materials for cathodes in SOFC due to their high electronic and

oxygen ion conductivity. However, there are several other required properties for an

ideal cathode: chemical stability with the electrolyte, the interconnect, operation

environments, and manufacturing process; similar thermal expansion coefficients with

other cell materials; low-as-possible operation temperature; and reasonably high

porosity, as not to limit the triple phase boundary reaction. Again, as with separation

membranes, substituted perovskites, such as La1-xAxM1-yM’y (A= Ca, Sr, Ba; M/M’ =

Cr, Mn, Fe, Co, Ga) [44] – [54], have been thoroughly researched for cathode

materials (discussed further in Chapter 4) and, with clever design, have achieved

power densities above 1 W/cm2 at 600°C.

1.3.3 Oxygen Sensors

Electrolyte gas separation membranes, as described in the previous section, can

also be used as “reverse oxygen pumps” to generate a small electric difference in the

presence of different partial-pressures of oxygen, which is again governed by the

relation

. Commercial oxygen sensors of this type have

frequently been various types of stabilized zirconia and zeolites and are still actively

13

researched as a component for oxygen sensors [55]. Transition metal–oxides, such as

perovskite and hexagonal ABO3+δ, that are prone to oxygen non-stoichiometry under

different oxygen partial-pressures are also potential materials for oxygen sensing

applications. This is because changes in oxygen content of these materials typically

also results in noticeable changes in electric conductivity (due to changes in structure,

d-shell occupancy and spin state, exchange interaction, etc.). Though not yet

commercially practical, proof of principle of this method has been demonstrated with

perovskite LaMnO3-δ with various levels Sr hole doping and SrFeO3-δ at temperatures

above 500°C [56] – [59].

1.3.4 Waste Heat Air Separation for High-Temperature Systems

Currently, roughly over 80% of commercially produced oxygen is used in

high-temperature industrial productions process [9]: smelting steel from iron ore, glass

production, creating ethylene oxide from ethylene for ethylene glycol production, etc.

Furthermore, potential uses of ceramic OSC and MIEC materials currently being

researched operate at or have components that operate at elevated temperatures. In

addition to the previous applications discussed here, OSC materials are also being

researched for components to improve automotive exhaust catalysts [60], solar water

splitting [61], non-solid oxide hydrogen fuel cells [62], various non-aerobic oxidation

processes [63], and the production of syngas (H2, CO) by partial oxidation of methane

[64]. For any of these current and potential systems, the redirection of the large

14

amounts of waste heat generated from all these methods to ceramic OSC or MIEC

materials for onsite air separation would undoubtedly have potential net energy,

economic, and waste advantages versus air separation by high-pressure (zeolites or

metal-organic frameworks) or low-temperature (cryogenic distillation) methods.

15

CHAPTER 2: EXPERIMENTAL METHODS

Synthesis methods were done by solid-state reaction (see Sections 3.2 and 4.2

for details). X-ray powder diffraction (XRD) measurements were made with a Rigaku

D/MAX powder diffractometer in the 2θ = 20-70° range with CuKα radiation. Room-

temperature neutron powder diffraction (NPD) data were collected both with time-of-

flight measurements, conducted at Argonne National Laboratory’s former Special

Environment Powder Diffractometer (SEPD) and General Purpose Powder

Diffractometer (GPPD) at the Intense Pulsed Neutron Source (IPNS), and with a

wavelength of 2.4395Å, carried out on the Echidna High-Resolution Powder

Diffractometer at the Bragg Institute. Structural refinements of diffraction data were

performed by the Rietveld method with GSAS/EXPGUI suite programs [65];

theoretical XRD patterns were generated with PowderCell v.2.4 and representational

drawings of crystal structures were made with the assistance of DRAWxtl v.5.1.

Thermogravimetric analysis (TGA) measurements were made with Cahn TG171 and

Cahn TherMax700 thermobalances. TGA reaction gases consisted of several different

mixtures of ultra-high-purity (99.999%) oxygen, hydrogen, and argon gasses and were

flowed at a rate of 100ccm using a MKS flow controller. TGA measurements were

done up to 1100°C at heating and cooling rates of 0.1 – 1.0°/min and were measured

16

with a 5 μg precision. TGA samples were approximately 1 g and were suspended in an

alumina crucible with a Pt, Au, or Mo wire (Au and Mo for hydrogen firings).

Response from the wire and crucible were subtracted from the raw data by conducting

empty runs with identical conditions. Dilatometry measurements were made with a

Linseis Differential Dilatometer L75 and samples were measured with a 1 μm

precision. Reaction gases consisted of ultra-high-purity (99.999%) oxygen and argon

and 21% oxygen balanced with argon. These gases were flowed at an approximate rate

of roughly 100ccm. Thermal behavior of the dilatometer’s alumina piston and sample

holder were subtracted from the raw data by conducting runs with a piece of alumina

that was close in length to sample lengths in identical temperature and atmosphere

profiles. DC susceptibility was measured on cooling in a 1 kOe magnetic field using a

Quantum Design Physical Property Measurement System 6000. Resistivity

measurements were made on a homemade apparatus by the four-point probe technique

with Pt electrodes embedded in dense bars of sample. This method provided superior

response at high temperature versus standard methods that make use of metallic paints

(e.g., Ag) due to reduced contact resistance and paint’s tendency to delaminate or melt

at higher temperatures. This apparatus was fitted into a tube furnace and heated up to

1100°C in different partial-pressures of oxygen. Electrochemical impedance

spectroscopy (EIS) measurements of half cells and full cells were conducted with a

Princeton Applied Research model 273 potentiostat/galvanostat and a Solartron model

1255 analyzer with a frequency range, AC current amplitude, and DC bias of 1 –

65535 Hz, 4 mA, and -4.1 mA, respectively. Section 4.5 contains further discussion on

17

EIS measurements of layered oxygen ion conductors and fabrication of test cells. EIS

measurements were supported by ZPlot and ZView software packages by Scribner

Associates.

18

CHAPTER 3: STUDY OF HEXAGONAL Dy1-xYxMnO3+δ

(-0.2 ≤ δ ≤ 0.4) MATERIALS FOR OSC APPLICATIONS

3.1 Introduction

Structural and physical properties of rare earth manganites have been

thoroughly studied over the past fifty years [66]. Figure 3.1 shows schematic drawings

of reported perovskite and hexagonal crystal structures and electronic occupation of

the 3d orbitals in their respective MnOn polyhedrons. The perovskite orthorhombic

Pnma structure is based on a three-dimensional network of corner-shared MnO6

octahedra. Distortion from the cubic structure, which can be explained by the low

value of the tolerance factor (

), is due to the difference in the (R-O)

and (Mn-O) bond lengths. Furthermore, Jahn-Teller distortion of the MnO6 octahedra,

which is caused by the two-fold degeneracy of the Mn3+ ion in a high-spin state of t3e1,

results in an elongated c-axis and three different (Mn-O) bond lengths [67]. These

distortions shift and rotate the octahedra along the ab plane and tilt and rotate the

octahedra about the c-axis, which results in considerably smaller than 180° Mn-O-Mn

bond angles that have large impact on the transport and magnetic properties of the

system [68][69]. The non-centrosymmetric P63cm hexagonal structure can be

described as close-packed layers of trigonal bipyramids of MnO5, which are centered

at Mn3+ sites and are separated by layers of R3+ ions. The MnO5 bipyramids are rotated

19

20

in the ab plane (planar bond angles Mn-O-Mn ≠ 180°) and tilted relative to the c-axis

due to the difference of the (R-O) and (Mn-O) bond lengths [70]. However, unlike in

the MnO6 octahedra, high-spin Mn3+ ions in the MnO5 bipyramids are not Jahn-Teller

active.

Recently, hexagonal manganites have been the subject of much investigation

due to their multiferroic properties. The rare coexistence of antiferromagnetic ordering

and ferroelectricity make these materials of particular interest. Long-range magnetic

ordering occurs in these materials for both the Mn3+ and R3+ ions at ~70 – 130 K (TMn)

and ~ 5 – 10 K (TR), respectively. Spin-spin interactions of the Mn3+ ions in the close-

packed basal planes are geometrically frustrated and form an antiferromagnetic

triangular structure in the (001) corner-sharing plane at TMn, where each spin is rotated

120° from its nearest neighbors in a P63’c’m symmetry. At lower temperatures, the

R3+ ions magnetically order along the c axis, which is also accompanied by a spin

rotation of the Mn3+ ions to a magnetic symmetry of P63cm. The type of long-range

ordering at TR is dependent on the R3+ ion and may be antiferromagnetic (Ho, Yb, and

Tm) or ferromagnetic (Er and Dy). Additionally, another spin rotation of the Mn3+ ion

occurs between these two temperatures (~40 – 60 K), which results in the magnetic

symmetry P63’cm’ (TSR). At elevated temperatures, RMnO3 remains ferroelectric with

a high Curie temperature (Tc ~ 300 – 650°C) [70]–[77]. It should also be noted that

additional, reversible transitions have been previously observed in situ among various

hexagonal RMnO3 phases at elevated temperatures in air. These studies reported, for

YMnO3, a displacement of the MnO5 bipyramids, which is associated with the Tc

21

ferroelectric transition and a transition to P63/mmc at ~650°C and ~950°C,

respectively [74][75]. Low-temperature magnetic studies of hexagonal DyMnO3 have

been reported for single-crystal and polycrystalline samples with TMn ~70 – 80 K,

TSR~57 K and TDy ~ 3 – 8 K [78]–[81]; however, elevated temperature studies of

DyMnO3 are currently limited to synthesis techniques.

Conventionally, the formation of the perovskite phase versus the hexagonal

phase is governed primarily by the size of the rare earth ion in RMnO3 (with constant

Mn3+size). During high-temperature solid-state synthesis in air, the perovskite phase

forms easily with larger rare earth elements (e.g., La, Pr, Nd, Sm, Gd, Tb, and Dy),

while smaller size rare earths (e.g., Ho, Er, Tm, Yb, Lu, and Y) favor the hexagonal

phase. It has been observed that the perovskite structure is stable for a tolerance factor

(calculated at room temperature using Shannon’s values [82]) in the range of 0.855 ≤

≤1 [83], whereas the hexagonal phase is stable for t < 0.855 [84]. Recently, Zhou et al

[71] suggested that the relative large difference in density between the perovskite and

hexagonal phases may have a larger impact on the formation of the perovskite phase

versus the hexagonal phase near the lower limit of the tolerance factor. DyMnO3 and

YMnO3 have tolerance factors of 0.857 and 0.854, respectively, and will tend to form

the perovskite and hexagonal phases, respectively, under normal solid-state synthesis

in air. Thus, the average (R-O) bond length of substituted samples causes Dy1-

xYxMnO3 to be on the cusp of this phase transition and, as will be further discussed,

results in a mixed state under synthesis in air.

22

Finally, hydrothermal synthesis in 3 kbar at 500°C has been shown to favor the

oxidation state of Mn4+, which results in the formation of Fd3m Dy2Mn2O7 pyrochlore

[85] (Figure 3.1, with Dy in 16d, Mn in 16c, O1 in 48f and O2 in 8b). Mn4+ octahedral

coordination is not a Jahn-Teller ion; therefore, the MnO6 octahedra in the pyrochlore

phase are not subject to the same distortions as in the perovskite phase. Though

transition to this state from the P63cm phase was not observed in our work here, it is a

reasonable assumption that such a transition would occur under high-pressure

conditions similar to this previous study. A transition of this nature, from Mn3+ to

Mn4+, would be ideal for achieving high OSC values.

Chapter 3 describes the synthesis of the P63cm hexagonal Dy1-xYxMnO3+δ in

Ar from its competing Pnma perovskite phase, which was guided by our previous

work on the temperature and oxygen vacancy dependence of the tolerance factor of

manganites. Hexagonal manganites have been largely believed to remain

stoichiometric in oxygen content at elevated temperatures; however, our

thermogravimetric measurements of oxygen-annealed hexagonal samples indicated

unusually large oxygen absorption over a narrow temperature range ~200 – 300°C,

which return to stoichiometric behavior above 275 – 375°C in O2 atmosphere. The

structures of these phases were studied with NPD and XRD. In addition to temperature

dependence, we have also found the oxygen content of Dy1-xYxMnO3+δ to be sensitive

to changes in partial-pressures of oxygen in these temperature ranges. Furthermore,

the hexagonal phase of this system was found to have considerable stability at high

temperature in partial-pressures of oxygen and to be recoverable from negative values

23

of δ from hydrogen reduction at 400°C. The chemical expansion properties resulting

from these large changes in oxygen are also reported, as well as the thermal expansion

coefficient of stable oxygen content regions. The transport and magnetic properties of

these phases were also studied. Finally, the observed properties of these materials are

discussed in context for possible applications as OSC materials and oxygen sensors.

3.2 Synthesis and Stability

Polycrystalline samples of Dy1-xYxMnO3+δ were synthesized by solid-state

reaction with appropriate amounts of Dy2O3, Y2O3, and MnO2 (all with >99.99%

purity). For all samples, reactants were thoroughly mixed in an agate mortar, and fired

in air in the temperature range of 800 – 1300°C with intermediate grindings followed

by pressing samples into high-density pellets at approximately 10 kbar. All steps of

the synthesis were monitored with XRD measurements and compared to previous

diffraction measurements in the literature of the hexagonal P63cm and perovskite

Pnma phases of DyMnO3 and YMnO3 (Figure 3.2) [69][86] –[88]. Dy1-xYxMnO3+δ

samples which formed the perovskite or a mixed phase in air instead of the single-

phase hexagonal structure (x = 0, 0.1 0.3, 0.5, 0.7) were then fired under ultra-high-

purity argon (99.999%) at 1300 and 1400°C. Dy1-xYxMn O3+δ samples (x = 0 and 0.1)

were then subsequently fired under ultra-high-purity argon with a hydroxyl purifier

(oxygen partial-pressures of 5 – 10 ppm) at 1400°C. All samples achieved the

hexagonal P63cm structure after these conditions.

24

Considerable effort was devoted to synthesizing Dy-rich, homogenous

hexagonal samples. The hexagonal DyMnO3 phase has been previously achieved by

epitaxially stabilized crystal growth with thin-films [80], thermal decomposition with

polynuclear coordination compound precursors [89], quenching methods from 1600°C

in air [90] or 1250°C in argon for three days with sol-gel methods [86]. Our work

confirmed that synthesis in argon at high temperature tends to favor the formation of

the hexagonal phase and synthesis in oxygen tends to favor the perovskite phase [91].

Figure 3.3 is a mapping of the phases that were measured with XRD after several

synthesis steps, which clearly shows that increasing reducing conditions are needed to

25

form the hexagonal phase as the average ionic radius of the R site increases. The

oxygen content dependence of the tolerance factor, which we have previously studied

for substituted SrMnO3 [92], is most likely responsible for this behavior. The

formation of oxygen vacancies in RMnO3+δ (δ < 0) causes a change in oxidation state

in some of the Mn3+ cations to Mn2+, resulting in a net Mn(3+2δ)+ cation, which

increases the (Mn-O) bond length with decreasing δ. TGA measurements in oxygen

show the reduced oxygen contents after synthesis of single-phase hexagonal samples

in argon. The resulting larger (Mn-O) bond lengths of these samples decrease their

tolerance factor below the lower limit of 0.855 and resulted in the perovskite phase

undergoing a phase transition to the hexagonal phase. Using Shannon’s room-

temperature values [82], the minimum necessary value of δ ranges from -0.023 – -

0.0027 to have t ≤ 0.855. We have observed, however, that samples with the

corresponding δ values did not transform completely to the hexagonal phase

(Table 3.1). Our previous in situ measurements with Ca and La substituted

26

SrMnO3[92][93] have shown that both (Ca,Sr,La-O) and (Mn-O) bond lengths

increase with temperature in a manner which increases the value of the tolerance

factor. Therefore, the transition from the perovskite phase to the hexagonal phase will

most likely occur in various oxygen pressures at δ, which is a function of temperature,

that occurs for DyMnO3+δ, for example, in ~10ppm O2 at 1400°C as we observed here

or in air at 1600°C as previously reported [90]. Further high-temperature in situ

structural measurements would be needed to completely substantiate this assertion;

however, the combination of our previous in situ measurements with similar

manganites and our XRD and NPD measurements of various oxygen contents after

progressive increased reducing conditions strongly support this conclusion. This

transition may also be enhanced by the difficultly of maintaining the twelvefold

coordination of R required for the perovskite phase in a high-temperature, oxygen

deficient atmosphere; thus, an eightfold coordination with hexagonal symmetry

results.

In any case, the reducing conditions needed for production of bulk

polycrystalline samples of hexagonal DyMnO3 and Dy0.1Y0.9MnO3 by standard firing

methods were very near to decomposition to simple oxides and many attempts were

27

needed to find the most favorable temperature and length of the firings. Increased

substitution of Y in DyMnO3 considerably eased the necessary reducing conditions to

synthesize the hexagonal phase.

The stability of hexagonal Dy1-xYxMnO3+δ compounds was also tested by

firing samples at high temperatures, 1100 – 1400°C, in oxygen. As reducing

conditions favor the hexagonal phase, atmospheres that allow samples to remain near

stoichiometric in oxygen content (or yield excess oxygen content δ > 0) at high

temperature will promote the perovskite over the hexagonal phase, due to the smaller

size of the Mn(3+2δ)+ cation in oxygen versus argon. Dy-rich samples (x = 0, 0.1) began

slight decomposition back to the perovskite phase at 1100°C and completely

transformed back to the perovskite phase at 1400°C. The remaining samples (x = 0.7,

0.5, 0.3, 0.1, 0) remained in the hexagonal structure with no signs of decomposition

back to the perovskite phase up to 1400°C, which suggests these materials may form

the hexagonal phase at high temperatures in air after a long duration in these

conditions (>3 days). These results are in agreement with the presented tolerance

factor arguments and may also explain why small rare earth manganites (R = Y, Ho,

Er, Tm, Yb, and Lu) have been observed to transition to the perovskite phase under

high-pressure oxygen [94][95], while smaller A-site cations (R = Sc and In) will not

transform to the perovskite phase under similar conditions [96].

Finally, many attempts were made to synthesize the hexagonal P63cm phase

with substitutions on the B-site for DyMn1-yMyO3+δ with M = Cr, Fe, Co, Ni and Al

(0.16 ≤ y ≤ 0.5); however, none were able to form single-phase hexagonal from the

28

perovskite phase with reducing conditions. The perovskite structure is stable to lower

values of the t-factor for these transition metals when compared to RMnO3, which is

most likely due to the absence of Jahn-Teller distortions in these non-degenerate

cations. Doping hexagonal YMnO3 with Cr, Fe, Co, and Ni under synthesis in air

causes a transition to the perovskite phase at approximately 20 – 30% substitution

levels [97] – [100]. These systems with higher levels of doping can however be

transitioned back to the hexagonal phase with reducing conditions. Furthermore,

hexagonal YFeO3 (synthesized by sol-gel methods with metal nitrate precursors and

careful pH control with citric acid) has been reported to transition back to the

perovskite phase at ~930°C in air [101], which is very similar in behavior as we

observed with DyMnO3+δ. We had particularly hoped that similar reducing conditions

for M = Co, which reduces more readily when compared to the Mn cation (M3+ to

M2+), would sufficiently increase the (B-O) bond lengths to form the hexagonal

DyMO3, but only perovskite and impurity phases were obtained. Low substitution was

also recently reported in single phase for hexagonal P63cm in YMn0.9M0.1O3 (M = Al,

Ru, and Zn) [102] and, subsequently, YMn0.9Re0.1O3+δ and YMn0.9Ru0.1O3+δ were

synthesized by similar methods as YMnO3+δ; however, we found these samples to

have unfavorable OSC and oxygenation/reduction behavior when compared to pure

YMnO3+δ. The 2+, 3+, and 4+ oxidation states easily available to the Mn cation may

make it ideal for the desired behaviors for synthesis and oxygen storage applications

and any substitution to the B-site may not be fruitful for enhancing OSC values.

29

3.3 Oxygen Storage Measurements

After initial synthesis of the hexagonal phase, all samples were annealed in

TGA up to 500°C with isothermal and 0.1 – 1°C/min heating and cooling in various

partial-pressures of oxygen and hydrogen to measure OSC values and to demonstrate

temperature and oxygen partial-pressure dependence of oxygen content. The oxygen

content after initial synthesis of DyMnO3+δ and YMnO3+δ were then determined with

TGA by the difference in weight between oxygenated samples and their respective

reduction products, Dy2O3, Y2O3, and MnO (verified by XRD), obtained by first

annealing at 1°C/min in O2 and followed by slow reduction at 0.1°C/min in 42%

H2/Ar (example of DyMnO3+δ in Figure 3.4). Thus, Figure 3.4 is normalized to

reduction products (δ = -0.5). DyMnO3 and YMnO3 were observed to reduce to stable

stoichiometric P63cm phase in oxygen above 375 and 275°C, respectively. Using this

information, stable weights of all samples above 400°C in O2 in TGA were normalized

30

to δ = 0. Table 3.2 is a compilation of OSC values achieved by the various methods

listed in this section.

Temperature dependence of oxygen content of Dy1-xYxMnO3+δ materials were

measured in TGA with heating and cooling rates of 0.1 and 1.0°C/min under high-

purity oxygen. The resulting TG curves (0.1°C/min, Figures 3.5 and 3.6) clearly show

the reversible absorption and desorption of oxygen below 400°C in a narrow

temperature range. OSC values were measured by the difference in oxygen content

between the stoichiometric P63cm phase observed above 400°C (δ = 0) and the final

oxygen content after cooling (δ = 0.01 – 0.29), which yielded a large range of values,

54 – 1200 μmol-O/g (Table 3.2). Comparing 0.1 versus 1.0°C/min, resultant TGA

curves, and OSC values indicate that oxygen absorption rates increased with Dy

content. Yet, samples (x = 0.1, 0.3, 0.5) were able to achieve higher oxygen content

than the pure Dy sample on 0.1°C/min cooling. Y-rich samples (x = 0.7, 0.9, and 1)

were also able to yield larger OSC values than observed in TGA with long isothermal

steps with slow cooling and indicate, if given enough time (>24 hours), would reach

excesses in oxygen content up to δ ≈ 0.25. Four different temperatures were also

31

32

identified from TGA runs in O2, which are plotted in Figure 3.7: the average

temperature of maximum oxygen absorption on heating and cooling (

), maximum oxygen desorption (

),

transition temperature from oxygen absorption to desorption, and the temperature

where samples return to stoichiometric behavior (

) (these can be

approximately identified on Figure 3.5 by inspection). A thermal swing absorption

process for air separation for each of these samples would most likely involve cycling

in between their respective temperatures slightly above “Ox = 3.0” and slightly below

“Ave. Max. Absorption,” which would yield cycling ranges of approximately 220 -

300°C (x = 1) to 310 - 390°C (x = 0).

Samples were also annealed at 250 bars of O2 at 400 – 500°C followed by

0.1°C cooling. These annealings were cooled from lower temperature than TGA due

to annealing at high temperature and pressure tends to favor a phase transition back to

33

the perovskite phase [94][95] or to the R2Mn2O7 pyrochlore phase [85]. The oxygen

content of these samples after annealing was determined in TGA by the difference in

weight between their starting weight and their weight at 375°C (1°C/min heating) in

21% O2 normalized to δ = 0 (Figure 3.8). All samples showed significant increase in

OSC (particularly with samples rich in Y content) under high pressure versus identical

cooling in 1 bar of O2 (Table 3.2). Figure 3.8 also shows increased stability of oxygen

content on reduction at ~300°C for all samples, which suggests the existence of a

stable phase at Dy1-xYxMn3+0.5Mn4+

0.5O3.25 and, possibly, the presence of another

stable phase at or above an oxygen content of 3.35. Though these samples show

increased oxygen content from atmospheric pressure oxygenations, the Mn3+ cation is

still not completely oxidized to the Mn4+ state, which would be ideal for maximum

OSC values. Table 3.2 also includes these theoretical values of OSC for a reversible

Mn3+ – Mn4+ (δ = 0 – δ = 0.5) transition. Higher temperature attempts were also made

to oxygenate samples (x = 0.5, 0.7, and 0.9) at 88 bars of O2 at 890°C (with a 12-hour

34

hold followed 10°C/min cooling), but these annealings yielded similar oxygen

contents and Hex2/Hex3 mixed phases as high pressure runs at 400 – 500°C at

comparable pressures.

Oxygen partial-pressure dependence of oxygen content of Dy1-xYxMnO3+δ and

absorption/desorption reversibility were demonstrated with TGA measurements at

isotherm in cycling O2 and Ar atmospheres every ~12 hours (Figure 3.9). Samples

were held at temperatures near their respective “transition temperatures” defined from

Figure 3.7 (for x = 0, 0.3, 0.5, 0.7, and 1; T = 330, 300, 280, 250, and 230°C,

respectively) and yielded OSC values of 95 – 1149 μmol-O/g (Table 3.2). Besides

DyMnO3+δ, which clearly comes to equilibrium in O2, these OSC values are

comparisons of absorption of 12 hours. Given more time, these samples can achieve

higher oxygen content; for example, δ ≈ 0.28 was obtained for Dy.3Y.7MnO3+δ after

~60 hours. Isothermal measurements also show oxygen content to have asymptotic

behavior significantly lower than achieved upon cooling (most noticeably for x = 1

and 0). Further isothermal TGA measurements at various temperatures have also

shown this kinetically oxygen-content limiting behavior, which increases equilibration

time at lower temperatures (this limiting behavior accounts for the significant

differences in absorption rates of Figures 3.5 and 3.9). Therefore, the OSC of samples

(x = 0 and 1) would probably improve at lower isothermal temperatures and the

desorption rate of x = 0.7 would most likely improve at slightly higher temperatures.

35

The nature of these transitions from the P63cm phase (δ = 0) to the Hex2 phase

(δ = 0.25) and from the Hex2 phase to the Hex3 phase (δ ≈ 0.40, see next section)

appears to easily equilibrate to intermediate oxygen content values. As a result, a

mixture of several phases will occur in various oxygen partial-pressures and

temperatures, where low temperatures, 150 – 200°C, favor the Hex3 phase;

intermediate temperatures, 230 – 330°C, favor the Hex2 phase; and high temperatures,

above ~275 – 375°C, favor the stoichiometric P63cm phase (these ranges are

dependent on oxygen partial-pressure and Dy/Y content). The slope of oxygen content

versus time during the P63cm – Hex2 phase transition at constant temperature (as well

as on cooling in Figure 3.5) decreases with increased Y content, which again indicates

slower absorption rates of Y-rich samples. Direct comparisons of these absorption

rates are, however, complicated by slower oxygen ion kinetics at lower temperatures,

which can be approximated by . The lower temperatures at which the

Hex2 – P63cm phase transition occurs for Y-rich samples prevents temperature-

independent absorption comparisons; thus, the differences in absorption observed in

36

Figure 3.9 are due to both differences in activation energy and temperature. This

increased rate of transition from the P63cm to the Hex2 phase may also be due to

increased distortion to the P63cm structure caused by larger average R-site anions. On

the other hand, the transition from the Hex2 to Hex3 phase (δ ≥ ~0.25) appears to favor

Y-doped DyMnO3.25 samples (x = 0.1, 0.3, 0.5) over pure DyMnO3.25, as seen on

cooling in Figure 3.5.

Hydrogen reductions in TGA for DyMnO3+δ and YMnO3+δ, which were

initially done to determine oxygen content, showed to have increased stability on

reduction at δ = -0.12 and -0.20, respectively (Figure 3.4). To test for recoverability of

the P63cm phase of DyMnO3+δ and YMnO3+δ, materials were heated to and held at

400°C in 42%H2/Ar in TGA until these respective values of δ were reached. These

samples were then cooled in Ar to 330 and 230°C, respectively, and held at these

temperatures under O2. Samples quickly returned to stoichiometric oxygen content (>1

hour) and continued to absorb oxygen, as seen during oxygen cycles in Figure 3.9.

XRD measurements after this process confirmed that samples did not decompose to

simple oxides. Thus, the addition of cycling to 400°C in hydrogen to either thermal or

oxygen partial-pressure cycling would yield an additional ~450 – 1050 μmol-O/g (for

x = 0 – 1) and would place these materials up to near-record levels of OSC, ranging

from 1150 – 2650 μmol-O/g (Table 3.2, where calculated values assume the stabilities

seen at δ = -0.12 to δ = -0.20 changes proportionally with x for intermediate samples).

While the values measured here do not surpass the best observed OSC in the

literature and the slow oxygen kinetics of Y-rich samples (x = 0.7, 0.9, 1) may be a

37

limiting factor for their potential use for OSC application, the Dy1-xYxMnO3+δ system

does have several key advantages for application over these other candidates. First and

foremost, the Dy1-xYxMnO3+δ system has the lowest reported reduction temperature,

being approximately 25 – 125°C lower than the record reduction temperature of

YBaCo4-xAlxO7+δ (with significant OSC values). On further comparison to YBaCo4-

xAlxO7+δ, which decomposes at 550 – 700°C, Dy1-xYxMnO3+δ has far superior stability,

remaining stable up to 1100 – 1400°C. Additionally, from a hazardous waste and cost

standpoint, mass production of manganese oxides is much preferable to that of cobalt

or chromium oxides. Finally, there is great potential for the Mn cation in hexagonal

RMnO3+δ to have large changes in oxidation state because, unlike the majority of OSC

materials, which depend on the creation of oxygen ion vacancies or interstitial sites at

high temperatures, the hexagonal Dy1-xYxMnO3+δ (as seen also with YBaCo4-xAlxO7+δ)

relies on reversible phase transitions between several structures containing transition

metal ions in variable coordination. The potential OSC of related hexagonal

manganites could easily surpass the current highest reported values, if they can be

modified to easily and reversibly transition in between phases with large amounts of

Mn2+ and Mn4+ at low temperatures.

Finally, apart from any possible OSC application, it should be noted that

hexagonal manganites have been largely believed, to the best of our knowledge, to

remain stoichiometric in oxygen content at elevated temperatures. In situ structural

measurements at high temperatures have reported a displacement of the MnO5

bipyramids and a transition to P63/mmc, which occur for YMnO3 at ~650°C and

38

~950°C, respectively [74][75]. Slight excesses of oxygen content (δ ≈ 0.01) have been

reported at 1200°C for YMnO3+δ and ErMnO3+δ [103] but did not show the non-

stoichiometric oxygen content behavior or the associated structural changes at lower

temperatures as we have observed with thermogravimetric and XRD measurements.

This behavior may not have been previously observed in other hexagonal manganites

due to the narrow range of temperature (~200 – 350°C) these new phases exist during

heating before returning back to δ = 0 above ~350°C and the slow cooling or high

oxygen partial-pressures they require. As discussed in the introduction, this

temperature range has not been of particular interest for structural studies of RMnO3,

as most of this work has been done at either low temperature to study magnetic

ordering (≤ 200 K) or high temperature to measure the rattling behavior of the MnO5

bipyramids or structural transitions (≥ 500°C). Our results indicate that the hexagonal

RMnO3+δ family is most likely prone to considerable oxygen non-stoichiometry and

also suggest a direct relation between reduction temperature and sorption rates of

oxygen to the average ionic size of R. If this is the case, other hexagonal RMnO3+δ

materials with rare earths that are close in ionic size to that of Y (e.g., Ho and Er) will

have similar non-stoichiometric behavior. It should be noted that our synthesis of

YMnO3+δ under fast cooling to room temperature yielded small, but measurable,

excesses in oxygen content (δ = 0.004). Many studies of RMnO3+δ use samples

prepared at elevated temperature followed by various cooling rates, which would yield

slightly non-stoichiometric samples for low-temperature measurements. Properties

associated with excess oxygen content (e.g., disruptions to the exchange interaction or

39

the presence of Mn4+) may very well have had a significant impact on the multiferroic

properties of these samples, as we have observed that even slight oxygen and cation

non-stoichiometry can have profound effects on magnetic and transport properties of

perovskite manganites [104][105]. In the following sections, we will show this effect

has a considerable impact of the structural, thermal/chemical expansion, transport, and

magnetic properties of Dy1-xYxMnO3+δ.

3.4 Crystal Structure

To study the structure of hexagonal oxygen-enriched phases (0 ≤ δ ≤ 0.4, x = 0

– 1), all samples were annealed after initial synthesis in varying conditions (in addition

to the TGA and high-pressure runs of the previous section) to achieve a large range of

oxygen contents. The oxygen content behavior of the DyMnO3+δ hexagonal sample

during annealing in oxygen (TGA, curve 1 of Figure 3.4) shows it to have stable

stoichiometric behavior above 350°C. Using this information, a stoichiometric P63cm

sample of DyMnO3 was synthesized by quenching in air from 420°C to liquid nitrogen

(verified by change in weight). Samples with δ > 0 were obtained on TGA by heating

to 400 – 500°C and then slow cooling to room temperature at 0.1 – 1.0°C/min in 21 –

100% O2 at ambient pressure. The final oxygen content of these samples was

determined by normalizing to stable weights above 400°C.

XRD measurements were made to verify the P63cm hexagonal structure after

synthesis and to obtain a preliminary structural understanding of annealed samples

before NPD measurements were conducted. Figure 3.10 is a compilation of XRD

40

patterns collected for DyMnO3+δ (δ = -0.037 – 0.35), which are representative of the

Dy1-xYxMnO3+δ series. Table 3.3 lists the synthesis conditions that produced these

samples. Peak positions and intensities of DyMnO2.963 and DyMnO3.0 were found to

be in good agreement with previously reported XRD patterns of P63cm DyMnO3 [81].

Furthermore, XRD data of the quenched sample (sample 2) confirmed that

stoichiometric samples are indeed P63cm after quenching from above 400°C as

observed with TGA data. XRD patterns of annealed samples (samples 3, 4, and 5) in

the δ range of 0.18 – 0.24 clearly show growth of a second phase (Hex2) and the

disappearance of the P63cm phase (arrows indicate growth and decrease of selected

peaks for the P63cm phase and the Hex2 phase, respectively). The pattern of sample 5

(δ = 0.24) is nearly single phase for this new set of peaks and is in agreement with the

stability seen in TGA at δ ~ 0.25 (Figure 3.8). Finally, the XRD pattern of the high-

pressure annealed sample 6 (δ = 0.35) shows a decrease of peak intensity of the Hex2

phase and the presence of an additional third phase (Hex3), which is again in

agreement with TGA observations. The relative intensities of these two phases suggest

the Hex3 phase could have an oxygen content of δ ≈ 0.40, though this is difficult to

approximate due to the high degree of peak positions overlap of the Hex2 and Hex3

41

42

phases. However, samples x = 0.3 and 0.5 achieved higher oxygen contents than x = 0

after high-pressure annealings (δ = 0.40 and 0.37, respectively) and their XRD

patterns showed increased ratio of the Hex3/Hex2 phases (Figure 3.11). To help clarify

the development of new peaks and peak overlap, Figure 3.12 shows an overlay of

XRD patterns of samples 2, 5, and 6 (δ = 0.0, 0.24, and 0.35) in the 2θ range of 26 –

35° (phases associated with δ = ~0.25 and ~0.40 in Figure 3.12 are referred to as Hex2

and Hex3, respectively). Figures 3.10 – 3.12 show similarities in the diffraction

patterns of the Hex2, Hex3, and P63cm phases, which suggest that the Hex2 and Hex3

phases are structurally similar to the P63cm phase. Furthermore, the increased number

of peaks in the Hex2 and Hex3 phases suggests a lowering of symmetry or the

formation of a superstructure. Finally, it should also be noted, though these

transformations involving rearrangement of cation-oxygen networks are unlikely at

these low temperatures under O2, that the Hex2 and Hex3 phases were compared to

patterns of other known RxMny4+Mny-1

3+O3+δ systems (e.g., pyrochlore R2Mn2O7,

perovskite R-3c, R2MnO4 and RMn2O5) and oxides (Mn2O3, MnO2), which could

account for the increase in oxygen content. No traces of these structures were

observed.

Guided by our initial XRD investigation, NPD measurements were conducted

for P63cm samples. High-resolution, backscattering data (2θ = 144°, Bank 1 of SEPD)

were collected for DyMnO2.963 and DyMnO3.0. High-resolution, backscattering data

(2θ = 164°, Bank 1 of Echidna) were also collected at room temperature for x = 0.5

and 1 after synthesis of the P63cm phase. Raw data for these samples were analyzed

43

44

with the Rietveld method in the space group P63cm based on previous reports for the

hexagonal RMnO3 system and our XRD measurements (Figure 3.13). The calculated

diffraction patterns of P63cm are in good match with the observed data for all samples

(see Appendix for structural parameters and agreement factors of Dy1-xYxMnO3+δ) and

their lattice parameters are in agreement with previous reports from XRD and NPD for

DyMnO3 and YMnO3, respectively [81][88]. Bond lengths in Appendix Tables A.1 and

A.2 were calculated using the geometric average and the values of <Mn-O>g for all

samples were calculated by assuming full site occupancy. For DyMnO2.963 and

DyMnO3.0, the average (Mn-O) bond length clearly increases from the stoichiometric

to the reduced state, while the average (Dy-O) bond length remains relatively

unchanged. Again, this is due to the enlargement of the Mn(3+2δ)+ cation with

increasing oxygen deficiency. Bond lengths of <Dy/Y-O>g were also observed to

decrease with increased Y content. These bond length results are in agreement with

the oxygen vacancy dependence of the tolerance factor and support our synthesis

arguments for forming the hexagonal phase by reduction of the perovskite phase in

RMnO3+δ. Furthermore, NDP data shows, in general, increased distortions of refinable

positions (increased Δ) with increased Dy content (Table 3.4, Mn-z and O1-z positions

were excluded from Table 3.4 due to their small variation). These increased in

distortions may be due to increased size of the Dy cations, which prompts transition

back to the perovskite at high temperatures in oxidizing atmospheres.

45

46

3.5 Thermal and Chemical Expansion

Expansion of the crystal lattice, in both the hexagonal and perovskite RMnO3

phases, can occur through two mechanisms: thermal and chemical expansion. Thermal

expansion (TE), as discussed in tolerance factor arguments, is caused by expansion of

the (R-O) and (Mn-O) bond lengths due to increased thermal energy at elevated

temperature. This effect has been extensively studied with both in situ structural

measurements and dilatometry for both the RMnO3 perovskite and hexagonal systems

and can have a large impact on their magnetic, transport, and structural properties

[69][70][74][75] [92][93][106]–[108]. Chemical expansion (CE) is expansion of the

lattice due to changes in oxygen stoichiometry. The effect of CE has also been heavily

studied for changes in structural properties and has great importance on the macro-

scale for applications such as films, coatings, and layer materials [109]–[113] . On the

other hand, CE measurements for the hexagonal manganites are currently nonexistent,

due to the belief that the system remains stoichiometric in oxygen content at elevated

temperatures in argon and oxygen atmospheres. It should also be noted that in some

cases the thermal expansion coefficient (TEC) is considered to be the net result of both

CE and TE; here we consider these to be separate effects, thus TEC in this report is

only attributed to TE.

Measurements of CE typically must be measured separately from TE with

RMnO3+δ perovskites, because both CE and TE change at similar rates as a function of

temperature. Absorption occurs in the RMnO3+δ perovskite phase during a Pnma – R-

47

3c transition, which creates equal number of A and B site vacancies while absorbing

oxygen to remain stoichiometric [114]. These changes in oxygen content are usually

relatively small (δ ≤ 0.15), and occur slowly over a wide range of temperatures (~500

– 1000°C) [69]. Thus, investigations of CE must be done at constant temperature over

long lengths of time (≥ 72 hours) with changes in oxygen partial-pressure to change

oxygen content and make it possible to separate the effects of TE from CE. However,

our TGA measurements for hexagonal DyMnO3+δ phase have shown large changes in

oxygen stoichiometry between two stable oxygen content regions, which occur over a

relatively short time scale (≤ 2 hours) and narrow range of temperatures (~100°C).

These characteristics allow us to measure the effective CE over a narrow range of

temperature by simply subtracting the relatively small value of TE from the observed

value of CE. Similarly precise measurements of TE, without any effect from CE, were

possible in temperature regions of stable oxygen content. The following equations

were used to calculate TE and CE:

, measured in K-1,

where L0, ΔL, and T are the sample starting length, the change in length, and

temperature, respectively, and m-n are the sets from the measured temperature ranges,

and

, measured in (moles of O)-1, where Δδ is the

absolute change in oxygen content from stoichiometric 3.0 and <TEC> is the average

TEC of the two oxygen content stable regions.

A perovskite sample of DyMnO3 for dilatometry was cut from a dense pellet

after initial synthesis in air of the perovskite phase (~ 5x3x2 mm in shape) and was

measured in 21% O2/Ar atmosphere with heating rates of 0.5°C/min to 900°C (Figure

48

3.14). Our previous studies of the perovskite DyMnO3+δ phase have shown that it

remains stoichiometric in 21% O2/Ar up to ~1000°C [69], thus the expansion seen in

Figure 3.14 is solely due to TE. TEC was measured from 50 – 850°C and was found to

be 7.3*10-6 K-1, which is in good agreement with a previous report [106]. Reported

dilatometry measurements of the perovskite YMnO3 phase, from 500 – 1000°C (again

stoichiometric), found the TEC to be approximately 6*10-6 K-1 [115].

Pellets were also cut from dense samples (x = 0, 0.3, 0.5, 0.7, 1) after synthesis

of the hexagonal material (~ 5x3x2 mm in shape) and were then annealed at 400°C

with 0.1°C/min cooling in O2 for dilatometry measurements. The oxygen contents of

these samples were also measured with identical starting samples and conditions in

TGA to determine the appropriate temperature ranges to separately extract TE and CE.

TEC values were measured for these samples in their respective temperature regions

of stable oxygen content observed in TGA for δ = 0.22 – 0.29 (~ 50 – 300°C) and for

δ = 0 (~600 – 850°C). CE values of these samples were measured during the reduction

between these stable oxygen contents, which again occurs over approximate

temperature gradient of ~100°C in the range of 240 – 390°C, where approximately

49

90% of the total oxygen reduction occurs. Figures 3.15 and 3.16 show these

measurements for DyMnO3 and illustrates how dilatometry and TGA were used in

combination to determine TE and CE for all samples, which is representative of the

measurements conducted for the Dy1-xYxMnO3+δ series. The lower starting oxygen

contents after annealing in oxygen and the slower reduction of dense pellets (as seen

for DyMnO3+δ in Figure 3.15) versus the small chucks of material observed in Figure

3.5 during TGA measurement are due to the differences in the samples’ density,

surface area, and diffusion distances. The TEC of the hexagonal phases in these two

temperature regions of stable oxygen content were found to be quite different, 8.2 –

10.2*10-6 K-1 (δ = 0.22 – 0.29) and 2.1 – 5.6*10-6 K-1 (δ = 0), which indicates the TEC

of the stoichiometric Hex2 (δ = 0.25, assumed) and P63cm phase are approximately

8.4 – 11.6*10-6 K-1 and 2.1 – 5.6*10-6 K-1, respectively (Figure 3.17). The values of

chemical expansion during loss of oxygen content are 0.82 – 3.48*10-2 mol-1 (Figure

3.18), which increase significantly with Dy content.

50

Previous reports of single-crystal hexagonal RMnO3 materials (R = Y, Ho, Sc,

and Lu) have shown to have lattice parameters that linearly increase in-plane and

decrease along c with increasing temperature [70][108]. The contraction of the c-axis

has also been shown to increase for larger R ions. Thus, the effect of substantial

51

contraction of the c-axis is responsible for the small change of volume of the unit cell

and significantly lowers TEC of our polycrystalline P63cm material when compared to

their hexagonal Hex2 and perovskite phases. It is also in agreement with the decrease

of net TEC with increase Dy content for P63cm materials as seen in Figure 3.17. This

tendency is, however, reversed for the Hex2 phase, which shows to have increased

TEC with increased Dy content. Finally, an increased rate of contraction along the c-

axis at the Curie temperature, ~650°C, was reported previously for YMnO3 and

HoMnO3 in one study [108] but was also not present in another report [70]. We did

not observe any anomalous behavior near this temperature; however, this effect may

be beyond the sensitivity range of our dilatometer for a polycrystalline sample, where

anisotropic effects are averaged out. On the other hand, if dense hexagonal RMnO3

materials are also prone to small non-stoichiometric behavior on heating, as seen here

for the temperature range of 400 – 600°C (0 < δ < 0.015), this effect could be due to

the CE associated with the reduction of a slightly oxygenated sample to stoichiometric

oxygen content. Our measurements show the importance of understanding oxygen

content behavior, as slight changes in oxygen content can have similar effects on net

expansion as structural changes not associated with changes in oxygen content (e.g.,

the P63cm to P63/mmc phase transition).

The CE during transition from the mixed-state Hex2/P63cm (~85 – 100%, δ ≈

0.22 – 0.25) materials to nearly single-phase P63cm has a larger effect on total

expansion than TE. Clearly, CE is not significantly affected by thermal expansion in

the indicated temperature ranges. Again, the primary cause of the CE seen during the

52

P63cm/Hex2 is due to the change in ionic radius of the Mn(3+2δ)+ cation as discussed

with the tolerance factor arguments. Finally, for comparison, the CE values reported

here are of the same order of magnitude as the CE associated with the absorption and

desorption of oxygen from stoichiometric perovskite LaMnO3 and various substituted

perovskite materials (~2.4*10-2 mol-1 and ~1 – 4*10-2 mol-1) [109] –[111]. However,

the effect of CE in the hexagonal structure is much more prominent than in the

perovskite phase due to the larger change in oxygen content occurring over a much

narrower temperature range.

3.6 Transport Properties

Rectangular bars for conductivity measurements (~ 8x5x0.7 mm in shape)

were pressed at 10 kbar with embedded Pt leads in powder ground from P63cm

hexagonal material (x = 0, 0.5, 1) after initial synthesis. Bars were then fired at 900°C

and annealed at 400°C with 0.1°C/min cooling in O2 before measurement.

Conductivity measurements were done from 130 – 900 °C in 21% O2/Ar (Figure

3.19a). Conductivity data was modeled both with and

(Figures 3.17 b and c, for x = 0, 1) and the activation energies of P63cm and Hex2

phases were determined (Tables 3.5 and 3.6).

Conductivity of the Hex2 phase (<200 – 325°C) was modeled with

based on phonon-assisted conductivity, which has been previously

applied to pevorskite manganites and the hexagonal YMnO3 phase hole-doped with Ca

[67][69][116]. In these systems, the observed strong electron-phonon coupling is due

53

54

to Jahn-Teller active 3d4 and 3d3 Mn ions (high-spin) in octahedral and bipyramidal

crystal field splitting, respectively. Figure 3.19b shows a plot of the natural logarithm

of the product of conductivity and temperature versus inverse temperature and also

shows excellent linear fits (linear regression, R2 = 0.9996 – 1.0000) of phonon-

assisted conductivity below 200 – 325°C (above ~2.11 – 1.67 1000/K-1). The slope of

these fits determines their respective activation energies. Activation energies were

found to slightly increase with increased Y from 45.73 to 48.72 kJ/mol. These values

are of the same order of magnitude as calculated from our previous work for

55

perovskite DyMnO3 of 25.86 kJ/mol. This behavior also suggests that the Hex2 phase

has electron-phonon coupling. If such behavior exists, the Hex2 phase may contain

MnOn octahedra or bipyramids, as they are Jahn-Teller active for Mn3+ and Mn4+,

respectively, and, when coupled with the high oxygen content of the Hex2 phase, may

suggest the presence of octahedra (plane sharing). The sharp decrease in conductivity

at 250 – 350°C is due to the Hex2-P63cm phase transition, as seen in TGA (e.g., Figure

3.16). This increase is due to 3d4 electrons filling the x2+y2 and xy orbitals, which

have the strongest overlap with O 2p orbitals for exchange interactions in the P63cm

structure. Two temperature regions where oxygen stoichiometric P63cm phases exist,

~370 – 550°C and ~660 – 900°C, show good linear fits (R2 = 0.9931 – 0.9997) with

activation energies of 4.63 – 11.96 and 109.03 – 126.40 kJ/mol, respectively.

However, the x = 0.5 sample appears to be relatively thermally independent or has a

very small activation energy for this intermediate temperature range. If phonon-

assisted conduction exists above 400°C, this may suggest the presence of another

mechanism for electron-phonon coupling. Furthermore, the significant increase in

activation energy above ~660 – 790°C may correspond to the increase in activation

energy associated with the ferroelectric-paraelectric transition. The upper limit of this

increase in activation energy for YMnO3 (~450 – 670°C) is in good agreement with its

reported Curie temperature of ~650°C [74]. If this correlation between the

temperatures of the ferroelectric-paraelectric transitions and increases in activation

energies holds, this would suggest that Tc’s for the P63cm phases of Dy0.5Y0.5MnO3

and DyMnO3 are near 660°C and 790°C, respectively. As such, these compounds

56

would remain ferroelectric to higher temperatures than any other hexagonal RMnO3

materials reported to date. As discussed in the introduction, ferroelectricity in the

P63cm structure is strongly correlated to the rattling behavior of the MnO5 bipyramids

and the phase transition of the P63cm phase to high-temperature P63/mmc phase. Thus,

the observed increased structural distortions of the P63cm phase with increased Dy

content may be responsible for the higher Curie temperatures of Dy-rich samples,

which are in agreement with conductivity observations.

Conductivity was also modeled with semiconducting behavior, ,

which has been previously applied to the hexagonal YMnO3 phase hole-doped with Ca

[116]. As shown in Figure 3.19c, the Hex2 phase can be fitted (R2 = 0.9997 – 1.0000)

with semiconducting behavior. These fits have a related activation energy of 41.39 –

45.25 kJ/mol (increases with Y content), which are smaller values than obtained from

phonon-assisted conductivity. Above ~310 – 400°C, the P63cm phase shows

semiconducting behavior, where notable thermal activation of conduction begins

between ~400 – 520°C and 680 – 800°C (temperature dependence of conductivity

seen here closely resembles extrinsic-type semiconductor behavior). The conductivity

of the P63cm phase for x = 0 and 0.5 from approximately 400 – 500°C is relatively

temperature independent; only the x = 1 sample has a small, measurable activation

energy of 6.66 kJ/mol. Above 680 – 790°C conductivities again have good

semiconducting fits (R2 = 0.9906 – 0.9980), which yield activation energies of 101.31

– 117.71 kJ/mol.

57

Based on the phonon-assisted and semiconducting fits, it is difficult to

determine the correct mechanism for conduction in the Hex2 and P63cm phases.

Additionally, σT behavior in the Hex2 phase could also be due to conduction by the

small polaron conduction mechanism associated with the mixed Mn3+/Mn4+ valence,

which has been extensively studied for perovskite, rare earth manganites doped with

divalent cations [117] and will be further discussed in the next chapter. Further

structural studies of the Hex2 phase and high-temperature in situ structural studies of

the P63cm Dy1-xYxMnO3+δ phase are required to fully understand their transport

properties; however, it is clear that any oxygen non-stoichiometry will have a

considerable impact on P63cm material’s transport properties.

The large changes in conductivity seen in Figure 3.19, on the Hex2-P63cm

phase transition during heating in 21% O2/Ar (when oxygen is released), are also

visible when oxygen partial-pressure is changed during isotherm at the transition

temperatures identified from thermogravimetric measurements (Figure 3.9). Figures

3.20 and 3.21 show the resistivity of DyMnO3+δ and YMnO3+δ material during cycling

between O2 and Ar atmospheres at 330°C and 230°C, respectively. Resistivity

measurements of this nature could be used for “in situ” monitoring of oxygen content

during an oxygen storage process. Even though cycling completely between the Hex2

and P63cm phases can take considerable time (particularly with Y-rich samples), the

minimal rate change in resistivity in the 5 – 95% Hex2/P63cm mixed phase ranged

from approximately 0.5 to 1.5 Ω*cm/s. Changes in resistivity of this order of

magnitude would be easy to detect if, for example, Dy1-xYxMnO3+δ material was

58

prepared as a thin separation membrane to act as a potential “fast-acting” oxygen

sensor. Futhermore, as was discussed in the previous sections, since the rate of

absorption is dependent on oxygen partial-pressure, the slope of the resistivity

response could easily be used to calculate the relative oxygen partial-pressures on each

side of such a membrane. The strong temperature dependence of these materials would

require any potential sensors to be held at precise isothermal temperature, which

would be impractical for most applications. Furthermore, prospective MIEC

perovskite/electrolyte layered materials have been shown to function as oxygen

detectors at lower temperatures (as low as 200°C) while operating in a much larger

temperature gradient [118].

59

3.7 Magnetic Measurements of DyMnO3+δ (x = 0)

Although magnetic properties are not the focus of this dissertation, a study of

new hexagonal manganites necessitates brief analysis relating to these properties,

which have been extensively studied for this family of materials because of their

multiferroic behavior. Thus, the focus of this section is to merely observe the impact

large non-stoichiometric oxygen content has on magnetic properties of hexagonal

manganites. Rectangular samples for DC susceptibility were cut from dense pellets of

sample after intial synthesis of the hexagonal phase, and subsequently quenched from

420°C in air or slow cooled from 400°C in O2 . All three samples of DyMnO3+δ (δ =

-0.037, 0.00, and 0.22) clearly exibit Curie-Weiss behavior above ~ 100 K,

(where χ0 is background susceptibility, μB is the

Bohr magneton, kb is the Boltzmann constant, μeff is the effective magnetic moment,

and Θ is the Curie-Weiss temperature) and have related parameters of Θ = {-16 K, -17

K, 19 K} and µeff = {11.4 µB, 11.3 µB, 11.0 µB} for oxygen contents of 2.963, 3.0 and

3.22, respectively (Figure 3.22). These values are in close agreement with the theorical

60

magnetic moment of the stoichiometeric P63cm phase, 11.6 µB (µth. eff2 = µDy

2 + µMn2,

high spin state assumed) and previous reports of -23 K and 10.8 µB, and -10 K and 11

µB for DyMnO3 [78][80]. Figures 3.23 and 3.24 show

versus temperature in the

temperature ranges of 20 – 120 K and below 20 K, respectively. In the higher

temperature range, TMn occurs at the discontinuity at ~71 K for DyMnO3.0 and appears

to be slightly lower, ~ 69 K, for DyMnO2.963 (Figure 3.23). The hexagonal RMnO3

family has shown to have a general lowering of the Néel temperature with increasing

ionic radius of R [70][71]. When comparing the TMn of DyMnO3 to the range of TMn

for smaller R radii of RMnO3 (70 – 130 K), it is in agreement with this trend. This is

also in good agreement with a recent report for single-crystal DyMnO3, for which the

observed peak in

was at 68 K when measured along the c-axis [79]. It is,

however, in slight disagreement with another reported measurement of 78 K for a

polycrystalline sample [81]. Our stoichiometric sample also appears to have a second

discontinuity at ~ 40 – 50 K, which is not present in the reduced sample. The

temperature of this peak is consistent with TSR for the RMnO3 family and its absence

for the reduced sample may be due to a disruption of the long-range spin rotation

ordering caused by properties relating to non-stoichiometric oxygen content (e.g.,

disruption to the exchange interaction due to oxygen vacancies or the presence of

Mn2+ cations). We have previously observed that slight changes of oxygen content can

have profound effect on the magnetic and resistive properties of perovskite manganites

[104][105] and similar behavior for the hexagonal phase may exist. However, the

difficulty of measuring the weak signal of TMn and TSR for DyMnO3 with

61

susceptibility measurements, due to the large magnetic moment of Dy, should be

noted. Of the four previous susceptibility measurement reports with single-crystal

DyMnO3 cited here, only two were able to observe TMn and none were able to detect

TSR. TSR has been observed at 57 K with specific heat measurements [78].

It is possible, however, that absence of TSR in these reports may not be just due

to the weak change in susceptibility at these transitions but due to slight oxygen non-

stoichiometry if samples were not fully stoichiometric after synthesis. Below 20 K

(Figure 3.24), there is a decrease in TDy from the stoichiometric (~9 K) to the slightly

62

reduced sample (~6 K). Previous reported values of TDy for single-crystal DyMnO3

have shown significant variation, 3 – 8 K [78] – [80], which could indicate that these

samples, which were synthesized by epitaxial thin film growth, are indeed subject to

large oxygen non-stoichiometry. Finally, measurements of the DyMnO3.22 sample,

with its positive Curie temperature, indicate slightly ferromagnetic behavior. Peaks in

occur at 6 and 310 k (Figures 3.22 – 3.24, inserts), suggesting that

antiferromagnetic and some other type of ordering are occurring below 6 and 310 K,

respectively. Clearly, the magnetic properties of the Hex2 phase differ significantly

from the P63cm phase and presence of any oxygen non-stoichiometry in the P63cm

structure will have considerable impact on these properties as well.

3.8 Conclusions

Our synthesis and structural measurements are in agreement with previous

work on perovskite manganites and suggest that the increased reducing conditions are

needed to form hexagonal Dy1-xYxMnO3+δ with decreasing x. This is in agreement

with previous reports of the perovskite forming from the hexagonal with smaller rare

earths (Ho, Er, and Y) under high pressure and support that the relative stability of

hexagonal and perovskite phases is due to the temperature, oxygen non-stoichiometry,

and compressibility dependence of the (R-O) and (Mn-O) bonds. Hexagonal

Dy1-xYxMnO3+δ materials were observed to reversibly absorb large amounts of oxygen

at ~200 – 300°C and to sharply desorb oxygen during transition back to the

63

stoichiometric P63cm phase above ~275 – 375°C or at lower temperatures in lower

partial-pressures of oxygen. Larger, reversible changes in oxygen content were

achieved by annealing at high pressures (δ = 0.25 – 0.38) and with hydrogen reduction

at 400°C (δ = -0.12 – -0.20), which, if combined, can yield OSC values up to ~2650

μmol-O/g. Rates of oxygen absorption were also observed to significantly decrease

with increasing Y content. However, regardless of any OSC potential applications, the

non-stoichiometric behavior of these hexagonal manganites’ oxygen content is, to the

best of our knowledge, newly reported for the RMnO3 family and was found to have

significant impact on the structural, thermal/chemical expansion, transport, and

magnetic properties. The TEC of the Hex2 and P63cm phases were determined to be

quite different, 8.4 – 11.6*10-6 K-1 and 2.1 – 5.6*10-6 K-1, respectively, and the

chemical expansion associated with the transition between these phases was found to

be 0.82 – 3.48*10-2 mol-1. Conductivity measurements at elevated temperatures

displayed thermal conduction of the Hex2 and the P63cm phases based either on

phonon-assisted conductivity or semiconducting behavior. The activation energies of

the Hex2 and the P63cm phase were found to be rather dissimilar, 43 – 47 and 105 –

120 kJ/mol, respectively. Proof of principle was also demonstrated that the large

changes in resistivity during the Hex2-P63cm transition could be coupled with oxygen

partial-pressure dependence of Dy1-xYxMnO3+δ to be used as a low elevated-

temperature oxygen sensor (230 – 330°C). Additionally, our conductivity

measurements of DyMnO3+δ may also suggest the P63cm phase of DyMnO3 has a Tc as

high as ~790°C, which may be due to the increased distortion of the P63cm structure

64

with increased Dy content. Finally, the magnetic properties of DyMnO3+δ were

demonstrated to change significantly with oxygen content. Strong oxygen content

dependence of the structural, magnetic, transport properties of Dy1-xYxMnO3+δ suggest

that the multiferroic properties of similar hexagonal RMnO3+δ manganites will change

significantly with oxygen stoichiometry.

65

CHAPTER 4: STUDY OF PEROVSKITE La1-xSrxFe1-yCoyO3+δ

(x ≥ 0.5), La0.2Sr0.8MnO3+δ, AND SrFe0.7Mn0.3O3+δ MATERIALS FOR

MIEC AND OXYGEN CARRIER APPLICATIONS

4.1 Introduction

As discussed in Section 1.2, the perovskite phases of La1-xAxM1-yM’yO3+δ (A=

Ca, Sr, Ba; M/M’ = Cr, Mn, Fe, Co, Ni, Ga) have been thoroughly studied for MIEC

applications. Traditionally, LaMnO3+δ doped with lower valance cations (such as Ca,

Sr, Ba) have been preferred materials for gas separation membranes and SOFC

cathodes applications. More recently, doped LaFeO3, and LaCoO3 materials have also

been extensively studied for MIEC applications due to their relatively higher oxygen

ion and electronic conductivities and faster surface exchange kinetics compared to

doped LaMnO3+δ compounds, but they also have large thermal expansion mismatch

and poor stability with YSZ and other common electrolytes for SOFC applications.

For most studies of these materials, cation doping is frequently less than 50% and

La.8Sr.2MnO3+δ and La.6Sr.4Fe.8Co.2O3+δ materials have been generally considered to

have the superior combination of properties (e.g., total ionic conductivity, TEC, and

stability) for application. At room temperature, LaMnO3+δ and LaFeO3 materials have

distorted orthorhombic perovskite structures (where this distortion, again, is larger for

LaMnO3+δ due to Jahn-Teller distortions). Both of these materials also have an

66

orthorhombic-rhombohedral phase transition at elevated temperatures (>400°C for Mn

and >900°C for Fe) and in general have been observed to transition to higher

symmetries with increased Sr substitutions [119]–[121]. The structure of LaCoO3

material is rhombohedrally distorted from the cubic perovskite; however, at high

temperatures or with increasing Sr substitution becomes increasingly cubic in

symmetry [122][123]. LaMnO3+δ material has p-type conductivity at high

temperatures (>1000°C) in air due to the formation of cation vacancies, as discussed in

Chapter 3 for RMnO3+δ (δ > 0) materials, and electronic conduction occurs by the

hopping of electron holes between the 2+ and 3+ states of the metal cations. LaFeO3

and LaCoO3 materials, however, remain nearly stoichiometric in such conditions. Hole

doping with divalent cations (Ba, Sr, Ca) can significantly improve electronic

conductivity for all of these materials over a wide range of temperatures. This

enhanced conduction is due to their mixed 4+/3+ valance state, which results in

electronic conduction by the small polaron conduction or metallic mechanisms

[119][124][125]. Substitutions with Sr have usually been preferred to other alkaline

earth metals because of its relatively good ionic size match with La and high

electronic conductivity. Sr substituted materials also show the best stability in

atmospheres containing H2O and CO2 and are able to accommodate large non-

stoichiometric oxygen contents at high temperatures in oxidizing and reducing

atmospheres [126]. Substitutions on the B-site can also increase the mixed valance

state and may further improve electronic conduction. In addition to the ratio of 4+/3+

B-site ion, the electronic conduction for substituted manganites, ferrites, and cobaltites

67

materials depends on several other factors, especially at elevated temperatures, such as

structural changes, oxygen non-stoichiometry, changes to the 3d-2p overlap with

structural distortions, band structure and filling, magnetic ordering, changes in the 3d

spin state, cation charge and size ordering, oxygen vacancy ordering, cation charge

disproportionation, etc. Experimentally, LaCoO3 material has shown to have better

electronic conductivity by comparison with LaMnO3+δ and LaFeO3 materials [127],

and studies of La.8Sr.2MxM’yM’’zO3+δ (M/M’/M’’ = Mn, Fe, Co; x+y+z = 1) and La1-

xSrxFe1-yCoyO3+δ (x = 0.2 and 0.4, 0 ≤ y ≤ 1) substituted compounds have also shown

pure Co samples to have considerably higher electronic conductivity (and

considerably higher thermal expansion) than Fe, Mn, or intermediate materials [128]–

[130]. The increased conductivity of Co-rich materials in these studies are primarily

attributed to charge disproportionation of Co3+ (2Co3+ → Co2+ + Co4+) at elevated

temperatures and increased 3d-2p overlap due to the higher electronegativity of Co.

However, the relative ease of reducing Co4+ compared to Mn4+ and Fe4+ can decrease

the mixed 4+/3+ valance ratio significantly below 50% (noting, 50% is considered

ideal for maximum conductivity) of Co-rich samples and hinder conduction at

elevated temperatures or in reducing atmospheres.

The ease of reduction of the metal cations is inherently not only important to

electrical conductivity but it is also extremely important to oxygen ion conduction as

well. Increased reduction increases fractional content of oxygen ion vacancies by the

need to charge balance and can result in enhanced oxygen ion conductivity (increased

σ0 in the Arrhenius relation). The ease of reducing from the 4+ to the 3+ valance states

68

with these cations increases with atomic number (Co > Fe > Mn); however, reduction

from the 3+ to the 2+ state changes to Co > Mn > Fe due to the stability of Fe3+ ion’s

3d5 high-spin configuration [131]. The ease of reducing from the 4+ to the 3+ state, in

the case for La1-xSrxFe1-yCoyO3+δ (x = 0.2 and 0.4, 0 ≤ y ≤ 1) materials, also results in

higher oxygen ion conductivity in Co-rich samples; however, oxygen ion conduction

has been shown to remain relatively high with significant Fe content for increased Sr

content (up to approximately an 80/20 Fe/Co ratio). Yet, the ease of reducing Co

samples also results in its poor stability and materials with large levels of Co tend to

begin decomposition to simple oxides at high temperatures in low partial-pressures of

oxygen (>1000°C), which occurs at lower temperatures with increased Sr content.

This behavior is especially undesirable for practical SOFC applications, which require

stability in low partial-pressures of hydrogen gas. Additionally, Co-rich samples have

comparatively higher TEC values, which can lead to thermal expansion mismatch for

layer material applications. Studies of La1-xSrxFe1-yCoyO3+δ materials have also shown

oxygen ion conductivity to improve significantly with increased Sr content, which

again can be partially attributed to it having higher fractional oxygen vacancies

[132][133]. However, for all these materials, oxygen ion vacancies must be disordered

in the crystal lattice to increase oxygen ion conductivity [134]. The resulting larger

decreases in oxygen content for high levels of Sr substitution significantly increases

the possibility of vacancy-ordered phases to occur (though vacancy ordering also

occurs in the in La-rich samples, e.g., LaMnO2.875 and LaMnO2.75 structures [135]).

For example, the case of the transition during reduction from the perovskite phase to

69

the Brownmillerite A2B2O5 (ABO2.5) structure, which occurs when the oxygen-

deficient perovskite structure begins vacancy ordering along the [101] direction at δ ≈

-0.5, has been frequently studied [136][137]. The Brownmillerite structure has an

orthorhombic unit cell and can be described as alternating sheets of corner-shared

octahedra and tetrahedral (MOn) along the c-axis (Figure 4.1). The vacancy ordering

of the Brownmillerite phase causes it to have significantly lower oxygen ion

conductivity than the reduced perovskite phase, despite the larger amount of oxygen

vacant-sites available for diffusion of the Brownmillerite structure [138]. The

Brownmillerite-type structure has been observed to occur the in large range of La1-

xSrxFe1-yCoyO2.5 compositions and it is generally believed to exist for its entire phase

diagram [139]–[141]. The Brownmillerite phase has also been observed in La-rich

La1-xSrxMnO2.5 (x = 0.2, 0.25, and 0.4) materials [142]. The Sr2Mn2O5 (SrMnO2.5)

structure, on the other hand, has the Ca2Mn2O5 structure type [143], which has ordered

Mn3+ pyramids instead of the alternating octahedra and tetrahedra layers as seen in the

Brownmillerite phase. This orthorhombic 225 structure (δ ≈ 0.5) has also been

previously observed for substituted La1-xSrxMnO2.5 samples (x = 0.8 and 0.9) [144].

Our previous work and other recent studies with related Sr-rich materials have found

additional vacancy-ordered phases between the stoichiometric perovskite phase and δ

= -0.5 and will be discussed further in Section 4.3.

70

Chapter 4 describes the synthesis of the perovskite phases of

La1-xSrxFe1-yCoyO3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ samples with high Sr

content (0.5 ≤ x), which was again guided by our previous work on the temperature

and oxygen vacancy dependence of the tolerance factor [92]. TGA measurements

were carried out to establish oxygen content behavior as a function of temperature and

atmosphere. TGA reductions were analyzed to identify potential oxygen vacancy-

ordered phases. EIS and four-point probe measurements were also conducted to

determine ionic conductivity and transport properties, respectively. Dilatometry

measurements determined values of TEC and CE for selected samples. The resulting

properties of these materials are also discussed in context throughout the chapter for

71

possible applications as MIEC materials for gas separation membranes, SOCF cathode

materials, and as oxygen carriers for CLC systems.

4.2 Synthesis and Stability

Polycrystalline samples of La1-xSrxFe1-yCoyO3+δ (0.5 ≤ x ≤ 1 for y = 0.5, 0 ≤ y

≤ 1 for x = 0.7), La1-xSrxMnO3+δ (x = 0.2, 0.8), and SrFe.7Mn.3O3+δ were synthesized

by solid-state reaction with appropriate amounts of La2O3, SrCO3, Fe2O3, Co3O4, and

MnO2 (all with >99.99% purity). Reactants were thoroughly mixed in an agate mortar

and fired in air in the temperature range of 500 – 1300°C in different partial-pressures

of oxygen with intermediate grindings followed by pressing samples into high-density

pellets at approximately 10 kbar. All steps of the synthesis were monitored with XRD

measurements for formation of the perovskite phase. Figure 4.2 is a compilation of

XRD patterns after the final synthesis step for selected samples of each series. Figure

4.3 shows an example of hydrogen reduction of the La.3Sr.7Fe.5Co.5O3+δ sample to

simple oxides to verify oxygen content of samples after synthesis by normalizing

samples’ oxygen content to their reduction products.

La1-xSrxFe1-yCoyO3+δ (0.5 ≤ x ≤ 1, 0 ≤ y ≤ 1 for x = 0.7) and La.8Sr.2MnO3+δ

samples formed reduced perovskite phase at 1100 – 1300°C in air and were then

subsequently slow cooled in air (1°C/min) from 500°C to achieve stoichiometric

oxygen content. Sr-rich samples of La1-xSrxFe1-yCoyO3+δ (x = 0.8, 0.9, 1), however,

required oxygenation at 180 – 200 bars at 500°C followed by 0.1°C/min to achieve

maximum oxygen content, which were determined by TGA to be δ = 0, -0.02, and

72

73

-0.11, respectively. The perovskite structure of La1-xSrxFe1-yCoyO3+δ (0.5 ≤ x ≤ 1, 0 ≤

y ≤ 1 for x = 0.7) is generally stable in 21 – 100% O2 for intermediate temperatures for

application (<900°C); however, Co-rich samples (x ≥ 0.7) may begin to decompose

as low as ~1000°C in Ar when held for extended periods of time (> 1day). Reports on

the solubility limit of Sr in La1-xSrxFe1-yCoyO3+δ materials for the perovskite phase are

conflicting and vary considerably from La.6Sr.4Fe.8Co.2O3 [145] (which is, in part,

responsible for the frequent choice of the composition for MIEC materials) to the

stoichiometric SrCoO3 perovskite phase [146]. The difficulty of formation is due to

the difficulty of achieving a high Co4+/Co3+ oxidation state ratio, which must increase

with Sr content (Co(3+x)+) to have stoichiometric oxygen content. Our high-pressure

annealings at low temperature confirm formation of the perovskite phase up to x = 0.9

(δ = -0.02) and formation of the stoichiometric perovskite phase for x = 0.3 from slow

cooling in air. Furthermore, high Sr content La1-xSrxFe1-yCoyO3+δ materials have been

confirmed to form the perovskite structure far beyond the frequently referenced x =

0.4 solubility limit several times over the past thirty years [147], which is in agreement

with our results.

SrMnO3+δ material, under standard high-temperature solid-state synthesis in

air, will form a four-layer hexagonal (4H) phase instead of the perovskite phase, due

to the small size of the Mn4+ cation, which raises the tolerance factor too high to form

the perovskite phase [92]. However, after the formation of the hexagonal 4H phase,

SrMnO3+δ material can be fired at 1400°C in UHP Ar to promote reduction of the

Mn(4+2δ)+ cations to lower the tolerance factor, which results in an oxygen-reduced

74

perovskite structure (δ ≥ - 0.43). This reduced sample can then be oxygenated at

500°C and slow cooled to room temperature at 1°C/min, which produces the

stoichiometric perovskite phase (verified by TGA and XRD). Pure SrMnO3 is only

kinetically stable at room temperature and begins to decompose back to the 4H phase

at 800°C in air. This “two-step procedure” was used to guide synthesis of

La.2Sr.8MnO3+δ and SrFe.7Mn.3O3+δ samples and the final steps of synthesis for these

materials were firings at 1400°C in Ar followed by oxygenation at 500°C. La and Fe

substitution considerably eases synthesis, due to increased Mn(3+x+2δ)+ cation size and

the relative ease of reducing Fe4+ (Fe4+ is also slightly larger than Mn4+ in octahedral

coordination), respectively. These substitutions also improve stability of the perovskite

phase to considerably higher temperatures. Because of the difficulty of forming Sr-

rich La1-xSrxMnO3 perovskites, only limited studies of these materials have been done

to date and the majority of previous MIEC studies have focused on materials below

~50% Sr substitution levels.

4.3 Oxygen Storage and Oxygen Content Behavior Measurements

As discussed in the introduction, oxygen flux of MIEC materials can be

approximated by σ

, where

. Thus, a

material’s fractional oxygen ion vacancies and activation energy of oxygen ion

conduction are primarily responsible for bulk oxygen ion conduction.

Thermogravimetric measurements were conducted to measure the fraction of oxygen

75

ion vacancies and to gain qualitative comparisons of activation energies of

La1-xSrxFe1-yCoyO3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ samples. Fractional oxygen

vacancies (FOV) were calculated by FOV = (-δ) / (stoichiometric oxygen content

value), where stoichiometric oxygen content is equal to 3 in all cases.

To measure fractional oxygen ion vacancies, assorted samples were heated up

to 500 – 1095°C at 1°C/min in various partial-pressures of oxygen (0 – 100%) in

TGA. Figure 4.4 shows oxygen deficiency of La.3Sr.7Fe.5Co.5O3+δ, La.2Sr.8MnO3+δ, and

SrFe.7Mn.3O3+δ samples on heating to 900°C with 1°C/min heating rate in 21% O2/Ar.

These conditions were chosen to simulate conditions for application. General behavior

of the La.3Sr.7Fe.5Co.5O3+δ sample seen here is typical for the La1-xSrxFe1-yCoyO3+δ

system, where increased Sr and Co content or lower partial-pressures of oxygen will

favor increased reduction, which results in various fractional oxygen vacancies in the

ranges of approximately FOV ≈ 0.033 – 0.1 (δ ≈ -0.1 – -0.3) for variations in Co

content (x = 0.7 and y = 0 – 1) and FOV ≈ 0.05 – 0.15 (δ ≈ -0.15 – -0.45) for variation

in Sr content (x = 0.5 – 1 and y = 0.5). Clearly, the fractional vacancies of

La1-xSrxFe1-yCoyO3+δ and SrFe.7Mn.3O3+δ samples are significantly higher than that of

the La.2Sr.8MnO3+δ sample. Again, for these vacancies to contribute to oxygen ion

conduction, they must be disordered. Our previous NPD studies have shown oxygen

vacancy ordering to occur between the perovskite and Brownmillerite/Ca2Mn2O5-type

phases (0 > δ > -0.5) for La1-xSrxFe1-yCoyO3+δ, La1-xSrxMnO3+δ, and SrFe1-yMnyO3+δ

(0.5 ≤ x) materials. High-temperature in situ NPD measurements of La1-xSrxMnO3+δ

samples showed pure Sr sample to begin vacancy ordering at δ = -0.07 and found

76

related vacancy-ordered phases of monoclinic Sr7Mn7O19 (SrMnO2.714) and tetragonal

Sr5Mn5O13 (SrMnO2.6) in addition to the orthorhombic Sr2Mn2O5 (SrMnO2.5) phase

(Figure 4.5). However, La substitution and increased temperatures were shown to

strain vacancy ordering considerably and the La.2Sr.8MnO3+δ sample’s oxygen ions

vacancies were fully disordered above δ ≈ -0.3 or above ~600°C [93][148]. Figure 4.6

shows increased stability of the La.2Sr.8MnO3+δ sample at these vacancy-ordered

phases during hydrogen reduction; however, as seen in Figure 4.4, its small amount of

oxygen ion vacancies at elevated temperature for application will certainly be

disordered. On the other hand, we have also found oxygen vacancy-ordered phases for

Sr8Fe4Co4O23 [149] and Sr8Fe8O23 (SrFe1-yCoyO2.875) structures (Figure 4.7) [139],

which would be in the reduced oxygen content range observed in Figure 4.4 (δ ≤

0.22). We have found that La , Co, and Mn substitutions in these materials disrupt the

formation of these ordered phases [150] and, with the addition of operating at elevated

temperatures, the δ = -0.125 oxygen vacancy phase for La1-xSrxFe1-yCoyO3+δ (x ≥ 0.2)

77

78

and SrFe.7Mn.3O3+δ samples will, most likely, not be present for application

conditions. This assertion is supported by the absence of any anomalous behavior

during reduction in Figure 4.4 at δ = -0.125 (FOV ≈ 0.0416) for these samples.

To qualitatively compare the activation energies of oxygen ion conduction,

samples were first reduced to approximately δ = -0.5 in TGA to 350 – 500°C with

1°C/min in 42% H2/Ar and then oxygenated by heating at 1°C/min in O2 (Figures 4.8

– 4.10). Hydrogen reductions were done to significantly increase the amount of

oxygen vacancies, which results in much higher σ0 constant in the Arrhenius equation

and permits oxygen diffusion at lower temperatures. Since bulk oxygen ion diffusion

is thermally assisted, the temperatures at which oxygenation of these samples begin is

directly related to the activation energies of these materials and lower temperatures of

oxygenation indicate lower activation energies. Table 4.1 lists the oxygenation

temperatures (TOx) for all samples. Comparing these oxygenation temperatures

suggest that the activation energies of La1-xSrxFe1-yCoyO3+δ and SrFe.7Mn.3O3+δ

samples are significantly lower than that of the La.2Sr.8MnO3+δ sample. Comparing the

oxygenation temperature of the La1-xSrxFe1-yCoyO3+δ series, activation energy appears

to be strongly correlated to the La/Sr content ratio, which is a minimum at x = 0.7, and

is lowest for an equal Fe/Co ratio (y = 0.5). Additionally, several La1-xSrxFe1-yCoyO3+δ

samples (0.28 ≤ x ≤ 0.32 and 0.3 ≤ y ≤ 0.7) oxygenated at room temperature. Thus, the

activation energies of these samples could not to be compared in similar fashion, as

they most likely begin oxygenation at various temperatures below room temperature.

The La.3Sr.7 Fe.5Co.5O3+δ sample oxygenated slightly faster than all other samples and

79

80

appears to be in the center of the phase diagram for the samples exhibiting

oxygenation at room temperature. For all of these comparisons to be valid, however, it

must be assumed that all samples have similar surface area, density, and diffusion

distances. These assumptions are reasonable for our samples, as they were synthesized

under identical conditions and broken into small chunks of similar size. On the other

hand, recent work on La.4Sr.6Fey-1CoyO3+δ materials have shown that grain size

decreases with Co content [129], which would result in shorter diffusion distances.

Such behavior may be present in our samples, which might explain why the

81

oxygenation temperature of the La.3Sr.7CoO3+δ sample is lower than that of the

La.3Sr.7FeO3+δ sample. Finally, it should be noted that attempts were made to

synthesize high-density samples (>95% crystalline density) with sol-gel synthesis to

calculate values of oxygen ion conduction from diffusion rates measured in TGA and

the samples’ surface area; however, no samples achieved sufficient density (all

attempts ranged from ~ 80 – 90%) to effectively discount molecular oxygen transport

through connected voids.

We have observed that these perovskite materials also have great potential as

OSC materials for CLC or certain types of separation membranes (e.g., synthesis gas).

In addition to determining activation energies of oxygen ion conduction, Figures 4.8 –

4.10 also demonstrate the potential of these materials for such applications. As

discussed with reversible hydrogen-oxygen cycling for oxygen storage in Chapter 1,

OSC materials for this type of application need to have considerable stability in

hydrogen and must be completely phase reversible upon reoxygenation. As shown in

Figures 4.9 and 4.10, SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ samples are reduced to their

Brownmillerite and (La.2Sr.8)5Mn5O13 phases, respectively. Their following oxidations

in TGA were verified to transition back to their stoichiometric perovskite phases (δ =

0) without decomposition to simple oxides (confirmed with XRD). However, while

these phases do show increased stability in TGA during reduction in hydrogen (see

Figure 4.6), they are relatively easy to further reduce on prolonged heating. Without

careful temperature control, these samples could further reduce to the point of

irreversible, partial decomposition, which is not an ideal property for application. On

82

the other hand, La1-xSrxFe1-yCoyO3+δ samples showed to have strong stability on

reducing its B-site cations to Fe3+ and Co2+ (calculated, -0.85 ≤ δ ≤ -0.35), which is

due to the relative stability of these oxidation states. Stability of samples’ Fe3+ and

Co2+ state was seen to increase significantly with La content from x = 0 to x = 0.5. We

believe this enhancement is due to the oxygen content of this mixed oxidation state

increasingly coinciding with the stability of ion vacancy-ordered structures at δ = -0.5

(e.g., at x = 0.5, results in δ = -0.5 due to charge neutrality). Figure 4.3 shows this

stable “plateau” for the La.3Sr.7 Fe.5Co.5O3+δ sample at Fe3+/Co2+ mixed oxidation state

(δ = -0.6), which was stable from approximately 350 to 500°C. These losses in oxygen

content are easily reversible back to the stoichiometric perovskite phase on

oxygenation (verified by XRD and TGA); however, high Sr and Co samples were not

fully reduced to their Fe3+/Co2+ state and samples of high Co content in this system

may begin decomposition under strong reducing conditions, though this behavior was

not observed at higher oxygen content. Table 4.1 includes the measured and calculated

OSC values of all samples. Calculated values for SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ

samples are for transitions between their fully oxygenated and δ = -0.5. Calculated

values for the La1-xSrxFe1-yCoyO3+δ series are for transitions between their fully

oxygenated phase and their reduced Fe3+/Co2+ mixed oxidation state (versus Δδ ≈ 0.35

– 0.55 for observed values). Though not directly observed, all materials, except

possibly La1-xSrxCoO3+δ, should be able to achieve these calculated OSC values by

increasing reducing conditions or extending the length of the 42% H2 firing. Finally, it

should be noted that the enhanced stability of the Fe3+/Co2+ mixed oxidation state

83

measured for La1-xSrxFe1-yCoyO3+δ samples is a rare property for perovskite materials

and, when coupled with its ability to reabsorb large amounts of oxygen at or near

room temperature seen in Table 4.1, makes this material a very strong candidate for

commercial development for oxygen carrier applications.

4.4 Electrical Conductivity

Rectangular bars for four-point probe measurements were pressed at 10 kbar (~

8x5x0.7 mm in shape) with embedded Pt leads in powder ground from the perovskite

phase after initial synthesis. Pressed pellets were then sintered at 1000°C with

3°C/min cooling in air before measurement. Conductivity measurements were done

with 1°C/min heating from room temperature to 900 °C in air (Figure 4.11).

84

As discussed in the introduction, the root causes of the temperature

dependence of La1-xSrxMO3+δ (M = Mn, Fe, and Co) materials’ conductivity are

extremely complex. Thus, the primary purpose of this section is to experimentally

compare the electrical conductivities of La1-xSrxFe.5Co.5O3+δ (x = 0, 0.1, 0.2, 0.3, 0.4,

and 0.5), La.2Sr.8MnO3+δ, and SrFe.7Mn.3 O3+δ samples with one another and with

previous reports of commonly cited MIEC materials. Activation energies of electrical

conduction were also attempted to be calculated based on small polaron conduction,

, which has been frequently used to model these systems as discussed in

the introduction. However, as will be shown, this model clearly fails for several

La1-xSrxFe.5Co.5O3+δ samples as these compositions become increasingly metallic with

La content.

La1-xSrxFe.5Co.5O3+δ samples were found to have higher conductivity, ranging

from ~200 – 1300 S/cm at 500°C with increased La content, compared to

La.2Sr.8MnO3+δ and SrFe.7Mn.3 O3+δ samples, which were found to be ~140 and 25

S/cm at 500°C, respectively. Previous reports of La1-xSrxFe1-yCoyO3+δ materials (0 ≤ x

≤ 0.4), when compared, show conductivity to generally increase with Sr content

[129][130][145][151] and reported conductivities of ~240 and 650 S/cm at 500°C for

La.8Sr.2Fe.5Co.5O3+δ and La.6Sr.4Fe.5Co.5O3+δ samples, respectively. Thus, the

maximum conductivity for the La1-xSrxFe.5Co.5O3+δ series appears to be at x ≈ 0.5,

which is in agreement with the ideal carrier concentration of a mixed 50/50 - 4+/3+

cation valance (for δ = 0) for polaron hopping and is also in agreement with a previous

report of the La1-xSrxFeO3+δ system [121]. Yet Sr-rich samples have increased oxygen

85

non-stoichiometry at elevated temperatures (e.g., x = 0.9 and 1 have oxygen

deficiencies of δ ≈ 0.35 and 0.45, respectively, at 900°C in 21% O2/Ar) which will

result in significant reduction of the 4+/3+ oxidation state below a one to one ratio.

Any such reduced samples should lower conductivity based on carrier concentration.

Thus, the high conductivity of Sr-rich versus La-rich samples suggests that other

mechanisms are enhancing the conductivity of x > 0.5 samples. Our previous NDP

studies of La1-xSrxFe.5Co.5O3+δ materials (0.5 ≤ x ≤ 1) have shown the system to

become increasingly cubic with Fe/Co-O-Fe/Co bond angles approaching 180° with

increased Sr content [149]. These structural changes may account for the increased

conductivity of Sr-rich samples, due to increased 3d-2p overlap, versus La-rich

samples with distorted perovskite phases of similar or increased carrier concentration.

Figure 4.12 shows conductivity data for all materials fitted with

based on small polaron conduction. Activation energies of these materials were

determined by the slope of this fit and are compared with common materials for MIEC

related applications in Table 4.2. La1-xSrxFe.5Co.5O3+δ samples (0.5 ≤ x ≤ 1) showed

decreasing activation energies up to x = 0.6 and slight increases from x = 0.6 to x =

1.0 (ranging 3.16 – 9.49 kJ/mol), which were notably lower than samples previously

reported with high La content and with similar Fe/Co ratio. However, it is clear on

inspection of Figure 4.11 that the conductivity of La1-xSrxFe.5Co.5O3+δ samples became

increasingly metallic in behavior with increased La content. This may be due to their

higher conductivities, and the activation energies for these samples may be less

accurate with increased La content. For the SrFe1-yMnyO3+δ system (0.1 ≤ y ≤ 1),

86

electrical conductivity and activation energies have been studied at lower temperatures

[152][153]. They have shown nearly constant activation energies and an increased

conductivity with increased Fe content. Our measurement of SrFe.7Mn.3 O3+δ at higher

temperatures is in good agreement with these reports. Finally, in addition to increased

conductivity, measurement of La.2Sr.8MnO3+δ material was shown to have

considerably smaller activation energy than that of previously reported La-rich, La1-

xSrxMnO3+δ materials.

Increased conductivity with temperature of La1-xSrxFe.5Co.5O3+δ,

La.2Sr.8MnO3+δ, and SrFe.7Mn.3 O3+δ samples ends approximately above 430 – 580,

190, and 530°C, respectively, and samples become “metallic” in behavior

(conductivity decreases with increased temperature). This may be due to Ea becoming

significantly smaller than RT at increased temperatures, which results in the pre-

87

exponential term, , effectively modeling conductivity because the exponential

term, , approaches 1. However, with the exception of La.2Sr.8MnO3+δ material,

the observed transition temperatures in Table 4.2 are not in good agreement with their

calculated values ( ). This indicates the common fitting of conductivity

with may be a fortunate model that averages several complex effects as

previously discussed. Decreases in conductivity with increased temperatures for

88

SrFeO3+δ material have been attributed to increased oxygen vacancies, which results in

a lower 4+/3+ oxidation state ratio [154]. Another possible mechanism of this

transition could be associated with spin transitions or magnetic ordering. Such

transition has been observed in LaCoO3, which is very similar in nature and

temperature to the change in the temperature dependence of conductivity seen in

Figure 4.11 for La1-xSrxFe.5Co.5O3+δ and SrFe.7Mn.3 O3+δ samples. This transition, at

380°C for LaCoO3 changes from thermally assisted semiconductor to thermally

impeded metallic behavior (noting there is no “jump” in conductivity as seen, for

example, with metal-insulator transitions for RMnO3 perovskites) is due to changes in

spin state that results from an ordered array of alternating high-spin (S = 2) and

intermediate-spin (S = 1) states of Co3+ [155]. Though the mechanism for thermal

conduction is different for undoped LaCoO3 (semiconducting), any such ordering

would also interfere with hole-doped polaron-assisted conduction. However, doping

LaCoO3 with Sr has shown to significantly complicate the spin structure at elevated

temperatures, due to the addition of mixed spin states and possible disproportionation

of Co3+, and continues to be an active area of debate [156]–[158]. The further addition

of Fe substitutions to this system naturally results in La1-xSrxFe.5Co.5O3+δ having

extremely complex high-temperature transport properties. Thus, the failure of thermal-

assisted behavior for SrFe.5Co.5O3+δ and SrFe.7Mn.3 O3+δ at 430 – 580°C may be

attributed to magnetic ordering due to spin-state transitions as seen with LaCoO3,

properties related to increased oxygen vacancies as seen in SrFeO3+δ or other changes

at elevated temperatures such as charge localization, band filling associated with a

89

spin transition, structural changes, etc. The disproportionation of Co3+, as discussed in

the introduction, may enhance conductivity and could possibly result in Tm-obs being

slightly greater than Tm-calc.

Deviation of La1-xSrxFe.5Co.5O3+δ (x = 0.2 – 0.5) samples from

behavior is interesting, as other studies with La-rich samples (x = 0.2 and y = 0 – 1,

and x = 0 – 0.4 and y = 0.2) have been shown to have reasonable agreement with this

model (though not with much justification) [130][145]. Clearly, Sr substitution in the

range of 0.5 ≤ x ≤ 0.8 has a strong impact on the temperature dependence of

conductivity. Conductivity data was also attempted to be fitted with

based on semiconducting (n = 0) and non-adiabatic polar hopping (n = 3/2); however,

these attempts yielded slightly worse linear fits and stronger disagreement of Tm-obs

and Tm-calc by comparison with the n = 1 case. It would seem the temperature behavior

of conductivity in these samples is on the cusp of metallic and semiconducting

behavior. Ultimately, however, the high electrical conductivity of Sr-rich samples at

elevated temperature is the most important property for application, no matter how

thermal conductivity is chosen to be modeled. Figure 4.11 gives a clear comparison of

conductivities at elevated temperatures of all samples for possible application.

4.5 Total Ionic Conductivity

Activation energies of total ionic conductivity of selected samples deposited on

YSZ were determined by EIS measurements by fabricating half and full cells, which

are common methods to test electrodes for separation membranes or for cathode

90

materials for SOFC applications. La.3Sr.7Fe.5Co.5O3+δ sample was chosen for EIS

measurements based on its relatively high electrical conductivity and TGA

measurements, which suggests it may have the system’s highest oxygen ion

conductivity. The methodologies of conducting half-cell measurements were largely a

continuation of former group member Stillwell’s M.S. thesis project [159]. Data from

this work for half cells with La.2Sr.8MnO3+δ and SrFe.7Mn.3O3+δ cathodes are compared

to these cells re-measured (in addition to La.3Sr.7Fe.5Co.5O3+δ) with the following

attempted improvements: sample inks were screen printed instead of drop evaporated

or brush painted, which created more uniform film thickness and planarity; the

addition of a thin spray-evaporated Ce.8Gd.2O2+δ (20CGO, δ < 0) buffer layer to

prevent Sr reactivity with YSZ, which were observed to have high reactivity in

stability experiments; replacing 8YSZ with 3YSZ electrolyte, to increase TEC of the

electrolyte to match higher TEC of samples (see Section 4.6); and the addition of a

Faraday cage around electrodes and the repositioning of electrode leads, which

significantly reduced signal noise observed in previous data. In addition to half-cell

measurements, La.3Sr.7Fe.5Co.5O3+δ material was also prepared as a full cell to function

as an operational SOFC. Prefabricated NiO-backed 3YSZ (0.55 mm thickness) was

first coated with a thin layer of CGO, which was deposited by spray evaporation at

~250°C and then sintered at 1000°C for 12 hours. La.3Sr.7Fe.5Co.5O3+δ material

suspended in ethanol/zirconia (ink) was screen printed on top of the CGO buffer layer

and then sintered again at 900°C for 12 hours. The cell was then secured between two

alumina cylindrical enclosures with alumina paste and high-temperature epoxy, which

91

were designed to flow pure oxygen and hydrogen over the cathode and anode sides of

the cell, respectively. Au and Ni electrodes were attached with La.3Sr.7Fe.5Co.5O3+δ and

NiO inks to the cathode and anode, respectively. Finally, the whole assembly was

loaded into a custom-built furnace for EIS measurements. Figure 4.13 shows diagrams

of completed half and full cells. Similar to the half-cell fabrication procedure, this

process was extremely time consuming (5+ days to produce a cell, with drying steps)

and frequently produced poor test cells despite best efforts and established practices

due to gas leaks, bad cathode/anode-electrode connections, layers delaminating or

cracking, cells fracturing, etc. The presented full-cell data for La.3Sr.7Fe.5Co.5O3+δ is

the one “good” run of four separate attempts, and half-cell measurements had about a

50% success rate. Full cells with other Fe/Co (x = 0.7, y = 0.4, 0.6) ratios were

attempted but were not successful. Additionally, half cells were also attempted to be

constructed with 20CGO wafers but proved to be too fragile and would crack when

mounted to the test stands. The combination of the length of time for fabrication and

the success rate of full and half cells, unfortunately, limited the number of samples

that could be tested based on the high demand for shared EIS instruments.

92

In this section, the basic theory of EIS measurements of layered oxygen ion

conductors will first be discussed. Following this overview, the activation energies of

cells with La.3Sr.7Fe.5Co.5O3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ cathodes are

compared to one another and these results are also compared to previously measured

Stillwell cells. Additionally, La.3Sr.7Fe.5Co.5O3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ

cells are also compared to similar cells with La.8Sr.2MnO3+δ and La.6Sr.4Fe.8Co.2O3+δ

cathodes, which are generally considered to be the best materials in their respective

series for layered MIEC applications. Finally, the area-specific resistance of our and

previously reported test cells are compared.

Electrochemical impedance is measured by applying AC potential,

(where ω is the radial frequency and t is time), and measuring the resulting

current response of the test cell, (where φ is the phase-shift of the

response), as function of ω. Applied AC potential must typically be small (here, 4 mA)

to ensure a linear response. Assuming ohmic behavior, the impedance of the cell is

, which can be expressed as a complex function with Eulers

relationship and simplified to . The behavior of the

electrochemical cell can be approximated with an effective circuit known as a Randles

cell (Figure 4.14), which is a resistor in series with a RC circuit. In this model, Rs is

the total electrical resistance of the cell, Cdl is the double-layer capacitance due to the

cell and electrodes, and Rp is the effective polarization resistance. The polarization

resistance is the property of interest, as it directly relates to total ionic conductivity.

93

Using this model and the derived expression for Z(ω), Zo (Rp is equivalent to Z0) can

be easily determined by plotting the real and complex impedance of the cell

(commonly referred to as a Nyquist plot). The effective diameter of the circular

response impedance yields the value of Z0. This parameter is typically multiplied by

the area of the active electrode (because the result is independent of electrode size)

and is referred to as the area-specific resistance (ASR) of the cell. The ASR is

inversely proportional to the Arrhenius relation: (where C is an arbitrary

constant). Thus, by plotting the natural log of ASR versus inverse temperature

(

), which is frequently called an Arrhenius plot, the

activation energy of the cell can be found by multiplying the slope of this line by the

gas constant R. Arrhenius plots are also useful for observing changes to the ideal

Arrhenius model (nonlinear behavior) and to quickly compare activation energies of

multiple materials. ASR is also an important Figure of merit for SOFC applications,

and should be less than 0.5 ohm*cm2 to achieve the common benchmark goal of

1kW/kg [160].

EIS half- and full-cell measurements were made at approximately 500 – 900°C

at 50 – 100°C intervals. Figure 4.15 is an example of a Nyquist plot for a “good” half

cell with La.3Sr.7Fe.5Co.5O3+δ material, which displays Randles cell-type behavior, and

is representative of the data collected for “good” cells. The resulting Arrhenius plots

for half-cell measurements with La.3Sr.7Fe.5Co.5O3+δ, La.2Sr.8MnO3+δ, and

SrFe.7Mn.3O3+δ cathodes and the full-cell measurement with La.3Sr.7Fe.5Co.5O3+δ

94

cathode are shown in Figure 4.16. Calculated activation energies from Arrhenius plots

are shown in Table 4.3. Activation energies decreased, in the order of largest to

smallest, for SrFe.7Mn.3O3+δ (128 kJ/mol), La.2Sr.8MnO3+δ (124 kJ/mol), and

La.3Sr.7Fe.5Co.5O3+δ (85.1 kJ/mol) half cells and ASR values were also observed to

decrease in the same order. The activation energy of the full-cell measurement of

La.3Sr.7Fe.5Co.5O3+δ is in good agreement with its half-cell measurement. Its increase

in ASR, seen in Figure 4.16, is mostly likely due to the full cell’s significantly larger

thickness than the half cell. The value of total ionic conductivity of

La.3Sr.7Fe.5Co.5O3+δ in comparison with the other two samples is in agreement with its

excepted lower value based on its comparative lower electronic conductivity and its

qualitative lower oxygen ion conduction determined from four-point probe and TGA

measurements, respectively. Furthermore, upon comparing the total ionic conductivity

of SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ cells, their close values are not surprising due to

the higher electrical conductivity of La.2Sr.8MnO3+δ (and its better TEC match with

20CGO and 3YSZ; see next section) and the implied, higher oxygen ion conductivity

of SrFe.7Mn.3O3+δ.

95

96

Figures 4.17 and 4.18 show Arrhenius plots for the improved half cells (as

described in section’s introduction) versus half cells from Stillwell data for half cells

with SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ cathodes and Table 4.3 shows their calculated

activation energies. Activation energies and ASR values of these materials with half-

cell “improvements” showed little change (minor decrease with SrFe.7Mn.3O3+δ and

small increase with La.2Sr.8MnO3+δ). However, the major enhancement of these new

cells was their length of operation. Previous measurements with both samples showed

significant increases in ASR and activation energies after short periods of time (<1

hour), which would eventually lead to complete failure of ionic conductivity. The

improved cells showed stable operation with minimal increase in ASR and activation

energy during the entire length of their measurements (> 3 hours) and no cell failures

were observed (La.3Sr.7Fe.5Co.5O3+δ full cell showed consistent operation for over 60

hours). This improvement is most likely due to the addition of the 20CGO buffer

layer, which almost certainly prevented reactivity of the cathode and electrolyte, as

YSZ and Sr-rich perovskite materials have been shown to have poor chemical stability

[159].

97

Table 4.3 also shows activation energies of commonly tested cathode materials

with similar full and half cells. By comparison, SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ

half cells had activation energies similar to that of the reported range of

La.8Sr.2MnO3+δ-8YSZ-Pt cells (100 – 160 kJ/mol) [159][161]. It appears the larger

thermal expansion of these materials effectively cancels their oxygen ion and electrical

conductivity advantages over La.8Sr.2MnO3+δ (see next section). Measurements of

La.3Sr.7Fe.5Co.5O3+δ half and full cells produced significantly lower activation energies

than these materials. More importantly, they also had lower activation energy than

La.6Sr.4Fe.8Co.2O3+δ -20SCO (Ce.8Sm.2O2+δ) - 3YSZ - Pt (154 kJ/mol), which was

prepared under similar methods [162] (20GSO and 20GCO should have nearly

identical material properties as a buffer material). These results indicate that the

La.3Sr.7Fe.5Co.5O3+δ composition is a superior cathode material for desired MIEC

applications. However, lower activation energies (50 – 115 kJ/mol) have been

achieved with La.6Sr.4Fe.8Co.2O3+δ cells by replacing the YSZ electrolyte entirely with

substituted CeO2+δ materials, which is due to their lower thermal expansion mismatch

[163] (see next section), but our data suggests that Sr-rich La1-xSrxFe1-yCoyO3+δ

98

cathodes in similar cells could have even lower activation energies with such

electrolytes.

Despite having improved or similar activation energies, the observed ASR

values of our cells are much higher than previous reports of various test cells with

La.8Sr.2MnO3+δ and La.6Sr.4Fe.8Co.2O3+δ materials. Our lowest measurement of ASR

was 0.21 Ω*cm2 at 900°C with a La.3Sr.7Fe.5Co.5O3+δ half cell. Similar ASR values

have been achieved with La.6Sr.4Fe.8Co.2O3+δ -20GCO -20ScSZ (20% Sc stabilized

ZrO2) - Ni/20ScSZ at 600°C (versus 3.74 Ω*cm2 at this temperature with

La.3Sr.7Fe.5Co.5O3+δ half cell) [54]; however, this improvement is principally due to

careful engineering of the cell and of the designed microstructure of the Ni/20ScSZ

anode. Measurements on ANL equipment of previously reported cathode-YSZ-Pt cells

(e.g., La.8Sr.2MnO3+δ) show them to have significantly larger ASRs (approximately an

order of magnitude) while having similar activation energies [164]. These results

indicate that the cells and test stand are poorly engineered by comparison with these

other reports (e.g., bad interconnects, larger thickness of cells, poor material

deposition methods, layers delaminating, etc.). The large values of Rs and its

temperature dependence seen in Figure 4.15 (which are typically < 0.1 Ω*cm2 and

nearly thermally independent for ideal behavior) are signs of deviation from ideal

Randles cell behavior. This type of behavior, though, can be modeled with a modified

Randles cell by adding another resistor in series with the effective polarization

resistance (ASR). Using this modified model, calculated activation energies seen here

are still valid and ASR of our cells can be compared qualitatively to one another.

99

Calculated activation energies suggest that La.3Sr.7Fe.5Co.5O3+δ-based cells may indeed

have lower ASR than similar cells with La.6Sr.4Fe.8Co.2O3+δ, if engineered with similar

methods. These results also clearly indicate how important careful engineering is for

potential SOFC application and could very well be a limiting factor for mass

production.

4.6 Thermal and Chemical Expansion

As is the case with hexagonal and perovskite RMnO3+δ materials, expansion to

the perovskite La1-xSrxFe1-yCoyO3+δ, La1-xSrxMnO3+δ, and SrFe1-yMnyO3+δ structures is

due to TE and CE. However, as also discussed in Section 3.5, separating these effects

in perovskite materials can be difficult due to gradual loss of oxygen content over

large changes in temperature in partial-pressures of oxygen. Again, oxygen content

behavior and expansion were compared with TGA and dilatometry data under

identical conditions to separate their effects. TE values were measured in temperature

regions below where significant reductions in oxygen content were observed (<200 –

300°C). CE was measured by subtracting the effect of TE during reduction observed

in TGA above 300°C. However, the effects of TE and CE are not as well separated as

experiments designed in Section 3.5 by virtue of the similar effects TE and CE have

on overall expansion, and a level of uncertainty exists for CE (by best estimation, 10-

20%). Thus, effective TEC values are also reported in this section, which were

calculated over the entire temperature range without separating the effects of CE from

TE.

100

Pellets were cut from dense samples for dilatometry measurements after

synthesis of perovskite material (~ 5x3x2 mm in shape) for La1-xSrxFe.5Co.5O3+δ (x =

0.5, 0.6, 0.7, and 0.8), La.3Sr.7Fe.6Co.4O3+δ, La1-xSrxMnO3+δ (x = 0.2 and 0.8), and

SrFe.7Mn.3O3+δ samples. The La.8Sr.2MnO3+δ sample was selected for dilatometry

measurement as a reference. The remaining samples were selected based on structural

(large fractional oxygen vacancies, which are mostly likely disordered), conductivity,

and EIS measurements, which indicated their potential as cathode materials for SOFC

or interconnects for separation membranes. Dilatometry measurements were measured

with 0.5°C/min heating to 900°C in 21% O2/Ar (Figure 4.19). TEC and CE values

were again calculated with the formulas

and

. These values are reported in Table 4.4 along with other

commonly used materials for MIEC applications [130][151]. Our measurements and

101

these reports are in agreement with general observations of effective TEC values for

the (La1-xSrxMnO3+δ)-( La1-xSrxFeO3+δ)-( La1-xSrxCoO3+δ) ternary system, which

indicated increased TEC with increased Sr/La ratio and with B-site cation content in

the order, from smallest to largest TEC, Mn, Fe, and Co. CE values reported in Table

4.4 do not appear to change with any component’s content and their differences may

be due largely to uncertainty. Still, their approximate value is in good agreement with

previous reports of substituted perovskite LaMnO3 materials (~2 – 4*10-2 mol-1) [109]

– [111].

In an effort to resolve this uncertainty and better separate the effects of TE and

CE, a reduced La.4Sr.6Fe.5Co.5O3+δ sample (δ = -0.19, from Ar reduction at 800°C) was

102

also oxygenated during dilatometry and TGA measurements under identical conditions

(Figure 4.20). Comparing Figures 4.19 and 4.20, the larger change in oxygen content

over a smaller temperature gradient of the reduced sample oxygenating versus the

previous dilatometry measurements of slight reduction in 21% O2/Ar is clearly

apparent on inspection. Thus, the CE calculated from oxygenation measurement was

more accurate, and yielded a CE of 1.83*10-2 mol-1 (which is in the estimated

uncertainty range of the 21% O2/Ar run’s value of 2.2*10-2 mol-1). This result may

suggest that the series has generally smaller CEs than values typically found in

perovskite manganites; however, it is important to remember that the larger changes in

oxygen content of the La1-xSrxFe.5Co.5O3+δ system in oxygen atmospheres results in

CE having a larger net effect on overall expansion than observed in most perovskite

manganites.

As previously discussed, thermal mismatch of expansion of layered materials

for separation membranes and SOFC application can severely limit total ionic

conductivity and cause a total failure in ionic conduction. Table 4.4 includes the TEC

values of typical electrolytes for separation membranes and SOFC applications

103

[165][166]. Clearly, the TEC values of our Sr-rich samples are much higher than these

materials and will be a limiting factor in their combined total ionic conductivity. The

small thermal expansion mismatch between La.6Sr.4Fe.8Co.2O3+δ and 20CGO, and

between La.8Sr.2MnO3+δ and 8YSZ materials (as seen in Table 4.4) is, in part,

responsible for their low activation energies of total ionic conductivity reported in

Section 4.5. However, as was also discussed in Section 4.5, our measurements of

La.3Sr.7Fe.5Co.5O3+δ material showed lower activation energies of total ionic

conductivity than reported measurements of La.6Sr.4Fe.8Co.2O3+δ with nearly identical

cell components and fabrication, despite the La.3Sr.7Fe.5Co.5O3+δ material’s increased

TEC. Clearly, the common choice of La.6Sr.4Fe.8Co.2O3+δ as the “premier” MIEC

material in the La1-xSrxFe1-yCoyO3+δ system for layered MIEC applications has

overcompensated for lower values of TEC (or because of the incorrectly reported

solubility limit of x = 0.4, as discussed in Section 4.2). Measurements of

La.3Sr.7Fe.5Co.5O3+δ material suggest that samples of higher Sr content can yield lower

activation energies of total conductivity, due to their significantly higher oxygen ion

and electrical conductivity, despite increased thermal mismatch with desired

electrolyte materials. However, comparing the similar activation energies of

SrFe.7Mn.3O3+δ and La.2Sr.8MnO3+δ cells to La.8Sr.2MnO3+δ cells, it appears any

advantage of increased oxygen ion conductivity of these Sr-rich materials is

effectively negated by their larger thermal mismatch with YSZ materials. Clearly,

ideal MIEC materials for layered applications must balance the advantages (increased

104

oxygen ion conductivity) and disadvantages (increased CE) of having increased

fractional oxygen ion vacancies.

It is possible that thermal mismatch is also partially responsible for the

significant increase in ASR we observed with our samples, which ultimately is the

most important factor in the cell’s total power density for SOFC application.

Unfortunately, the quality of cell and test stand engineering also has a large impact on

ASR, which makes it extremely difficult to isolate the impact of cathode-electrolyte

thermal mismatch on ASR. Repeated temperature cycling of the La.3Sr.7Fe.5Co.5O3+δ

full cell did not show decreases in performance over time, which may be an indicator

that engineering, and not thermal mismatch, is principally responsible for increased

ASR.

4.7 Conclusions

TGA measurements indicated that La1-xSrxFe1-yCoyO3+δ and SrFe.7Mn.3O3+δ

samples have large oxygen ion conductivity based on their high fractional oxygen ion

vacancies in air and their ability to oxygenate at low temperatures (<100°C) when

highly reduced (δ ≈ -0.5). TGA measurements also demonstrated

La1-xSrxFe1-yCoyO3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ samples have considerable

OSC (1500 – 3500 μmol-O/g) with hydrogen-oxygen cycling. Several La1-xSrxFe1-

yCoyO3+δ samples, with stoichiometries near x = 0.7 and y = 0.5, were able to reabsorb

oxygen at room temperature, making them excellent candidates for chemical looping

combustion or related air separation applications. Four-point probe measurements of

105

La1-xSrxFe.5Co.5O3+δ, La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ samples found electrical

conductivities of ~ 200 – 1300, 140 and 25 S/cm at 500°C, respectively. The

conductivity of La.2Sr.8MnO3+δ and SrFe.7Mn.3O3+δ samples showed temperature

dependence, , which would nominally indicate thermally assisted

behavior. EIS measurements of 3YSZ-20CGO-Pt/Ni cells with La.3Sr.7Fe.5Co.5O3+δ,

La.2Sr.8MnO3+δ, and SrFe.7Mn.3O3+δ cathode materials yielded activation energies of

total ionic conductivity of ~ 82.5, 128, and 125 kJ/mol, respectively. The activation

energy of cells with La.3Sr.7Fe.5Co.5O3+δ cathode material was considerably lower

(approximately 50%) than previously reported in similar cells with

La.6Sr.4Fe.8Co.2O3+δ cathode material, which has generally been considered to be the

superior MIEC material for air separation and SOFC applications in the

La1-xSrxFe1-yCoyO3+δ system. On the other hand, cells with La.2Sr.8MnO3+δ and

SrFe.7Mn.3O3+δ materials were found to have comparable activation energies with

previously reported similar cells with the commonly cited La1-xSrxMnO3+δ cathode

material, La.8Sr.2MnO3+δ, despite our measurements which indicate their higher

oxygen ion and electronic conductivity. All measured cells had ASR values

significantly greater than previously reported similar cells with La.6Sr.4Fe.8Co.2O3+δ

and La.8Sr.2MnO3+δ cathodes primarily due to poor cell and test stand engineering.

Dilatometry measurements agreed with previously reported trends of La1-xSrxMnO3+δ,

La1-xSrxFeO3+δ, and La1-xSrxCoO3+δ systems, which showed that effective TEC

increases significantly with increased Sr content. These measurements also found

effective TECs of 20.8 – 27.5, 21.5, and 14.5*10-6 K-1 for La1-xSrxFe1-yCoyO3+δ (0.2 ≤

106

x ≤ 0.5), SrFe.7Mn.3O3+δ, and La.2Sr.8MnO3+δ samples, respectively. While our results

suggest that the increased ASR values of our materials are mainly due to their inferior

cell design and engineering, some of the increase in ASR is also undoubtedly due to

increase thermal expansion mismatch from the higher TEC of our materials with

3YSZ. The similar activation energies of total ionic conductivity of cells with

La.2Sr.8MnO3+δ and SrFe.7Mn.3O3+δ materials compared to La.8Sr.2MnO3+δ material

may also be due to the relatively higher thermal expansion mismatch of our samples

with 3YSZ. Nevertheless, our results clearly indicate that Sr-rich La1-xSrxFe1-yCoyO3+δ

materials can yield superior activation energies of total ionic conductivity, despite its

thermal mismatch with 3YSZ, and are strong candidates for SOFC and layered

material applications. The performance of all measured and related Sr-rich materials

would also most likely improve significantly when supported by better electrolyte

materials (e.g., CGO and CSO) designed for use with the La1-xSrxFe1-yCoyO3+δ system,

which have similar thermal expansion and better chemical stability.

107

REFERENCES

[1] C. Linde US Patent 727650 1903

[2] M. Davies, History of Science 22 (1989)

[3] M. Kanoglu, I. Dincer, M.A. Rosen, I.J. of Energy Research 32 (2008) 35

[4] B.Z. Maytal, Cryogenics 46 (2006) 49

[5] N. Greenwood, A. Earnshaw, Chemistry of the Elements 2nd

ed., Butterworth Heinemann (2002) 604

[6] Compressed Gases Association, Handbook of Compressed Gases, Van Nostrand Reinhold Co, 1990

[7] S.P. Nandi, P.L. Walker Jr., Sep. Sci. and Tech. 11 (1976) 441

[8] L. Murray, M. Dinca, J. Yano, S. Chavan, S. Bordiga, C. Brown, J. Long, Amer. Chem. Soc. 132 (2010) 7856

[9] J. Emsley, Nature’s Building Blocks: An A-Z Guide to the Elements, Oxford University Press 2001

[10] S. Shelley, CEP 105 (2009) 6

[11] Y. Lin, D. MacLean, Y. Zeng US Patent 6059858 2000.

[12] H. He, H.X. Dai, C.T. Au, Catal. Today 90 (2004) 245

[13] R. DiMonte, P. Fornasiero, M. Graziani, J. Kašpar, Alloys and Comp. 275 (1998) 887

[14] Y. Nagai, T. Yamamoto, T. Tanaka, S. Yoshida, T. Nonaka, T. Okamoto, A. Suda, M. Sugiura, Catal. Today 74 (2002) 225

[15] P. Singh, M. Hegde, J. Gopalakrishnan, Chem. Mater. 20 (2008) 7268

[16] M. Karppinen, H. Yamauchi, S. Otani, T. Fujita, T. Motohashi, Y. Huang, M. Valkeapää, H. Fjellvag, Chem. Mater. (2006) 490

108

[17] T. Motohashi, S. Kadota, H. Fjellvag, M. Karppinen, H. Yamauchi, Mater. Sci.

Eng. B 148 (2008) 196

[18] S. Kadota, M. Karppinen, T. Motohashi, H. Yamauchi, Chem. Mater. 20 (2008) 6378

[19] S. Räsänen, T. Motohashi, H. Yamauchi, M Karppinen, Solid State Chem. 183 (2010) 692

[20] O. Chmaissem, H. Zhen, A. Huq, P. Stephens, J. Mitchell, Solid State Chem.

181 (2008) 664

[21] W. Nernst, Z. Elektrochem. 6 (1899) 41

[22] J. Kilner, Solid State Ionics 129 (2000) 13

[23] C. Wagner, Prog. Solid State Chem. 10 (1975) 3

[24] P. Gellings, H.J.M. Bouwmeester, The CRC Handbook of Solid State

Electrochemistry 1997

[25] D. Yuan, F. Kroger, Electrochem. Soc. 116 (1969) 594

[26] Y. Teraoka, H. Zhang, S. Furukawa, N. Yamazoe, Chem. Lett. (1985) 1743

[27] H. Dong, Z. Shao, G. Xiong, J. Tong, S. Sheng, W. Yang, Cata. Today 67 (2001) 3

[28] C. Tsai, A. Dixon, W. Moser, Y. Ma, Ceram. Process. 43 (1997) 2741

[29] S. Badwal, F. Ciacchi, Int. Ion. 6 (2000) 1

[30] V. Kharton, A. Yaremchenkop, A. Kovalevsky, A. Viskup, E. Naumovich, P. Kerko, Membr. Sci. 163 (1999) 307

[31] T. Ishihara, T. Yamada, H. Arikawa, H. Nishiguchi, Y. Takita, Solid State

Ionics 135 (2000) 631

[32] C.L. Mariz, Can. Petrol. Tech. 37 (1998) 42

[33] B. Buhre, L. Elliott, C. Sheng, P. Gupta, T. Wall, Prog. Energy Combust. Sci. 31(2005) 283

[34] M. Kanniche, R. Grosbonnivard, P. Jaud, J. Valle-Marcos, J.M. Amann, C. Bouallou, C. Appl. Therm. Eng. 30 (2009) 53

109

[35] J. Ciferno, et al. DOE/NETL-2007/1291 2007

[36] J. Figueroa, T. Fout, S. Plasynski, H. McIlvried, R. Srivastava, I. J. of

Greenhouse Gas Control 2 (2008) 21

[37] J. Klara, et al. DOE/NETL-2008/1307 2007

[38] M. Rydén, A. Lyngfelt, T. Mattisson, D. Chen, A. Holmen, E. Bjørgum, I. J.

Greenhouse Gas Control 2 (2008) 21

[39] A. Abad, T. Mattisson, A. Lyngfelt, M. Rydén, Fuel 85 (2006)1174

[40] J. Readman, A. Olafsen, Y. Larring, R. Blom, Mater. Chem. 15 (2005) 1937

[41] E. Baur, H. Preis, Elektrochem. 43 (1937) 727

[42] A. Appleby, F. Foulkes, Fuel Cell Handbook, Van Nostrand Reinhold, 1989

[43] S. Singal, Solid State Ionics, 135 (2000) 305

[44] O. Yamamoto, Y. Takeda, R. Kanno, T. Kojima, Solid State Ionics 22 (1987) 241

[45] K. Huang, M. Feng, J. Goodenough, M. Schmerling, Electrochem. 143 (1996) 3630

[46] S. Adler, Chem. Rev. 104 (2004) 10

[47] K. Chang, B. Ingram, B. Kavaipatti, B. Yildiz, D. Hennessy, P. Salvador, N. Leyarovska, H. You, Mat. Res. Soc. Symp. Prec. 1126 (2009) 27

[48] V. Dusastre, J.A. Kilner, Solid State Ionics 126 (1999) 163

[49] J. Yan, H. Matsumoto, M. Enoki, T. Ishihara, Electrochem. Solid State 8 (2005) A389

[50] A. Esquirol, N. Brandon, J. Kilner, M. Mogensen, Electrochem. 151 (2004) A1847

[51] M. Rozumek, P. Majewski, T. Maldener, F. Aldinger Mater. Werkst. 33 (2002) 348

[52] T. Jardiel, M. Caldes, F. Moser, J. Hamon, G. Gauthier, O. Joubert, Solid State

Ionics 181 (2010) 894

[53] S. Li, Z. Lü, X. Huang, W. Su, Solid State Ionics, 178 (2008) 1853

110

[54] T. Suzuki, A. Hasan, Y. Funahashi, T. Yamaguchi, Y. Fujishiro, M. Awano, Science 325 (2009) 852

[55] N. Izu, W. Shin, I. Matsubara, T. Itoh, M. Nishibori, N. Murayama, Ceram.

Soc. Jap. 118 (2010) 175

[56] Y. Shimizu, Y. Fukuyama, H. Arai, T. Seiyama, ACS Symp. Series 309 (1987) 83

[57] M. Post, B. Sanders, P. Kennepohl, Sens. Act. B: Chem. 13 (1993) 272

[58] J. Ramirez-Salgado, P. Fabry, Sens. Act. B: Chem. 82 (2002) 34

[59] N. Izu, N. Oh-hori, W. Shin, I. Matsubara, N. Murayama, M. Itou, Sens. Act.

B: Chem. 130 (2008) 105

[60] J. Kašpar, P. Fornasiero, N. Hickey, Catal. Today 77 (2003) 419

[61] T. Kodama, N. Gokon, Chem. Rev. 107 (2007) 4048

[62] Z. Xu, Z. Qi, A. Kaufman, Power Sources 115 (2003) 40

[63] B. Sakakini, Y. Taufig-Yap, K. Waugh, J. Catal. 189 (200) 253

[64] S. Pei, M. Kleefisch, T. Kobylinski, J. Faber, C. Udovich, V. Zhang-McCoy, B. Dabrowski, U. Balachandran, R. Mieville, R. Poeppel, Cata. Lett. 30 (1995) 201

[65] C. Larson and R.B. Von Dreele, Los Alamos National Laboratory Reports LAUR 86-748, 1994.

[66] H.L. Yakel, W. Koehler, E. Bertaut, E. Forrat, Acta. Cryst. 16 (1962) 957

[67] J.S. Zhou, J.B. Goodenough, Phys. Rev. B 68 (2003) 144406

[68] J.B. Goodenough, Phys. Rev. 100 (1955) 564

[69] B. Dabrowski, S. Kolesnik, A. Baszczuk, O. Chmaissem, T. Maxwell, J. Mais, Solid State Chem. 178 (2005) 629

[70] T. Katsufuji, M. Masaki, A. Machida, M. Moritomo, K. Kato, E. Nishibori, M. Takata, M. Sakata, K. Ohoyama, K. Kitazawa, H. Takagi, Phys. Rev. B 66 (2002) 134434

[71] J.S. Zhou, J.B. Goodenough, J.M. Gallardo-Amores, E. Morán, M.A. Alario-Franco, R. Caudillo, Phys. Rev. B 74 (2006) 014422

111

[72] M. Fiebig, T Lottermoser, R.V. Pisarev, Appl. Phys. 93 (2003) 8194

[73] O.P. Vajik, M. Kenzelmann, J.W. Lynn, S.B. Kim, S.W. Cheong, Phys. Rev.

Lett. 94 (2005) 087601

[74] Th. Lonkai, D.G. Tomuta, U. Amann, J. Ihringer, R.W.A. Hendrikx, D.M. Többens, J.A. Mydosh, Phys. Rev. B 69 (2004) 134108

[75] I. Jeong, N. Hur, T. Proffen, App. Cryst. 40 (2007) 730

[76] C.N.R. Rao, C.R. Serrao, Mater. Chem. 17 (2007) 4931

[77] W.S. Choi, D.G. Kim, S.S.A. Seo, S.J. Moon, D. Lee, J.H. Lee, H.S. Lee, D.Y. Cho, Y.S. Lee, P. Murugavel, J. Yu, T.W. Noh, Phys. Rev. B 77 (2008) 045137

[78] S. Harikrishnan, S. Rößler, C.M.Naveen Kumar, H.L. Bhat, U.K. Rößler, S. Wirth, F. Steglich, S. Elizabeth, J. of Phys: Condens. Matter 21 (2009) 096002

[79] S. Nandi, A. Kreyssig, J.Q. Yan, M.D. Vannette, J.C. Lang, L. Tan, J.W. Kim, R. Prozorov, T.A. Lograsso, R.J. McQueeney, A.I. Goldman, Phys. Rev. B 78 (2008) 075118

[80] V.Y. Ivanov, A.A. Mukhin, A.S. Prokhorov, A.M. Balbashov, L.D. Iskhakova, Phys. of Solid State 48 (2006) 1726

[81] N. Kamegashira, H. Satoh, S. Ashizuka, Mater. Sci. Forum 449 (2004) 1045

[82] R.D. Shannon, Acta. Cryst. A 32 (1976) 751

[83] H.L. Yakel, Acta. Cryst. 8 (1955) 394

[84] H.L Yakel, W.C. Koehler, E.F. Bertaut, E.F. Forrat, Acta. Cryst. 16 (1963) 957

[85] M.A. Subramanian, C.C. Torardi, D.C. Johnson, J. Pannetier, A.W. Sleight, Solid State Chem. 72 (1988) 24

[86] N. Kamegashira, H. Satoh, S. Ashizuka, Mater. Sci. Forum 449 (2004) 1045

[87] J. Park, J.G. Park, G.S. Jeon, H.Y. Choi, C. Lee, W. Jo, R. Bewley. K.A. McEwen, T.G. Perring, Phys. Rev. B 68 (2003) 104426

[88] S. Lee, A. Pirogov, J.H. Han, J.G. Park, A. Hoshikawa, T. Kamiyama, Phys.

Rev. B 71 (2005) 180413(R)

[89] O. Carp, L. Patron, A. Ianculescu, J. Pasuk, R. Olar, Alloys and Comp. 351 (2003) 314

112

[90] G. Szabo, R.A. Paris, Seances Academy Sci. C 268 (1969) 517

[91] H.W. Brinks, H. Fjellvág, A. Kjekshus, Solid State Chem. 129 (1997) 334

[92] B. Dabrowski, O. Chmaissem, J. Mais, S. Kolesnik, J.D. Jorgensen, S. Short, Solid State Chem. 170 (2003) 154

[93] L. Suescun, B. Dabrowski, J. Mais, S. Remsen, J.W. Richardson Jr., E.R. Maxey, J.D. Jorgensen, Chem. Mater. 4 (2008) 1636

[94] A. Waintal, J. Chenavas, Mater. Res. Bull. 2 (1967) 819

[95] M. Tachibana, T. Shimoyama, H. Kawaji, T. Atake, E. Takayama-Muromachi, Phys. Rev. B 75 (2007) 144425

[96] K. Uusi-Esko, J. Malm, N. Imamura, H. Yanauchi, M. Karppinen, Mat. Chem.

Phys. 112 (2008) 1029

[97] C. Nivot, J. Bernard, C. Lelievre, J. Haussonne, D. Houivet, Ceram. Internat. 36 (2010) 929

[98] S. Samal, W. Green, S. Lofland, K. Ramanujachary, D. Das, A. Ganguli, Solid

State Chem. 181 (2008) 61

[99] D. Gutierrez, O. Peña, P. Duran, C. Moure, Euro. Ceram. Soc. 22 (2002) 1257

[100] D. Gutierrez, O. Peña, P. Duran, C. Moure, Euro. Ceram. Soc. 22 (2002) 567

[101] O. Yamaguchi, H. Takemura, M. Yamashita, Electrochem. Soc. 138 (1991) 1492

[102] J. Park, M. Kang, J. Kim, Phys. Rev. B 79 (2009) 064417

[103] K. Kamata, T. Nakajima, T. Nakamura, Mater. Res. Bull. 14 (1979) 1007

[104] B. Dabrowski, P.W. Klamut, Z. Bukowski, R. Dybzinski, J.E. Siewenie, Solid

State Chem. 144 (1999) 461

[105] Z. Bukowski, B. Dabrowski, J. Mais, P.W. Klamut, S. Kolesnik, O. Chmaissem, App. Phys. 9 (2000) 5031

[106] I.O. Troyanchuk, A.I. Akimov, L.A. Bliznjuk, N.V. Kasper, Alloys and Comp.

228 (1995) 83

[107] J. Rodríguez-Carvajal, M. Hennion, F. Moussa, A. H. Moudden, L. Pinsard, A. Revcolevschi, Phys. Rev. B 57 (1998) R3189

113

[108] H.D. Zhou, J.C. Denyszyn, J.B. Goodenough, Phys. Rev. B 72 (2005) 224401

[109] X. Chen, J. Yu, S.B. Adler, Chem. Mater. 17 (2005) 4537

[110] S. Miyoshi, J. Hong, K. Yashiro, A. Kaimai, Y. Nigara, K. Kawamura, T. Kawada, J. Mizusaki, Solid State Ionics, 161 (2003) 209

[111] S. McIntosh, J.F. Vente, W.G. Haije, D. Blank, H. Bouwmeester, Chem.

Mater. 18 (2006) 2187

[112] S. Pei, M. Kleefisch, T. Kobylinski, J. Faber, C. Udovich, V. Zhang-McCoy, B. Dabrowski, U. Balachandran, R. Mieville, R.Poeppel, Catal. Lett. 30 (1994) 201

[113] A. Atkinson, T. Ramos, Solid State Ionics 129 (2000) 259

[114] J. Van Roosmalen, E. Cordfunke, R. Helmhodt. H. Zandbergen, Solid State

Chem. 110 (1994) 100

[115] B. Fu, W. Huebner, Mater. Res. 9 (1994) 2645

[116] C. Moure, M. Villegas, J. Fernandez, J. Tartaj, P. Duran, Mat. Sci. 34 (1999) 2565

[117] G. Subba Rao, B. Wanklyn, C. Rao, Phys. Chem. Solids 32 (1971) 345

[118] T. Inoue, N. Seki, K. Eguchi, H. Arai, Electrochem. Soc. 137 (1990) 2523

[119] J. Kao, O. Anderson, D. Spartin, Solid State Chem. 87 (1990)

[120] A. Fossdal, M. Einarsrud, T. Grande, Euro. Ceram. Soc. 25 (2005) 927

[121] M. Patrakeev, J. Bahteeva, E. Mitberg, I. Leonidov, V. Kozhevnikov, K. Poepelmeier, Solid State Chem. 172 (2003) 219

[122] R. Raccah, J Goodenough, Phys. Rev. 155 (1967) 932

[123] A. Mineshige, M. Inaba, T. Yao, Z. Ogumi, K. Kikuchi, M. Kawase, Solid

State Chem. 121 (1996) 423

[124] M. Senarís-Rodríguez, J. Goodenough, Solid State Chem. 116 (1995) 224

[125] W.H Jung, Physica B 299 (2001) 120

[126] A. Hammouche, E. Schouler, M. Henault, Solid State Ionics 28 (1988) 1205

114

[127] O. Yamamoto, Y. Takeda, R. Kanno, M. Noda, Solid State Ionics 22 (1987) 241

[128] F. Tietz, I. Raj, M. Zahid, A. Mai, D. Stöver, Prog. Solid State Chem. 35 (2007) 539

[129] D. Huang, Q. XU, F. Zhang, W. Chen, H. Liu, J. Zhou, Wuhan Uni. Tech.

Mater. Sci. 23 (2007) 80

[130] L. Tai, M. Nasrallah, H. Anderson, Solid State Ionics, 76 (1995) 259

[131] T. Nakamura, G. Petzow, L. GAukler, Mat. Res. Bull. 14 (1979) 649

[132] Y. Teraoka, H. Zhang, N. Yamazoe, Chem. Letters (1985) 1743

[133] Y. Teraoka, T. Nobunaga, N. Yamazoe, Chem. Letters (1988) 503

[134] B. Dabrowski, J. Greedan, M. Greenblat, C. Grey, A. Jacobson, D. Keszleer, J. Li, M. Subramanian, Y. Xia, Prog. Solid State Chem. 36 (2007) 1

[135] M. Ruiz-Gonzalez, R. Cortes-Gil, J. Alonso, A. Hernando, M. Vallet-Regi, J. Gonzalez-Calbet, Chem. Mater. 18 (2006) 5756

[136] V. Caignaert, N. Nguyen, M. Hervieu, B. Raveau, Mater. Res. Bull. 20 (1985) 479

[137] P. Gallaher, J. MacChesney, D Buchanan, Chem. Phys. 41 (1964) 2429

[138] J. Goodenough, J. Ruiz-Diaz, Y. Zhen, Solid State Ionics, 44 (1995) 21

[139] J. Hodges, S. Short, J. Jorgensen, X. Xiong, B. Dabrowski, S. Mini, C. Kimball, Solid State Chem. 151 (2000) 190

[140] P. Battle, T. Gibb, P. Lightfoot, Solid State Chem. 76 (1998) 334

[141] O. Hansteen, H. Fjellvag, B Hauback, Solid State Chem. 141 (1998) 411

[142] P. Casey, D. Barker, M. Hayward, Solid State Chem. 179 (2006) 1375

[143] K. Poeppelmeier, M. Leonowicz, J. Scanlon, J. Longo,W.B. Yelon, Solid State

Chem. 45 (1982) 71

[144] T. Mori, N. Kamegashira, Alloys and Comp. 408–412 (2006) 1210

[145] L. Tai, M. Nasrallah, H. Anderson, Solid State Ionics, 76 (1995) 273

115

[146] S. Balamurugan, K. Yamaura, A. Karki, D. Young, M. Aria, E. Takayama-Muromachi, Phys. Rev. B 74 (2006) 172406

[147] Y. Teraoka, H. Zhang, S. Furukawa, N. Yamazoe, Chem Letters Chem. Soc.

Jap. (1985) 1743

[148] L. Suescun, B. Dabrowski, S. Remsen, J. Mais, Solid State Chem. 182 (2009) 187

[149] K. Swierczek, B. Dabrowski, L. Suescun, S. Kolenik, Solid State Chem. 182 (2009) 280

[150] K. Swierczek, B. Dabrowski, L. Suescun, S. Kolenik, S. Remsen, J. Mais, unpublished data.

[151] H. Tu, Y. Takeda, N. Imanishi, O. Yamamoto, Solid State Ionics 117 (1999) 277

[152] S. Kolesnik, B. Dabrowski, J. Mais, D. Brown, R. Feng, O. Chmassiem, R. Kruk, C. Kimball, Phys. Rev B 67 (2003) 144402

[153] I. Fawcett, G. Veith, M. Greenblatt, M. Croft, I. Nowik, Solid State Sci. 2 (2000) 821

[154] C. Solís. M. Rossell, G. Garcia, A. Figueras, G. Tendeloo, J. Santiso, Solid

State Ionics 179 (2008) 1996

[155] M. Senarís-Rodríguez and J. Goodenough, Solid State Chem. 116 (1995) 224

[156] T. Takami, J. Zhou, J. Goodenough, H. Ikuta, Comb. Sci. Tech. 19 (1981) 124

[157] J. Burley, J. Mitchell, S. Short, Phys. Rev. B 69 (2004) 054401

[158] S. Kolesnik, B. Dabrowski, J. Mais, M. Maijiga, O. Chmaissem, Phys. Rev. B 73 (2006) 214440

[159] B. Stillwell, “Evaluation of Potential Cathode Materials for Reduced-

Temperature Solid Oxide Fuel Cells,” Northern Illinois University, Dept. of Physics M.S. Thesis (2007)

[160] B. Steele, A. Heinzel, Nature 414 (2001) 345

[161] A. Barbucci, R. Bozzo, G. Cerisola, P. Costamagna, Electrochimica Acta 47 (2002) 2183

[162] S. Jiang, Solid State Ionics 146 (2002) 1

116

[163] A. Mai, V. Haanappel, F. Tietz, I. Vinke, D. Stöver, Electrochem. Soc. Proc. 7 (2003) 525

[164] D. Carter, B. Dabrowski, B. Stillwell, unpublished data.

[165] H. Hayashi, T. Saitou, N. Maruyama, H. Inaba, K. Kawamura, M. Mori, Solid

State Ionics 176 (2005) 613

[166] H. Hayashi, M. Kanoh, C. Quan, H. Inaba, S. Wang, M. Dokiya, H. Tagawa, Solid State Ionics 132 (2000) 227

117

APPENDIX: STRUCTURAL PARAMETERS AND AGREEMENT FACTORS

FOR Dy1-xYxMnO3+δ

118