+ All Categories
Home > Documents > Molly Gentleman Thesis 2006

Molly Gentleman Thesis 2006

Date post: 13-Apr-2017
Category:
Upload: molly-m-gentleman
View: 491 times
Download: 8 times
Share this document with a friend
147
UNIVERSITY OF CALIFORNIA Santa Barbara High Temperature Sensing of Thermal Barrier Materials by Luminescence A Dissertation submitted in partial satisfaction of the Requirements for the degree Doctor of Philosophy In Materials By Molly Maureen Gentleman Committee in charge: Professor David R. Clarke, Chair Professor Ted Bennett Professor Anthony K. Cheetham Professor Anthony G. Evans Professor Carlos G. Levi September 2006
Transcript
Page 1: Molly Gentleman Thesis 2006

UNIVERSITY OF CALIFORNIA

Santa Barbara

High Temperature Sensing of Thermal Barrier Materials by Luminescence

A Dissertation submitted in partial satisfaction of the

Requirements for the degree Doctor of Philosophy

In Materials

By

Molly Maureen Gentleman

Committee in charge:

Professor David R. Clarke, Chair

Professor Ted Bennett

Professor Anthony K. Cheetham

Professor Anthony G. Evans

Professor Carlos G. Levi

September 2006

Page 2: Molly Gentleman Thesis 2006

ii

The dissertation of Molly Maureen Gentleman is approved.

___________________________________________ Ted Bennett

___________________________________________ Anthony K. Cheetham

___________________________________________ Anthony G. Evans

___________________________________________ Carlos G. Levi

___________________________________________ David R. Clarke, committee chair

August 2006

Page 3: Molly Gentleman Thesis 2006

iii

ACKNOWLEDGEMENTS

It is a privilege to acknowledge the many people who have contributed to

my graduate career.

First, I would like to thank my sister Eileen. She and I spent many years

hammering away in our Grandpa’s basement confident that we could construct

radios out of the junk that was lying around his workspace. Upon graduating from

junior-high, she embarked on a new academic journey that shaped both of our lives.

Eileen attended the Illinois Mathematics and Science Academy and soon learned

that advanced education was truly accessible to anyone regardless of their

background. Upon completion of high school, she went off to Tulane to pursue a

degree in biomedical engineering and her plan to get me to do the same. During

my senior year of high school, she demanded that I apply for college and get a

degree in engineering. It sounded like a good idea to me, and so I did.

Eileen graduated from Tulane summa cum laude as the first woman in my

family to graduate from college and went on to pursue a masters and doctorate

degree in her field. As I approached my college graduation, she again insisted that

I go to graduate school (only really good ones) and get an advanced degree. That

brings me here. Over the past five years, Eileen has been my primary source of

support for everything from the best way to approach an experiment to when and

where to meet her for vacation. I could not have made it here without her.

Page 4: Molly Gentleman Thesis 2006

iv

I must also thank the many friends I have met here in Santa Barbara. The

Clarkers are a particularly wonderful group of people who I feel privileged to have

met and worked with. I would like to thank Vladimir Tolpygo specifically as he

has contributed to my development as a researcher more than any other member of

the group has. There has also been a large number of staff here at the university

who I would have been pained to do without. In particular, I would like to thank

Deryck Stave who has provided technical experience support that I have not seen

matched elsewhere.

Finally, I would like to thank my advisor David Clarke and the rest of my

committee. All of them have been a great source of technical and scientific support

over the last five years.

Page 5: Molly Gentleman Thesis 2006

v

MOLLY M. GENTLEMAN

Materials Department Bldg. 503 Rm 1355 University of California, Santa Barbara Santa Barbara, CA 93106-5050 (805) 893-7509 [email protected] EDUCATION

Ph.D., Materials, University of California, Santa Barbara. Degree expected Summer 2006. Dissertation: High Temperature Sensing of Thermal Barrier Materials by Luminescence Advisor: Dr. D.R. Clarke

Bachelor of Science, Metallurgical and Materials Engineering with honors, Illinois Institute of Technology, August 1998-May 2001.

AREAS OF SPECIALIZATION

Luminescence spectroscopy, optical characterization of materials, high temperature luminescence, phase characterization, Raman spectroscopy, thermal barrier coatings.

RESEARCH EXPERIENCE

Doctoral Research, UCSB, 08/2001-present. Conceived and developed novel temperature sensing technique for

the measurement of temperature of embedded layers within a thermal barrier coating. Analysis of crystal structure/luminescence relationship as well as concentration/luminescence relationship for yttria stabilized zirconia and pyrochlore zirconate materials.

Dr. D.R. Clarke, Materials Department

Research Assistant, Argonne National Laboratory, 03/2000-08/2001. Evaluation of the relationship between acoustic emission and

mechanical measurements of elastic modulus for ceramic matrix composites with a focus on development of a non-destructive technique for measuring accumulated damage. Dr. W.A. Ellingson, Energy Technology Division.

TEACHING EXPERIENCE

Teaching Assistant, Introduction to materials, UCSB 01/2002-03/2002. Prepared questions and homework solutions. Prepared and lead

recitation lecture weekly.

Page 6: Molly Gentleman Thesis 2006

vi

PUBLICATIONS Gentleman MM, Eldridge JI, Zhu DM, Murphy KS, and Clarke DR. In Press. Non-contact sensing of TBC/BC interface temperature in a thermal gradient. Surf. Coat. Technol. Gentleman MM, Lughi V, Nychka JA, and Clarke DR. 2006. Non-contact methods for measuring thermal barrier coating temperatures. Int. J. Appl. Ceram. Technol. 3 [2] 105-112. Gentleman MM and Clarke DR. 2005. Luminescence sensing of temperature in pyrochlore zirconate materials for thermal barrier coatings. Surf. Coat. Technol. 200: 1264-1269. Gentleman MM and Clarke DR. 2004. Concepts for luminescence sensing of thermal barrier coatings. Surf. Coat. Technol. 188-189: 93-100. Clarke DR, Tolpygo VK, Gentleman M. 2004. Luminescence-based characterization of protective oxides: from failure mechanisms to non-destructive evaluation. Mat. Sci. Forum 461-464: 621-629.

ABSTRACTS AND PRESENTATIONS Gentleman MM, Eldridge JI, Zhu DM, Murphy KS, and Clarke DR. 05/2006. Non-contact sensing of TBC/BC interface temperature in a thermal gradient. International conference on metallurgical coatings and thin films, San Diego, CA. Gentleman MM, Eldridge JI, Zhu DM, Clarke DR, and Murphy KS. 01/2006. Novel, direct sensing of the TBC/BC interface temperature in a thermal gradient (poster). MURI/TBC meeting, Santa Barbara, CA. Gentleman MM, Eldridge JI, and Clarke DR. 05/2005. Luminescence sensing of temperature in pyrochlore zirconate materials for thermal barrier coatings. International conference on metallurgical coatings and thin films, San Diego, CA. Gentleman MM and Clarke DR. 01/2005. Non-contact temperature measurements by Eu luminescence (poster). MURI/TBC meeting Santa Barbara, CA. Gentleman MM and Clarke DR. 12/2004. Non-contact temperature measurements by luminescence. International symposium of research students on materials science and engineering, Madras, Chennai, India.

Page 7: Molly Gentleman Thesis 2006

vii

Gentleman, MM and Clarke, DR. 04/2004. Concepts for luminescence sensing of thermal barrier coatings. International conference on metallurgical coatings and thin films, San Diego, CA. Gentleman M, Sandberg D, Deemer C, Ellingson WA, and Todd J. 01/2002. Correlation of destructive and non-destructive elastic modulus measurement methods in ceramic composites. The American Ceramic Society Conference and Exposition on Advanced Ceramics and Composites, Cocoa Beach, FL.

HONORS Recipient, Madras Metallurgical Services, Best Technical Presentation Prize, International symposium of research students on materials science and engineering, Madras, Chennai, India. 12/2004. Recipient, Bunshah Award for Best Paper in Symposium A, International conference on metallurgical coatings and thin films, San Diego, CA. 04/2004. Recipient, NSF IGERT fellowship on Advanced Optical Materials. 12/2001-12/2003. Recipient, The Engineering Ceramics Division, student technical presentation First Place, The American Ceramic Society Conference and Exposition on Advanced Ceramics and Composites, Cocoa Beach, FL. 01/2002. Recipient, CAMRAS/NEXT full-tuition scholarship, Illinois Institute of Technology, Chicago, IL. 08/1998-05/2001.

Page 8: Molly Gentleman Thesis 2006

viii

ABSTRACT

The basis for a number of concepts for in situ monitoring of thermal barrier

coating heath and temperature using non-contact luminescence techniques is

described. Restrictions imposed by phase compatibility and the optical properties

of the coating materials are discussed as guidelines for the selection of rare-earth

ions as chromophore dopants in these sensors. Extensive studies of type and

concentration of rare-earth doping lead to the selection of Eu-doped materials as the

preferred material for temperature sensing. Sensor layers, consisting of 1 atomic

percent europia-doped yttria stabilized zirconia were deposited and exposed to high

heat flux conditions similar to those seen today’s turbine environments and were

shown to be capable of measuring temperature at different depths within the

coating.

Page 9: Molly Gentleman Thesis 2006

ix

TABLE OF CONTENTS

Foreword…………………………………………………………………………...1

1. Chapter 1: Concepts for Luminescence Sensing of…………………………..3

Thermal Barrier Materials

1.1. Introduction

1.2. Thermal barrier coating systems

1.3. Sensing concepts

1.4. Measuring temperature using luminescence

1.5. Phase compatibility of sensors

1.6. Selection of luminescent dopants

1.7. Demonstration of sensor concepts

1.8. Summary

2. Chapter 2: Experimental Investigation……………………………………38

2.1. Introduction

2.2. Materials characterization

2.2.1. Materials processing

2.2.2. Raman spectroscopy

2.2.3. X-ray diffraction

2.3. Luminescence characterization

2.3.1. Room temperature luminescence characterization

2.3.2. High temperature luminescence characterization

2.3.3. Amplification and electronics

3. Chapter 3: Luminescence sensing of temperature in………………………54

pyrochlore zirconate materials

3.1. Introduction

3.2. Materials

3.3. Room temperature luminescence

Page 10: Molly Gentleman Thesis 2006

x

3.4. High temperature luminescence

3.5. Discussion

4. Chapter 4: Concentration Effect on Luminescence in……………………74

Eu-doped Yttria Stabilized Zirconia Intensity

4.1. Introduction

4.2. Luminescence spectra

4.3. Luminescence lifetime

4.4. Experimental procedure

4.5. Results and discussion

4.6. Summary

5. Chapter 5: Luminescence Thermometry of (Gd,Eu)2Zr2O7……………....95

5.1. Introduction

5.2. Experimental

5.3. Room temperature luminescence

5.4. High temperature luminescence

5.5. Discussion

5.6. Summary

6. Chapter 6: Non-contact Sensing TBC/BC Interface……………………...116

Temperature in a Thermal Gradient

6.1. Introduction

6.2.

6.3. Sensor layer deposition

6.4. Luminescence lifetime calibration

6.5. Luminescence lifetime measurement in a thermal gradient

6.6. Results

6.7. Discussion

7. Conclusions and Future Work……………………………………………134

Page 11: Molly Gentleman Thesis 2006

1

Foreword

The goal of this thesis is the development of non-contact sensors capable of

measuring the temperature at the thermal barrier coating, near its interface with the

bond coat at temperatures applicable to current turbine systems. The requisite of

“prime reliance” on today’s engines is discussed as the driving force for the

development of non-contact heath monitoring tools for thermal barrier coatings.

To this end, luminescence sensing of temperature and damage is proposed as one

such non-contact technique. Chapter 1 begins with the introduction, description,

and ultimate demonstration of two basic sensor concepts, “visualization,” and

“measurement” sensors. Here the use of the optical properties and phase stability

of common thermal barrier materials (yttria stabilized zirconia and Gd2Zr2O7) as

well as the thermally grown oxide, another vital component in thermal barrier

systems, are discussed as selection criteria for the development of embedded sensor

materials. The choice of rare-earth ions as chromophore dopants is described and

the technique of luminescence lifetime decay for non-contact temperature

measurements is illustrated.

Chapter 2 departs from sensor development to discuss, in detail, the

development of the experimental apparatus used in this work and all the techniques

implemented in processing and analysis of the materials discussed. Several of the

most common problems encountered by others working in this area are reviewed

and their complications resolved. The third chapter discusses the use of Eu3+ as a

dopant ion in several thermal barrier coating hosts. Issues involving the selection

Page 12: Molly Gentleman Thesis 2006

2

of excitation energy for these studies and the differences seen in the luminescence

lifetime for a series of materials with slightly different hosts, but the same doping

conditions are discussed. The work done in this chapter provides a basis for the

studies done subsequently in chapters 4 and 5.

Chapters 4 and 5 are an in-depth assessment of the effects that dopant and

defect concentrations have on the temperature sensing capabilities of the Eu-doped

yttria stabilized zirconia as well as (Eu,Gd)2Zr2O7. In these chapters, the changes

in luminescence lifetime are attributed to Eu3+ site symmetry and the extent of

disorder and chromophore ion-ion interaction in the two systems. Both of these

chapters provide a basis of composition selection for temperature sensors.

Finally, chapter 6 describes the implementation of a sensor based on the

work of the previous five chapters. In this chapter, a temperature sensor, of

optimum composition (determined in chapter four), was deposited by electron

beam physical vapor deposition. Temperature calibrations were done using the

techniques described in chapter 2. Following deposition and calibration, the sensor

was then exposed to a thermal gradient similar to that seen in current turbine

environments and, combined with pyrometry, was used to determine the thermal

conductivity and heat fluxes in the coating.

Page 13: Molly Gentleman Thesis 2006

3

Chapter 1 Concepts for Luminescence Sensing of Thermal Barrier

Materials

1.1 Introduction

The use of ceramic coatings for thermal insulation of metal blades and hot

sections of aerospace and power generation turbines has become integral in modern

engine design. These coatings, commonly known as thermal barrier coatings,

enable metallic engine components to run in gases whose temperatures are often

above the melting temperature of the superalloys, resulting in greater engine

efficiency [1]. One result of increasing turbine temperatures is a risk of

catastrophic coating failure leading to substrate overheating and potentially failure

of the engine. Consequently, “prime reliance” of coatings has become an industry

standargoal for these systems. The ability to predict the remaining life of a coating

has hence become a sort of “holy grail,” allowing for component replacement to be

based on actual damage accumulation instead of scheduled replacement.

Monitoring of the metal temperature beneath the coating is a particularly promising

means to predict the life of the coating because there is an exponential relationship

between the metal temperature and the evolution of damage that leads to failure of

the coating. In this chapter, a novel technique for monitoring this temperature, as

well as damage in the coating, will be presented along with the selection criteria for

material schemes for developing such sensors.

Page 14: Molly Gentleman Thesis 2006

4

1.2 Thermal barrier coating systems

Thermal barrier coatings are multilayer systems (Figure 1.2.1) that consist

of a thin overcoat of low thermal conductivity ceramic on top of a thin layer of

metallic “bond-coat” formed on a superalloy component[2-4]. During exposure to

high temperatures in air (and in typical turbine combustion environments), a thin-

layer of aluminum oxide forms at the interface between the “bond-coat” and the

thermal barrier coating. The composition of the “bond-coat” is chosen to form

aluminum oxide, commonly called the thermally grown oxide (TGO), upon

oxidation since aluminum oxide has the slowest oxygen diffusion rate of any oxide

and hence leads to the lowest oxidation rate. The most common material presently

used as a thermal barrier coating is yttria-stabilized zirconia (YSZ) but there is

some use of the rare-earth zirconates, such as gadolinium zirconate, Gd2Zr2O7,

(GZO) on account of their lower thermal conductivity at high temperature [4, 5].

Pertinent to optical sensing applications, these coatings are typically deposited by

either atmosphere plasma-spraying (APS) or electron-beam physical vapor

deposition (EB-PVD). The former results in the formation of “splat” boundaries

parallel to the coating that tend to scatter light and decrease optical transmission.

The EB-PVD coatings have a columnar microstructure, which tends to cause

preferential channeling of light into and out of the coating and scattering of light

perpendicular to the direction of coating growth. Both methods of deposition also

result in the coatings having microscopic porosity, a microstructural feature

desirable for further decreasing the low thermal conductivity of these coatings

Page 15: Molly Gentleman Thesis 2006

5

below that of even the bulk material. It is important to note that YSZ coatings have

the metastable tetragonal prime crystal structure [4] and are quite distinct from

fully stabilized cubic zirconia (“synthetic diamond”) used in applications such as

optical windows and substrates for thin-film growth[6]. Their composition is also

distinct, usually 7-8 w/o Y2O3 as opposed to the 20 w/o of the cubic single crystals.

Both YSZ and GZO coatings are translucent in the visible due to extensive

entrapped porosity. The optical band-gap of YSZ is reported to be ~5 eV[7]. The

optical absorption spectrum of GZO is unknown, but is believed to be similar to

YSZ and other complex oxides, with band-gap absorption in the mid-ultraviolet

and absorption in the mid-infrared. Although the far-infrared (6 – 25 µm) optical

properties of tetragonal-prime yttria-stabilized zirconia have been reported[8] as a

function of temperature, the optical absorption properties of both materials in the

visible and near-infrared at elevated temperatures have not been reported. The

presence of porosity within the coating causes optical scattering, particularly in

plasma spray coatings. The frequency dependence of the optical absorption and

optical scattering are shown for a typical YSZ coating in Figure 1.2.2[9]. As will

be demonstrated, the difference in absorption in the visible and the UV, illustrated

in Figure 1.2.2, provides an opportunity to use lasers of different frequencies to

probe different depths into YSZ coatings for a variety of sensor applications.

Also pertinent to the issue of sensing is that the primary modes of failure in

a TBCs are spallation and delamination of the coating[10]. In both modes, the

coating separates from the component along a failure path within the ceramic

Page 16: Molly Gentleman Thesis 2006

6

coating, typically close to the interface with the thermally grown oxide. (In large

thermal gradients, the failure path can shift to different planes in the coating.)

Failure thus causes a local reduction in the thickness of the TBC with the

consequence that the temperature of the metal under the spalled region increases

because of the lower thermal resistance afforded by the thinner coating.

1.3 Sensing Concepts

Several concepts for monitoring the “health” of thermal barrier coatings can

be envisaged. These range from the possibility of visually assessing damage and

wear to in situ measurement of coating temperatures. Broadly speaking the

concepts can be classified as being either “visualization” sensors or “measurement”

sensors.

Possibly the simplest “visualization” sensor is as an inspection method at

room temperature, for instance standard boroscope inspection during periodic

maintenance, to assesses whether the coating has locally spalled away. In cases

where the spalled regions are very large, they can be seen visually. This is the

basis for boroscope inspection today. However, when the spalled regions are

smaller, the concept of a “red-line” sensor facilitates greater contrast. The concept

is shown in Figure 1.3.1 and consists of an inner layer of the coating doped with

chromophore adjacent to the bond-coat. Where the coating is intact, the doped

region is buried and so can only be probed with a laser wavelength to which the

coating is transparent. Where the coating has spalled away or has been eroded, the

Page 17: Molly Gentleman Thesis 2006

7

doped layer is exposed and can be illuminated with a laser in the UV as well as in

the visible.

An elaboration of the “red-line” sensor is what might be termed a

“rainbow” sensor, shown schematically in Figure 1.3.2. It consists of a series of

layers, each with a different dopant that luminescence at a different frequency.

Under optical excitation in the visible, characteristic luminescence from each of the

layers can be collected. As the thermal barrier coating erodes away, successive

doped layers will be removed and the visible luminescence will characterize the

remaining layers, enabling the remaining thickness of the coating to be assessed.

Alternatively, using UV excitation above the optical band gap of the coating, it

should be possible to probe the outermost layer in the “rainbow” sensor exclusively

and thereby directly monitor the progress of erosion.

Both sensor configurations also form the basis of temperature measurement

using dopants whose luminescence characteristics are temperature-dependent. The

“red-line” sensor configuration then provides a means of monitoring the

temperature of the coating in contact with the bond-coat. More ambitious, and also

highly desirable, is the measurement of the temperature gradient through the

thickness of the coating by monitoring the luminescence simultaneously from

layers at both the inner and outer portions of the coating.

1.4 Measuring Temperature Using Luminescence

Traditionally, three properties of the luminescence emission of a material

have been used to measure temperature: observed shifts in peak position, relative

Page 18: Molly Gentleman Thesis 2006

8

peak intensities, and luminescence lifetime decay[11]. The first is the measurement

of observed line shifts of a single luminescent transition with temperature. It has

been noted that luminescence lines from chromophore ions show an observed blue

shift with increasing temperature [12] and that these shifts are attributed to

interactions between the electronic states and thermally activated lattice phonons.

At first glance, measuring temperature by peak shifts appears to be an appealing

method due to the small spectral range required (peak shifts are on the order of ~10

cm-1 for 0 – 300 K) but the technique becomes increasingly unreliable due to

extensive spectral broadening that occurs at temperatures much above room

temperature. Therefore, peak shift measurements are inappropriate for thermal

barrier coating applications where the temperature range of interest is well above

room temperature (in excess of 1000 °C) and spectral broadening and decreased

intensity would make any such measurement very difficult.

The second technique is the measurement of peak intensity ratios. This

technique monitors changes in the relative intensities of two distinct luminescence

peaks resulting from different excited states of the chromophore ion for a specific

concentration. Because the relative populations of the excited states are strongly

dependent on temperature, the relative intensities are also temperature dependent,

providing a basis for the use of relative intensities to measure temperature. By

measuring the relative intensities of two different luminescence lines over the

temperature range of interest and developing an intensity-ratio temperature

relationship, it is possible to derive the temperature of a material from observed

Page 19: Molly Gentleman Thesis 2006

9

luminescence. This technique has been applied to various materials systems at

temperatures approaching those applicable to TBC systems [13-16], but caution

must be used when applying the intensity ratio technique. The technique is very

sensitive to small changes in luminescence intensity due to compositional changes

and the reabsorption of emitted light. While the application of this technique is

currently being explored for use on thermal barrier systems, it remains to be seen if

small deposits of calcium-magnesium-aluminosilicate (CMAS) or other debris

commonly found on engine hardware may affect the observed spectral intensities

and invalidate the temperature measurement.

The third scheme, luminescence-lifetime decay, is the technique chosen for

temperature measurements in this work. In this scheme, the luminescence from the

dopant ions is excited with a very short pulse of light and the intensity of the

luminescence from a single excited state is recorded as a function of time after the

end of the pulse. From this measurement, the decay time can be extracted to yield

a lifetime–temperature relationship, which can subsequently be used to calibrate the

temperature based on the material’s luminescence lifetime. In the simplest cases,

the logarithm of intensity decreases linearly with time and the decay can be

identified as the slope of the decay curve (Figure 1.4.1) fit using a simple

exponential decay

( )τtII −= exp0 Equation 1.4.1

where I is the intensity of light, I0 is the intensity of light at time t=0, and τ is the

luminescence lifetime of the material. In some of the materials studied here, the

Page 20: Molly Gentleman Thesis 2006

10

luminescence decays with a more complex function, characterized by the sum of

exponential decays (Figure 1.4.1) and are fit using equation 1.4.2.

( ) ( )2211 expexp ττ tItII −+−= Equation 1.4.2

Materials used to measure temperature using the luminescence lifetime decay

technique can often be chosen to yield long lifetimes (> 1 µs) avoiding the use of

complicated high-speed electronics. The use of simple electronics, coupled with

the small spectral range required for the measurement, relative insensitivity to

thermally induced line broadening and background variations, and the techniques

limited susceptibility to uncertainties introduced by small changes in concentration

makes the luminescence lifetime technique particularly appealing for very high

temperature measurements.

The temperature sensitivity of luminescence lifetime can be described as

follows. As the temperature of a luminescing body increases, the equilibrium

populations of the excited and ground states follow a Boltzmann distribution,

( )[ ]kTEEgg

NN

abb

a

b

a −−= exp Equation 1.4.3

where N is the equilibrium population of the level, g is the dengereracy of the level,

E is the energy of the level, k is the Boltzmann constant, and a and b denote the

ground and excited states respectively. The rate of decay from the excited state, w,

is given as a sum of the rate of radiative decay for the isolated ion (wr), the rate of

non-radiative decay (wnr), and the rate of energy transfer between like ions (wET).

This can be described simply as

Page 21: Molly Gentleman Thesis 2006

11

ETnrrtotal wwww ++= . Equation 1.4.4

Where τ1=w . For dilute solutions at low temperatures, the radiative rate

dominates. As the temperature of the system is increased, phonon modes within

the host material are activated and non-radiative decay plays an increasingly

important role in the total decay rate of the material. Moos and Weber [17, 18]

describe this process in detail through the mean of a multi-phonon relaxation model

where the multiphonon rate ( )Twmp varies with temperature according to

( ) ( ) pmpmp nwTw )1(0 += Equation 1.4.5

where ( )0mow is the rate of multiphonon emission at 0 K, the phonon occupancy

factor ( )[ ] 11exp −−= kThn ν , hν is the energy of the accepting phonon mode, and p

is the number of phonons required to bridge the gap between the excited and

ground states. Following activation of the phonon modes in the host and thus

multiphonon de-excitation of the excited state, the lifetime decreases rapidly with

increasing thermal energy. An example of typical luminescence lifetime

dependence on temperature is seen in Figure 1.4.2. As the thermal energy of the

system approaches the energy gap between the ground and excited state, complete

quenching of the luminescence is observed. This effect will be explored in the

following chapters.

1.5 Phase Compatibility

As a thermal barrier coating is in direct physical contact with the thermally

grown aluminum oxide, phase compatibility of the coating material with aluminum

Page 22: Molly Gentleman Thesis 2006

12

oxide places very severe restrictions on the selection of possible luminescent

materials and their microstructural location within a coating system. Luminescent

materials that are not thermodynamically compatible with the coating will react and

decompose with inevitable degradation in luminescence and, possibly coating life.

For the two most common thermal barrier coatings materials, YSZ and GZO, the

pertinent phase compatible diagrams with aluminum oxide are shown in Figure

1.5.1. drawn from Leckie et al. [19] and Leckie, Kramer, Ruhle, and Levi [20].

The figure consists of two pseudo-ternary phase diagrams joined along their

common Al2O3-ZrO2 pseudo-binary at 1200 °C.

Of the compounds in the compatibility diagrams of Figure 1.5.1, the three

best known to be hosts for luminescence ions [21-23], yttria (Y2O3), yttrium

aluminate (YAlO3), and yttrium aluminum garnet (YAG), are seen to be

incompatible with the tetragonal-zirconia (YSZ) of thermal barrier coating and in

the case of yttria and yttrium aluminate, the thermally grown oxide, aluminum

oxide (Al2O3), as well. This indicates that none of these materials can be used as

stable luminescence hosts in direct contact with the thermally grown oxide formed

by oxidation of the bond-coat and the YSZ. If the thermal barrier coating were to

be made from the pyrochlore phase, such as Gd2Zr2O7, then it would be possible to

use the GdAlO3 phase as a host for the luminescence sensor, but the pyrochlore

itself is not compatible with alumina. The lack of stability in oxidizing

atmospheres at high temperatures obviously precludes the use of other well-known

phosphors, such as the oxy-sulphides.

Page 23: Molly Gentleman Thesis 2006

13

These phase compatibility restrictions indicate that the coating materials

themselves must be used as the hosts for the luminescing chromophore.

Fortunately, both YSZ and GZO can accommodate trivalent luminescent ions such

as the rare-earth ions (by partial substitution for the Y3+ and Gd3+ ions respectively)

because of their similar ionic size and valence [24] as will be discussed in the next

section. Indeed, there are multiple reports of the photoluminesence of a number of

rare-earth ions in YSZ [16, 25-27]. It should be noted that the phase compatibility

restrictions are less severe for sensing the outer portions of the coating since the

luminescence sensing material does not have to be compatible with alumina as well

as the coating material. With that in mind, it would be possible for the perovskite

phases, such as YAG, to be deposited on the top of the coating for temperature

sensing.

1.6 Selection of luminescent dopants

The rare-earth ions ranging from cerium to ytterbium display a partial

filling of the 4f shell. In their elemental state, the outer 5d and 5s electrons act to

shield the inner 4f electrons. This is important because the 4f electrons are those

responsible for the optical transitions in these ions. The observed energy levels for

several of the trivalent rare-earth ions in LaCl3 are seen in the Dieke diagram in

Figure 1.6.1. Shielding by the outer shells results in sharp absorption and emission

peaks as well as splitting of the free ion energy levels by much less than the

spacing between levels (denoted by the thicknesses of the energy levels in the

Dieke diagram Figure 1.6.1).

Page 24: Molly Gentleman Thesis 2006

14

As mentioned previously, the trivalent rare-earth ions have some limited

solubility in both the YSZ and GZO phases. As they are also well known to be

luminescent ions [22, 28] and are widely used in phosphors and displays, the

trivalent rare-earth ions are the natural choice for the class of dopants used in

luminescence sensing of thermal barrier coatings. For room temperature sensor

applications, such as the “red-line” sensor, there are several choices of rare-earth

dopant. For the rainbow sensors, the dopants in the different layers must be chosen

so that there is little overlap between the luminescence lines of the different

dopants. Guidance in selecting potential rare-earth ion dopants for YSZ and GZO

coatings is provided by the Dieke diagram (Figure 1.6.1) which tabulates the

luminescence from different energy levels of the indicated rare-earth ions in

lanthanum chloride [28]. This diagram indicates that for room temperature sensor

applications almost all the rare-earth ions can serve because potential luminescence

dopants as they luminesce over the entire visible spectrum and into the infrared.

For sensing applications at temperature, the selection of appropriate dopants

is more restrictive since the luminescence signal must be detectable against the

blackbody radiation from the surroundings and hot gas as well as the coating itself.

Therefore, to be detectable, the luminescence must be at wavelengths below the

high-energy edge of the blackbody curve. This is shown in Figure 1.2.2 for a

number of representative blackbody temperatures and indicates that that ions that

luminesce at wavelengths below ~650 nm are required in order to avoid being

swamped by blackbody radiation. In addition, to allow the luminescence from a

Page 25: Molly Gentleman Thesis 2006

15

buried layer to reach the surface for collection, the ion must luminesce at

wavelengths longer than the band gap of the material (~ 300 nm)., thus enabling

propagation through the coating

To measure temperature, the choice of rare-earth ions is still more

restrictive, particularly if the luminescence lifetime method is used. Throughout

the course of this work, europium has been shown to be a particularly good dopant

for temperature sensing in YSZ and GZO. The selection of europium for this study

is based on it having electron energy levels with a large energy gap between the

states that can be excited in the visible (5D) and the lower energy levels (7F), and

because the resulting luminescence emission is not only in the visible, but also at a

wavelength shorter than the edge in the black-body emission curve for temperatures

up to ~1500 °C (Figure 1.6.1). The large energy gap found in Eu3+ is desirable

because the rate of non-radiative decay by multiphonon relaxation, which competes

with radiative decay limiting the maximum temperature where luminescence is

observed, depends on the size of this gap [29] .

1.7 Concentration Effects

Once an appropriate chromophore ion has been identified, it is essential to

establish its useful range of concentration for the particular sensor application. For

visualization sensing, the most desirable doping concentration is the one that

maximized the luminescence signal intensity (subject to phase stability

considerations). On the other hand, for in situ temperature measurements, although

maximizing the luminescence intensity is clearly desireable, it is important that the

Page 26: Molly Gentleman Thesis 2006

16

selected concentration is one for which the lifetime decay varies over the

temperature range of interest.

The concentration at which the luminescence intensity is at its maximum is

usually set by the phenomenon of concentration quenching. This is illustrated by

the intensity of different luminescence lines from Sm-doped YSZ excited using an

argon-ion laser at 514 nm (Figure 1.7.1). The intensity data as a function of

concentration, normalized to the maximum intensity in each spectrum, is plotted in

Figure 1.7.2. The curves through the data correspond to a fit using the Johnson-

Williams equation [30].

( )( )cAccc z

−+−

=1

1η Equation 1.7.1

Where η is the quantum efficiency of the chromophore and c is the mole fraction of

luminescent ions. Notationally, the exponent z represents an effective coordination

number of isolated, like ions around the excited ion and the parameter A represents

a ratio of optical cross sections but, in practice, they are usually used as fitting

parameters since usually neither is known a priori. The values of z and A vary for

the different spectral transitions in Figure 1.7.2 and are listed in Table 1.7.1 and are

typically on the order of 38 and 0.008.

Also, as indicated by the data in Figure 1.7.2, the Sm-dopant concentration

at which the luminescence intensity peaks depends on the particular spectral

transition. Interestingly, there are dopants and combinations of excitation and

luminescence wavelengths over which no apparent concentration quenching occurs.

Page 27: Molly Gentleman Thesis 2006

17

This unexpected behavior occurs, for example, for the 606 nm line of Eu-doped

YSZ when excited with a 514 nm laser (Figure 1.7.3), as well as for the 654 and

678 nm lines of Er-doped YSZ, whereas the lines at 543 and 562 nm or Er-doped

YSZ do exhibit concentration quenching (Figure 1.7.4.) In the particular case of

Er-doped YSZ, the continuing increase in intensity as a function of concentration is

believed to be due to cross-relaxation because the second set of peaks (654 and 678

nm) are significantly lower in energy and intensity than the primary transitions.

This is a process where the increase in the concentration dopant ions leads to

increased energy transfer between ions and the energy transfer between like ions

results in the luminescence lines observed at 654 and 678 nm. Consequentally, the

increase in concentration causes an increase in intensity. The reason that

concentration quenching is not seen in Eu-doped YSZ has not yet been explained.

1.8 Demonstration of Sensor Concepts

In this section, the basic concepts described in the preceding sections are

illustrated. Two types of samples were used. One was a series of 7YSZ thermal

barrier coatings deposited by electron beam deposition at UCSB. These coatings

consisted of a thin (~ 10 µm) layer of Er-doped YSZ deposited on an oxidized

metal alloy and then over-coated with ~120 µm of undoped YSZ to make a

standard thickness YSZ coating. On one of these coatings a second layer of Eu-

doped YSZ was deposited on top of the 120 µm of undoped YSZ. For this

demonstration, the composition of the Er-doped sensor layer was

Zr0.965Y0.07Er0.005O1.9625 while the Eu-doped sensor layer had a composition of

Page 28: Molly Gentleman Thesis 2006

18

Zr0.96Y0.07Eu0.01O1.96. The first of the EB-PVD samples was used to demonstrate a

“red-line” sensor (Figure 1.3.1) and the second a simple “rainbow” sensor (Figure

1.3.2) consisting of only two sensor layers: one at the top of the coating an the

other at the bottom (layers 1 and 5 in Figure 1.3.2) with undoped YSZ between the

two layers. This second configuration will be referred to as a “heat-flux” sensor

since temperature measurements from the two layers taken simultaneously can be

used to measure the heat flux in the coating. The other types of samples were a

series of ceramic multilayer materials of differently doped layers created by

sequential deposition of rare-earth-doped powders and sintering at 1500 °C. Both

doped YSZ and doped GZO multilayer materials were produced this way to

simulate the “rainbow” coatings based on both material systems.

First the “red-line” sensor was examined. When the buried Er-doped sensor

was illuminated using the 514 nm excitation, the characteristic luminescence from

the Er3+ was detected through the full thickness of the coating (Figure 1.8.1). In

contrast, using the illumination at 248 nm, no luminescence from the Er-doped

layer was detectable since the absorption length at 248 nm is substantially smaller

than the coating thickness. This straightforward demonstration illustrates the

combined use of a buried luminescence layer and selection of excitation

wavelengths to differentiate between the surface and inner-layers of a thermal

barrier coating. Incidentally, the presence in the spectrum of the R-line

luminescence from the thermally grown oxide confirms that the 514 nm excitation

penetrates through the TBC, substantiating that the Er luminescence is from the

Page 29: Molly Gentleman Thesis 2006

19

TBC next to the TGO. Figure 1.8.2 shows the spectra from the “heat flux” sensor

when excited in the visible. The spectra demonstrates that both the Er3+

luminescence from the buried layer as well as the Eu-doped top-layer in the “heat-

flux” sample can be excited simultaneously and that the lines do not overlap.

Figure 1.8.3 is an example of the luminescence spectra from a prototypical

“rainbow” sensors based on rare-earth doped gadolinium zirconate powders excited

at 514 nm, consisting of ErO1.5 and EuO1.5 doped layers. As indicated by the

labeling, the spectral lines from the individual layers are distinct and discernable.

1.9 Summary

The restrictions of sensor layer phase compatibility with the thermally

grown aluminum oxide and the TBC coating itself place severe constrains on the

possible luminescence sensor materials and suggest that the best chromophores are

likely to be those that are capable of being incorporated into the crystal structure of

the thermal barrier coating. For both YSZ and the rare-earth pyrochlore coatings,

rare-earth dopants can serve as chromophores. A number of concepts for

“visualization” and “quantitative” luminescence sensors are demonstrated based on

the optical properties of these coating materials and the luminescence properties of

the rare-earth dopants. For temperature sensing, the choice of rare-earth dopants

and concentrations is much more restrictive, as will be described in the following

chapters.

Page 30: Molly Gentleman Thesis 2006

20

1. Miller, R.A., NASA Report CP 3312 17. 1995. 2. Stern, K.H., Metallurgical and Ceramic Protective Coatings. 1996, New

York: Chapman and Hall. 3. Evans, A.G., et al., Mechanisms controlling the durability of thermal

barrier coatings. Prog. Mater. Sci., 2001. 46: p. 505-553. 4. Clarke, D.R. and C.G. Levi, Materials design for the next generation of

thermal barrier coatings. Annu. Rev. Mater. Res., 2003. 33: p. 383. 5. Maloney, M., Thermal barrier coating systems and materials, U.T.

Corporation, Editor. 2000: USA. 6. Heuer, A.H. and L.W. Hobbs, Science and Technology of Zirconia. 1981,

Columbus, OH: Am. Ceram. Soc. 7. Harris, D.C., Materials for Infrared Windows and Domes. 1999: SPIE

Optical Engineering Press. 8. Nychka, J.A., et al., Temperature-Dependent Optical Reflectivity of

Tetragonal-Prime Yttria-Stabilized Zirconia. J. Am.Ceram. Soc, 2006. 89(3): p. 908.

9. Wahiduzzaman, S. and T. Morel, Final Report to Oak Ridge National Laboratory, #ORNL/sub/88-22042/2 available as DE92-041384 from NTIS.

10. Hillary, R.E., NRC Report. 1996, National Academy Press. 11. Grattan, K.T.V. and Z.Y. Zhang, Fiber Optic Fluorescence Thermometry.

1995, London: Chapman & Hall. 12. Sovers, O.J. and T. Yoshioka, Fluorescence of Trivalent-Europium-Doped

Yttrium Oxysulfide. J. Chem. Phys., 1968. 49(11): p. 4945. 13. Wickersheim, K.A. and R.V. Alves, Fluoroptic Thermometry: a New RF-

Immune Technology, in Biomedical Thermology. 1982, Alan Liss: New York. p. 547.

14. Allison, S.W., L.A. Boatner, and G.T. Gillies, Characterization of high-temperature themographic phosphors: spectral properties of LuPO4:Dy(1%), Eu(2%). Appl. Optics, 1995. 34(25): p. 5624.

15. Feist, J.P. and A.L. Heyes, The characterization of Y2O2S:Sm powder as a thermographic phosphor for high temperature applications. Meas. Sci. Tech., 2000. 11: p. 942.

16. Feist, J.P., A.L. Heyes, and J.R. Nicholls, Phosphor thermometry in electron beam physical vapor deposition produced thermal barrier coating doped with dysprosium. Proc. Instn. Mech. Engrs., 2001. 215(G): p. 333.

17. Moos, H.W., Spectroscopic relaxation processes of rare earth ions in crystals. J. Lum., 1970. 1(2): p. 106.

18. Weber, M.J., Multiphonon relaxation of rare-earth ions in yttrium orthoaluminate. Phys. Rev. B, 1973. 8(1): p. 54.

Page 31: Molly Gentleman Thesis 2006

21

19. Leckie, R.M.R., et al. Microstructure stabliity issues in emerging TBC materials. in International Symposium on Thermal Barrier Coatings and Titanium Aluminides. 2002. Bonn, Germany.

20. Leckie, R.M.R., et al., Thermochemical compatibility between alumina and ZrO2-GdO3/2 thermal barrier coatings. Acta Mat., 2005. 53: p. 3281.

21. Blasse, G., Luminescent Materials. 1994, New York: Springer-Verlag. 22. Henderson, B. and G.F. Imbusch, Optical Spectroscopy of Inorganic Solids.

1989, New York: Oxford University Press. 23. Kaminskii, A., Laser Crystals: Thier Physics and Properties. 1981, New

York: Springer Verlag. 24. Shannon, R.D., Revised effective ionic radaii and systematic studies of

interatomic distances in hallides and chalcogenides. Acta Crystallogr., 1976. A(32): p. 751.

25. Feist, J.P. and A.L. Heyes, Europium-doped yttria-stabilized zirconia for high-temperature phosphor thermometry. Proc. Instn. Mech. Engrs., 2000. 214(L): p. 7.

26. Sergo, V. and D.R. Clarke, unpublished research. 1994. 27. Eldridge, J., In Press. Surface & Coatings Technology, 2003. 28. Dieke, G.H., Spectra and Energy Levels of Rare Earth Ions in Crystals, ed.

H.M. Crosswhite and H. Crosswhite. 1968, New York: Interscience Publishers.

29. Riseberg, L.A. and M.J. Weber, Relaxation phenomena in rare-earth luminescence, in Progress in Optics, E. Wolf, Editor. 1976, North-Holland. p. 89.

30. Johnson, P. and F. Williams, Interpretation of the dependence of luminescence efficiency on activator concentration. J. Chem. Phys., 1950. 18(11): p. 1477-1483.

Page 32: Molly Gentleman Thesis 2006

22

Figure 1.2.1 Schematic diagram of thermal barrier system. The system consists of a thin overcoat of low thermal conductivity ceramic on top of a thin metallic bond-coat on a superalloy substrate. The thermally grown oxide (TGO) results from oxidation of the bond-coat on exposure to high temperatures in air. The microstructure represented in this schematic is typical of an electron-beam physical vapor deposited (EB-PVD) coating.

Page 33: Molly Gentleman Thesis 2006

23

Figure 1.2.2 The optical absorption and scattering cross-sections for YSZ

material, redrawn following reference [9]. Superimposed are the wavelengths of two of the lasers used to excite luminescence from the sensors constructed. Superimposed are the blackbody radiation curves for the indicated temperatures. The shaded band indicates the wavelength window over which any temperature sensor must operate in the presence of thermal radiation. At shorter wavelengths, absorption and the optical band-gap (Eg) of YSZ limit both luminescence excitation and emission. At longer wavelengths, sensing will be limited by the ability to distinguish the luminescence signal above the thermal radiation.

Page 34: Molly Gentleman Thesis 2006

24

Figure 1.3.1 Schematic diagram of the use of a buried luminescence layer as a

“red-line” sensor. The same configuration, with a temperature sensitive chromophore, can be used to monitor the temperature of the coating in contact with the underlying bond-coat metal.

Page 35: Molly Gentleman Thesis 2006

25

Figure 1.3.2 Schematic diagram of the “rainbow” sensor, consisting of a

multilayer of luminescent dopants. As individual doped layers are eroded, worn away or spalled, the spectrum changes with the characteristic luminescence form those layers disappearing as they are eroded away as shown schematically.

Page 36: Molly Gentleman Thesis 2006

26

Figure 1.4.1 Typical luminescence decay curves illustrating (a) a characteristic

single exponential decay and (b) a double exponential decay fit with Equation 1.4.1 and Equation 1.4.2 respectively.

Page 37: Molly Gentleman Thesis 2006

27

Figure 1.4.2 Schematic diagram of the temperature dependence of luminescence

lifetimes.

Page 38: Molly Gentleman Thesis 2006

28

Figure 1.5.1 Diagram representing the thermodynamic phase stability of alumina with zirconia containing yttria or gadolinia. The symbols refer to the crystal structures of the compounds: P – perovskite, C – cubic, G – garnet, F – fluorite, Py – pyrochlore. (Redrawn after Leckie et al. [19] and Leckie, Kramer, Ruhle, and Levi [20]).

Page 39: Molly Gentleman Thesis 2006

29

Figure 1.6.1 Dieke diagram for the energy levels of four of the candidate rare-

earth ions that are soluble within the crystal structure of both YSZ and GZO coatings. For reference, the energies of candidate excitation lasers and the peak in blackbody radiation at 1500 °C are also indicated.

Page 40: Molly Gentleman Thesis 2006

30

Figure 1.7.1 Luminescence spectrum in the visible for YSZ doped with 0.5 – 5

m/o SmO1.5 and excited at 514 nm.

Page 41: Molly Gentleman Thesis 2006

31

Figure 1.7.2 Luminescence intensity of Sm3+ ions in YSZ as a function of the

concentration illustrating the phenomenon of concentration quenching. The different curves correspond to the different luminescence lines at the wavelengths labeled. The curves through the data correspond to Equation 1.7.1. Excitation was done at room temperature with and argon-ion laser operating at 514 nm.

Page 42: Molly Gentleman Thesis 2006

32

Table 1.7.1 Johnson and Williams fitting parameters for the data in Figure 1.7.2

Page 43: Molly Gentleman Thesis 2006

33

Figure 1.7.3 Luminescence intensity verses concentration of Eu3+ in YSZ

illustrating the absence of concentration quenching for all lines with 514 nm excitation.

Page 44: Molly Gentleman Thesis 2006

34

Figure 1.7.4 Luminescence intensity versus concentration of Er3+ in YSZ

illustrating the absence of concentration for the emission lines at 645 and 678 nm but quenching for the lines at 543 and 562 nm. The lack of concentration quenching of the 645 and 678 nm lines is attributed to cross-relaxation.

Page 45: Molly Gentleman Thesis 2006

35

Figure 1.8.1 Luminescence from a buried ~10 µm layer of Er-doped YSZ in an

EB-PVD YSZ coating and excited with a 514 nm laser. The presence of the R-line luminescence from the TGO in the same spectra confirms that the excitation laser penetrated to the buried sensor layer. No luminescence is excited with the 248 nm laser as it cannot penetrate into the TBC.

Page 46: Molly Gentleman Thesis 2006

36

Figure 1.8.2 Luminescence from a buried ~10 µm layer of Er-doped YSZ at the

base of an EB-PVD YSZ coating with a ~10 µm layer of Eu-doped YSZ at the top of the coating and excited with a 532 nm laser. The luminescence from the two layers illustrates how a “heat-flux” sensor could be constructed by choosing luminescence ions whose luminescence spectra do not overlap.

Page 47: Molly Gentleman Thesis 2006

37

Figure 1.8.3 Demonstration of two prototype “rainbow” sensors containing

ErO1.5 and EuO1.5 on fabricated out of YSZ and the other out of GZO. Both were excited at 514 nm.

Page 48: Molly Gentleman Thesis 2006

38

Chapter 2 Experimental Investigation 2.1 Introduction

As described in the previous chapter, there are many materials, materials

processing techniques, and luminescence techniques that are available for use in

development of temperature sensors for thermal barrier materials. The

experimental investigation of these materials presented in this chapter will be

broken into two categories: the characterization of the materials and the calibration

and use of luminescence to measure temperature. Characterization studies

consisted of crystal structure analyses as the materials’ dopant concentrations and

aging conditions were varied. These studies established the validity of the use of

bulk materials in such tests as they apply to thermal barrier coatings. Calibrations

of luminescence lifetime as a function of temperature provide a basis for the

sensors described in subsequent chapters. This chapter outlines techniques for

processing both bulk and coated materials as well as describes methods for

examining crystal structure and luminescence at both room and elevated

temperatures.

2.2 Materials Characterization

7-8 w/o Y2O3 stabilized zirconia is the preferred material for thermal barrier

coating, although there is some use of rare earth zirconates (M2Zr2O7 where M is

the rare-earth ion). In the YSZ system, under equilibrium conditions, the yttria

stabilizes the tetragonal phase at temperatures above around 1050 °C. After

extended exposure to high temperatures, the material partitions into the equilibrium

Page 49: Molly Gentleman Thesis 2006

39

tetragonal and cubic phases, which can then transform to monoclinic upon cooling.

In TBCs, the deposition process is not an equilibrium process and results in a

metastable tetragonal (t’) crystal structure, which does not transform upon

cooling[1]. In order to develop luminescence sensors applicable to thermal barrier

coatings, the bulk materials used in these studies must have the same crystal

structure (t’) as that of the coating. The constraint is not so restrictive in the case of

the zirconates, but the extent of ordering in the materials still needs to be taken into

consideration. It is also vital that there be a random distribution of luminescence

ions in both bulk and coating TBC materials to prevent changes in the

luminescence behavior due to clustering and highly doped materials effects[2].

2.2.1 Materials Processing

The bulk rare-earth doped yttria stabilized zirconias as well as rare-earth

doped zirconates used in these studies were prepared from powders synthesized by

reverse co-precipitation of aqueous rare-earth nitrate solutions (RE(NO3)3) and

zirconium acetate (ZrO(C2H3O2)2) in dilute acetic acid solutions into ammonium

hydroxide (NH4OH) following the work of Ciftcioglu and Mayo[3]. Reverse co-

precipitation (the addition of acidic solution to ammonium hydroxide solution) was

selected for the preparation of these materials to avoid segregation of the cations

and ensure molecular mixing of the dopant into the host material. The random

dispersion of the dopant material is particularly important in the case of processing

luminescent materials because clustering of luminescent ions can change the

characteristic luminescence of the material. An ammonium hydroxide solution

Page 50: Molly Gentleman Thesis 2006

40

with constant PH of nine was used to guarantee instantaneous precipitation of the

mixed cation solution. After precipitation and drying under heat lamps, the

powders were calcined at 950 °C for 2 hours to obtain the oxide, pressed into

pellets, and sintered at 1200 °C in air for 2 hours. After initial heat treatments,

some of these materials underwent secondary exposures at varying temperatures to

examine the microstructural evolution and are noted where appropriate.

Several EB-PVD coatings were made of the rare-earth doped YSZ

materials. The ingots for EB-PVD of the sensor layers were prepared by

infiltration of a standard commercially supplied 7YSZ ingot with aqueous rare-

earth nitrate (RE(NO3)3). The nitrate solution was gelled into the pores of the ingot

with NH4OH, dried under heating lamps, and calcined at 950 °C for 2 hours. This

process was repeated until the desired weight of rare-earth oxide was achieved in

the pores of the ingot. This technique for the doping of the ingot was chosen in

order to achieve a coating of homogenous composition often a problem for dual

source evaporation [4] or bored ingot deposition [5]. The layered structure of the

sensor layers was achieved in a single run of EB-PVD by layering the doped ingot

with a standard undoped 7YSZ ingot. The final coatings consisted of a series of

different layered structures with two different rare earth dopants. An example of

the final structure as illustrated by cathodoluminescence is seen in Figure 2.2.1.1.

2.2.2 Raman Spectroscopy

Zirconia exhibits one of the strongest Raman signals reported in the

literature. As a result, it is a powerful tool in the analysis of zirconia based

Page 51: Molly Gentleman Thesis 2006

41

systems[6]. Room temperature Raman spectroscopy was used to confirm that the

yttria stabilized zirconia had the tetragonal crystal structures for all dopant

concentrations and heat treatments. The Raman spectra for the ordered and

disordered pyrochlore structures are also well known[7] and so were used in these

studies to determine the structure and ordering of the pyrochlore zirconates.

Raman spectra were collected using an Instruments SA spectrometer fitted with an

Argon-ion laser emitting at either 514 or 488 nm. An example of the Raman

spectra for each of these crystal structures is seen schematically in Figure 2.2.2.1

with arrows denoting the peaks used to identify the respective crystal structures.

Raman spectra were also taken for some of the materials at temperatures up to 1200

°C using a Linkam 1500 heating stage to examine the change in phonon energies as

a function of temperature (work of Vanni Lughi). These values were then used for

fitting in the multiphonon model discussed previously (Equation 1.4.5).

2.2.3 X-ray Diffraction

Powder x-ray diffraction was carried out on the YSZ materials to confirm

the tetragonal-prime crystal structure as well as to determine the c/a ratios for the

different dopant concentrations. X-ray analysis was performed over the range of 72

≤ °2θ ≤ 76, which isolates the [400] reflections and very clearly indicates the extent

of tetragonality of the material. Initial x-ray scans were done on a Panalytical

MRD PRO thin film x-ray diffractometer with a monochromic source to provide a

baseline for kα1 stripping of the subsequently taken peaks. All other scans were

completed using a Phillips XPERT powder x-ray diffractometer (these data are

Page 52: Molly Gentleman Thesis 2006

42

examined in chapter 4). X-ray analysis was also attempted on the pyrochlore

zirconates, but in all cases, the peaks characteristic of pyrochlore ordering were

absent.

2.3 Luminescence Characterization

As described in the previous chapter, the luminescence lifetime of a rare-

earth doped material is sensitive to temperature, according to the multiphonon

model described in chapter 1. In order to use the temperature sensitivity of

luminescence to measure temperature of a thermal barrier material, we carried out

the following set of optimization experiments. The first, summarized below, is the

use of excitation spectroscopy to select the best possible excitation wavelength for

the illumination of the sensor material. The next step is to examine the effects of

concentration on intensity and then compare that with the effects that concentration

has on the luminescence lifetime as a function of temperature. By taking all of

these aspects into consideration, the best sensor materials were identified. In this

chapter, in addition to materials issues, several issues regarding the development of

the experimental methodology are discussed. The development of these techniques

has been vital in the development of sensors capable of temperature measurements

at temperatures applicable to thermal barrier systems.

2.3.1 Room temperature luminescence characterization

Excitation spectra of each of the materials were recorded at room

temperature using a PerkinElmer LS 55 spectrometer in order to identify the

excitation wavelengths that would maximize the intensity of the spectral emission

Page 53: Molly Gentleman Thesis 2006

43

at different wavelengths. The PerkinElemer LS 55 spectrometer was capable of

excitation from 200 – 800 nm and collecting luminescence emission from 200 –

900 nm. In all the measurements, both the excitation and emission slit widths were

set to 5 nm. Cutoff filters were placed after the sample in the light path to avoid

scattering of the excitation. All major luminescence lines were examined for each

of the materials but only those for the strongest transitions are reported. These

excitation spectra were then compared with the laser wavelengths at our disposal

(248 nm KrF Excimer, 355 nm tripled YAG:Nd, 532 nm doubled YAG:Nd) to

determine the most efficient laser frequency for excitation used in both spectral and

luminescence lifetime measurements.

For Eu3+ doped materials used most extensively in this work, the

luminescence studies presented were all performed using a frequency-doubled,

Continuum Model NY81C-1, Q-switched YAG:Nd laser emitting at 532 nm and

with a pulse length of ~10 ns for a Q-switch time of 250 µs, based on the excitation

spectra and the fact that the excitation yields the strongest intensity at 530 nm. The

luminescence emission from the rare-earth doped materials was collected using a

single crystal sapphire light pipe with a polished end and fitted at the opposite end

with a standard SMA fiber-optic coupler. A silica fiber optic cable with a 500 µm

core was connected to the SMA end of the light pipe and the luminescence was

directed to a series of collimating lenses with a 532 nm Semrock Raman filter in-

line to remove the scattered laser light from the signal. A second fiber-optic cable

was used to transmit the signal to an Ocean Optics USB 2000 grating spectrometer

Page 54: Molly Gentleman Thesis 2006

44

fitted with a 300 nm cutoff filter with a photodiode array for spectrum collection

(Figure 2.3.1.1). The same configuration was used for luminescence collection

with other excitation frequencies, with the exception that the 532 nm Raman filter

was replaced with a filter of appropriate wavelength to block the operating laser

frequency.

2.3.2 High temperature luminescence characterization

For measurements of luminescence at elevated temperatures, a similar setup

to that for low temperatures was used. The samples were placed inside a box-

furnace with two access ports, one for the free-space propagation of the laser beam

excitation into the furnace and the other for collecting the luminescence signal

(Figure 2.3.2.1). To enable the luminescence to be collected at all temperatures up

to 1300 °C, a single crystal sapphire rod was used to gather the emitted light. The

sapphire light pipe was 3 mm in diameter and 10 cm long, polished at both ends,

and fitted with an SMA fiber-optic coupler at one end. The light pipe was

connected to a 500 µm core fiber-optic cable and passed through the same series of

collimating lenses and filters as described for the room temperature measurements.

This was coupled into a standard silica fiber-optic cable that was bifurcated along

its length so that a portion of the collected light could be spectral analyzed and a

portion used for luminescence decay measurements. An Ocean Optics USB 2000

grating spectrometer with a photodiode array was again used for spectral collection

and analysis. The luminescence lifetime-decay measurements were made using an

Acton SpectraPro 2150i monochromator fitted with a Hammamatsu R928

Page 55: Molly Gentleman Thesis 2006

45

photomultiplier tube (PMT) or using a 610 nm bandpass filter and a Hammamatsu

H6780-20 photosensor module. In both cases the PMT bias was kept between 800

– 900.

The lifetime measurements were made under computer control with the

luminescence intensity collected after the laser pulse was triggered. The laser

trigger was connected directly to the oscilloscope. To ensure good signal to noise,

an average of the luminescence signal from 1000 laser pulses was recorded with the

signals accumulated in a LeCroy 9310AM digital oscilloscope. In some cases, a

trans-impedance amplifier was used to amplify the luminescence signal before the

oscilloscope recorded it.

2.3.4 Amplification and Electronics

One problem that has been experienced by other groups has been the

mastering the electronics for collection. Figures 2.3.4.1 and 2.3.4.2 illustrate

several common artifacts associated with inappropriate impedance matching and

bandwidths. Most common is the mismatching of impedance between the PMT

(typically 50 Ω) and that of either the amplifier (ranging from 500 Ω to 1 MΩ) or

the oscilloscope (50Ω or 1 MΩ). Mismatches of several orders of magnitude

between the impedance of these two components result in a significant

amplification of the signal intensity but have a detrimental effect on the speed of

the measurement. As is illustrated in Figure 2.3.4.1, the signal can be significantly

distorted when the PMT is set to 50 Ω and the oscilloscope to 1 MΩ. Although the

signal is greatly amplified when compared to that of a 50 Ω -50 Ω connection, the

Page 56: Molly Gentleman Thesis 2006

46

resolution provided by the higher speed measurement, including the faster initial

decay, is lost and a time constant of ~ 300 µs is introduced. Similar results are seen

for the oscilloscope set at 1 MΩ, 100 kΩ, and 1 kΩ in Figure 2.3.4.2. This

phenomenon has been reported incorrectly in the literature as actual luminescence

lifetime behavior [8]. To reach the fastest speed capable by the PMT it is

imperative to match the impedance of the oscilloscope exactly to the PMT (50 Ω).

This was the condition used for the very fastest lifetime measurements reported in

these studies.

Page 57: Molly Gentleman Thesis 2006

47

1. Clarke, D.R. and C.G. Levi, Materials design for the next generation of thermal barrier coatings. Annu. Rev. Mater. Res., 2003. 33: p. 383.

2. Eldridge, J.I., Unpublished work. 3. Ciftcioglu, M. and M.J. Mayo, Processing of Nanocrystalline Ceramics.

Mat. Res. Soc. Symp. Proc., 1990. 196: p. 77. 4. Allison, S.W., et al. Use of phosphor coatings for high temperature

aerospace applications. in 29th AIAA/ASME/SAE/ASEE Joint Propulsion Conference. 2003. Huntsville, AL: AIAA.

5. Feist, J.P., A.L. Heyes, and J.R. Nicholls, Phosphor thermometry in electron beam physical vapor deposition produced thermal barrier coating doped with dysprosium. Proc. Instn. Mech. Engrs., 2001. 215(G): p. 333.

6. Clarke, D.R. and F. Adar, Measurement of the Crystolographically Transformed Zone Produced by Fracture in Tetragonal Zirconia Containing Ceramics. J. Am.Ceram. Soc, 1982. 65: p. 284.

7. Michel, D., M. Perez Y Jorba, and R. Collongues, Study by raman spectroscopy of order-disorder phenomena occurring in some binary oxides with fluorite-related structures. J. Raman Spec., 1976. 5(3): p. 163.

8. Feist, J.P. and A.L. Heyes, Europium-doped yttria-stabilized zirconia for high-temperature phosphor thermometry. Proc. Instn. Mech. Engrs., 2000. 214(L): p. 7.

Page 58: Molly Gentleman Thesis 2006

48

Figure 2.2.1.1 Cathodluminescence image of an EB-PVD coating with a ~10 µm

Eu-doped layer at the interface next to the bond-coat. The brightest contrast is due to the Cr3+ impurities in the alumina thermally grown oxide. Immediately above the TGO the luminescence from the Eu3+ ions in the sensor layer are visible as a light gray band.

Page 59: Molly Gentleman Thesis 2006

49

Figure 2.2.2.1 Representative Raman spectra for (a) tetragonal-prime 7 w/o Y2O3

stabilized zirconia and (b) ordered and (c) disordered pyrochlore crystal structures. The * mark the peaks used to identify the tetragonal crystal structure and arrows denote the ordering peaks in the pyrochlore.

Page 60: Molly Gentleman Thesis 2006

50

Figure 2.3.1.1 Schematic drawing of the optical arrangement used for measuring

the luminescence lifetime and spectrum of sensor materials at room temperature.

Page 61: Molly Gentleman Thesis 2006

51

Figure 2.3.2.1 Schematic drawing of the optical arrangement used for measuring

the luminescence lifetime and spectrum of sensor materials at elevated temperatures.

Page 62: Molly Gentleman Thesis 2006

52

Figure 2.3.4.1 The luminescence lifetime decay of 1 a/o Eu:YSZ at 600 °C using

two different impedances at the oscilloscope. Although the 1 MΩ termination increases the signal intensity, it slows the time resolution of the measurement and the fine features seen in the 50 Ω signal are not apparent.

Page 63: Molly Gentleman Thesis 2006

53

Figure 2.3.4.2 The luminescence lifetime as a function of temperature for 1 a/o

Eu:YSZ using several different input impedances at the oscilloscope. Larger impedances result in time constants that limit the maximum temperature that can be measured using the luminescence technique.

Page 64: Molly Gentleman Thesis 2006

54

Chapter 3 Luminescence sensing of temperature in pyrochlore zirconate

materials

3.1 Introduction

One method of determining the temperature of a component without

making physical contacts is by measurement of the luminescence lifetime decay of

a phosphor either embedded in the component or painted on its surface[1, 2]. This

method, which dates back to the work of Bradley in 1953 [3], is particularly

attractive in measuring temperature in aggressive environments but has generally

been limited to materials that are both stable and exhibit temperature-dependent

luminescence decay. Although there have been recent reports of temperature

dependent lifetimes up to 1500 °C in Tm and Eu-doped YAG (yttrium aluminum

garnet)[4], these compounds are neither phase compatible with the oxide formed on

the bond-coat in air at high temperatures, nor do they exhibit low thermal

conductivity required for thermal barrier coatings[5].

To ensure thermodynamic compatibility, it is desirable that the luminescent

species be individual ions in solid solution in the crystal structure of the thermal

barrier coating material itself so that the TBC also acts as the sensor. In this

concept, various practical issues with previous luminescent sensors, such as

attachment, adhesion, and high-temperature compatibility, are circumvented. The

first task is to identify suitable dopants that can be taken into solid solution within

the TBC material without destabilizing it and also have the luminescence

Page 65: Molly Gentleman Thesis 2006

55

characteristics necessary for sensing and temperature measurement in the presence

of thermal radiation from other hot surfaces. These and related selection criteria

have been presented previously, together with results of luminescence doping of

yttria-stabilized zirconia[6, 7].

Observations of the principal luminescence characteristics of four different

europium-containing materials are presented in this chapter. Three are zirconates

with the pyrochlore structure and the fourth, for comparison purposes, is YSZ with

1 a/o of the Y3+ ions replaced with Eu3+, and having the formula

Y0.06Eu0.01Zr0.93O1.965. Two of the zirconates were doped with a small

concentration of europium, (Sm0.99Eu0.01)2Zr2O7 and (Gd0.975Eu0.025)2Zr2O7. The

third was europium zirconate, Eu2Zr2O7.

3.2 Materials

All materials used in this study were prepared as bulk ceramics by reverse

co-precipitation outlined in section 2.1. The Raman spectra recorded following

sintering at 1200 °C for 2 hours for all four materials examined in this study are

seen in Figure 3.2.1. The spectra from the zirconates not only confirm that they

were single-phase materials but also that they exhibit different degrees of ordering.

By comparison of the peaks with those reported in the literature [11], the

gadolinium zirconate appears disordered following initial heat treatments whereas

the samarium and europium zirconates both exhibit some degree of ordering as

evidenced by the presence of distinct Raman lines at approximately 310, 397, 530,

and 594 cm-1. The Raman spectrum of the Eu-doped YSZ confirms that the

Page 66: Molly Gentleman Thesis 2006

56

material was also single-phase with no indication that the Eu-doping had caused

any destabilization of the tetragonal phase to monoclinic zirconia.

Excitation spectra for each of the materials were recorded at room

temperature to identify excitation wavelengths that would maximize the intensity of

spectral emission at different wavelengths. The excitation spectra for emission at

610 nm are reported in Figure 3.2.2 and indicate that several wavelengths,

including 390, 470, and 530 nm, could be used to excite luminescence from all of

these materials. Superimposed on these spectra are the reported[12] observed

energy levels for Eu3+ as well as the wavelengths of lasers available for excitation:

248 nm KrF excimer, 355 nm frequency tripled YAG: Nd, and 532 nm frequency

doubled YAG: Nd lasers.

On the basis of the excitation spectra and the fact that the excitation yields

the strongest intensity at 530 nm, the luminescence studies reported in this work

were all performed using a frequency-doubled YAG: Nd laser (section 2.1). Room

temperature as well as high temperature luminescence spectra measurements were

recorded according to the specifications outlined in chapter 2 (Figure 3.2.3). The

luminescence lifetime measurements were recorded from room temperature, until

either the decay time of the luminescence reached the detectability limit of out

current electronics or the intensity of the signal dropped to a level below the noise

floor of the measurements, making further measurements impossible.

Page 67: Molly Gentleman Thesis 2006

57

3.3 Room temperature luminescence

The room temperature luminescence of the four materials from 550 to 680

nm (the wavelength region that encompasses the principal emission lines of Eu3+

ions in the visible) is shown in Figure 3.3.1. Both the three zirconates and the

doped YSZ exhibit similar luminescence, although that from the

(Sm0.99Eu0.01)2Zr2O7 zirconate is significantly weaker in intensity. The similarity of

major features of the luminescence spectra is in accord with expectation based on

the transitions of the 4f electrons responsible for luminescence in trivalent rare-

earth ions, and the relative insensitivity of the emission wavelengths is

characteristic of the 4f electrons being shielded by the outer shell 5s and 5p

electrons [13]. Thus, although there are small differences in the wavelengths of the

peaks from one compound to another, the principal transitions can be readily

identified as being due to the 5D0à7F0 (~582 nm), the 5D0à

7F1 (~592 nm), and the

5D0à7F2 (~610 and 635 nm) transitions. The most notable distinction between the

luminescence from the zirconates and the doped YSZ is that the splitting of the

5D0à7F2 results in a much weaker peak at 635 nm that is red-shifted by ~5 nm in

the YSZ.

The 5D0à7F2 (~610 nm) transition was chosen for lifetime measurements

because of both its high intensity and narrow full-width-half-maximum (FWHM),

as the intensity will decrease and the FWHM increase with temperature (Figure

3.2.3). This transition is also the only one of the observed transitions that is a

strong electric dipole only allowed transition [12] meaning it will have a strong

Page 68: Molly Gentleman Thesis 2006

58

intensity in low symmetry crystals such as the D2d ( )m24 symmetry of the Eu3+ ion

in tetragonal zirconia. Its presence and relative strength in the pyrochlore materials

suggest that there is a significant deviation from the D3d ( )m23 symmetry of a

perfectly ordered pyrochlore as the presence of a inversion center in this symmetry

would forbid the 5D0à7F2 transition.

3.4 High temperature luminescence

Measurements were taken of luminescence lifetimes up to temperatures at

which the lifetime decay times fell below our current detectability limit (~10 ns) or

the intensity of the luminescence signal fell below the noise floor of the

measurement. The form of the luminescence intensity decay (as a function of time

after the excitation ceased) varied from one material to another. For both the

Eu2Zr2O7 and (Sm0.99Eu0.01)2Zr2O7 zirconates, the decays were single exponentials.

In contrast, the luminescence decays were more complex for the doped gadolinium

zirconate and YSZ. Both materials displayed a short exponential decay followed

by a longer exponential decay (Figure 3.4.1). In analyzing the luminescence

decays, they were fit using either a single (Equation 1.4.1) or double (Equation

1.4.2) exponential decay depending on the nature of the decay. The decay times

derived from the fitting are presented in the subsequent graphs.

The variation with temperature of the decay times of all the materials

exhibited the same characteristic features of a relatively constant, almost

temperature-independent, value, from room temperature up to an intermediate

temperature and thereafter an exponential decrease with increasing temperature.

Page 69: Molly Gentleman Thesis 2006

59

As the focus of this work is primarily on the high temperature characterization of

the luminescence lifetime, the majority of the data shown is for the temperatures

where there is a temperature-dependent effect on lifetimes.

The luminescence intensity decay from the pure europium zirconate showed

the simplest behavior. It exhibited a single exponential decay over the entire

temperature range investigated with the decay time being almost temperature-

independent up to ~700 °C and then decreasing exponentially up to 1250 °C

(Figure 3.4.2). At the very highest temperature, the luminescence intensity had

reached the limit of detectability with our present measurement system, suggesting

that the exponential variation with temperature may well be continued to even

higher temperatures.

Rather similar luminescence decay characteristics were shown by the

europium-doped samarium zirconate ((Sm0.99Eu0.01)2Zr2O7), although the numerical

values for the lifetime decay were different. Interestingly, although the overall

luminescence intensity was lower at room temperature, luminescence decays could

be measured up to almost 1000 °C (Figure 3.4.3). In this figure, the data points

oscillate about a straight exponential line. This is attributed to difficulties

experienced with our furnace calibrations.

As mentioned above, the Eu-doped GZO and YSZ materials exhibited more

complex, double exponential luminescence decays. The high temperature variation

of both the fast and the slow lifetime decays recorded from the

(Gd0.975Eu0.025)2Zr2O7 material is plotted in Figure 3.4.4. Superimposed are the data

Page 70: Molly Gentleman Thesis 2006

60

for the slow decay recorded from the Eu-doped YSZ. This comparison by

superimposition of data illustrates the surprising result that the luminescence

lifetime is essentially identical for these two materials even though they have

different crystal structures. This phenomenon will be addressed in the following

section.

3.5 Discussion

The principal finding of technological interest is that Eu-doping of the

pyrochlore zirconates investigated can be used as the basis for temperature

measurements in samarium zirconate up to at least 1000 °C and for both the

europium-doped gadolinium zirconates and YSZ up to at least 1150 °C.

Luminescence from Eu2Zr2O7 can be used to measure temperatures approaching

1300 °C. These temperatures are similar to those in current gas turbine systems

and thermal barrier coatings, but it is also likely that these materials can be used for

luminescence sensing of temperature at still higher temperatures with the

implementation of superior collector and detector equipment. Additionally, the

optimum doping concentrations in the pyrochlore have not yet been identified.

This is the subject of the following two chapters.

The lifetime data presented reveal some interesting correlations between the

crystal structures of the materials investigated (Figure 3.5.1). At first

consideration, one might conclude that because the Eu3+ ions occupy the same

position in all the zirconate structures, the luminescence characteristics would be

the same. The emission spectra are indeed very similar but the lifetime

Page 71: Molly Gentleman Thesis 2006

61

characteristics clearly are not. Furthermore, the lifetime decays of 5D0 à 7F2

radiative transition for the Eu-doped YSZ and Eu-doped GZO are almost identical,

despite the fact that these two compounds have different, albeit related, structures.

We believe that part of the explanation lies in the fact that both these oxides exhibit

extensive site disorder as well as site defects such as oxygen vacancies, whereas the

Raman spectra indicate that both the Eu2Zr2O7 zirconate and (Sm0.99Eu0.01)2Zr2O7

zirconate are ordered to a larger extent. In contrast to the disordered GZO, the

Raman spectra of these last two zirconates exhibit well defined Raman lines at

approximately 310, 397, 530, and 594 cm-1, lines previously identified as being

indicative of ordering in pyrochlore structures [11].

At dilute concentrations of a luminescent ion in a homogenous solid, the

luminescence lifetime is largely controlled by the probability of non-radiative

transitions of the excited state transitioning back to the ground state. The dominant

non-radiative transitions in defect-free materials are temperature-dependent

multiphonon relaxation processes, and therefore the observed radiative lifetime can

be expressed as:

mpiR τττ111

+= Equation 3.1

where iτ is the intrinsic lifetime, and mpτ is the lifetime of a multiphonon

relaxation of the excited state[14]. Equation 3.1 was used to fit the data in Figure

3.4.4, but the same fit was unsuccessful for the Eu2Zr2O7 and Eu-doped Sm2Zr2O7

materials. At higher concentrations, such as the conditions in the Eu2Zr2O7 and Eu-

Page 72: Molly Gentleman Thesis 2006

62

doped Sm2Zr2O7 materials, the excited state can diffuse from one luminescent ion

to another through the material until transitioning, non-radiatively at some defect

site, to the ground state. This energy diffusion process has the effect of increasing

the observed lifetime since it is a competitive process to the multiphonon relaxation

process. In effect, it can be considered as introducing a concentration-dependent

intrinsic lifetime and effectively adding a third term ⎟⎟⎠

⎞⎜⎜⎝

ETτ1 to Equation 3.1 to

indicate the rate of energy transfer. Unfortunately, without more information about

the nature of this additional term, there are too many parameters to meaningfully fit

the data using this form. The effect of disorder is more complex. One of the major

effects at low concentrations is to create two principal populations of luminescent

ions: isolated ions near defects and clusters of ions close to one another.

Luminescence from the former population is believed to be responsible for the

initial rapid luminescence lifetime decay since defects effectively decrease the size

of the energy gap between excited and ground states, resulting in faster decay times

and increasing the probability of non-radiative transitions. In contrast, energy

transfer can occur amongst ions in the latter populations, because they are close

together and consequently result in longer luminescence lifetimes.

Many of these effects can be seen in the data presented here. The lifetime

of the ordered Eu2Zr2O7 zirconate is longer at high temperature than that exhibited

by either the YSZ or GZO materials even though the effective activation energy is

similar. Both the Eu-doped GZO and Eu-doped YSZ exhibit characteristic double

Page 73: Molly Gentleman Thesis 2006

63

exponential decays and, furthermore, the lifetime versus temperature curves almost

overlap. The luminescence from the Eu-doped samarium zirconate,

(Sm0.99Eu0.01)2Zr2O7, is somewhat unusual due to the similarity in the excited state

energies of the Sm3+ and Eu3+ levels. This similarity has two consequences. One is

that the luminescence intensity from the Eu3+ ion is reduced because the excited

state energy can transfer to the non-radiative Sm3+ ions. This energy transfer path,

however, also enables the energy, with some probability, to transfer back to a Eu3+

ion and then radiate, thereby extending the observed radiative lifetime.

The results obtained so far suggest that these Eu-doped materials could be

employed in a number of different configurations for temperature sensing in

thermal barrier coating systems. For instance, a thin layer of Eu-doped YSZ could

be deposited on a component and then undoped YSZ deposited on top to form a

temperature sensor at the bond-coat interface. A second configuration, for

measuring the temperature at the surface of the TBC, could be made by depositing

a thin layer of either Eu-doped YSZ or Eu2Zr2O7 on an existing TBC. A third

configuration, for measuring the temperature difference through a coating, would

consist of two differently doped layers, one on the inside and the other on the

outside of the coating. During the course of this study, such sensors have been

deposited by electron beam evaporation and their analysis will be reviewed in

chapter 6.

Page 74: Molly Gentleman Thesis 2006

64

1. Allison, S. and G. Gillies, Remote thermometry with thermographic phosphors: Instrumentation and applications. Rev. Sci. Instrum., 1997. 68(7): p. 2615-2650.

2. James, K.A., W.H. Quick, and V.H. Strahan, Control Eng., 1979. 26: p. 30. 3. Bradley, L.C., Rev. Sci. Instrum., 1953. 24: p. 219. 4. Allison, S.W., et al. Use of phosphor coatings for high temperature

aerospace applications. in 29th AIAA/ASME/SAE/ASEE Joint Propulsion Conferrence. 2003. Huntsville, AL: AIAA.

5. Clarke, D.R. and C.G. Levi, Materials design for the next generation of thermal barrier coatings. Annu. Rev. Mater. Res., 2003. 33: p. 383.

6. Gentleman, M.M. and D.R. Clarke, Concepts for luminescence sensing of thermal barrier coatings. Surface & Coatings Technology, 2004. 188-189: p. 93.

7. Eldridge, J.I., et al., J. Thermal Spray Tech., 2004. 13: p. 44. 8. Gentleman, M.M. and D.R. Clarke, Luminescence sensing of temperature in

pyrochlore zirconate materials for thermal barrier coatings. Surface & Coatings Technology, 2005. 200: p. 1264.

9. Weber, M.J., Multiphonon relaxation of rare-earth ions in yttrium orthoaluminate. Phys. Rev. B, 1973. 8(1): p. 54.

10. Riseberg, L.A. and M.J. Weber, Relaxation phenomena in rare-earth luminescence, in Progress in Optics, E. Wolf, Editor. 1976, North-Holland. p. 89.

11. Michel, D., M. Perez Y Jorba, and R. Collongues, Study by raman spectroscopy of order-disorder phenomena occurring in some binary oxides with fluorite-related structures. J. Raman Spec., 1976. 5(3): p. 163.

12. Dieke, G.H., Spectra and Energy Levels of Rare Earth Ions in Crystals, ed. H.M. Crosswhite and H. Crosswhite. 1968, New York: Interscience Publishers.

13. Henderson, B. and G.F. Imbusch, Optical Spectroscopy of Inorganic Solids. 1989, New York: Oxford University Press.

14. Moos, H.W., Spectroscopic relaxation processes of rare earth ions in crystals. J. Lum., 1970. 1(2): p. 106.

Page 75: Molly Gentleman Thesis 2006

65

Figure 3.2.1 Raman spectra of the three zirconate materials and of the Eu-doped

YSZ material. Ordering in the zirconates is indicated by the presence of additional Raman peaks indicated.

Page 76: Molly Gentleman Thesis 2006

66

Figure 3.2.2 Excitation spectra for the emission at 610 nm. All the spectral

features can be attributed to either electronic transitions of the Gd3+ or Eu3+ ions. Uniformly, the strongest excitation for all four samples occurs at a wavelength corresponding to the 5D0 and 5D1 transitions of the 4f electrons of the Eu3+ ions.

Page 77: Molly Gentleman Thesis 2006

67

Figure 3.2.3 Spectra of Eu-doped YSZ taken at several temperatures with 532 nm

excitation.

Page 78: Molly Gentleman Thesis 2006

68

Figure 3.3.1 Emission spectra for the three Eu-containing zirconates and Eu-

doped YSZ materials studied. The spectra were recorded at room temperature using excitation at 532 nm. The arrow denotes the emission line (5D0à

7F2) used for luminescence lifetime measurements.

Page 79: Molly Gentleman Thesis 2006

69

Figure 3.4.1 Luminescence decay curves illustrating the double exponential for

the (Gd0.975Eu0.025)2Zr2O7 at several temperatures.

Page 80: Molly Gentleman Thesis 2006

70

Figure 3.4.2 Luminescence lifetime decay time as a function of temperature for

Eu2Zr2O7. A line has been drawn though the data to aid the eye.

Page 81: Molly Gentleman Thesis 2006

71

Figure 3.4.3 The variation of the luminescence decay with temperature of

(Sm0.99Eu0.01)2Zr2O7. A line has been drawn though the data to aid the eye.

Page 82: Molly Gentleman Thesis 2006

72

Figure 3.4.4 Temperature dependence of the luminescence decay times for both

the fast and slow decays of Eu in (Gd0.98Eu0.02)2Zr2O7. Superimposed are the lifetime decays data for the Eu-doped YSZ material. The line through the data is a fit of the data using the multiphonon model. The scatter in the data at ~400 °C is due to difficulties in the fitting of the lifetime decays due to the emergence of the faster exponential decay.

Page 83: Molly Gentleman Thesis 2006

73

Figure 3.5.1 Schematic drawing of the rare-earth zirconate and zirconia crystal

structures. The introduction of trivalent ions on the zirconia structure results in oxygen vacancies on the anion sub-lattice.

Page 84: Molly Gentleman Thesis 2006

74

Chapter 4: Concentration Effect on Luminescence in Eu-doped Yttria

Stabilized Zirconia

4.1 Introduction

In the previous chapters, europia-doped YSZ was shown to be a promising

material for temperature sensing. In chapter three a composition of 1 a/o Eu:YSZ

was investigated and shown to exhibit temperature sensitivity ranging from 500 to

1150 °C[1]. This measurement was limited by both the speed of the decay time

(~10ns) and the decrease in intensity caused by thermal quenching of the

luminescence. Although this composition was not chosen arbitrarily, very little

was known about the choice of optimum dopant compositions for europium in YSZ

a priori. In this chapter, we will explore the effects of Eu3+ and Y3+ composition

on the luminescence and temperature-sensing capabilities of europia-doped yttria

stabilized zirconias.

The interest in Eu3+ as a phosphor in thermal barrier materials has grown

considerably due to its promise as efficient non-contact temperature sensor.

Traditionally, Eu3+ has been used to probe the site symmetry and chemical bond

differences of the surrounding host materials in the neighborhood of the Eu3+ ion[2-

7]. In the work of Dexpert-Ghys, site selective spectroscopy of the Eu3+ ion was

used to explore changes in symmetry induced by the partial stabilization of zirconia

by yttria[8]. In those studies it was determined not only that the luminescence from

the europium ions was characteristic of lower symmetry than expected for

tetragonal zirconias, but also that there was no significant change in the

Page 85: Molly Gentleman Thesis 2006

75

luminescence spectra as the crystal structure evolved from tetragonal to cubic. In

these studies we go a step further and look at the luminescence from Eu3+ in

partially stabilized zirconias as it pertains to changes in luminescence intensity and

luminescence lifetime. The development of a luminescence temperature sensor for

use in YSZ thermal barrier coatings requires the understanding of both the

luminescence intensity as a function of temperature as well as concentration and

simultaneous calibrations of the luminescence lifetime as a function of temperature.

Both of these behaviors are intimately linked to the crystal structure of the material

as well as the intrinsic rates of radiative and non-radiative luminescence decay for

the ion in YSZ. To determine if it is possible to increase the sensor sensitivity, we

carried out a series of experiments looking at the effects of Eu3+ doping in

conjunction with the effects of stabilizer content to examine the role ion-ion

interaction and crystal structure on the luminescence behavior of the material.

4.2 Luminescence Spectra

The luminescence from the trivalent Eu ions consists primarily of

transitions from the upper 5DJ=0-3 multiplet to the lower 7FJ=0-4 levels. Of particular

interest, are the 5D0à7F1 and 5D0à

7F2 transitions. The former being allowed as a

magnetic dipole transition and forbidden as an electronic dipole transition and the

latter allowed only as a forced electronic dipole transition induced when the Eu3+

site does not contain a center of symmetry[7]. As the deviation from inversion

symmetry increases, the intensity of the electric dipole transition increases with

respect to the intensity of the magnetic dipole transition. 7YSZ used in thermal

Page 86: Molly Gentleman Thesis 2006

76

barrier coatings, consists of a metastable tetragonal crystal structure, (P42/nmc).

This yields a D2d ( )m24 symmetry at the site of the Eu3+ site, which lacks a center

of symmetry. Based on selection rules on the total quantum number J [4, 7]both

the 5D0à7F1 and 7F2 transitions would be expected. Luminescence from a structure

containing D2d ( )m24 symmetry at the Eu3+ site is expected to also result in a

splitting of the 5D0à7F2 transition into two lines at ~610 and 630 nm along with a

concomitant splitting of the 5D0à7F1 level into two peaks[2]. If the symmetry is

further distorted to symmetry less than D2d ( )m24 due to the distortions caused by

oxygen vacancies in the crystal, both the 5D0à7F1 and 5D0à

7F2 transitions may

exhibit many more lines. For example, lowering to a Cs (m) symmetry for instance,

would result in one line observed for the 5D0à7F0 line, three for the 5D0à

7F1 line,

and five for the 5D0à7F2 line similar to what is seen in the spectra recorded for the

zirconia materials studied here (Figure 4.2.1).

Concentration quenching of the luminescence intensity, as discussed in

chapter 1, is a phenomenon often seen in luminescent materials where the

luminescence intensity increases linearly with increasing dopant concentration and

then begins to drop in intensity as ion-ion energy transfer results in energy transfer

to luminescence quenching sites. The decrease in intensity is the result of the

nearness of one luminescent ion to another significantly reducing the thermal

activation energy for non-radiative transitions from those ions resulting in

extinction of the luminescence. This behavior can be understood through use of

Page 87: Molly Gentleman Thesis 2006

77

Equation 1.7.1 described by Johnson and Williams [9]. This equation has been

used to successfully fit the data in Figure 1.7.2 as well as the data in Figure 4.2.2

but begins to break down at high dopant concentrations when the distortion caused

by large volume fraction of dopant ions distorts the crystal symmetry of the host

material. This is important in these studies because the intensity of the signal can

significantly hinder or increase the ability to measure the luminescence lifetime and

hence the temperature of a material.

4.3 Luminescence Lifetime

Although previous work has shown that there is no significant change in the

luminescence spectra as the dopant levels and crystal structures of partially

stabilized zirconias are varied [8], this is not expected to be the case for

luminescence lifetime behavior under the same conditions. When a photon excites

an electron in a luminescent ion into an excited state, multiple competing processes

dictate the rate of de-excitation of the electron. These processes include radiative

emission, energy transfer to other ions, energy dissipation to phonons, and energy

dissipation to defects and traps and are dependent on concentration of the dopant

ion, host crystal structure, and temperature of the system[10].

The rate of radiative emission of the isolated Eu3+ ion should remain

constant for a given host material, but all other processes are expected to vary with

both composition and temperature. The rate of energy transfer to other ions was

most appropriately described by Forster [11]. The dissipation of energy to phonons

has two important components. The first is the low temperature rate of

Page 88: Molly Gentleman Thesis 2006

78

multiphonon transitions. This value can be determined experimentally for an

isolated ion in a host and is exponentially dependent on the distance between the

excited state and next lower available electron level [12, 13]. The second

component depends on the number and energy of available phonon modes for the

host material and the strength of ion-phonon coupling and can be determined

experimentally from the high temperature luminescence lifetime. The processes

involved in ion-ion interaction also play a part in the luminescence lifetime of a

material. This is manifested in a change in luminescence lifetime with

concentration where 1/τ=1/τ0+1/τD where τ0 is the lifetime in an infinitely dilute

composition and τD is the lifetime of ion-ion energy transfer and dependent on the

average ion-ion distance. As the ion-ion energy transfer rate increases, there is a

corresponding decrease in the luminescence lifetime. This process may be

exaggerated with increasing temperature leading to even larger effects of ion-ion

interactions for high temperature luminescence measurements.

4.4 Experimental Procedure

All materials used in this study were made from reverse-coprecipitation of

aqueous zirconium acetate, yttrium nitrate, and europium nitrate as described in

chapter two. The materials were calcined at 950°C before being pressed into

pellets and sintered at 1200°C for 2 hours. The compositions made are represented

on the ternary phase diagram in Figure 4.4.1. Raman spectroscopy was used to

confirm the presence of a tetragonal phase in all materials (Figure 4.4.2). X-ray

analysis was then used to determine the c/a ratio for the tetragonal phase or

Page 89: Molly Gentleman Thesis 2006

79

determine if the cubic phase had developed if appropriate (Figure 4.4.3). The

materials consisted of two series: one where the total dopant and hence total

vacancy concentration was held at 7a/o and a second where the yttria content was

held constant and europia was added in excess of the yttria with Eu3+ ranging from

0.05 a/o to 5 a/o. This allowed two studies: the first to look at the effect of doping

concentration while keeping the total doping concentration at the level of doping

commonly used in thermal barrier materials, and the second to examine the effects

of defects introduced by oxygen vacancies.

Luminescence lifetimes and spectra in this work were all excited using a

frequency-doubled, Q-switched YAG:Nd laser emitting at 532 nm and with a pulse

length of ~10 ns. Both room and elevated temperature measurements were carried

out according to the guidelines outlined in chapter two. All Eu-doped zirconias

studied exhibited double exponential decays and were fit using Equation 1.4.2. The

long luminescence lifetime (τ2) as a function of temperature is plotted in Figures

4.4.5 and 4.4.6 for the two doping regimes.

4.5 Results and Discussion

In this work a series of parallel experiments on two types of europia-doped

yttria stabilized zirconia systems were carried out to determine the effects of dopant

concentration and oxygen vacancies on the luminescence intensity and lifetime.

The luminescence spectra reveal several interesting features about the nature of all

of the materials used in these studies. Luminescence spectra recorded for all

materials exhibit luminescence from all three transitions discussed previously

Page 90: Molly Gentleman Thesis 2006

80

(5D0à7FJ=0-2) as is illustrated in Figure 4.2.1. Specifically, in the EuxY0.07-

xZr0.93O1.965 “substitutional series”, where the total dopant concentration is

maintained at 7 a/o, the symmetry has distorted from a regular tetragonal material

resulting in as many as five peaks in the 5D0à7F2 transition in addition to the

presence of the 5D0à7F0 transition and the splitting of the 5D0à

7F1 level into three

peaks instead of the two expected. Correspondingly, the Raman and x-ray spectra

(Figures 4.4.2 and 4.4.3) illustrate that these materials are single-phase with

increasing tetragonality (Figure 4.4.4) as the europium concentration was increased.

The increase in concentration of europium ions (ionic size of 1.066 Å) for yttrium

(ionic size 1.019 Å [14]) causes an increase in the distortion in the crystal and

hence an increase in the distortion on the Eu3+ site. This is the source of the

additional peaks seen in the luminescence spectra of Figure 4.2.1.

The second, EuxY0.07Zr0.93-xO1.965-x/2 “additive series” of materials,

consisting of EuO1.5 added in addition to 7 a/o YO1.5, does not show an increase in

the number of observed peaks in the luminescence spectra with concentration. In

contrast to the substitutional-series, as the europium concentration increases, the

5D0à7F2 peak splitting becomes more distinct consisting of two primary peaks at

606 and 630 nm suggesting an increase, rather than decrease, in symmetry at the

Eu3+ site. Although the intensity continues to increase for all peaks with

concentration, the relative intensity of the 5D0à7F2 transition with respect to the

5D0à7F1 transition decreases with concentration also indicating that the symmetry

at the Eu3+ site is increasing with concentration. Both the Raman and X-ray spectra

Page 91: Molly Gentleman Thesis 2006

81

for these materials support this assertion. The Raman spectra, although apparently

exhibiting tetragonal structure at first glance, reveals the presence of a cubic peak

emerging at ~600 cm-1 upon peak fitting and spectral analysis, which strengthens as

the total dopant level is increased and the composition moves towards the cubic

field on the phase diagram. X-ray analysis of these materials likewise confirms the

emergence of a cubic phase as the dopant concentration is increased.

Luminescence spectra for all materials is this study are characterized by

large line widths on the range of fifty to hundreds of wavenumbers. These values

are typical of line-widths measured for Eu3+ in glassy matrixes[8, 15]. This

suggests that although these Eu-doped zirconias exhibit diffraction patterns, the

fluctuations in the local environment from one europium ion to the next due to the

defect structure of the material yields a mean measurement of Eu3+ environment by

luminescence. This is opposed to specific information of the Eu3+ ion site that is

often gleaned from site selective spectroscopy of rare earth materials in crystalline

hosts. This must be taken into consideration when studying all luminescence data

from these materials because strict rules of symmetry do not apply because of this

averaging over all the Eu3+ sites.

The luminescence intensity as a function of concentration for these two

series of materials show that for the substitutional-series, with constant vacancy

concentration, has a maximum in luminescence intensity for the 5D0à7F2 transition

at approximately 2 a/o EuO1.5. The additive-series, doped to approach the cubic

field in the ternary phase diagram, does not reach a maximum in intensity even at 5

Page 92: Molly Gentleman Thesis 2006

82

a/o EuO1.5. The intensity of the luminescence spectra must be considered in

conjunction with the luminescence lifetime as a function of temperature as it

appears that both are capable of being the limiting factor in the luminescence

lifetime measurement.

The luminescence lifetime as a function of temperature data for the two

series of materials also show dissimilar results (Figures 4.4.5 and 4.4.6). Both

series exhibited the same basic variation of the lifetime decay time with

temperature; a relatively constant, almost temperature independent, value from

room temperature up to an intermediate temperature followed by and exponential

decrease with increasing temperature. The low temperature luminescence lifetime

of the substitutional series shows a trend similar to that seen in the intensity as a

function of concentration studies with an increase in luminescence lifetime with

concentration up to 2 a/o EuO1.5 and then decreases with increasing concentration

(Figure 4.4.5). The high temperature behavior displayed uniform lifetime with

temperature for concentrations up to 1 a/o where upon the rate of change of the

luminescence lifetime with temperature began to increase with increasing

concentration and can be fit using a multiphonon model. The high-temperature

luminescence lifetime concentration-dependence is dependent on the increased rate

of ion-ion interactions at those elevated concentrations and temperature.

The behavior of the additive-series is simpler to explain (Figure 4.4.6). The

low-temperature luminescence lifetime increases monotonically with concentration.

This is in contrast to behavior seen in other concentration studies where the lifetime

Page 93: Molly Gentleman Thesis 2006

83

generally decreases as the dopant concentration is increased[16], but may be

attributed to in increase in decay time as the excitation energy is able to transfer

from one ion to another effectively increasing the amount of time spent in an

excited state. The high temperature behavior of these materials shows no

sensitivity to concentration and has the same values of luminescence lifetime as a

function of temperature as for Eu0.05Y0.02Zr0.93O1.965, the highest dopant

concentration of the substitutional-series. This suggests that the limiting factor in

the luminescence lifetime at high temperature for these highly doped materials

appears to be not only the amount of Eu3+, but more importantly, the total amount

of trivalent ions and hence oxygen vacancies in the material. The presence of

additional Y3+ atoms appears to have the same effect as the additional distortion of

the larger Eu3+ atoms in the substitutional-series in terms of increasing the rate of

non-radiative decay and hence, decrease the luminescence lifetime.

4.6 Summary

There are several findings of technological interest uncovered in the course

of this chapter. The first is that Eu-doped YSZ zirconia systems can be used to

measure temperatures in thermal barrier systems ranging from ~500 to 1150 ºC.

These temperatures are similar to those seen in current gas turbine system and

thermal barrier coatings. Second, and more specific to the studies of this chapter, is

a basis for the choice of compositions for specific temperature sensing applications

is outlined.

Page 94: Molly Gentleman Thesis 2006

84

The substitutional-series of doping provides several benefits over the

additional-series. The first is the higher temperature sensitivity of the lower

concentration materials. Compositions ranging from 0.1 a/o to 1 a/o EuO1.5 have

exhibited temperature sensitivity up to ~1150 ºC allowing temperature sensing in

that range assuming the signal is strong enough to be efficiently collected.

Considering the intensity data of Figure 4.2.2, the highest temperature sensing

capabilities would be for a materials composition of Eu0.01Y0.06Zr0.93O1.965 both in

terms of intensity as well as temperature sensitivity. Luminescence lifetime of this

composition is also insensitive to small changes in composition making it appealing

as small changes in composition due to deposition conditions or aging of the

material would not invalidate the temperature measurement.

The additive-series of materials poses some interesting questions in terms of

selecting materials for sensor applications. Although the intensity of the

luminescence increases continually as the dopant level increases, so does the

instability of the tetragonal phase vital to TBC life. This effect coupled with the

fact that there is no change in the luminescence lifetime at elevated temperature

with concentration and the temperature sensing capabilities are limited to ~1100 ºC

suggest that this series of materials may not be well suited for temperature sensing

in thermal barrier materials.

The results so far suggest that Eu-doped materials can be employed with the

proper doping concentrations as temperature sensing materials in thermal barrier

materials. The work in this chapter has shown that doping must be done in a way

Page 95: Molly Gentleman Thesis 2006

85

to preserve the tetragonal prime crystal structure as well as in a doping regime with

high concentration and low lifetime sensitivity to concentration. This generally

consists of materials in the range of Eu0.005Y0.065Zr0.93O1.965 to

Eu0.015Y0.055Zr0.93O1.965. Although doping in addition to 7 a/o YO1.5 yields

materials with many appealing features for temperature sensors, they cannot be

applied to TBC systems for the reasons discussed above. Sensors of the

compositions discussed here have been deposited by electron beam physical vapor

deposition and used to measure temperature in thermal gradients and will be

discussed in detail in chapter 6.

Page 96: Molly Gentleman Thesis 2006

86

1. Gentleman, M.M. and D.R. Clarke, Luminescence sensing of temperature in pyrochlore zirconate materials for thermal barrier coatings. Surface & Coatings Technology, 2005. 200: p. 1264.

2. Blasse, G. and A. Bril, On the Eu3+ flourescence in mixed metal oxides. V. The Eu3+ flourescence in the rocksalt lattice. J. Chem. Phys., 1966. 45(9): p. 3327.

3. Blasse, G., A. Bril, and W.C. Nieuwpoort, On the Eu3+ fluorescence in mixed metal oxides Part I. J. Phys. Chem. Solids, 1966. 27: p. 1587-1592.

4. Judd, B., Optical absorption intensities of rare-earth ions. Physical Review, 1962. 127(3): p. 750-761.

5. McCauley, R.A. and F.A. Hummel, Luminescence as an indication of distortion in A2 3+ B2 4+ O7 type pyrocholores. J. Lum., 1973. 6: p. 105.

6. Nieuwpoort, W.C. and G. Blasse, Linear crystal-field terms and the 5D0 - 7F0 transition of the Eu3+ ion. Sol. Stat. Commun., 1966. 4: p. 227.

7. Ofelt, G., Intensities of crystal spectra of rare-earth ions. J. Chem. Phys., 1962. 37(3): p. 511-519.

8. Dexpert-Ghys, J., M. Faucher, and P. Caro, Site selective spectroscopy and structural analysis of ytrria-doped zirconias. J. Solid State Chem., 1984. 54: p. 179-192.

9. Johnson, P. and F. Williams, Interpretation of the dependence of luminescence efficiency on activator concetration. J. Chem. Phys., 1950. 18(11): p. 1477-1483.

10. Riseberg, L.A. and M.J. Weber, Relaxation phenomena in rare-earth luminescence, in Progress in Optics, E. Wolf, Editor. 1976, North-Holland. p. 89.

11. Forster, R., Ann. Physik, 1948. 2(55). 12. Weber, M.J., Multiphonon relaxation of rare-earth ions in yttrium

orthoaluminate. Phys. Rev. B, 1973. 8(1): p. 54. 13. Yamada, Shionoya, and Kushida, J. Phys. Soc. Jap., 1972. 32: p. 1577. 14. Shannon, R.D., Revised effective ionic radaii and systematic studies of

interatomic distances in hallides and Chalcogenides. Acta Crystallogr., 1976. A(32): p. 751.

15. Dieke, G.H., Spectra and Energy Levels of Rare Earth Ions in Crystals, ed. H.M. Crosswhite and H. Crosswhite. 1968, New York: Interscience Publishers.

16. Ryba-Romanowski, W., Effect of temperature and activator concentration on luminescence decay in erbium-doped tellurite glass. J. Lum., 1990. 46: p. 163.

Page 97: Molly Gentleman Thesis 2006

87

Figure 4.2.1 Emission spectra for Eu-containing YSZ materials for several

different compositions over two doping regimes (a) substitutional EuxY0.07-

xZr0.93O1.965 and (b) additional-series EuxY0.07Zr0.93-xO1.965-x/2. All materials were excited with a 532 nm frequency doubled YAG:Nd laser.

Page 98: Molly Gentleman Thesis 2006

88

Figure 4.2.2 Luminescence intensity verses concentration of Eu3+ in YSZ with

different amounts of Y3+ illustrating a peak in intensity around 2 a/o for Eu-doped materials with a constant amount of 7 a/o total stabilizer while the Eu-doped zirconias with 7 a/o YO1.5 in addition to europia doping have not reached a peak in intensity by 5 a/o EuO1.5. The data are fits according to Equation 1.7.1 with fitting parameters of z=0.74 and A=38 for the substitutional series and z=0.032 and A=1.5 for the additive series.

Page 99: Molly Gentleman Thesis 2006

89

Figure 4.4.1 Diagram represented the thermodynamic phase stability of the

zirconia, yttria, europia system illustrating the compositions synthesized for this work. Where t and c denote the tetragonal and cubic phases respectively.

Page 100: Molly Gentleman Thesis 2006

90

Figure 4.4.2 Raman spectra of representative materials from the (a)

substitutional- and (b) additional-series showing the characteristic peaks for tetragonal zirconia. The data have been offset for comparison. There is a shoulder in the additional-series spectra at ~600 cm-1 denoting the emergence of a cubic phase in the materials with concentration.

Page 101: Molly Gentleman Thesis 2006

91

Figure 4.4.3 X-ray spectra of the (a) substitutional- and (b) additional-series of

yttria stabilized zirconias illustrating the tetragonality of the materials or the presence of the cubic peak in the additional-series.

Page 102: Molly Gentleman Thesis 2006

92

Figure 4.4.4 Calculated c/a ratio for the x-ray data seen in Figure 4.4.3.

Page 103: Molly Gentleman Thesis 2006

93

Figure 4.4.5 Luminescence lifetime decay as a function of temperature for the

substitutional-series. The data are fit using a multiphonon model.

Page 104: Molly Gentleman Thesis 2006

94

Figure 4.4.6 Luminescence lifetime decay as a function of temperature for the

additional-series. The lines through the data have been drawn to aid the eye and are not an actual fit of the data.

Page 105: Molly Gentleman Thesis 2006

95

Chapter 5: Luminescence Thermometry of (Gd,Eu)2Zr2O7

5.1 Introduction

Rare earth zirconates are an alternative to yttria stabilized zirconia as

thermal barrier materials and therefore must be studied in the context luminescence

temperature sensors. Initial studies discussed in chapter 3 have revealed a

considerable divergence of the luminescence lifetime behavior between Eu-doped

gadolinium zirconate and fully concentrated europium zirconate[1]. Gadolinium

zirconate doped with 1 a/o EuO1.5 has been shown to have a luminescence lifetime

with temperature sensitivity from 400-1150°C. Europium zirconate has been

demonstrated to show temperature sensitivity of the luminescence in excess of

1280°C. The large disparity between the temperature sensitivity of these two

compounds and more generally the concentration dependence of luminescence

lifetime of the 5D0 state on temperature in (EuxGd1-x)2Zr2O7 the solid solution series

are the subject of this chapter. Spectral and Raman analysis are utilized to

examine the relationship between crystal structure, ion-ion interaction,

luminescence intensity, and the luminescence decay behavior as the composition is

systematically varied across the phase diagram from (Eu0.01,Gd0.99)2Zr2O7 to

Eu2Zr2O7.

Luminescence lifetime measurements of temperature traditionally rely on

luminescence from dilute solutions of chromophore-doped materials where

luminescent ions can be assumed to be isolated. In dilute materials such as those

Page 106: Molly Gentleman Thesis 2006

96

discussed in chapter four, the temperature dependence of luminescence lifetime is

dependent on the radiative rate of the ion as well as the rate of multiphonon de-

excitation of the luminescence[2, 3]. As the concentration of luminescent ion

increases, such as in the case of moving from (Eu0.01Gd0.99)2Zr2O7 to Eu2Zr2O7,

competing processes such as ion-ion interactions and excitation energy migration

begin to play in increasing role in the temperature dependence of the luminescence

lifetime[4]. Ion-ion interactions dominated by excitation energy migration, a

resonance energy process where an excited rare earth ion shifts from the excited

state to the ground state while an unexcited coupled ion makes the opposite

transition. Because there is no removal of excitation energy between the new

excited-state and the former excited-state, no distinction can be made between the

two states. This process, similar to concentration quenching, can accelerate the

decay rate of the excitation by energy transfer to ions acting as energy sinks. The

rate of the quenching depends on the separation between interacting ions and

therefore the doping level of the material and the likelihood of finding the excited

energy landing on a defect. When the transition between ions is resonant, such as

direct energy transfer between ions with the same excited states, the rate at which

the luminescence is quenched is very high because energy transfer can occur

without the assistance of phonons. For energy transfer between states that are of

slightly different energies, such as in the case or luminescent ions in a distorted

crystal structure similar to those that will be discussed here, the process is slow at

low temperatures and increases as temperature increases and the population of

Page 107: Molly Gentleman Thesis 2006

97

phonons required to bridge the energy mismatch is increased. When Gd3+ ions are

introduced into a matrix of Eu3+ ions, they can serve to distort the energies of the

Eu3+ ions slightly and hence increase the non-radiative decay rate of the material.

One would expect that this would result in a decrease in the luminescence lifetime

in these materials.

5.2 Experimental

The europium-doped zirconates used in this study were prepared from

powders synthesized by reverse co-precipitation of aqueous rare earth nitrate and

zirconium acetate solutions. Compositions were made at 10 a/o intervals along the

solid solution line between Gd2Zr2O7 and Eu2Zr2O7 (Figure 5.2.1). Following

precipitation, the powders were calcined at 950 °C for 2 hours, pressed into pellets,

and sintered at 1200 °C for two hours. Raman spectroscopy was used to determine

the extent of ordering of the pyrochlore crystal structure in the materials. Raman

spectra for compositions from (Eu0.01Gd0.99)2Zr2O7 to Eu2Zr2O7 after sintering at

1200 °C are seen in Figure 5.2.2. The arrows in Figure 5.2.2 indicate the peak

positions for an ordered europium zirconate pyrochlore[5]. The Raman spectrum

for an ordered gadolinium zirconate pyrochlore would show a similar peak

structure with the peak energies shifted to larger Raman shifts by approximately 2-

10 cm-1. The spectra from the zirconates confirms not only that all the materials

examined are single phase, but also that there are different degrees of ordering for

different compositions. All compositions appear to be disordered after initial heat

treatments, but the shift of peaks to smaller Raman shifts, as the europium

Page 108: Molly Gentleman Thesis 2006

98

concentration is increased illustrate the trend towards increased ordering with the

increase in europium concentration. This is expected since the order-disorder

transition temperature is higher in Eu2Zr2O7 than in Gd2Zr2O7.

The luminescence studies reported in this work were all performed using a

frequency double, Q-switched YAG:Nd laser emitting at 532 nm and with a pulse

length of ~10 ns. Room temperature spectra were collected using a standard silica

fiber-optic cable fitted with an in-line 532 nm laser filter to block scattered laser

light and recorded using an Ocean Optics USB 2000 (Figure 2.3.2.1). High

temperature luminescence lifetime measurements were taken according to the

details outlined in chapter two.

The spectra were fit and the normalized peak intensities were recorded and are

plotted as a function of Eu composition (Figures 5.2.3 and 5.2.4).

5.3 Room temperature luminescence

Luminescence spectra of six representative compositions in Figure 5.3.1

show the three principal luminescence lines from Eu3+ illustrating the 5D0 à 7F0,

7F1, and 7F2 transitions as labeled. All compositions show similar spectra with

variation in the relative peak intensities and splittings as a function of composition.

The similarity in luminescence spectra is in accordance with expectation that

luminescence from 4f electrons is relatively insensitive to small changes in the host

as the 4f electrons are shielded from the host by the outer 5s and 5p electrons[6].

As the composition of the europium is increased, the luminescence spectra show a

decrease in the relative intensity of the 5D0à7F0 transition with respect to the much

Page 109: Molly Gentleman Thesis 2006

99

stronger 5D0à7F1 and 5D0à

7F2 transitions. This is not a surprise as this transition

is highly forbidden as it is a J=0 à J=0 transition and its presence in the

pyrochlore crystal structure suggests that the symmetry around the Eu3+ site has

been broken allowing a linear term in the crystal-field expansion [7-9]. The

breaking of D3d symmetry at the europium atom site allowing the above transition

is likely associated with the disorder seen in the Raman spectra of these materials.

The decrease in intensity of the 5D0à7F0 transition corresponds with the ordering

of the pyrochlore with composition as seen in the Raman spectra.

In the pyrochlore crystal structure a D3d symmetry at the site of the trivalent

rare-earth ion results in a splitting of the 7F1 level into two components of A2g and

Eg symmetry resulting in a singlet and doublet respectively from the 5D0 level (A1g)

as observed in Figure 5.3.1. Consequently, this splitting can then be used as a

direct indication of the distortion of the pyrochlore[10]. In Figure 5.3.1, the

splitting of this transition decreases as the total amount of europium is increased

indicating a total decrease in the distortion at the Eu3+ in the as the europium

concentration is increased. This confirms the results of the same effect observed by

the ordering in the Raman spectra.

The 5D0à7F2 transition, like the 5D0à

7F0 transition, is allowed only as a

forced electric-dipole transition in the pyrochlore crystal structure. The presence

and high intensity of this transition in all spectra recorded in this study suggest that

there is a deviation from inversion symmetry at the Eu3+ site resulting in a lowering

of symmetry to D2d. This explains not only the intensity of the peaks, but also the

Page 110: Molly Gentleman Thesis 2006

100

splitting of the 7F2 level. The intensities of the 5D0à7F2 peaks as a function of

concentration are seen in Figure 5.2.3 and are fit using the Johnson and Williams

equation (Equation 1.7.1). The data shows an increase in luminescence intensity as

a function of concentration up to 30-40% europium where the intensity quickly

begins to drop with concentration. The initial rise in intensity with concentration is

due to increase in Eu3+ concentration coupled with a high level of disorder in the

material as seen by the Raman analysis. The decrease in intensity following this

maximum is not fit well using Equation 1.7.1 and is likely related to the ordering of

the pyrochlore that leads to a diminished intensity as the distortion in the

pyrochlore decreases forbidding the transition as evidenced by both the

luminescence and Raman spectra. The high concentration region of this data can

more accurately be fit using an exponential decay Figure 5.2.4.

5.4 High temperature luminescence

Measurements of luminescence lifetime of the 610 nm line as a function of

temperature were made for all compositions from room temperature until the

luminescence lifetime decay was faster than the detectability limit (~10 ns) of our

system or the intensity of the luminescence signal fell below the noise floor of our

current detection system. The 610 nm line was chosen for luminescence lifetime

measurements for its strong intensity even in the materials that exhibited ordering

in the Raman spectra. All materials, with the exception of europium zirconate,

exhibited a double exponential decay and were fit according to Equation 1.4.2.

Europium zirconate exhibited a single exponential decay over the entire

Page 111: Molly Gentleman Thesis 2006

101

temperature range examined. The longer of the two exponentials (τ2), or the single

decay time in Eu2Zr2O7, are plotted as a function of temperature for all the

materials in Figure 5.4.1 a and b.

The luminescence lifetime as a function of temperature exhibited the same

characteristic features for all materials examined. From room temperature to some

intermediate temperature, the luminescence lifetime was relatively independent of

temperature. The range over which the lifetime was independent of temperature

varied with composition having the greatest extent for pure europium zirconate

ranging from room temperature to ~700 °C. At higher temperatures, the

luminescence lifetime displayed an exponential dependence of lifetime on

temperature. The rate of change of the luminescence lifetime as a function of

temperature in this region also displayed dependence on europium concentration.

The slope of the luminescence lifetime as a function of temperature varied without

any discernable trend for compositions consisting of less that 50 percent europium,

but for compositions from 60 to 100 percent the rate of change of the luminescence

lifetime as a function of temperature decreased with increasing concentration

(Figure 5.4.2). While for 60% europium the lifetime detection limit of ~10 ns was

reached by 1100 °C, pure europium zirconate retained lifetimes as long as 1 µs at

temperatures above 1200 °C.

To evaluate the effects of ordering on the luminescence lifetime and

luminescence intensity, the europium zirconate sample was examined as a function

of Raman ordering. The as-sintered pellet was given a secondary heat treatment at

Page 112: Molly Gentleman Thesis 2006

102

1400 °C for 5 hours. Figure 5.4.3 clearly illustrates the increase in ordering of the

pyrochlore materials. The pyrochlore Raman peaks 311 and 397 cm-1 are seen to

have sharpened significantly upon aging and the peak at 530 cm-1 has emerged

while the shoulder at 650 cm-1, characteristic of the disordered structure, has

receded. Following the heat treatment, the luminescence lifetime as a function of

temperature was also re-examined Figure 5.4.4. The most significant change in the

data from the as-sintered state was the diminished intensity of the luminescence

signal. This is illustrated by the increase in scatter of the data at elevated

temperatures where the signal to noise ratio was decreased by ordering of the

material. Surprisingly, while the overall luminescence intensity of the signal was

decreased for intermediate temperatures, it was possible to extend the luminescence

lifetime measurement by approximately 100 °C following heat treatment allowing

temperature sensitivity up to ~1300 °C.

5.5 Discussion

The luminescence spectra reveal some interesting correlations with the

crystal structure of the materials. As the total amount of europia is increased in the

materials both the Raman spectra as well as the luminescence spectra show that

there is a continuing decrease in the distortion and disorder of the pyrochlore

crystal structure as is illustrated by the narrowing of the Raman peaks and the

decrease in the splitting of the 5D0 à 7F1 peak. This is not a surprise as the order-

disorder transition temperature is higher in Eu2Zr2O7 than in Gd2Zr2O7, but neither

luminescence intensity nor luminescence lifetime as a function of temperature

Page 113: Molly Gentleman Thesis 2006

103

shows a corresponding monotonic trend. As seen in Figure 5.4.1 compositions

with between 1 and 50 a/o europia in the (EuxGd1-x)2Zr2O7 the luminescence

lifetime as function of temperature varies without a discernable trend while the

intensity increases almost linearly over that range (Figure 5.3.2). Part of the

explanation for this is likely due to the fact that these oxides are exhibiting

extensive site disorder as well as site defects whereas the more ordered materials

discussed later show more distinct trends. In this range of compositions, we

believe that there are several processes competing for control of the luminescence

lifetime as a function of temperature. These include the radiative rate of the Eu3+

ion in the host material, ion-ion interaction caused by the addition of more Eu3+

ions, and the multiphonon relaxation of luminescence and the changes in that

process as the crystal structure evolves changing phonon energies. This results in a

very complicated picture that will take further work to resolve more completely.

For compositions above ~60 a/o EuO1.5 the phenomena appears to be

simplified as the luminescence lifetime is dominated by the ion-ion interaction as a

function of temperature. This results in a linear relationship between the slopes of

the luminescence lifetime as a function of temperature over the exponential region

with composition (Figure 5.4.2). This linear relationship allows for the engineering

of compositions capable of measuring temperature over specific temperature ranges

within this range. The simplest of these oxides is europium zirconate. As

discussed in chapter two, this material, in addition to having temperature sensitivity

to the highest temperatures recorded in this study also exhibited the simplest decay

Page 114: Molly Gentleman Thesis 2006

104

behavior of a single exponential decay for all temperatures observed. This suggests

that the presence of Gd3+ may be the source of the defects that cause the fast initial

decay. We propose the Eu3+ ions that are nearest to Gd3+ ions act as ions sitting

next to a defect and have a faster decay time as observed in the fast initial

luminescence decay observed in all Gd3+ containing materials. This may also

explain the decrease in luminescence lifetime as a function of temperature for these

materials as the introduction of defects would be expected to decrease the observed

lifetime by providing addition paths for non-radiative de-excitation.

These results suggest that the use of europium zirconates is necessary to

achieve the highest temperature sensitivity of from the materials. Conversely, as

show in the case of Eu2Zr2O7 this may not be the best choice in term of intensity

either because of concentration quenching as discussed in chapter one or as the

result of changes in intensity due to the evolution of the crystal’s site symmetry.

Both of these factors must be considered when choosing a material for a particular

application. For example, when choosing a sensor for the outermost surface of the

TBC, the luminescence will not need to pass through the coating and therefore

intensity will be less important than the ability to measure the high temperatures

experience by this part of the TBC therefore, the use of Eu2Zr2O7 would be the

preferred material for this sensor. Conversely, a sensor located deeper within the

coating, requiring higher intensity to pass its signal through the thickness of the

coating but with less stringent temperature sensing requirements, may be better

served with a lower Eu-containing composition.

Page 115: Molly Gentleman Thesis 2006

105

5.6 Summary

The principal findings of technical interest of this work are that there is a

significant dependence of both the luminescence lifetime and intensity of these

materials on europium concentration. Gadolinium zirconate doped with 1 a/o

europia can be used to measure temperatures ranging from ~500 to 1150 °C while

pure europium zirconate has a temperature sensitivity range of ~ 700 to 1300 °C.

These temperatures are similar to those seen in current gas turbine systems and

thermal barrier systems, but as discussed in chapter three there is potential for

temperature sensing with these materials to even higher temperature through the

use of higher sensitivity detection and collection equipment. The trends governing

the temperature sensitivities of the intermediate compositions are extremely

complicated. There are two significant considerations when choosing a dopant

concentration for the temperature sensing of a (EuxGd1-x)2Zr2O7 material. The first

is the temperature range of interest and the second is the importance of the intensity

of the signal in the measurement as this study has shown that these two features of

the measurement are at odds in this system.

Page 116: Molly Gentleman Thesis 2006

106

1. Gentleman, M.M. and D.R. Clarke, Luminescence sensing of temperature in pyrochlore zirconate materials for thermal barrier coatings. Surface & Coatings Technology, 2005. 200: p. 1264.

2. Moos, H., Spectroscopic relaxation processes of rare earth ions in crystals. J. Lum., 1970. 1(2): p. 106-121.

3. Weber, M.J., Multiphonon relaxation of rare-earth ions in yttrium orthoaluminate. Phys. Rev. B, 1973. 8(1): p. 54.

4. Ryba-Romanowski, W., Effect of temperature and activator concentration on luminescence decay in erbium-doped tellurite glass. J. Lum., 1990. 46: p. 163.

5. Michel, D., M. Perez Y Jorba, and R. Collongues, Study by raman spectroscopy of order-disorder phenomena occurring in some binary oxides with fluorite-related structures. J. Raman Spec., 1976. 5(3): p. 163.

6. Henderson, B. and G.F. Imbusch, Optical Spectroscopy of Inorganic Solids. 1989, New York: Oxford University Press.

7. Dieke, G.H., Spectra and Energy Levels of Rare Earth Ions in Crystals, ed. H.M. Crosswhite and H. Crosswhite. 1968, New York: Interscience Publishers.

8. Blasse, G. and A. Bril, On the Eu3+ fluorescence in mixed metal oxides II. the 5D0 - 7F0 emission. Phillips Res. Repts., 1966. 21: p. 368-378.

9. Blasse, G., A. Bril, and W.C. Nieuwpoort, On the Eu3+ fluorescence in mixed metal oxides Part I. J. Phys. Chem. Solids, 1966. 27: p. 1587-1592.

10. McCauley, R.A. and F.A. Hummel, Luminescence as an indication of distortion in A2 3+ B2 4+ O7 type pyrochlores. J. Lum., 1973. 6: p. 105.

Page 117: Molly Gentleman Thesis 2006

107

Figure 5.2.1 Diagram representing the compositions synthesized for these

studies.

Page 118: Molly Gentleman Thesis 2006

108

Figure 5.2.2 Raman spectra of the zirconate materials examined in this work after

heat treatment at 1200 ºC for 2 hours. Ordering in the zirconates is indicated by the presence of additional Raman peaks denoted by the arrows. The spectra illustrate an increase in ordering of the materials as the composition of EuO1.5 is increased.

Page 119: Molly Gentleman Thesis 2006

109

Figure 5.2.3 Luminescence intensity of Eu3+ ions in (Eu,Gd)2Zr2O7 as a function

of concentration illustrating concentration quenching. The high concentrations at which quenching occurs coupled with the ordering data colleted in the Raman data suggests that pyrochlore ordering is primarily responsible for the quenching of luminescence at high concentrations.

Page 120: Molly Gentleman Thesis 2006

110

Figure 5.2.4 Luminescence intensity of of Eu3+ ions in (Eu,Gd)2Zr2O7 as a function of concentration. The data suggests that the intensity decreases

exponentially with concentration.

Page 121: Molly Gentleman Thesis 2006

111

Figure 5.3.1 Luminescence spectra in the visible for six representative

compositions of the (EuxGd1-x)2Zr2O7 system showing the three principal luminescence transitions of the trivalent europium ion.

Page 122: Molly Gentleman Thesis 2006

112

Figure 5.4.1 Luminescence lifetime decay time as a function of temperature for

(EuxGd1-x)2Zr2O7 for (a) x = 0.01 to x = 0.50 and (b) x = 0.60 to x = 1.0. The lines have been drawn to aid the eye and are not a fit of the data.

Page 123: Molly Gentleman Thesis 2006

113

Figure 5.4.2 Slope of the luminescence lifetime as a function of temperature

plotted as a function of concentration.

Page 124: Molly Gentleman Thesis 2006

114

Figure 5.4.3 Raman spectra of Eu2Zr2O7 taken before and after heat treatment at

1400 ºC for 5 hours. Following aging the Raman spectra shows clear evidence of increased ordering of the pyrochlore structure.

Page 125: Molly Gentleman Thesis 2006

115

Figure 5.4.4 Luminescence lifetime decay time as a function of temperature for

Eu2Zr2O7 before and after heat treatment at 1400 ºC for 5 hours. Aging resulted in decreased luminescence intensity resulting in the increased scatter in the data.

Page 126: Molly Gentleman Thesis 2006

116

Chapter 6 Non-contact Sensing of TBC/BC Interface Temperature in a Thermal Gradient

6.1 Introduction

The use of phosphors to measure temperature dates back to the work of

Bradley [1] and has been reviewed in detail by Allison and Gillies [2]. More

recently the concept has been shown to be applicable to TBC materials and

coatings [3-6]. The previous chapters of this work explored the use of rare-earth

ions as temperature sensing dopants in thermal barrier materials. Luminescence

lifetimes of the 5D0à7F2 transition of trivalent europia ions in yttria-stabilized

zirconia exhibited the ability to measure temperature from 500-1150 °C[3, 4]. The

luminescence decay times of these materials at high temperatures have also been

shown to be sufficiently short that the measurements can be made at speeds

considerably faster than the rotational speed of gas turbines[7].

Of particular interest is the deposition of temperature-sensing layers

consisting of rare-earth doped TBC materials at the coating/bond-coat interface

used to measure the temperature at that interface. In situ temperature

measurements from such a layer can provide vital information about the remaining

life of the coating as the oxidation life of a TBC system is exponentially dependent

on the temperature. In chapter 1, it was described that thermal barrier coatings with

rare-earth doped layers were deposited at Howmet using electron-beam physical

vapor deposition (EB-PVD) and the luminescence was excited through undoped

zirconia coatings with several laser frequencies [4]. In this chapter, thin (~10 µm)

Page 127: Molly Gentleman Thesis 2006

117

layers of europia-doped yttria-stabilized zirconia were deposited by EB-PVD onto

bond-coated CMSX-4 superalloy buttons to achieve sensor layers located next to

the TBC/BC interface. These coatings were then used to measure the interface

temperature while under thermal gradients similar to those in typical turbine

environments. The use of superalloys and bond coats was chosen to mimic the

conditions of a turbine environment as closely as possible. Combined with

pyrometric measurements of the coating-surface temperature and metal-surface

temperature, continuity of heat flux calculations yielded the thermal conductivity of

the coating (1.5 W/mK) and heat flux (~1 MW/m2) experienced in the tests. These

values were then used to confirm the validity of the sensor temperature

measurement. These measurements are a vital component in the integration of

sensor layers into commercial turbine coatings for health monitoring.

6.2 Sensor Layer Deposition

The coatings were deposited by electron beam evaporation at Howmet. The

ingot for EB-PVD of the sensor layers was prepared by infiltration of a standard

7YSZ ingot with aqueous Eu(NO3)3. The nitrate solution was gelled into the pores

of the ingot with NH4OH and calcined at 950°C to yield a composition of

Y0.06Eu0.01Zr0.93O1.965 as described in chapter two. The layered structure was

achieved in a single run of EB-PVD by layering the doped ingot with a standard

undoped 7YSZ ingot. The final coating consisted of two layers, an Eu-doped layer

approximately 10 µm in thickness at the base of the coating, and an undoped layer

on top, approximately 140-170 µm thick (Figure 6.1.1). The coatings were

Page 128: Molly Gentleman Thesis 2006

118

deposited on CMSX-4 buttons, each 2.54 cm (1 inch) diameter, 3mm thick with

MDC-150L bond coats.

For comparison, bulk ceramic materials of the same composition were

fabricated by reverse co-precipitation with final sintering at 1200°C for 2 hours.

Raman spectroscopy was used to confirm that the bulk material was single phase

and had the tetragonal-prime crystal structure found in the coated material.

6.3 Luminescence Lifetime Calibrations

On the basis of the work in chapter three [3], the luminescence emission

spectra were excited for both the bulk and coated materials using a frequency

doubled, Q-switched, YAG:Nd laser emitting at 532 nm and having a pulse length

of ~10 ns. To develop the calibration curves relating luminescence lifetime-decay

to temperature, luminescence measurements were first recorded under isothermal

conditions inside a box furnace with a thermocouple placed next to the coatings and

bulk samples to record temperatures. The luminescence lifetimes were recorded

using the high temperature luminescence setup described in detail in chapter two

(to measure the luminescence lifetime of the 606 nm line of the Eu3+

luminescence). Isothermal measurements of luminescence intensity as a function

of time were taken for temperatures between room temperature and the

detectability limit of the collection electronics (approximately 0.01µs decay time).

These measurements were repeated for selected temperatures after both the thermal

gradient testing discussed below as well as after a series of one hour cycles at 1150

°C. These subsequent studies were done to confirm that there was no change in the

Page 129: Molly Gentleman Thesis 2006

119

luminescence lifetime as a function of temperature as a result of the thermal

gradient testing or upon more extensive aging of the materials used in this study.

6.4 Luminescence Lifetime Measurement in a Thermal Gradient

Several coatings were subjected to high heat-flux exposure at NASA Glenn

Research Center and luminescence lifetime measurements were made under

different heat flux conditions. In the NASA rig, a 3.0 kW CO2 laser (10.6µm

wavelength) was used to heat the outer surface of the coating and the back of the

superalloy was cooled with a high-pressure compressed air jet (Figure 6.3.1). This

generated a heat flux through the coating and the underlying superalloy. The

instrumental details of the high heat-flux laser rig have been described elsewhere

[8]. The temperature at the top surface of the coating was monitored using an 8 µm

infrared pyrometer while a series of two-color pyrometers were used to monitor the

temperature of the metal surface. The CO2 laser delivered a constant power of

1060 W to the top surface of the coating and by adjusting the compressed air flow,

the metal temperature, the top surface temperature, and the heat flux could be

varied.

To interrogate the sensor layer while the sample was exposed to the thermal

gradient, a solid-state frequency doubled YAG:Nd laser emitting at 532 nm was

directed to illuminate the center of the coating surface while mounted in the high

heat-flux rig. As in the isothermal furnace tests, the end of a sapphire light pipe

was positioned to collect the excited luminescence signal and the luminescence

intensity was recorded and averaged in the same way as described in the previous

Page 130: Molly Gentleman Thesis 2006

120

section. The one difference was that for safety reasons the measurements were

made remotely under computer control.

6.5 Results

First, we present the luminescence lifetime as a function of temperature

calibration data. As described in the previous section, the calibration curves of

luminescence lifetime as a function of temperature were determined for both bulk

ceramic samples and coatings. Both forms of the sensor material exhibited double

exponential decays over the entire temperature range and were therefore fitted

using the function:

2121

ττ tt eIeII −− += Equation 1.4.2

where I is the intensity and τ the luminescence lifetime for the individual decays

(Figure 6.4.1.). The second, longer luminescence lifetime (τ2) was used for

calibration purposes and its value is shown as a function of temperature for both the

bulk and coatings in Figure 6.4.2. The second, longer decay was chosen for its

length, which allowed measurement to higher temperatures than the shorter decay,

and because it was not dependent on the excitation intensity. The luminescence

lifetime data was consistent with a temperature dependent multi-phonon relaxation

model[9]:

( ) pmprtotal nwww )1(0 ++=

Equation 1.4.5

where wr is the rate of radiative emission, wmp(0) is the rate of multiphonon

Page 131: Molly Gentleman Thesis 2006

121

emission at 0 K, ( )[ ] 11/exp −−= kThn ν is the phonon population density, v is the

frequency of the phonon, and p is the number of phonons required to bridge the

energy gap corresponding to the energy of the emitted luminescence, 16,500 cm-1.

The phonon energy was taken to be that of the highest energy Raman band for

tetragonal YSZ. Its energy and temperature dependence have been determined

recently [7] and this information was used in fitting the calibration data. The fitting

of the data to the multiphonon model was then used to represent the calibration

curve in converting lifetime measurements made under the thermal gradient tests to

the TBC/BC interface temperature. The fit of the data resulted in values of

wr=561 s-1 and wmp=0.111 s-1 for the sensor layers deposited here. Superimposed

on Figure 6.4.2 are the lifetime as a function of temperature data taken after the

heat flux measurements at NASA as well as after thermal cycling done at UCSB.

They show that there is no change in the luminescence lifetime as a result of the

thermal cycling indicating that the sensor will remain accurate under similar turbine

cycling. Additionally, the sample failed after ~550 1 hour cycles at 1150 °C which

is typical of similar materials without sensor layers under the same conditions.

Two examples of our temperature measurements are shown in figure 6.4.3

for a variety of heat flux conditions produced by changing the cooling rate on the

back of the superalloy. One plot is for a coating that is 127 µm thick and the other

for a 178 µm thick coating. In all cases, the sensor layer is 10 µm thick. In these

plots, the pyrometric measurements of the top surface of the coating and the

Page 132: Molly Gentleman Thesis 2006

122

luminescence measurements are both plotted as a function of the pyrometric

measurements of the alloy surface temperature.

In addition to the measurements of the TBC/BC temperature, the steady

state, one-dimensional heat flow in the NASA rig enabled the thermal conductivity

of the coating, κTBC, to be determined. Assuming one-dimensional heat flow, the

heat flux, q, through the coating and the superalloy are the same:

alloy

alloyalloy

TBC

TBCTBC h

ThTq

Δ=

Δ= κκ .

Equation 6.3.1

ΔT is the temperature drop across either the coating or alloy, κ is the thermal

conductivity of the layer, and h is the thickness of the layer. The temperature drop

across both the coating and metal were determined using the temperature

measurement of luminescence at the interface and the pyrometry measurements of

the top and bottom surfaces. Using the reported temperature dependence of the

thermal conductivity of CMSX-4 alloy [10], the thermal conductivity of the coating

was calculated to be 1.5 W/mK over the range of 850 to 1150oC. The

corresponding heat fluxes were determined from equation 3. These ranged from

0.4 to 1 MW/m2.

6.6 Discussion

The principal finding of this work is that luminescence of a Eu-doped YSZ

sensor layer buried under a standard YSZ coating can be used to measure the

temperature at the TBC/BC interface in the presence of a thermal gradient using

Page 133: Molly Gentleman Thesis 2006

123

typical thermal gradients and temperatures of coatings in use in engines today.

These results demonstrate that the luminescence lifetime method of measuring

temperature can be implemented using europium ions incorporated within the

crystal structure of current EB-PVD thermal barrier coatings and the temperature

measurement can be localized by controlling the deposition process. The fact that

the luminescence was localized to the TBC/BC interface was confirmed by

observations of the cross-section of one of the coatings after the thermal gradient

measurements were taken. This is illustrated in Figure 6.1.1, which is a

superimposition of a normal optical micrograph, to reveal the coating and the alloy,

over a luminescence image recorded using the luminescence from the Eu-ions.

There is also no apparent diffusion of the Eu3+ ions into the rest of the coating after

either the isothermal treatments or cycling in the high heat flux rig. This verifies

that the temperatures measured by the luminescence lifetime decay technique

provide us with the temperature of the TBC/BC interface and not some other depth

of the coating. This is vitally important since the technique is an average of the

luminescence from all ions excited. If the ions had been distributed at different

depths then the luminescence lifetime would have been a convolution of the

lifetimes at different depths within the thermal gradient. In principal, this can be

detected, as the luminescence decay would have been a “stretched” exponential

rather than a pure exponential. This and other higher magnification micrographs

also illustrates that there is no discontinuity in the coating caused by the switch

from the Eu-doping to the undoped ingot during deposition. The presence of

Page 134: Molly Gentleman Thesis 2006

124

uninterrupted single crystal columns suggests that the addition of the sensor layer

did not disturb the integrity of the coating.

Our current detection system limits us to measurements of interface

temperatures at about 1150°C. With more efficient collection optics and detectors

as well as higher speed electronics, higher interface temperatures can be measured,

allowing the technique and the Eu-doped coatings to be used at still higher

temperatures. The fitting to the multi-phonon model indicates that increasing the

detection limit by only a factor of ten would enable us to measure temperatures as

high as 1250°C. The other attribute of the Eu-doped coatings is that the

luminescence lifetime is sufficiently short that the technique has the potential for

measuring the temperature of rotating blades and other components[3].

The lifetime data presented also shows several interesting features

apparently peculiar to the Eu3+ luminescence in yttria-stabilized zirconia. First, the

luminescence lifetime of the second, longer exponential decay, τ2, is the same for

both the bulk and coated materials over the temperature range explored up to

1150oC. Rare-earth ion luminescence is known to have very little sensitivity to the

host material because the optically active 4f electrons of the ions are shielded by the

outer 5s and 5d shells[11]. This shielding may also create an insensitivity of the

luminescence lifetime in the slightly different environments of the bulk and coated

material. By contrast, the fast decay, τ1, which is usually attributed to

luminescence quenching by energy transfer to defects, shows sensitivity to the

changes in local environment from bulk to coated material [12]. Specifically in the

Page 135: Molly Gentleman Thesis 2006

125

coating, the intensity of the fast decay is higher with respect to the that seen in the

bulk and the luminescence decay time from the faster decay in the coating is

approximately 25% faster than the decay recorded from the bulk. The origin of this

difference is not known but may be related to the presence of the fine porosity

within the coatings. As mentioned earlier, the temperature dependence is attributed

to non-radiative transitions associated with the excitation of lattice phonons. In this

mechanism, the energy of the excited state, instead of being radiated as a detectable

photon, excites phonons through an ion-phonon coupling mode. The simplest

model ascribes this to phonons all of the same energy and assumes that this is the

highest energy phonon. In the case of YSZ, the B1g phonon mode, with room

temperature energy of 648 cm-1 determined from Raman spectroscopy, is likely to

be the primary phonon mode responsible for the multiphonon relaxation. Using the

measured temperature dependence of the Raman spectra, between 25 and 27

phonons are required for the non-radiative transition over the temperature range

examined. Although the energy of the B1g mode provides a good fit of the data for

the dilute concentrations of Eu3+ used in the sensors described in this work, work

done in chapter 4 illustrates that ion-ion energy transfer is likely more important

than the multiphonon relaxation rates in the case of more concentrated sensor

materials. Nevertheless, the fact that the lifetime data reported here fits the multi-

phonon relaxation model suggests that the Raman spectra provide a valuable guide

to selecting other combinations of luminescent ions and coating materials for future

high-temperature thermometry.

Page 136: Molly Gentleman Thesis 2006

126

Another interesting feature of the luminescence of these sensor layers is that

there may be an effect on the emissivity of the TBC and hence the ability to

measure temperature of the TBC using infrared pyrometry when these sensor layers

are in place. Figure 6.4.4 illustrate the measurements taken by both pyrometry as

well as luminescence for a 146 µm coating deposited in the same run as other

samples examined. Superimposition on this data are the calculated values of top

surface temperature assuming the luminescence measurements are correct and the

thermal conductivity is 1.5 W/m/K as confirmed by Ted Bennett’s group by the

phase shift method[13]. This calculation illustrates that at lower temperature the

pyrometer underestimates the temperature of the surface of the TBC and

overestimates that temperature at higher temperatures. The underestimate at lower

temperatures is likely due to an increase in the optical penetration depth of the

pyrometer at those temperatures. At higher temperatures the overestimate may be

due to thermal activation of 7F levels of the europium ions in the sensor layer

leading to luminescence at wavelengths where the pyrometer is reading. This

would result in an increase in the measured temperature by pyrometery and an

increasing error as the temperature as is seen in Figure 6.4.4. This phenomenon is

currently under examination.

Page 137: Molly Gentleman Thesis 2006

127

References 1. Bradley, L.C., Rev. Sci. Instrum., 1953. 24: p. 219. 2. Allison, S.W., L.A. Boatner, and G.T. Gillies, Characterization of high-

temperature themographic phosphors: spectral prperties of LuPO4:Dy(1%), Eu(2%). Appl. Optics, 1995. 34(25): p. 5624.

3. Gentleman, M.M. and D.R. Clarke, Luminescence sensing of temperature in pyrochlore zirconate materials for thermal barrier coatings. Surface & Coatings Technology, 2005. 200: p. 1264.

4. Gentleman, M.M. and D.R. Clarke, Concepts for luminescence sensing of thermal barrier coatings. Surface & Coatings Technology, 2004. 188-189: p. 93.

5. Allison, S. and G. Gillies, Remote thermometry with thermographic phosphors: Instrumentation and applications. Rev. Sci. Instrum., 1988. 68(7): p. 2615-2650.

6. Feist, J.P., A.L. Heyes, and J.R. Nicholls, Phosphor thermometry in electron beam physical vapor deposition produced thermal barrier coating doped with dysprosium. Proc. Instn. Mech. Engrs., 2001. 215(G): p. 333.

7. Gentleman, M.M., et al., Non-contact methods for measuring thermal barrier coating temperatures. Int. J. Appl. Ceram. Technol., 2006. 3(2): p. 105.

8. Zhu, D.M. and R.A. Miller, Thermal conductivity and elastic modulus evolution of thermal barrier coatings under high heat flux conditions. J. Thermal Spray Tech., 2000. 9(2): p. 175.

9. Moos, H.W., Spectorscopic relaxation processes of rare earth ions in crystals. J. Lum., 1970. 1(2): p. 106.

10. CMSX Property Data. 1994, Cannon-Muskegon Corporation: Michigan. 11. Henderson, B. and G.F. Imbusch, Optical Spectroscopy of Inorganic Solids.

1989, New York: Oxford University Press. 12. Kallendonk, F. and G. Blasse, Luminescence and energy transfer in

EuAl3B4O12. J. Chem. Phys., 1981. 75(2): p. 561. 13. Yu, F.L. and T.D. Bennett, Phase of thermal emission spectroscopy for

properties measurements of delaminating thermal barrier coatings. JOURNAL OF APPLIED PHYSICS, 2005. 98: p. 10.

Page 138: Molly Gentleman Thesis 2006

128

Figure 6.1.1 Optical image of the cross-section of a thermal barrier coating

illustrating the location of YSZ:Eu sensor layer appearing red between the Superalloy and YSZ coating. The image is a superimposition of a white light image of the coating and a luminescence image from the Eu-layer obtained with UV excitation of the same region.

Page 139: Molly Gentleman Thesis 2006

129

Figure 6.3.1 Experimental arrangement used to measure the temperature in contact

with the superalloy using the luminescence sensor in a temperature gradient.

Page 140: Molly Gentleman Thesis 2006

130

Figure 6.4.1 Representative luminescence lifetime decays for the bulk and coated

samples at 700°C. While the long decays have the same slope, the faster decay varies in the two materials. The longer decay is used for temperature measurement.

Page 141: Molly Gentleman Thesis 2006

131

Figure 6.4.2 Comparison of the luminescence lifetime decays as a function of

temperature for the bulk and coated materials. The decays are fit using a multiphonon relaxation model, equation 1.4.5 in the text. Also included are lifetime data after thermal gradient testing and thermal cycling illustrating that the luminescence lifetime as a function of temperature does not change with aging of the material.

Page 142: Molly Gentleman Thesis 2006

132

Figure 6.4.3 Temperatures measured at the top of the TBC and bottom surface of

the CMSX4 alloy by pyrometry and the TBC/BC interface temperature by luminescence for three different coatings. (a) 127 µm and (b) 178 µm thick coatings.

Page 143: Molly Gentleman Thesis 2006

133

Figure 6.4.4 Temperatures measured at the top of the TBC and bottom surface of

the CMSX4 alloy by pyrometry and the TBC/BC interface temperature by luminescence for a 146 µm coating. The blue line illustrates the calculated values for the 8 µm pyrometer assuming a thermal conductivity of the coating at 1.5 W/m/K.

Page 144: Molly Gentleman Thesis 2006

134

Chapter 7 Conclusions and Future work

The first and most compelling conclusion of this thesis is that it is possible

to develop non-contact heath monitoring and temperature sensors for thermal

barrier coatings using luminescence techniques. Thermal barrier materials doped

with rare-earth ions have been shown to not only exhibit an ability to substitute into

thermal barrier material without negative effects on the coating life as seen in the

thermal cycling data in chapter 6, but also that they are capable of being used to

measure temperatures up to those observed in typical turbine environments. Yttria

stabilized zirconia doped with europia can be used to measure temperature up to at

least 1150 °C while Eu2Zr2O7 has temperature sensing capabilities approaching

1300 °C. This work has also shown that measurements made on bulk materials are

consistent with those made on coatings of the same composition and can therefore

be used to examine new dopant compositions easily through precursor routes

without the use of expensive deposition of all compositions of interest. Finally, the

measurement of temperature from a buried sensor layer within an EB-PVD coating

while under a thermal gradient was demonstrated resulting in not only a way to

monitor the temperature at the TBC/BC interface, a major driving force in this

work, but also a direct measurement of the thermal conductivity and heat fluxes in

the thermal barrier coating. Simple calculations have also shown that the

luminescence lifetime decays of these Eu-doped sensors are consistent with those

needed for in situ measurement on rotating engine components.

Page 145: Molly Gentleman Thesis 2006

135

Future work in this area should focus on understanding the sources of

uncertainty in these experiments and on ways to diminish them. Although the

current temperature measurement techniques available for thermal barrier coatings

(infrared pyrometry) have an error similar to that of this technique, luminescence

provides us with an opportunity to decrease the error through a series of

improvements to the technique. This thus makes luminescence temperature

measurements much more valuable in terms of their potential for high sensitivity.

Possible avenues for these improvements include the exploration of more efficient

optics and electronics as well as room for the development of a more precise means

for the automated collection and analysis of the luminescence lifetime signal.

Additionally, the use of other dopant ions, including Tb3+ and Dy3+ show

promise and in each case need to be examined more thoroughly. In early studies,

terbium exhibited a tendency to substitute into the YSZ system as a tetravalent

species as opposed to the desired trivalent ion. This resulted in a weak

luminescence signal with a relatively short luminescence lifetime decays (~1 µs) at

room temperatures. It has been suggested that the use of a chloride based precursor

as opposed to the nitrates used in these studies, or co-doping with pentavalent ions

may be helpful in resolving this issue and should be explored further. In the case of

Dy-doped systems, the main concern is finding an efficient excitation source for the

luminescence. For these studies, short of purchasing new lasers of appropriate

laser frequencies (~440 nm), it may be possible to use broadband flash lamps for

excitation.

Page 146: Molly Gentleman Thesis 2006

136

The effects of the sensor layer on both the oxidation characteristics of the

TGO as well as the life of the coating must be examined. Preliminary work has

suggested that the presence of the Eu-doped layer next to the TGO may have

changed the degree or rumpling observed during oxidation of the material, but

insufficient data was collected to quantify the extent of this effect if any.

Moreover, the long-term stability of the sensors in terms of phase stability and

luminescence lifetime must be established before implementation.

Finally, as Figure 7.1 illustrates, there is a significant change in the

luminescence lifetime as a function of temperature for the same materials with two

different excitation sources. This should be examined to determined if the effect

can be used to further increase the temperature sensing capabilities of rare earth

doped zirconias.

Page 147: Molly Gentleman Thesis 2006

137

Figure 7.1 1 a/o Eu-doped YSZ luminescence lifetime as a function of

temperature for two different excitation energies 532 nm and 248 nm. The reason for the increase in lifetime in the case of 248 nm excitation has yet to be determined.


Recommended