Sie sind auf Seite 1von 196

HIGH TEMPERATURE

CORROSION OF CERAMICS

J.R. Blachere
F.S. Pettit
Department of Materials Science and Engineering
University of Pittsburgh
Pittsburgh. Pennsylvania

NOYES DATA CORPORATION


Park Ridge, New Jersey,U.S.A.
Copyright @ 1989 by Noyes Data Corporation
Library of Congress Catalog Card Number: 88-38242
ISBN: O-8155-1188-4
Printed in the United States

Published in the United States of America by


Noyes Data Corporation
Mill Road, Park Ridge, New Jersey 07656

10987654321

Library of Congress Cataloging-in-Publication Data

Blachere, J.R.
High temperature corrosion of ceramics / by J.R. Blachere and F.S.
Pettit.
P. cm.
Bibliography: p.
Includes index.
ISBN O-8155-1188-4 :
1. Ceramic materials--Corrosion. I. Pettit, F.S. (Frederick
S.), 1930- . II. Title.
TA455C43B57 1989
620.1’404217--dc19 88-38242
CIP
Other No yes Publications

CERAMIC RAW MATERIALS

Edited by

D.J. De Renzo

The broad-based ceramics Industry encom- provides chemical and physlcal property data
passes all types of glass; refractones; abrasives; for more than 1000 products supplied by 181
whitewares. such as porcelain and pottery, ceramic raw material suppliers in the U.S. and
structural clay materials; etc. Increasing use of Canada.
advanced ceramic materials in the automotive Part I is an alphabetical listing of 99 raw
and aerospace Industries, as well as in such material categories. Raw material suppliers are
diverse areas as electronics and medical devices, then included alphabetically under each cate-
IS expected to push the demand for raw materials gory, along with their product information. The
far beyond that associated with the traditional categories in Part II include additives; semi-
ceramics industry. processed materials, some of which are available
Prepared directly from manufacturers’ data as unfinished shapes or substrates; and materi-
sheets and tables at no cost to, nor influence als intended for specific end uses. Raw materials
from, the contributing companies, this book categories are:

Parr I Corundum Potassium Compounds Yttrium Oxide


Alumina Diatomaceous Earlh Pyrophyllite Zinc Oxide
Alumina Chrome Dolomite Duartz Zircon
Alumina Zirconia Feldspar Rutile Zirconates
Aluminum Nitride Ferrites Sand Zirconia
Aluminum Titanate Fireclay Silica Zirconium Boride
Andalusite Flint Clay Silicon Zirconium Nitride
Antimony Compounds Fluorspar Silicon Carbide
Ball Clays Fluxes Silicon Nitride
Barium Compounds Frits Sillimanite Part II
Bauxite Garnet Slags Binders
Beryllium Oxide Graphite Soapstone Ceramic Additives
Bismuth Compounds Iron Oxide Soda Ash Ceramic Adhesives,
Bone Ash Kaolin Sodium Silicate Potting Materials,
Borates Kyanite Sodium Sulfate and Putty
Borax Lanthanide Compounds Spars Ceramic Coatings
Boric Acid Lead Compounds Spine1 Ceramic Colors
Boron Lime Spodumene Ceramic Fibers
Boron Carbide Limestone Stannates and Whiskers
Boron Nitride Lithium Compounds Stannic Oxide Ceramic Materials
Calcium Compounds Magnesia Strontium Carbonate Ceramic Precursors
Carbon Magnesite Talc Dielectric Compositions
Celestite Manganese Compounds Titanates Dispersing Agents
Chlorite Mineral Mica Titanium Boride Electronic Ceramics
Chrome Ore Mullite Titanium Carbide Reagents
Chromium Oxide Nepheline Syenite Titanium Dioxide Glazes
Clays (Miscellaneous) Nickel Compounds Titanium Nitride Glaze Stains
Cobalt Compounds Ochre Ulexite Pie20 Compositions
-. .
Cofemamte Periclase Vermiculite Refractory Materials
Copper 8 Copper Oxide Petalite Whiting Sealing and Solder
Cordierite Phosphates Wollastonite Glasses

ISBN O-8155-1143-4 (1987) 8%” x 11” 900 pages


Foreword

This book describes high temperature corrosion of ceramics. The materials in-
vestigated in this particular study were silica, alumina, silicon nitride and silicon
carbide. In addition to the pure single crystals or CVD materials, typical engi-
neering materials of various purities were included in the study. The corrosion
conditions were ‘hot corrosion’ in which gaseous corrosion was enhanced by
Na2S04 deposits. Some gaseous corrosion and oxidation experiments were also
performed. The hot corrosion was studied at IOOO’C and at lower tempera-
tures, in the presence of pure oxygen and oxygen containing SO2 and SOs. The
changes in morphology of the surfaces were observed in the scanning electron
microscope. This instrument and the related x-ray microanalysis were the
major tools of research. A method of measurement of oxide thickness in the
electron microprobe was developed for the experiments on silicon nitride and
silicon carbide.

In the use of materials at elevated temperatures in harsh environments, it is ap-


parent that, in most instances, ceramics are the best choice to provide corrosion
resistance. While ceramics may be more corrosion resistant than metallic alloys
and polymers, ceramics can react with certain environments. The purpose of
this study, then, was to systematically investigate the corrosion of ceramics and
to develop a theory generally applicable to all ceramics.

A great number of ceramic materials are available for use in a variety of corrosive
environments. In order to develop a theory applicable to the corrosion of cer-
amics in general, it was necessary to investigate a variety of representative cer-
amic materials exposed to a number of different corrosive environments. In
order to keep the number of experiments at a reasonable number, the ceramic
materials used in this study were selected on the basis by which resistance to
corrosion was developed, in particular, by being immune to the environment
or by developing passivity. Furthermore, the environments used to produce
corrosion were selected based upon the likelihood of their being encountered in
practice and their severity.

V
vi Foreword

The information in the book is from High Temperature Corrosion of Ceramics,


prepared by J.R. Blachere and F.S. Pettit of the University of Pittsburgh for the
U.S. Department of Energy, December 1987.

The table of contents is organized in such a way as to serve as a subject index


and provides easy access to the information contained in the book.

Advanced composition and production methods developed by


Noyes Data Corporation are employed to bring this durably
bound book to you in a minimum of time. Special techniques
are used to close the gap between “manuscript” and “completed
book.” In order to keep the price of the book to a reasonable
level, it has been partially reproduced by photo-offset directly
from the original report and the cost saving passed on to the
reader. Due to this method of publishing, certain portions of
the book may be less legible than desired.

NOTICE

The materials in this book were prepared as an account


of work sponsored by the U.S. Department of Energy.
Neither the United States Government nor the Depart-
ment of Energy, nor any of their employees, nor any of
their contractors, sub-contractors, or their employees,
nor the Publisher, makes any warranty, express or im-
plied, or assumes any legal liability or responsibility for
the accuracy, completeness, or usefulness of any infor-
mation, apparatus, product or process disclosed or rep-
resents that its use would not infringe privately-owned
rights.

Final determination of the suitability of any informa-


tion or procedure for use contemplated by any user,
and the manner of that use, is the sole responsibility of
the user. The reader is warned that caution must always
be exercised when dealing with ceramics at high temper-
atures, and expert advice should be sought at all times
before implementation.
Contents and Subject Index

INTRODUCTION.........................................l

EXPERIMENTAL PROCEDURES. . . . . . . . . . .......... .6


Experimental .
Conditions. . .. . . . . . . . . . .......... .6
Special Techniques . .. . .. . . . . . . . . . . . . .......... .7
Measurement of Oxide Thickness . . . . . . .......... .7
Measurement of Contact Angles. . . . . . . . . .......... .9
Cross Sections and Related Techniques .......... .9

GASEOUS CORROSION ....................... ...... . . . . 11


Introduction. ............................. ...... . . . . 11
Gaseous Corrosion of Silica and Alumina. .......... ...... ... 11
Silica. ................................ ...... . . . 11
Alumina .............................. ...... . . 13
Gaseous Corrosion of Silicon Nitride and Silicon Carbide ...... . . . . 14

HOT CORROSION OF SILICA. . . . . . . . . . . .. . . . . . . .. . . . . . . .16

HOT CORROSION OF ALUMINA. . . . . . . . . . . . . . . .. . . . . . . . .21

HOT CORROSION OF SILICON CARBIDE AND SILICON NITRIDE. . . .26

REFERENCES..........................................30

APPENDIX A-GASEOUS CORROSION OF SILICA AND ALUMINA


IN SULFUR OXIDE ENVIRONMENTS . . . . . . . . . . . . . . . . . . . . . . . . 31
H. R. Kim, J. R. Blachere, F.S. Pettit
Introduction. .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Gaseous Corrosion of Ceramics . . . . . . . . . . . . . . . . . . . . . . . . . 33

vii
viii Contents and Subject Index

Experimental Procedure. .............. . . . . . . . . .35


Materials ....................... . . . . . . . . . . .35
Experiments. .................... . . . . . . . . . .35
Results and Discussion ................ . . . . . .37
Silica. ......................... . . . . . . . . .37
Alumina ....................... . . . . . . . .39
General Results. ................ . . . . . . .39
Thermodynamics of Sulfate Formation . . . . . . . . .41
Single Crystal .................. . . . . . . .43
Polycrystalline Aluminas. .......... . . . . . . .43
General Discussion. .................. . . . . . : :53
52
Conclusions ....................... . . . . . .
References ........................ . . . . . . . . .55

APPENDIX B-HOT CORROSION OF SILICA. ................. . .5B


M. G. Lawson, H. R. Kim, F.S. Pettit, J. R. Blachere
Introduction. .................................... . 58
Hot Corrosion. ................................... : :60
59
Experimental Procedure .............................
Materials ..................................... : :60
60
Gaseous Corrosion. ..............................
Hot Corrosion. ................................. . .60
Results and Discussion .............................. 63
Gaseous Corrosion. .............................. . .63
Wetting by Sulfates .............................. . .64
Hot Corrosion. ................................. . .66
Kinetics of Crystallization. ......................... : 176
72
General Discussion. ..............................
Discussion of Crystallization Kinetics .................. . .82
Hot Corrosion of Silica Formers. ..................... . .85
Conclusions ..................................... . .86
References. ..................................... . .87

APPENDIX C-HOT CORROSION OF ALUMINA . . . . . . . . . . . .89


M.G. Lawson, F.S. Pettit, J.R. Blachere
Introduction. . . . . . . . .
. . . .. . . .. . . . . .. . . . . . . .89
Experimental Procedure. . . . . . . . . . . . . .. . . . . . . .95
Materials . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . .95
Hot Corrosion Experiments. . . . . . . . . .. . . . . . . .97
Results and Discussion . . . . . . . . . . . . . . .. . . . . . . . .97
Wetting . . . . . . .. . . . . . . . . . . . .. . . . . . .97
Hot Corrosion of Al*Os .. . . . . . . . . .. . . . . . . 102
Results, Acidic Conditions . . . . . . . . . . . . . . . . . 107
Single Crystal Alumina. . . . . . . . . . . . . . . . 107
High Purity Polycrystalline Alumina. . . . . . . . . 109
Medium Purity Polycrystalline Alumina. . . . . . . . . 112
Low Purity Polycrystalline Alumina . . . . . . . . . . . 113
Results, Basic Conditions . . . . . . . . .. . . . . . . . 115
Contents and Subject Index ix

Single Crystal . . . . . . . . .. . . . .
. .. . . . 116
High Purity Polycrystalline . . .
Alumina . . 116
Medium Purity Polycrystalline Alumina. . . . 118
Low Purity Polycrystalline Alumina .. . . . . 119
Discussion of Results . . . . . . . . . . . . . . . . 122
References. . . . . . . .. . . .. . . . . . . .. . 136

APPENDIX D-HOT CORROSION OF SILICON NITRIDE AND


SILICON CARBIDE ............................... . . . . 137
J. R. Blachere, D. F. Klimovich, F.S. Pettit
Introduction. ................................ . . . . . 137
Experimental Procedure ......................... . . . 137
Results and Discussion .......................... . . . 138
Acidic Hot Corrosion of Silicon Nitride. ............ . . . 152
Discussion of the Hot Corrosion of Silicon Nitride. ..... . . 157
Acidic Hot Corrosion of Silicon Carbide ............ . . . . . 161
Model for the Oxidation and Hot Corrosion of Silicon
Carbide Under Acidic Conditions .............. . . . 168
Proposed Model .......................... . . . . . 169
Defect Structure and Stoichiometry of Silica ....... . . . . 170
Oxidation of Silicon Carbide .................. . . . . 172
Acidic Hot Corrosion. ...................... . . . 178
References. ................................. . . . . 182

APPENDIX E-PUBLICATIONS ............................. 184


References. ....................................... 185
Introduction
In the use of materials at elevated temperatures in harsh environments, it

is apparent that in most instances ceramics are the best choice to provide

corrosion resistance. While ceramics may be more corrosion resistant than

metallic alloys and polymers, ceramics can react with certain environments.

The purpose of this program was to systematically investigate the corrosion of

ceramics and to develop a theory generally applicable to all ceramics.

A great number of ceramic materials are available for use in a variety of

corrosive environments. In order to develop a theory applicable to the corrosion

of ceramics in general, it is necessary to investigate a variety of representative

ceramic materials exposed to a number of different corrosive environments. In

order to keep the number of experiments at a reasonable number, the ceramic

materials used in this study were selected on the basis by which resistance to

corrosion was developed, in particular, by being immune to the environment or

by developing passivity. Furthermore, the environments used to produce corrosion

were selected based upon the likelihood of their being encountered in practice

and their severity.

A material is considered to be immune to a particular environment when it

is in equilibrium with that environment. Total or complete immunity is rare in

practice but in some cases the amount of reaction required for equilibrium to be

achieved is very small, and consequently the changes in the properties of the

material are very small. For example, when Al303 is heated in oxygen at

elevated temperatures, the oxygen activity in the Al.303 may not be the same as

that in the gas, and oxygen will be incorporated into or removed from the Al303

depending upon the oxygen activities in the Al303 and in the gas. Upon obtaining

equilibrium between the Al303 and the gas, which may require extremely long
2 High Temperature Corrosion of Ceramics

times depending upon the temperature, the Al303 does not exhibit any significant

changes in mechanical properties. The concentrations of point defects in the

Al303 however will change and there could be significant changes in properties

such as electrical and ionic conductivities. Nevertheless, in terms of a corrosion

reaction the Al303 may be considered to be immune to this gaseous environment.

Alumina and silica are two ceramic materials which can be considered to be

immune to many environments which are extremely corrosive to metallic systems,

and therefore these two materials were studied in the present investigation. The

purity and structure of these ceramics can also affect their behavior in different

corrosive environments and therefore these two parameters were also examined

in the present investigation. Four types of crystalline Al303 were studied of

varying purity as defined in Table I. Only one type of silica was studied. As

described in Table I it was of relatively high purity and was vitreous.

Passivity to corrosive environments is achieved by the formation of a

reaction product barrier through which the reactants involved in the corrosion

process must diffuse. The development of passivity is the principal means by

which metallic alloys achieve resistance to corrosive environments, but passivity

is also an important mechanism in certain ceramic materials, depending upon the

ceramic and upon the environment causing corrosion. For example, Si3N4 and

Sic react with most environments encountered at elevated temperatures and

resistance to corrosion is achieved via the formation of a passive reaction product

barrier. Furthermore, Al303 can react with gaseous environments containing

sufficient SO3 to form sulfates and again the properties of the sulfate reaction

product play a significant role in the corrosion properties. In the present studies

Si3N4 and SIC specimens with purities as defined in Table I were therefore used.
TABLE I - MATERIALS UNDER STUDY

Designation Material Purity (Wt%) Supplier

S.C. Sapphire -99.99 Saphikon

998 poly A120; (a) 99.8(0.1MgO, O.llSiO2) Lucalox (G.E.)

995 poly A1203 (a) 99.6 - 99.5(0.17 M@ AlSi Mag 772 (3M)
0.17 Si02) or ADS 995 (Coors)

975 poly A1203 (a) 97.4 (0.75 MgO, 1.6SiO2 S-697 (Saxonburg)
0.1 Na20)

Silica fused silica 99.99 Corning 7940

SIN CVD CVD*Si3N4 99.99 Deposits & Composites

SiN H.P. H.P.** Si 3N4 93(6Y203Fe, Al, 02) Airesearch

Sic SC. s.c.*** (a6H S.C.) 99.99 W.J. Choyke

Sic CVD CVE (Sic) 99.98 W. J. Choyke

Sic H.P. H.P. (aSiC) 94(3Al203, 2.5W, Fe, 02) Norton NC-203

* Chemical Vapor Deposition *** Single Crystal


** Hot Pressed + Polycrystalline Alumina
4 High Temperature Corrosion of Ceramics

When ceramic materials are used in practical applications, a variety of

corrosive environments can be encountered. Most of these environments will

contain oxygen but other reactants such as sulfur, nitrogen, carbon and chlorine

can also be present. Moreover, deposits such as metallic sulfates, carbonates or

chlorides may also accumulate upon exposed surfaces and substantially affect

the corrosion processes. In the case of studies concerned with the corrosion of

metallic alloys it has been useful to examine corrosion reactions in environments

of increasing severity extending from gaseous environments containing oxygen,

to mixed gases containing one or more reactants in addition to oxygen, and finally

considering the effects of deposits such as Na3S04, Na3C03 or NaCl. The

environments used in the present studied are identified in Table II. These

environments consisted of gases containing oxygen, and other reactants such as

sulfur, carbon and hydrogen at temperatures of 700, 1000 and 14OOOC. The

effects of deposits were studies by using Na3S04 deposits in O3-SO3-SO3 gas

mixtures at temperatures of 700 and 1000°C. These conditions were selected

because they are frequently present in many environments encountered in

practice. Furthermore, as established from studies using metallic alloys, the

principles established from studies using these selections should be generally

applicable to other corrosive systems.

In the following sections of this report the experimental procedures will be

described and then a summary of the results will be presented and discussed.

Some of these results have been presented in previous reports for this

program.(l) Other results are included in student theses and are also presented

in drafts of papers about to be submitted for publication. Consequently, some of

the results will not be repeated in this report but will be referred to by

references or by draft papers included in appendices in the present report.


introduction 5

Results obtained from studies using gaseous environments will be discussed first

and then results obtained from studies using deposits on silica, alumina, Sic and

Si3N4 will be presented in sequence.


Experimental Procedures

Experimental Conditions

The materials that were studied in this program and the environments that

were used to produce gaseous corrosion and hot corrosion attack have been

briefly discussed previously in this report (Tables I and II). The ceramics that

were studied (Table I) consisted of Al3O3, SiO3, Sic and Si3N4. A range of

purities were examined in the case of Al3O3, Sic and Si3N4. Single crystals of

SIC and Al303 were also studied and compared to polycrystalline specimens.

The gas compositions that were used are presented in Table II. Gas mixtures

consisting of O3-SO3-SO3 were used in both gaseous corrosion studies and the

hot corrosion studies. Hot corrosion experiments were also performed in pure

oxygen. The gaseous corrosion studies were performed at temperatures of 700,

1000 and 1400°C whereas the hot corrosion investigations used temperatures of

700 and 1000°C. When O3-SO3-SO3 mixtures were used at 700°C, this mixture

was passed over a platinum catalyst to ensure that equilibrium was achieved.

The catalyst was not required to achieve equilibrium at 1OOOOC.The flow of the

gas was 1 cm3/s.

The experimental procedures have been discussed in detail in previous

reports and are also discussed in Appendices A, 0, C and D. The procedure

usually consisted of exposing specimens in a horizontal tube furnace at a fixed

temperature to a flowing gas stream of fixed composition. In the case of the hot

corrosion studies the specimens were coated with deposits of Na3S04 for

6
Experimental Procedures 7

investigations at 1000°C, or a NaSS04-CoSO4 equimolar solution for studies

performed at 700C. These deposits were applied on specimens by spraying warm

specimens with an aqueous solution saturated with NaSS04 or the sulfate

mixture. Most specimens were diamond polished and they were ail cleaned

before the experiments.

The characterization techniques consisted of morphological studies in the

light microscope and particularly the Scanning Electron Microscope (SEM) with

microanalysis of salient features by EDS (Energy Dispersive X-ray Spectroscopy)

and WDS (Wavelength Dispersive X-ray Spectroscopy used particularly for light

elements). These techniques were supplemented by X-ray diffraction, weight

change measurements and in some cases surface analysis techniques (ESCA). A

large number of other techniques were used in special cases they are SIMS, ISS,

FTIR, Laser beam ellipsometry.

Furthermore, since after the corrosion tests some deposited salt often

remained on the sample, the Nag.504 was washed off to reveal the sample

surface. The samples were characterized before and after this washing.

Experience has shown that the examination of the sample before washing is

extremely fruitful, and the observed morphologies have taken many forms. The

washwater was analyzed as described in a previous report (l); this analysis is now

a routine semi-quantitative method which allows the identification of the

elements dissolved in the salt and the stoichiometry of the salt.

Special techniques

Measurement of oxide thickness

A method using WDS measurements in the electron microprobe (EPMA) was

adapted from that described by Yakovitz and Newbury to estimate the

thickness of coatings from the intensities of emitted characteristic X-rays. The


8 High Temperature Corrosion of Ceramics

intensity of the oxygen Ka line is used in the present research to measure the

thickness of oxide layers generated on non-oxide materials. This intensity is

measured under constant conditions (say 1OkVaccelerating voltage and 2 x lo8

A beam current) for a scale and a bulk standard of pure fused SiO2. Both

samples are coated with 200A of carbon. The ratio k of their intensities

corrected for background is correlated to the thickness of the oxide layer.

The calibration curve for the relationship between the k ratio and the oxide

thickness was calculated with a computer program written for this research

following the semiempirical approach of Yakovitz and Newbury(2) to generate

the Q (pz) curve which is the intensity of X-ray generated at a weighted PZ depth

into the sample. This calculation includes many corrections and correlations

used to predict the characteristic X-ray intensity. It must be understood that

the data available for these corrections (such as absorption) for light elements is

relatively poor and that this calibration can only be approximate. Measurements

based on light element are usually not used quantitatively. However the method

developed in this research gives good reproducibility ? 2% for thickness

measurements on silica layers up to about 0.8pm - lum. It was published

recently(3) and more details will be found in Appendix E which contains a copy of

the article on this measurement. It must be emphasized that since the

characteristic X-ray intensity for oxygen is used to calculate the oxide

thickness, a stoichiometric silica (SiO2) was assumed in the calculations. This

may lead to significant error if the composition of the scale deviates markedly

from that assumed in the calculations. However the method is quite

reproducible and has a high spatial resolution.

The direct measurement of scale thicknesses on cross sections in the SEM

is probably not better than + 10% in accuracy considering magnification


Experimental Procedures 9

calibration and other sources of error under the best conditions. It is good down

to about lum and depends greatly on the thickness of the layer (constancy and

magnitude). In some cases the scales are difficult to separate from the substrates

in cross sections in the SEM.

Measurement of contact angles

The contact angles of the deposits have been measured in the SEM at room

temperature using a method previously described by Murr(4). The wetting angle

is particularly important since deposits often break into droplets decreasing the

area of interaction between the sample and the melt. After coating with carbon

(about ZOOA)the samples are placed in a tilted position in the SEM which allows

the location of the drop under the electron beam and then they are tilted further

to a vertical position to record the profile of the drop. The edges of the drops

are enlarged for the measurement. A number of drops are measured under those

conditions on the same sample. Experience with the measurements show that

their reproducibility is about f: 2O. Since this reproducibility is very good, it has

been possible to establish when two wetting behaviors occurred simultaneously

on the same sample.

Cross sections and related techniques

The study of cross sections is necessary for a number of measurements.

However the samples are often damaged (silica and silica formers) during exposures

and processing. Also, cross sections are often desired with salt on the sample, so

that sawing or polishing in wet media has been avoided to preserve the soluble

sodium compounds. Most cross sections were either dry-cut with a diamond saw

or fractured. Fractured specimens show structural features well as has been


10 High Temperature Corrosion of Ceramics

established in many studies of the microstructure of ceramics. The multilayer

nature of the scale and sometimes its apparent layered growth is shown clearly

in cross sections. The cross sections can also be used directly for the measurement

of oxide layer thickness (~l~rn). In some cases salt drops have been crossed by

the fracture giving valuable information. However the lack of flat cross sections

has limited the use of the crystal spectrometers of the electron microprobe.

In many cases the deposit drops do not adhere welt to the sample surface

after cooling so that the underside of the drops and the area of the sample which

was under the drops can be examined and analyzed at least qualitatively.

Mechanical bursting of bubbles in surface layers and removal of drops are also

used to study the underlying microstructure prior to washing of the specimens.


Gaseous Corrosion

Introduction

The results from the gaseous corrosion studies will be discussed by considering

first the studies of silica and alumina, and then the studies performed using Sic

and Si3N4.

Gaseous Corrosion of Silica and Alumina

The gaseous corrosion of silica and alumina were performed at temperatures

of 700, 1000 and 14000C in a number of different gas environments which included

oxygen, 02-H20-H2, 02-CO2-CO and O2-SO2-SO3 gas mixtures (Table II). The

results obtained from these studies are discussed in detail in Reference 1 and

Appendix A. The specific conclusions developed from these studies are as follows:

Silica

(1) Devitrification of silica glass to cristobalite took place rapidly under all

atmospheres studied at 14OOOC. The rate of crystallization increases with

increasing temperature and time.

(2) The devitrified layer undergoes a displacive transformation with a large

volume change on cooling which causes cracking. Such a transformation in crystalline

SiO2 is important with regards to the use of crystalline SiO2 scales on Sic, Si3N4

and metallic alloys as protective barriers for high temperature applications.

(3) The silica is significantly affected when exposed to low oxygen pressure

at 1400°C. Silica weight losses occurred after exposure to either wet hydrogen

or a CO-CO2 gas mixture and are related to the decomposition of silica in the

low oxygen pressure and the reaction of either hydrogen or CO with silica to

form SiO vapor and either H20 or CO2 gas. Weight losses increase with increasing

temperature and decreasing oxygen pressure.

11
12 High Temperature Corrosion of Ceramics
Gaseous Corrosion 13

(4) No significant reaction of silica in CO2 or 02 was observed at all

temperatures except for devitrification at 14OOOC.

(5) Silica is very resistant to attack by SO2-SO3-02 gas mixture under the

test conditions.

Alumina

(1) Alumina is resistant to attack by H2-H20-02 gas mixture but impurities

in the alumina materials such as SiO2 and Na20 can result in volatilization. The

volatilzation is favored by low oxygen pressures and high temperatures.

(2) Alumina is very resistant to attack by CO2 gas.

(3) Alumina is resistant to corrosion by SO2-SO3-02 gas mixtures. Some

reactions occurred especially at high SO3 pressures (e.g. 7~10~~ atm. SO3) and

low temperature (7OOOC).

(4) Corrosion of alumina in SO3 containing gas can occur even on the highest

purity alumina where reaction products with activities less than unity are formed.

It may be due to the formation of a solid solution of Al2(SO4)3 with Al203 or a

nonstoichiometric sulfate. The observed sulfur was identified as a ~+6 (as in

sulfate) by ESCA.

(5) Degradation of alumina in SO3 containing gas becomes more severe

when Mg or Ca containing impurities are present, and it increases as impurity

content increases. The formation of products occurs preferentailly along the

grain boundaries (i.e., on the impurity second phases at the grain boundaries).

(6) Corrosion of alumina in SO3 containing gas becomes more favorable at

higher SO3 pressures and lower temperature (e.g. at 7000 than 1000°C), since a

lower SO3 pressure is necessary to form sulfate at 700°C than at 1000°C. The

severity of corrosion increases with time.


14 High Temperature Corrosion of Ceramics

Gaseous Corrosion of Silicon Nitride and Silicon Carbide

The gas induced corrosion of Si3N4 and SIC was studied only at 1400°C in

oxygen and at 1OOOoC in an 02 - SO2 - SO3 gas mixture with an initial SO2

pressure of 0.01 atm. These limited studies were performed to determine if

devitrification of the silica scales formed on these materials occurred at 1400°C,

and to provide baseline data for comparison with the hot corrosion studies at

1ooooc.

The studies in oxygen at 1400°C showed varying results depending upon the

form of the substrate material. In the case of the single crystal Sic

crystallization of the silica scales was observed. These scales which were about

1 urn thick after 12 hours of oxidation and cracked upon cooling to form star

patterns that were believed to radiate from nucleation centers indicating the

original nuclei of the crystallization of the amorphous layer into cristobalite.

The silica scales which formed on the polycrystalline samples were thicker in

most cases than those formed on the single crystals of Sic being about 10 urn

after 12 hours of oxidation. The hot pressed silicon carbide exhibited a glazed

surface and contained bubbles resulting from CO and CO2 evolution. The lack of

crystallization was attributed to impurities, particularly Al2O3, stabilizing the

glass structure. Some cristobalite was identified in these scales however by

XRD along with a large “glass” peak. The hot pressed silicon nitride also formed

a 10 urn thick scale which was composed of cristobalite and enstatite (MgSiO3).

The CVD silicon nitride, which was much purer than the other silicon nitrides

developed an extremely thin silica scale after 12 hours of oxidation. There was

virtually no change in surface morphology of the oxidized specimens compared to

specimens prior to oxidation and the weight change of specimens after 12 hours

of oxidation was below the detection limits of the techniques used in this program
Gaseous Corrosion 15

to measure weight change. No impurities were detected in these silica scales on

CVD silicon nitride which XRD analyses showed to be a mixture of cristobalite

and glass. The sintered silicon nitride developed a scale composed of glass

containing some cristobalite. Yttrium silicate crystals were observed to protrude

out above the surface of the silica scale. Impurities influenced the oxidation of

this material substantially.

The results obtained from the studies performed at 1400°C show that the

oxidation of SIC and SiSN4 is dependent upon the structure and composition of

the silica scales that are formed upon these materials. Glassy silica provides a

more protective reaction product barrier than crystalline silica, however, the

incorporation of impurities into the glassy silica can cause the protectiveness of

the glass structure to be decreased very substantially by promoting devitrification.

At 1000°C the pressure of SO2 and SO3 in the gas mixture along with oxygen

did not significantly affect the oxidation of pure silicon nitride or pure silicon

carbide compared to oxidation in pure oxygen. The major effect of SO2 and SOS

occurred when the specimens contained impurities. While the effects of impurities

were significant but not documented extensively, these effects were not as

substantial as those observed at 1400°C, since the impurities cannot concentrate

in the oxide scale at 1000°C and thereby affect the protective properties of the

glassy silica scales.


Hot Corrosion of Silica

The results obtained from the studies on the hot corrosion of silica are

presented and discussed in more detail in Appendix B of this report. In the

following the important results from these studies are briefly summarized.

Specimens of fused silica (Corning 7940) about 1 cm x 1 cm square and 1

mm thick were exposed to a variety of conditions known to cause the hot corrosion

of metallic alloys. The experimental procedures have been described in the

experimental section of this report. Results from four different sets of

experimental conditions will be discussed in this summary. These experiments

were performed at 700 and 1000°C. The deposits of Na3S04 which were applied

to the specimens’ surfaces were liquid at 1OOOoC. At this temperature two

different gas compositions were used. One consisted of an SO3-03 gas mixture

and the other was pure oxygen. The SO3 pressure in the gas mixture was 1.5 x

10e3 atm. When Na3S04 is exposed to gases containing SO3 the activity of

Na90 in the sulfate is inversely proportional to the SO3 pressure. If the activity

of Na30 in the sulfate is taken as a measure of the basicity of the Na3S04,

higher SO3 pressures established less basic, or more acidic melts, whereas the

gas containing only oxygen causes the more basic liquid sulfate to develop.

Similar sets of experiments were performed at 700°C however at this

temperature the SO3 pressures in the gas mixtures was 7 x 10m3atm.

Furthermore the deposit applied to the specimen surfaces was Na3S04 - 50 mole

percent CoSO4 since pure Na$304 melts at about 883OC and the sulfate mixture

is liquid at 700°C.

16
Hot Corrosion of Silica 17

The liquid sulfate deposits wetted the silica specimens in varying amounts

depending upon the experimental conditions. Time at temperature and the

thickness of the deposit also affected wetting. Generally the liquid deposits

wetted the silica more completely under basic conditions. At 1000°C under

basic conditions the salt wetted most of the coupon after 1 hour with a wetting

angle of 20. After 24 hours the wetting was continous. At this temperature and

under acidic conditions droplets of liquid were formed with wetting angles Of

24O, and 130 on large droplets (’ 0.2 mm dia), after 24 hours. Wetting was not as

complete at 7OOoC. In the case of basic conditions droplets with wetting angles

between 13 -240C were observed after 24 hours, whereas under acidic conditions

the angles ranged between 36 - 49O.

At 7000C under acidic conditions some limited localized attack of the

silica was observed under of the salt droplets. No evidence of devitrification of

the silica was observed. After water washing to remove the salt, small weight

losses (0.1 - 1 mg/cm2) were detected and small voids were evident in the silica

where salt droplets had been present prior to water washing. The voids were

more concentrated near the perimeter of the droplets. Analysis of these results

has been complicated by the decomposition of the CoSO4 in the liquid via the

reaction,

coso4 * coo + so3

since the SC3 pressure in the gas was lower than the equilibrium pressure

required to maintain the initial liquid deposit. The principal reaction of the

deposit with the silica should involve the Na20 component of liquid since

sulfates and sulfides of silicon are not stable under the experimental conditions

that were employed. Hence, a reaction of the following type seema plausible,

XSiO2 + YNa2S04 = YNaZO-XSi02 + ~~03


18 High Temperature Corrosion of Ceramics

Inspection of this reaction shows that the formation of sodium silicate phases

becomes less favorable as the SO3 pressure is increased. It is believed that the

SOS pressure in the melt over most of the specimen surfaces is too high to

permit silicate formation. However, some dissolution of silica occurred in the

sulfate at localized regions under the drops, in particular along the sulfate -fused

silica interface which caused voids to be evident upon water washing. This

localized dissolution is believed to result from impurities in the silica which

affect its stability.

Under basic conditions at 700°C the complication from decomposition of

the CoSO4 component of the liquid was more severe since there was no SOS in

the gas phase. Cristobalite spherulites were observed beneath the salt drops.

Weight change measurements after water washing were not meaningful because

some cobalt oxide resulting from decomposition of CoSO4 remained upon the

specimen surfaces. Since dissolution of the fused silica was observed under the

acidic conditions, and this process is believed to involve NaSO, such a reaction

would also be expected in the more basic melt. The important result obtained at

7000 under basic conditions however is that crystallization of the fused silica

was observed beneath the salt droplets. Since crystallization was not observed

at 700°C under acidic conditions, the activity of NaSO in the liquid deposit must

be established at some specific level in order for crystallization to proceed.

At 1OOOoCunder acid conditions cristobalite was observed to form beneath

all of the droplets, but no devitrification of the fused silica was evident away

from the droplets. The weight losses of specimens under acidic conditions at

1000°C were less than at 700°C. This can be accounted for by less sodium

silicate being formed at the higher temperature.


Hot Corrosion of Silica 19

The most extensive degradation of the vitreous silica occurred under basic

conditions at 1OOOoC. A layered reaction product was formed over the total

surface of the fused silica specimens. Proceeding from the salt-specimen

interface the following sequence of phases was observed after very long

exposures: sodium silicate, tridymite, cristobalite, unaffected fused silica. The

thickness of the crystallized silica conformed to a parabolic rate law under

isothermal conditions. Under cyclic conditions this crystallization proceeded

more rapidly and conformed to a linear rate law since the crystalline products

spa&d from the specimen aa a result of thermally induced stresses. The

observed parabolic rate has been accounted for by assuming the crystallization

of the fused silica is caused by sodium from the liquid deposit. There is no

question that some sodium silicate was formed. Analyses of wash water,

however, shows that more water soluble corrosion products are formed early in

the corrosion process than after long reaction times when the crystalline phases

have been formed. Such results suggest that the reaction to form silicates is

dependent upon the composition of the liquid deposit and the structure of silica.

Also as the silicate becomes richer in silica it is less soluble in water.

As discussed previously in the section on gaseous corrosion, silica was not

significantly affected by any of the gas environments used in these studies. No

corrosion nor devitrification was detected at 700 or 1OOOOC. On the other hand

significant degradation of fused silica was observed in SOS-02 gas mixtures, and

in oxygen, when liquid sulfate deposits were present. This degradation occurred

by two different, but related, processes both of which were dependent upon the

activity of NaSO in the deposit. The process which caused the most severe

degradation was devitrification. This process increased as the Na70 activity in

the liquid was increased and it occurred at both 700 and 1000°C. It requires a
20 High Temperature Corrosion of Ceramics

threshold Nag0 activity. It was especially severe when the specimens were

thermally cycled since the crystallized products spalled under the influence of

thermally induced stresses. The other process involved the formation of sodium

silicate as a reaction product on the surfaces of fused silica. It appears that this

latter process is less prevalent when the liquid is reacting with crystalline silica

than vitreous silica, nevertheless this reaction also increases as the activity of

Nag0 in the liquid is increased. This reaction also appears to be affected by

impurities in the fused silica when its driving force is low.


Hot Corrosion of Alumina

The results obtained from the studies on the hot corrosion of alumina are

presented and discussed in detail in Appendix C. In the following important

results are summarized.

The weight changes measured after hot corrosion of the aluminas were

small but not negligible. They were due to offsetting reactions such as the

solution of Al203 into the sulfate melt, the silica and silicate precipitation

mostly due to impurity phases and the precipitation of Co0 at 7OOoC in pure

oxygen. In general the weight changes appear greater than for gaseous

corrosion.

The sulfate tended to wet the aluminas partially at 7OOoC under acidic

conditions (~20~ after 24 hours). The wetting of the purer aluminas decreased as

a function of time apparently as some alumina dissolved into the sulfate and

reduced the affinity of the sulfate for the alumina substrates. At higher

temperature (lOOO°C) under similar conditions the wetting improved a little.

The wetting was better at 1000°C under basic conditions than under acidic

conditions as the impurities tended to increase the wetting tendency. The wetting

is an indication of the reaction tendency of the sulfate with the substrate. Good

wetting results in more contact area between the sulfate and the substrate which

also promotes the corrosion.

The single crystal alumina tended to react very little with the Na2S04 and

in line with the gaseous corrosion results more reaction occurred under acidic

conditions than under basic conditions. The single crystal has basal orientation.

The corrosion of the polycrystalline aluminas indicated that the corrosion of the

grains is a function of orientation, with greater attack of planes away from the

21
22 High Temperature Corrosion of Ceramics

basal orientation. This result is expected for solution of a single crystal in a

melt.

Under acidic conditions, at 7OOoC sulfates of aluminum and magnesium

were formed and after long exposures globular silica was observed on all

polycrystailine materials. This shows that under the most acidic conditions the

silicate impurities are attacked by the melt generating sulfates and precipitating

silica. The required solubility gradient is set for continued dissolution of alumina

by acid fluxing and precipitation of silica. The precipitation of silica globules,

probably cristobalite, occurs in the sulfate melt away from the interface with

the substrate and is not related to the microstructure of the substrate. The

corrosion is concentrated near the grain boundaries at 700°C and occurs on a

wider scale at 1OOOoC. At the higher temperature the regions under the sulfate

melt are smooth with no marked preferential attack at the grain boundaries. At

1000°C magnesium sulfate and calcium sulfate are formed and alumina is

incorporated in sodium aluminum silicates.

Under basic conditions significant corrosion occurred at 700 and 1000°C.

At 1000°C sodium magnesium aluminum silicates and sodium calcium aluminum

silicates were formed. The salt on cooling contained Mg, Al and some Ca. The

attack of the alumina grains was limited under the more basic conditions overall

at 1000°C and significant intergranular corrosion was evident in the

micrographs.

The impurities played a major role on the corrosion behavior of the

polycrystalline aluminas, particularly at high temperature (lOOO°C), in pure

oxygen (basic conditions). While there was little evidence of basic fluxing of the

single crystal, the high Nag0 activities promote reaction with the silicates

present at the grain boundaries of the polycrystalline materials. Sodium

aluminum silicates grew from the melt with transport of silica and other oxides
Hot Corrosion of Alumina 23

along the grain boundaries. This is illustrated by the perfect decoration of the

grain boundaries of high purity polycrystalline alumina with crystals of this

silicate after long term cyclic exposures. Some magnesium was present in the

silicates formed near triple points. With the more impure aluminas more silicate

was formed and the crystals contained various alkaline earth elements and

potassium which were present as impurities in the materials. While this fluxed

growth of silicate crystals feeding from grain boundaries and impurity grains is

interesting, it plays a fundamental role in the corrosion of the alumina. It is

proposed that the impurity reactions which form sodium silicates lower the Na30

activity and raise the SO3 locally in the melt. This more acidic Na3S04 then can

dissolve the alumina grains by formation of sulfate in the melt by acidic fluxing.

The two reactions, the grain reaction and the grain boundary reaction will

proceed cooperatively, as the ion needed for one is produced by the other.

Therefore a fundamental mechanism is proposed for the hot corrosion of

polycrystalline alumina in which the impurities play a major role. As shown

above they have a major influence even on 99.8% purity alumina, which means on

most technical ceramics. This may apply also to alumina scales grown on

coatings on superalloys but in a general manner since they do not contain silica-

based impurities for which the following mechanism applies.

Under acidic conditions, the sulfate tends to dissolve the alumina even

though conditions are not favorable for the formation of aluminum sulfate at

unit activity. The alumina is dissolved by acid fluxing. The formation of

aluminum sulfate in the field of stability of alumina was already observed in the

gaseous corrosion experiments. In a sulfate melt a wide range of activities can

be established locally. Under basic conditions little or no attack of single crystal

alumina was observed although basic fluxing should be possible, particularly at

high temperature (more basic conditions). However this is not promoted in


24 High Temperature Corrosion of Ceramics

presence of silica-based impurities which are present at the grain boundaries and

as second phases in the polycrystalline aluminas. The impurities modify the local

conditions so that intermediate activities favoring the attack of the silicates and

the dissolution of the alumina in the sulfate prevail under the acidic and the

basic conditions of this study. Under acidic initial conditions, set by the gaseous

environment, the sulfate tends to dissolve alumina and as this is done the

activity of Nag0 in the sulfate is increased. This promotes the attack of the

silicates which are dissolved in the sulfate thus decreasing the Nag0 activity.

The two reactions can proceed cooperatively. At 700°C the activity of Na70

was always too low to form sodium silicates and SiO7 is precipitated. At

1000°C, higher Nag0 activities are generated by the same mechanism and

sodium aluminosilicates are formed. Under basic initial conditions, the sulfate

melt tends to attack the acidic impurity phases, thus promoting the dissolution

of neighboring alumina by acid fluxing. The two reactions proceed cooperatively

as discussed earlier.

The mechanisms just presented explain that the polycrystalline aluminas

were attacked under both acidic and basic conditions of the experiments,

however the processes are extremely slow and there was no experimental

evidence of catastrophic attack even after 400 hours exposure. Under the

circumstances one may ask if any of the proposed mechanisms would lead to slow

continuous attack without replenishment of the salt although acidic conditions

provide some of the requirements. However in view of the rate of attack by

sodium sulfate it appears that in many industrial processes in which it could be a

factor the sulfate will be replenished before it might become saturated or

depleted. Higher temperatures might increase the rate of corrosion and basic

conditions might become predominant under usual (percent or less) sulfur


Hot Corrosion of Alumina 25

concentrations in the atmosphere, however the sulfate vapor pressure then will

limit the corrosion since sulfate deposits are no longer formed at higher

temperatures.
Hot Corrosion of Silicon Carbide
and Silicon Nitride

The results of recent experiments are presented and discussed in detail in

Appendix D. Important results are summarized below.

The hot corrosion of silicon nitride and silicon carbide has been studied in

presence of Na9SO4, under acidic, 1% SO9-balance oxygen initially and basic,

pure oxygen, conditions in the temperature range 900-1000°C. During basic hot

corrosion the salt wets completely the samples while during acidic corrosion it

breaks up in droplets. The hot corrosion increases the rate of oxidation and the

thickness of the oxide layers formed increases markedly from acidic (measured

between the drops) to basic hot corrosion and both are greater than for dry

oxidation. The oxide layers formed tend to be vitreous and devitrify rapidly

under the liquid sulfate. For the purer materials devitrification was sparse and

very limited in between the sulfate droplets under acidic conditions. In general,

clearly different behaviors are observed for the pure materials under the two

conditions. They are controlled by the activity of sodium oxide in the sulfate

melt near the interface with the substrate. In all cases the materials oxidize and

the oxide dissolved into the sulfate. For acidic conditions, the sulfate does not

wet the oxide, and a surface activity of sodium oxide in equilibrium with the

atmosphere is set up in between the sulfate droplets. The sodium diffuses into

the vitreous silica and modifies it.

26
Hot Corrosion of Silicon Carbide and Silicon Nitride 27

Under basic conditions, a thick product layer was formed and the NagSO4

was consumed slowly in the reaction. The ceramics oxidize at their surfaces and

the oxide dissolves into the sulfate melt. The Nag0 activity builds up at the

interface and a silicate layer is formed. As the silicate enriches in silica,

cristobalite is nucleated at the interface. After long exposures, the silicate

phase remains on top of the silica and the sulfate left is in small isolated drops

on top of the silicate. The melt contains a high concentration of silica initially

as indicated by the wash water analysis and it has also been well documented by

otherst5). The observations are consistent with the mechanism proposed by

Mayer and Riley(S) except that the reaction is much slower with Nags04 than

for NagCOS which they studied. Some protection is offered by this complex

product layer since greater degradation was observed after preoxidation of the

samples.

Under acidic conditions (1.5~10-~ atm SOS), poor wetting and little

reactivity with the salt were observed. However the oxidation was enhanced

even between the sulfate drops. This oxide growth between the drops was studied

in detail. The oxide formed was mostly vitreous. In these regions a very thin

layer rich in sodium is detected (~10 A thick) and sodium diffuses into the silica

formed. The silica layers formed under the sulfate droplets devitrified rapidly

into cristobalite by spherulitic crystallization or random globular formation.

They were thicker than the vitreous layer formed outside the drops except for

the C-side silicon carbide for which these thicknesses were similar. The

thicknesses of oxide formed under the drops tended to be similar for all three

surfaces (CVD silicon nitride, C-side silicon carbide and Si-side silicon carbide)

studied but they were still smaller than the product layers formed under basic
28 High Temperature Corrosion of Ceramics

conditions. This is generally consistent with a mechanism determined by the

liquid phase and therefore independent of the nature of the substrate surface.

Indeed for the basic conditions and under the droplets for acidic conditions there

was no more differentiation in the behavior of C-side and Si-side silicon carbide.

A model was proposed for the formation of random cristobalite globules which

correlates with the results obtained for the hot corrosion of bulk silica. Selective

attack of the substrate occurs by transport through the liquid phase between the

spherulite fibrils.

The acidic hot corrosion which occurs between the sulfate droplets is

essentially enhanced oxidation. The kinetics of this hot corrosion of silicon

nitride are complicated apparently by the formation of oxynitride. Some evidence

of this formation was obtained in this research. It is not clear presently how this

thin layer formed during the oxidation slows down the oxidation so strongly. The

silicon carbide has different behavior for the carbon-side and for the silicon-side.

During the acidic hot corrosion they oxidize parabolically at different rates.

Kinetic data and apparent activation energies were obtained for the hot corrosion
of each side. In order to interpret these results, it was logical to modify a model

for the oxidation of silicon carbide. However, a satisfactory explanation of the

oxidation of silicon carbide had not been proposed.

Based on the oxidation results of others, a model was developed on the

premise that the parabolic oxidation is controlled by diffusion of oxygen through

the oxide layers to the substrate interface. Although vitreous silica is formed in

all cases, the different rates of oxygen transport are associated with different

oxygen deficient vitreous structures produced in the oxidation of silicon, silicon

carbide C-side and silicon carbide Si-side. Tentative mechanisms were proposed

for the oxidation of the two sides of silicon carbide in which the oxides are formed
Hot Corrosion of Silicon Carbide and Silicon Nitride 29

under different oxygen pressures and therefore are expected to have different

defect structures. The observed trends in the variation of the parabolic constants
for the acidic hot corrosion of these materials compared to their oxidation and
that of silicon are predicted qualitatively by these models. High apparent

activation energies such as those measured for hot corrosion are generally

consistent with the models although that for the silicon-side of silicon carbide

could not be estimated.

The behavior of engineering materials is strongly influenced by the

impurities they contain. MgO and Y3O3 tend to segregate toward the surface

during oxidation and hot corrosion. They lower the acidity of the sulfate under
acidic conditions so that wetting is improved and thicker oxide layers are formed

as the conditions are more basic than promoted by equilibrium with the

atmosphere. Conversely, the wetting is not as good under basic corrosion for
these materials as for the purer ones. The impurities set conditions intermediate

between those promoted by the equilibria with the atmospheres selected for the
basic and acidic corrosion experiments. Alumina impurities in the silicon carbide
do not segregate to the product layer formed on oxidation or hot corrosion. The
alumina entering the oxide layer appears to stabilize it against devitrification.

Great progress has been made in the understanding of the hot corrosion of
silicon nitride and silicon carbide. The corrosion varies with environmental

conditions and it was shown to be controlled by the activity of sodium. The

behavior observed correlates well with the results on the hot corrosion of silica.
In particular the major role played in the degradation of the scales by their

devitrification is emphasized by these results. It was shown also that

preferential attack occurred locally under acidic corrosion. Overall the hot

corrosion of these materials is well understood qualitatively. The atomic

mechanisms proposed for the oxidation and hot corrosion of silicon nitride and
silicon carbide suggest directions for further studies.
1. J.R. Blachere and P.S. Pettit “High Temperature Corrosion of Ceramics”
a) DOE Report ER45117-2, March 1988
b) DOE Report ER45117-1, June 1985
C) DOE Report ER10915-4, June 1984

2. a) H. Yakovitz and D.E. Newbury, SEM/1976/1, IIT Research Institute,


Chicago, IP, p. 151.
b) J.I. Goldstein et al.,References
Scanning Electron Microscopy and X-ray
Microanalysis, Plenum, No. 4, 1981, p. 354.

3. J.R. Blachere and D.F. Klimovich, J. Am. Ceram. Sot., 70 [ll] C324-C326
(1987).

4. L.E. Murr, Interfacial Phenomena in Metals and Alloys, Addison-Wesley,


Reading, Mass., 1976, p. 67.

5. N.S. Jacobson, J. Am. Ceram. Sot., 69 [l] 74-82 (1986).

6. M.I. Mayer and F.L. Riley, J. Mat. Sci., l3, (1978) 1319-1328.

30
Appendix A-Gaseous Corrosion of Silica
and Alumina in Sulfur Oxide Environments
H.R. Kim, J.R. Blachere and F.S. Pettit

Abstract:

A fused silica of high purity and alumina in single crystal and poly-

crystalline forms with different impurity levels xiere exposed to mixtures of

SOj- SO2~02 at 700°C and 1000°C. The silica and alumina are very resistant

to corrosion in these environments. No reaction was observed with the silica.

Sulfates formed on the aluminas particularly at low temperature and high SO3

pressures. The reaction occurred under less severe conditions and with greater

intensity with increasing impurity content, particularly Ca and Mg.

Supported by the U.S. Department of Energy under Agreement Number

DE-ACOZ-8 lER109 15-A000

Presented in part to the 1982 Fall meetings of American Ceramic

Society in Cambridge, Mass.

31
32 High Temperature Corrosion of Ceramics

I. INTRODUCTION

Ceramics are considered actively for components of advanced power systems

operating at high temperatures. (1) This is because they have high refractori-

ness ( chemical stability and mechanical properties much improved thrcugh

extensive research over the past twenty years. They have been partially

successful in recent demonstration projects (*) and it may be anticipated that

the needed improvements in reliability of components will result from the

present emphasis on ceramic processing. Ceramics may be used at high tem-

peratures as monolithic parts rePlacing superalloys, as thermal barrier

coatings or as coatings protecting metallic alloys from environments which cause

them severe corrosion problems.

One should not assume however that ceramics are totally corrosion resistant.

Corrosion of refractories by oxide melts such as slags and glasses are costly

to the steel and glass industries. As a result research has been performed on

the corrosion of ceramics in deep melts, (3) but little is known about the

corrosion df these materials in gaseous atmospheres. The environments encountered

in advanced energy systems may contain corrosive products of combustion.

In addition, deposits of salts such as NaZSO4 mpy form on the components and

tend to enhance the gaseous corrosion. It has been shown recently that ceramics

such as AlZO3, Si3N4 and yttria-stabilized ZrO2 can react with molten deposits

of Na2S04 in SO3 gas.(b) The gaseous corrosion of ceramics has been studied

in this research (5,6) not only to establish the susceptibility of ceramics to

the corrosion by gaseous atmospheres representative of those encountered in

energy systems but also as the baseline for research on the hot corrosion of

crramics( ‘1 C i n presence of salt deposits). The corrosion of silica and alumina is

discussed here. These materials are of great importance since they perform

as protective scales on superalloys and SiOZ forms on silicon nitride and silicon

carbide in oxidizing environments.


Appendix A-Gaseous Corrosion 33

II. GASEOUS CORROSION OF CERAVICS

Gaseous corrosion of ceramics can occur by a variety of mechanisms

just as in the case of the gaseous corrosion of metals and alloys. Moreover,

the corrosion process are ultimately controlled by the nature of the products

that are formed as a result of reaction between the gas and the ceramic. In

metals and alloys reaction with the gas is almost always characterized by

metallic elements being oxidized via transfer of electron to reducible species

in the gas. The cations of the ceramics are already in oxidized states and the

following possible corrosion processes can be proposed. First, the cations

may be oxidized to higher states which in the case of oxygen as an oxidant

would result in oxygen being incorporated into the ceramic. Such a process

could result in the formation of new phases or merely changing of stoichrometry

of the phase being oxidized. This plays a major role with degradation of chrome-

based refractories with temperature and particularly atmosphere cyclings.

Second, the atoms or molecules of the gas may form more stable compounds with

the cations of the ceramic, for example silicon nitride reacts with oxygen to

form silicon dioxide and nitrogen. This type of reaction will depend on the

properties of the products formed upon the surface of the ceramic but effects

produced by the elements displaced by the oxygen can also be very important.

h third possibility involves the formation of volatile species such as the re-

duction of SiOz (s) to SiO (g) or the oxidation of CrpOj (s) to CrOj (g). Such

reactions probably would involve diffusion through gaseous boundary layers and

could proceed at rates fast compared to those of processes involving the form-

ation of condensed phases on the surface of the ceramics. Finally, molecules

in the gas such as SOj. CO2 or Hz0 may possess a significant affinity for the

cationic component or its oxide and react to form compounds such as sulfates

carbonates or hydrates.
34 High Temperature Corrosion of Ceramics

The rates of such processes will be dependent upon transport through

corrosion products and rather complicated mechanisms may prevail.

The gaseous corrosion of ceramics is expected to be similar in some

respects to the corrosion of ceramics in melts which has been studied extensively

in the literature (gv 3l It is usually modelled as a dissolution process.

The kinetics involve the transport of reactants and products which often

determine the overall rate of the corrosion process. The corrosion reaction

rate depends on the relative corrosion rate of the various phases, the viscosity

of the liquid, the density gradient in the liquid, the concentration gradient

in the liquid, the porosity of the solid, the wettability of the solid phases

by the liquid, the boundary layer at the interface and the geometry of the system.

The grain size and the pore structure are major factors in controlling the rate

of dissolution. Fine grain materials dissolve faster than coarse grain materials.

The liquid may penetrate into the grain boundaries resulting in engulfment of the

(9,101
solid grains by the liquid .The influence of the microstructure on the

corrosion complicates greatly the interpretation of corrosion experiments so that

for fundamental studies Simple systems with simple geometries and microstructures

must be used. This is valid also for gaseous corrosion studies since except for

wetting effect, the grain size, the pore structure and the grain boundaries should

play a similar role as in liquid corrosion.

There is significant literature on the oxidation of silicon nitride and

silicon carbide (*l! the only literature on gaseous corrosion of simple oxides

appears to be on their reduction and volatilization particulary for silica.

Silica at high temperature under low oxygen pressures volatilizes forming

of SiO (g) by thermal decomposition and/or by reaction with hydrogen or CO (12*13!

For instance, significant weight losses were observed in wet hydrogen (51 at
Appendix A-Gaseous Corrosion 35

1400oc. Alumina is much more resistant to conditions of this tvpe and tends

to lose impurities such as sodium at high temperatures. Significant losses of

alumina are only observed at very high temperatures, of the order of 19OOOC. (14)

Steam at low temperatures can form hydrates with silica resulting in degradation

of silicate refractories. (15) However, there are no reports on the Influence

of sulfur oxide gases on alumina and silica; this is discussed below.

III. EXPERINENTAL PROCEDURE

(1) Materials

They include four aluminas and a fused silica which were obtained from

commercial suppliers in relatively high densities and flat configurations.

A brief description is given in Table I. The aluminas were selected to investigate

the influence of microstructure and impurities on the corrosion. The single crystal

is a transparent substrate with (0001) orientation. The (L) material is

in the form of almost transparent washers; (S) and (M) are white substrates.

The more impure (S) alumina was included to show the gross effects of impurities.

The (S) and (MI aluminas have a much finer average grain size L-kin than (L)
I’ 1
alumina (19Urn). Their microstructure will be discussed later as needed.

(‘2) Experiments

Flat specimens ( z l~l.cm) were exposed to sulfur oxide eases at ml and lno~~c

in either as received, polished or relief polished conditions. Prior to exposure

they were cleaned thoroughly in acetone and alcohol and then heated in flowing

oxyeen atmosphere at IOOOoC for one hour. The ~02 and 02 mixtures were passed

over a platinized catalyst at the temperature of the experiments before contact

with the samples. The conditions were selected to have relevance to gas turbine

operations. The SO3 pressures were high enough in most instances to prevent the

decomposition of Na2Sfl4 in hot corrosion exueriments performed in the same

environments. The initial SO2 contents of the gases were “s~ally 1%. 0.1% and
TABLE I. MATERIALS UNDER STUDY

Designation Materials % Purity (Imp) Suppliers


3
Silica fused silica 2.20 99.89 (0.1 H20) Corning 7940(l)

8.C. sapphire 3.97 99.99 Saphikon(2)

L a-A1203* 3.97 99.79 (0.1 MSO. 0.11 Lucalox (3)


SiO2)
M u-A1203* 3.89 99.6 (0.17 t&O, AlStiS 772 (4)
a-A1203* 0.17 SiO2)
S 3.74 97.4 (0.75 ugo,
+l.6SiO2, O.lNa20 S-697 (5)

* Polycrystalline

(1) Corning Glass

(2) Tyco Laboratories

(3) General Electric, Lamp Div.

(4) 3M, Technical Ceramic Prod. Div.

(5) Saxonburg Ceramics. Inc.


Appendix A-Gaseous Corrosion 37

0.01% with the balance 02 for a total pressure of one atmosphere and a flow race

of about lcm3/sec. The corresponding equilibrium SO3 pressures are shown in Table

II. The exposure times were usually one week but ware varied from one day to one

month. Some experiments were performed in sequential exposures to follow the

product development on the same area of the sample, examined also before exposure.

Since the materials are electrically non-conductive the samples were coated with
o
about 200A of carbon prior to any observation in the Scanning Electron Microscope

(SEM). This carbon was oxidized at 700°C in 02 before any additional exposure

of the samples in sequential experiments. It was checked by comparison of

duplicate samples exposed only once for the same total time that this procedure

did not interfere with the results of the sequential experiments.

The major investigative tools were the SM’s equipped with Energy

dispersive X-ray Spectrometers (EDS) for elemental analysis. X-ray diffraction

(XRD) was used to identify the products when they formed in sufficient quantities.

Some specimens were also analyzed by Electron Soectroscopy for Chemical Analysis

(ESCA) after exposure. The weight of the specimen was checked on an analytical

balance before and after the experiments.

IV. Results and Discussion

No significant weight change was measured in any of the experiments. It can

be shown, considering the size of the samples, that a weight change of D.Img/

cm2 of exposed surface corresponds to the formation of a continuous layer of

product about lum thick. The corrosion products were usually discrete and often

with little coverage of the surface, they could be detected using the SEM , EDS

and ESCA, but in no case were they equivalent to a continuous layer of the order

I urn thickness. These products are described in the following.

(1) Silica

Silica is very resistant to atta’ck by SD2- SO3 atmospheres at 700 or 1DDO’C.

No visible product formed and no sulfur was detected by EDS or ESCA after exposure
38 High Temperature Corrosion of Ceramics

Tr\BLEII: Environmental Conditions in Experiments on


A1203 and Silica

Initial
Gas Pressure Pressure (Atm) at
(Atm) 7oooc 1ooooc

02 Sl. ‘Ll . -1.

SO2 10-2 2.9 x 10-3 8.7 x 1O-3

SO3 7.2 x 10-3 1.3 x 10-3

02 %I. Sl. %I.

SO2 10-3 2.9 x 10-4 8.7 x 10-4

SO3 7.2 x 1O-4 1.3 x 10-4

02 -1. Sl. %I.

SO2 10-4 2.9 x 10-5 8.7 x 10-5

SO3 7.2 x 10-5 1.3 x 10-5


Appendix A-Gaseous Corrosion 39

up to one month. where is no evidence of sulfate formation and there is *o thermo-

dynamic data for silicon sulfate. The oxygen pressure of the experiments was too

high for the formation of silicon sulfide. No experiments in C3-SO3-So3

atmospheres were performed above 1OOOoC since Sulfate formation is less favorable

as the temperature increases. .No devitrificatioa of the silica glass was observed

after the exposures at 700 or 1OOOoC.

(2) Alumina
(a) General Results

The alumina samples were exposed to all three atmospheres at 700cC

and 1000°C in extensive experiments including sequential exposures to observe

the evolution of the products. The results are summarized in Table III. This

cable is based on three types of results: the observation of products in the

SEM. the detection of sulfur by EDS and the identification of sulfates

by ESCA. Since only very small amounts of products were formed in many cases and

the analytical techniques were close to their limit of detection, the formation

of products (P) is reported in the table only when it was indicated by at least

two techniques. If only one technique suRRested the presence of products, Su

was placed in the table.

It was established that the products were sulfates by the combined observations

of sulfur in the products by EDS and the identification of the valency of this

sulfur as S-6 as in sulfates by ESCA. Moreover, on the purer aluminas the

1
elements de:ec:cd in the products were only aluminum and sulfur. Extraction replicas

on exposed (L) samples showed also the same combination, thus leading to the

conclusion that aluminum sulfate was formed on the purer aluminas, Mg and Ca

sulfates were identified in a similar manner on the more impure samples.

Table111 clearly shows that there is an increasing tendency to form

sulfates with decreasing temperature, with increasing impurity contents and

with increasinR SO3 pressure in the atmosphere. These general trends are
TABLE III

Results of Alumina Corrosion Experiment in SO3 Gas at 700° add 1000°C for One Week

Tempe'rature SO3 pressure (atm.) Single Crystal Polycrystalline Aluminas


(OC) in approx. 1 atm 02 Alumina (L)99.8% M)99.5% S)97.4%"

700 7.2 x 10-3 P P P P

7.2 x IF4 P P P 1'

7.2 x IO-' N P P P

1.3 x 10-3 N P - P

1.3 x 10-4 N SU P P

1.3 x 10-5 N N SU SU

P: Sulfur containing products present


N: No sulfur containing products
su: Presence of sulfur suggested but not confirmed
* Purity
Appendix A-Gaseous Corrosion 41

consistent with thermodynamic analysis.

(b) Thermodynamics of sulfate formation

Considering the reaction:

Al*03+ 3SO3(g)= Al*(S04)3

at equilibrium one has:


g_ a Al2 W4)3

a Al203 (pso3j3

and the SO3 pressure for the equilibrium between hl2(SO4) 3 can be calculated.

The stability diagrams of Figure 1 for A1303 in SO3-SO3 atmospheres were

generated bv considering the appropriate equilibrium reactions assuming

unit activity for the solid phases. The test conditions and the relevant

boundaries for the HgO- Mg SO4 and CaO- CaSO4 equilibria are shown on the

diagrams.

The oxide-sulfate equilibrium boundaries shift to lower SO3 pressures ac

lower temperatures. Furthermore, for a given initial mixture of 02- SO3 the

eqdlibfiumSO3 pressure increases with decreasing temperature. Therefore,

sulfates have a greater tendency to form at lov temperatures as shown by comparison

of Figures la and 16 and observed experimentally: It is also obvious from Figure 1

that the SO3 pressures required for sulfate formations increase from CaSO4

CO ?IgSO4 and finally co A12(SO4; 3 so thar under similar activity conditions

CaSC4 and YgSO4 will form more readily than Al?(SO4)3 this explains the increase

in product formation with increasing contents of CaO and YgO.

*Preliminary results obtained by Raman spectroscopy also indicated that

sulfates formed on S alumina (16)


42 High Temperature Corrosion of Ceramics

a 0

cao C&m*
4- WN Me

log POY
-lo - A12(so4
Al203

-1a -
A12S3
-1(1 -10 4 0 I

log Psoa

b Ow

4-

log Por
-10 *

log PSOl
Figure 1. Stability diagrams for A1203 in SO@03 atmospheres at: (a)
700°C (b) 1OOOoC. The experimental conditions (+) and the
boundaries for CaO and MgO are also shown.
Appendix A-Gaseous Corrosion 43

Cc) Single Crystal

Very little reaction of the single crystal alumina was observed during

exposure to any of the gas mixtures at 700°C and IOCWC. A few thin tetragons

about 0.5 pa on a side were detected after an exposure of one week to the highest

SO3 pressures of 7000C (Table III). For the same time of exposure to the inter-

mediate pressure of SC3 at 70C°C no products could be found but sulfur was

detected by EDS and sulfates were indicated by ESCA analysis. NO sulfur was

found after exposure to the lower SD3 pressure at 700°C nor for any of the

gas compositions at IOOOoC. These results show that the single crystal material

was affected by the environment when the experimental conditions were most

favorable for reaction. The stability diagrams of Figure 1 show that Al2(SC4)3

should not be formed in any of the experiments. These diagrams have been

constructed however for unit activity of the phases. Solid solutions (e.g.

Al2(SO4)3 at less than unit activity dissolved in Al2O3) or other phases not
considered in these diagrams may be formed.

(d) Polycrystalline Aluminas

The sintering aids added to the (L) material degrade its purity but it is

anticipated that the MgO isilithin its solubility limits in A120,.

No magnesium spine1 was found by optical metallography or the electron microprobe.

The role of Sin2 is less clear and it may have enhanced solubility in presence

of MgD.(L6) If its solubility is exceeded, it would be expected to segregate

to the grain boundaries, but this has not been observed. MgO apparently does not

segregate greatly at grain boundaries but Ca does even though its average concentrat_

ion in this lnrterial is very srr.alL. (17)


44 High Temperature Corrosion of Ceramics

More products were evident on the (L) alumina compared to the single

crystal material. Products were evident after exposures of one week to all of

the three gas mixtures at 7OO’C. As shown in Figure 2 the amount of products

increased with increasing SO3 pressure. Corrosion products were also observed

at the higher SO3 pressure at 1OOOnC (Table III), but in less quantity than at

7oonc. Exposure to the intermediate SO3 pressure at 1OOW’C did not produce any

observable products but sulfates were identified on specimens by EDS and ESCA.

The products formed on the (L) alumina did vary for apparently, the

same exposure conditions. Two types of morphologies were detected which are

shown in Figure 2 and 3 for as received (L) altiina. One is small cuboids.

The other contains many cuboids but also other shapes are evident such as laths

and films. The prodlicts were concentrated on certain grain faces of the same

grain while other faces of these grains exhibited no products indicating a

strong influence of crystallographic orientation of the grain. This nay be

due in part to surface segregation of impurities but in no case could any

impurities such as Ca or Mg be detected on these surfaces or in the products.

?iore corrosion products formed on the aluminas

L. M, S than on the single crystal. These materials are all polycrystalline and of

greater impurity content than the single crystal. Products were formed on all

polycrystalline materials under all conditions at 700°C and at least under the

higher SO3 pressure at 1000%. In general. the products formed more readily

(Table III) and were larger and in greater quantity (Figure 4) with increasing

impurity content. The samples in Figure 4 were relief polished and exposed in

tha same axpari-nt (7 x 10-3atm.S03at 700°C for one week). It shows also that

the products form preferentially at grain boundaries.


Appendix A-Gaseous Corrosion 45

Figure 2. (L) alumina surfaces (A) as received, B,C,D, exposed for one
week at 700oC to 7 x 10-4 atm. 503 and 7 x 10-3 atm. 503
respectively. More products form at higher P503 (5EM)
46 High Temperature Corrosion of Ceramics

Figure 3. L alumina surfaces exposed to 7 x 10-3 atm. SQJ at 700oC


for one week. A shows various morphologies of reaction
products and 8 only a few discrete ones. (SEM).
Appendix A-Gaseous Corrosion 47

Figure 4. Relief polished polycrystalline A1203 surfaces exposed to 7 x


10-3 atm. 803 at 700oC for one week (a) L , (b) M, (c)
another M-type (d) 8 aluminas. Products increase in size and
quantity with decreasing A1203 purity. They form
preferentially along grain boundaries. (8EM).
48 High Temperature Corrosion of Ceramics

Two types of products were found on the (M) alumina, exposed to 7 ~10~~

atm SO3 for one week at 700°c. one was larger and was a magnesium sulfate

with no ca1ci.m in it and the other smaller in size was a CaSOh with little

or no magnesium.

The lowest purfty (S) alumina had the mosf product in both number and

size after exposure. EDS analysis of the products Indicated S, Al, Si, Mg,

Ca. They appear to be mainly magnesium sulfate, calcium sulfate and their

solid solutions. The larger products were mostly MgSDq which formed on impurity

magnesium silicate grains identified by EDS before exposure. The impurity

phase( s’) in (S) occur as large grains of the size of the alumina grains

and along grain boundaries. Some smaller products, mostly CaSO4, appeared to

be on the grain boundaries. (Fi_eure 51. As indicated by analysis of the large

grains of products and of products extracted by replica techniques. silicon

and aluminum are included in the products. It is believed that the silicon

is present as silica formed by reaction of SO3 with the silicate impurities

while Al is probably dissolved in the sulfates. In the (S) alumina, this

aluminum is expected to come mostly from the impurity phases Since talc and clay

are used as sintering aids.

The evolution of the corrosion products as a function of time is shown

in Figure 6, for the same area of a relief-polished sample (S) after a sequence

of exposures from three days to 30 days. The products nucleate on the impurity

phases which are darker on the initial micrograph due to a weak acomic number

contrast and the relief polishing. The products grow wirh coalescence and

coarsening occurring ddring the latter stages. The product marked by the arrow

is a calcium-magnesium sulfate of unknown aluminum content.


Appendix A-Gaseous Corrosion 49

Figure 5. Relief-polished 5 alumina exposed to 7 x 10-3 atm. 503 at


700oC for one week. EDS spectra (1) whole micrograph, 2,3,4
areas indicated by numbers. (5EM).
50 High Temperature Corrosion of Ceramics

Figure 6(a). Same area of relief-polished (S) alumina exposed to 7 x 10-3


atm. SQ3 at 700oC. (A) before exposure, (6) for 3 days, (C)
for 10 days, (D) for 30 days. (SEM).
Appendix A-Gaseous Corrosion 51
~
tD...
QJ CJ
..:3
~'C
.-£
0.
00
.cVI QJ
0...
~ ~
/;;,~
0 ~
CJ ~ .
VI
E ~QJ
'C:3
0 QJ '6'
~ ~ 0.
QJ..~
..~QJ
~ E
..r/)0.
~0.U
QJ~r5
..~
~.,;<
o.~..
~:3°
UVl
~o~
l1Jo.o
<~~..QJ ..
'C0.~
QJ ~...
~U~
~ ~bD
E ~C::
r/) <.- ~
0...0
~~/;;,
:3
~
II
..
~
~
52 High Temperature Corrosion of Ceramics

V. GENERAL DISCUSSION

‘Ihe results that have been obtained in the present studies show that

silica and alumina are considerably more resistant to sulfur-oxygen gas mixture

than metallic alloys. Nevertheless. reaction of such materials do occur when

the thermodynamic and kinetic conditions are appropriate. In the present studies

the experimental conditions were such that no reactions of the gas mixtures

with silica were possible. In the case of alumina, however, reactions did

occur and these reactions were significantly affected by the purity of the

specimens.

On the aluminas it appears that a layer of sulfate is first absorbed

on the sample building perhaps to thicknesses giester than a monolayer. The

reactions are possible by solid solutions of the sulfate probably

with alumina in the case of the single crystal and with the impurities in

the other materials. Calcium sulfate and magnesium sulfate form at lower PSO3

than aluminum sulfate so that their solid solutions maybe anticipated to form

at lower PSO3. The formation of intermediate compounds cannot be ruled out at

this point although no such sulfates are known with Ca and Mg. Intermediate

compounds occur in the sulfation of K20.3 A1203 in which above 300% a mixed

sulfdte K~u(Su4)3 is formed. (lg).Sulfates were produced on the impurity phases

and the grain boundaries as documented for the more impure aluminas; however,

the control of the sulfate formation by the impurities appears to play a major

role in all the polycrystalline aluminas studied. This explains at least in

part the prominent role of the grain boundaries in this corrosion. The

impurities may not only increase the driving force for sulfate formation but

also provide nuclei for the growth of sulfates. A major point is that when

reaction products involving the impurities are formed, due to their greater

thermodyhamic stability, reaction of the alumina itself becomes possible.


Appendix A-Gaseous Corrosion 53

The observation that corrosion of alumina is dependent upon impurity


(19)
content is consistent with data in the literature for reaction of Sic and Si3N4

with oxygen where impurities such as Mg and Ca have been found to accumulate

in the oxidation products and thereby significantly modify the protective

properties of the scales that are formed on such materials.

In view of the results that have been obtained with the alumina samples,

deposits of Na7S04 can be expected to significantly influence the corrosion of

alumina (hot corrosion) because these products should have significant solubilities

in NazS.04. This effect will be especially pronounced when temperatures are

such that the Na7304 is liquid. Moreover, liquids can be expected to be formed

below the melting ooint of Na7SOh due to the formation of phases involving

solution of Na7S04 and impurity phases. This research is continuing with major

emphasis on the hot corrosion of ceramics.

IV CONCLUSIONS

(1) Silica and alumina are very resistant to corrosion by SO3 gases.

(2) No interaction was found between SO3 and fused silica.

(3) Small amounts of products were formed on the aluminas.

(4) The products identified as sulfates formed only under the more favorable

conditions on the high purity sapphire. They formed in greater quantities

and at lower PSC3 on the more impure polycrystalline materials.

(5) The general trends of the sulfate formation are in qualitative agreement

with predictions from thermodynamics which are that the corrosion of alumina

in SO3 gases becomesmore favorable at higher SO3 pressures and lower

temperatures. It is suggested that the aluminum sulfate formed on the

Single crystal is stabilized by solid solution with alumina.

(6) The severity of the corrosion increases with the impurity content.

(7) The severity of the corrosion increases with time.

(8) In the polycrystalline aluminas the impurities dominated the corrosion


54 High Temperature Corrosion of Ceramics

by formation of various solid-solutions of CaSO;- MgS04 and Alp (So,)-, _

The corrosion products formed preferentially on impurity phases and at grain

boundaries.

Acknowledgements:

The authors thank the 3M and Saxonburg Ceramics Companies for donating

materials, B. Draskovich, M. Zedar and L. Fisher for contributions to the

preparation of the manuscript.


Appendix A-Gaseous Corrosion 55

References:

1. Anon., " Reliability of Ceramics for Heat Engine Applications".

National Materials Advisory Board, Commission in Sociotechnical Systems

NMAB-357.

2. A.F McLean, "Ceramic Technology for Automotive Turbines" ~~11. AIII. &ram.

Sot. 61 (8) 861-71 (1982)

3. A.R. Cooper, "Kinetics of Refractory Corrosion" Ceram. Eng. Sci. Proc.

2. 1963-89 (1981)

4. R.H. Barkalow and F.S. Pettit, "Mechanisms of Hot Corrosion Attack

of Ceramic Coating Materials" Conf. on Advanced Materials for Alternate

Fuel Capable Directly Fired Heat Engines, Castine. Maine, Aug. 1979.

5. H.R. Kim u Gaseous Corrosion of Oxide Ceramics" M.S. Thesis

University of Pittsburgh, 1983.

6. H.R. Kim, J.R. Blachire and F.S. Pettit "Gaseous Corrosion of Ceramics",

paper 11411.164th Meeting Electrochemical Society, Washington, D.C. Oct. 1983.

7. B. Draskovich, J.R. Blachere and F.S. Pettit, "Hot Corrosion of Silicon

Nitride and Silicon Carbide", Basic Sci. Div. Fall Neeting Am. Gram.

sot., Columbus, Ohio, 1983

8.a W.D. Kingery, H.K. Bowen and D.R. Uhlmann, "Introduction to Ceramics",

2nd edition, Wiley, 1976.

b. A.R. Cooper, Jr and W.D. Kingery, "Dissolution in Ceramic Systems:

I Molecular diffusion. Natural Convection and Forced Convection Studies

of Sapphire Dissolution in Calcium Aluminum Silicate", J. Am. Ceram.

Sot. 41 (1) 37-43 (1964)


56 High Temperature Corrosion of Ceramics

c. L. Reed and L.R. Barrett, "The Slagging of Refractories

II lhe Kinetics of Corrosion", Trans. Brit. CSUII. SOC. 63 ~~‘h

534 (1964)

9. M. Millard. H. Wuttig and H.U. Anderson, "Influence of Grain Boundaries

on Liquid Corrosion of Al203 and N&l" paper No. 76-SII-82, An. Gram.

sot. May 1982

10. K.K. Kappmeier, ItThe Importance of Microstructural Considerations in

the Performance of Steel Plant Refractories" in Ceramic Microstructures,

R.M. Fulrath and J.A. Pask eds. Wiley, 1968, p.22

lla..D. Cubicciotti and K.H. Lau "Kinetics of Oxidation of Hot-Pressed

Silicon Nitride Containing Magnesia", J. Am. Gram. Sot. 61 (11-12)

512-17 (1978)

b. S.C. Singhal "Oxidation Kinetics of Hot Pressed SIC".

J. Mat. Sci l_l, 1246-53 (1976)

12. R.A. Gardner, "The Kinetics of Silica Reduction in Hydrogen"

J. Solid State Cliem9 336-44 (1974)

13. M.S. Crowley, "Hydrogen-Silica Reactions in Refractories"

Bull. Am. Ceram. Sot. 46 (7) 679-82 (1967)

14. L.J. Trostel, Jr, "Stability of Alumina and girconia in Hydrogen",

Am. Ceram. Sot. Bull 46 (12) 950-52 (1965).

15. D.E. Day and F.S. Gac. "Stability of Refractory Castables in Steam-

N2 and Steam-CO Atmospheres at 199OC, "Bull Amer. Ceram. Sot. 5f!

(7 1 644-48 (1977)

16. A. Negelberg, private communication.

17. I.K. Lloyd and H.K. Bowen "Iron Tracer Diffusion in Aluminum Oxide"

J. Amer. Ceram. Sot. 64 (12) 744-47 (1981)


Appendix A-Gaseous Corrosion 57

18s. W.C. Johnson "Magnesium Distributions at Grain Boundaries in Sintered

A1203 Containing Mg A1204 Precipitates," J. Am. Gram. Sot. 61 (5-6)

234-237 (1978).

b. P.E.C. Franken and A.P. Cehring 'Grain Boundaries Analysis of ?!gO-Doped

A1203", J. Nat. Sci., l6, 384-88 (1981)


19. N.G. Krlshnan and R.W. Bartlett "Kinetics of Sulfation of Reduced Alunite"

Environ. Sci. Tech, ,1,923-929 (1973)


Appendix B- Hot Corrosion of Silica
M.G. Lawson, H.R. Kim, F.S. Pettit and J.R. Blachere

INTRODUCTION

There is a great emphasis on higher working temperatures for heat engines

which is dependent on advanced materials. As progress has been made in the

science and technology of ceramics leading to greater reliability of ceramic

parts, they are beginning to be incorporated in car engines and other high

temperature-high stress device&). Since ceramics are becoming serious

candidates for these applications, understanding their corrosion resistance under

environmental conditions related to these uses becomes important. While

ceramic compounds, in particular oxides are quite stable at high temperatures in

air, the stability of ceramics under various conditions which might be

encountered, e.g. in a heat engine, must be established. At high temperatures a

material may not be stable in presence of some gases, i.e. metals in oxygen, and

the degradation of a material by reaction with gases is called gaseous corrosion.

In combustion systems, various compounds such as NagC03, Nags04 may be

generated in addition to combustion gases t2). Below some temperatures these

compounds condense in the engine and often enhance the gaseous corrosion in a

complex process named ‘hot corrosion’. This type of degradation is well known in

superalloys but has not been studied in detail in ceramics.

Silica forms as a protective scale at high temperatures in oxidizing

atmospheres on silicon carbide and silicon nitride and on some coatings on

superalloys. Vitreous silica is protective in clean oxidizing environments because

it is continuous with some ability to self heal at higher temperatures. However

above llllO-12000C it is less protective as it begins to crystallize@). Although

growing scales may differ in properties from the same materials in bulk, it was

To be submitted to the Journal of the American Ceramic Society

58
Appendix B-Hot Corrosion of Silica 59

believed that valuable information could be obtained from a study of hot

corrosion of silica as a part of a systematic investigation of the hot corrosion of

ceramics (4). This paper covers the study performed under conditions relevant to

gas turbines, with deposits of sodium sulfates.

HOT CORROSION

Hot corrosion is dependent upon temperature and the gas atmosphere since

they affect the stability of the deposit and its reactivity with high temperature

materials. If the deposit is in equilibrium with the atmosphere, equation (1)

applies:

Na2S04 (1) = SO3 (g) + Na20 (1) (1)

Kl = PSO3. a(Na20)

The activity of Na20 and the pressure of SO3 in the melt are fixed by the

temperature and PSO3 in the atmosphere as defined by Kl. For pure Na2S04, its

activity can be taken as unity, and as a(Na20) increases, PSO3 decreases and

viceversa. An Na20 activity of 2~1O-l~ was calculated from equation (1) for the

equilibrium between the melt and the S03-containing atmosphere selected for

the hot corrosion experiments at 1OOOoC. The corresponding activity at 7OOoC is

lower. In presence of pure oxygen, the activity of Na20 in the melt is not

defined by reaction (l), it is expected to increase to values of the order of 10-6

to 10s4 since SO3 and SO2 are evolved from the Na2S04 to the gas. These

values are high and such Na2S04 melts are considered basic. However

equilibrium of the deposit with the atmosphere may be established only at the

beginning of the hot corrosion process and acid or basic behavior depends on the

specific reactions that occur between the sulfate and the ceramics.
60 High Temperature Corrosion of Ceramics

EXPERIMENTAL PROCEDURE

Materials

The experiments reported here were all performed on a high purity fused

silica (Corning 7940). The major impurities are 1OOOppmof ‘water’ and lo-100

ppm of chlorine. All other impurities were in the ppm or subppm range. The

samples were cut to about 1x1 cm in area and polished prior to exposure.

Gaseous Corrosion

The samples were exposed in tube furnaces at 700 and 1000°C. SO3-03 gas

mixtures flowed through a platinized catalyst in the furnace before passing over

the samples. The initial SO3 contents of the gas were usually 1, 0.1 or O.Ol%,

balance oxygen at a total pressure of 1 atmosphere. The corresponding equilibrium

PS03 and PSO3 are given in table I. The gas flow rate was 1 cm3/s. The exposure

time was usually one week but varied from one day to one month. Weight changes

were checked on an analytical balance. The major tool was an SEM equipped

with X-ray spectrometers (Energy Dispersive, EDS and crystal spectrometers

WDS) for elemental microanalysis. These experiments were supplemented by

ESCA. (X-ray photoelectron spectroscopy).

Hot Corrosion

The polished specimens were placed on a hot plate and sprayed with an

aqueous solution of Na3S04 to form a continuous layer of sulfate usually with a

loading of 5mg/cm2. The samples were placed in a boat lined with platinum foil

and exposed in tube furnaces for times of 1 to 495 hours. Flow conditions and

catalysts were as described for gaseous corrosion. The extreme conditions of


Appendix B-Hot Corrosion of Silica 61

Table I

SO3 Pressures for Corrosion Experiments

Temp. (OC) Initial (SO2) (1) S03(atm.) (1)

700 1 7x10-3 (2)


0.1 7x10-4
0.01 7x10-5

1000 1 1.5x10-3 (2)


0.1 1.5x10-4
0.01 1.5x10-5

(1) 96 S02-Balance 02 - total pressure 1 atm.

(2) Environments for Hot corrosion experiments.


62 High Temperature Corrosion of Ceramics

table I were selected since they correspond to relatively acidic and basic

conditions. They are the initial mixture of 1% S02-balance oxygen (acidic) and

pure oxygen (basic) at 700 and 1000°C. Since Nags04 is not molten at 700°C, an

equimolar mixture of CoSO4 and Na2SO4 was used at that temperature. Cycling

experiments were also performed with or without water washing and reapplying

the salt usually on a 45 hours cycle.

The weight of the coupons was checked on an analytical balance with a

sensitivity of 0.1 mg before salt application and after exposure once the salt was

washed off in 10 cm3 of high purity water per cm2 of exposed area. The washing

lasted 20 minutes in an ultra sonic cleaner. The washwater of some of the

samples was analyzed by EDS in the SEM using a procedure adapted previously

(5~6). A single drop of the washwater was deposited on a beryllium coupon and

dried. The deposit left on the coupon was analyzed and the spectrum was

processed semi-quantitatively* after subtraction of a water spectrum obtained

under the same conditions and scaled to the same background. The specimens

were examined in the light microscope and the SEM before and after removal of

the salt, however the SEM with its X-ray spectrometers for microanalysis were

the major tools of this investigation. When possible, crystalline products were

identified by X-ray diffraction with a diffractometer. The wetting angle of the

salt on the substrates was measured on photographs of the droplet profiles taken

in the SEM at room temperature. Contact angles have been measured previously

by this method(7). The crystallized thicknesses were measured on SEM

micrographs of cross sections of the coupons.

* SSQ software by Tracer-Northern


Appendix B-Hot Corrosion of Silica 63

RESULTS AND DISCUSSION

Gaseous corrosion

The gaseous corrosion(*) was studied for comparison with the hot

corrosion(g). No weight changes were detected, no products were visible in the

SEM and no sulfur was detected by EDS or ESCA on the surface of the specimens

after exposures up to one month under the conditions given in table I. While the

smallest weight change that could be measured (0.1mg/cm2) would require a

uniform deposit of the order of 1 urn, the SEM with careful examination and

ESCA are much more sensitive. An estimate of the detection limits of discrete

products in the SEM is

10m7 to lo-* g/cm2. Under these assumptions* sulfur in the products seen in the

SEM could be detected by EDS. ESCA is a surface technique, not a

microanalysis technique, and its detection limits were of the order of lo-* to

10mg g/cm2.

No evidence of degradation or formation of solid products containing sulfur

was found in these experiments. This might have been expected since there are

no reports of the existence of a silicon sulfate, and no silicon sulfide was

expected to form under the experimental conditions since the oxygen pressure

was always around one atmosphere. The formation of silicon sulfide has been

discussed recently(lOl. No experiments were performed above 1000°C because

sulfate deposition becomes less favorable at higher temperatures. No

devitrification of the silica was observed after any of these exposures at 700 and

1OOOoCin many experiments for times as long as 720 hours. Others have

reported that SO2 has a fluxing action on fused silica.

* A 1% coverage with islands 20nm thick containing 20% sulfur was assumed for
the calculation of the limits of detection using SEM and EDS. In ESCA, 0.1% of
a sampling depth of 5nm was assumed.
64 High Temperature Corrosion of Ceramics

Wetting by sulfates

The wetting angles of the salts used in the hot corrosion experiments varied

with the environments. They are given in table II. The general wetting morphology

developed in a few minutes at the temperature of the experiments. Some evolution

occured with time with the wetting increasing from 1 hour to 24 hours. In general

the sulfate wetted the silica better under basic conditions than acidic conditions

under which discrete small droplets were formed. After 24 hours under acidic

conditions at 1OOOoCthe wetting angle was 24O on the droplets and 13O on an

exceptionally large drop about 0.02cm across. Under basic conditions at 1000°C

the salt wetted most of the coupon after 1 hour, and a wetting angle of 2O was

measured; after 24 hours the wetting was continuous.

In some cases, the salt formed drops of several sizes and different wetting

angles were observed for different size-range (table II). The wetting angle

measurements were very reproducible for given size range. The angles reported

are called wetting angles and not contact angles since the latter terminology

may imply that equilibrium was reached. It is clear from the increased wetting

with time and the influence of atmosphere on the wetting that reactions are

occuring between the sulfate and the silica surface, even under acidic conditions.

The wetting tends to increase with the affinity between the liquid and the

substrate surface and the greater wetting of the silica by the salt under basic

conditions reflects the higher affinity of the basic salt for the acidic silica.

It was established by careful microscopy of the same areas of samples

which were cycled, with washing off the salt and reapplication between cycles,

that the droplets did not reform preferentially on previous droplet sites. This

suggested that no preferential local attack would occur under acidic conditions.

This conclusion was substantiated by a long term experiment reported below.


Appendix B-Hot Corrosion of Silica 65

Table II

Wetting Angles of Na2 SO4 on Silica (O)

7oooc (1) 1ooooc (2)


lhr. 24hr. lhr. 24hr.

Acidic 52 36-49 t3) 13-24 t3)

Basic - 26 t4) 2 0

(1) (Nag Co) SO4

(2) Na2 SO4

(3) Range of angles with drop sizes - smaller


angles for larger drops.

(4) Salt phase separated.


66 High Temperature Corrosion of Ceramics

Hot Corrosion

Under acidic conditions, the salt did not wet well the silica and it formed

droplets. At 700°C the weight changes were small, for example, after removing

the deposit by washing there were losses of 0.1 mg/cmz after 1 hour and 1.1

mg/cm2 after 24 hours. No evidence of devitrification was apparent. Some

localized corrosion occurred under the droplets, as shown in figure 1. It is

suggested that this ‘wormy’ void texture formed by dissolution of sodium silicate

in water. This corrosion was very limited as indicated by the small weight

losses. The attack occurred preferentially under the salt at the perimeter of the

drops suggesting an interaction with the atmosphere. No spalling occurred after

495 hours of cyclic exposure of the silica. The sample was washed and salt

reapplied (5 mg/cm2) between cycles. The substrate had a strong texture after

exposure and a region 33 urn thick contained sodium. This attack over the whole

surface of the coupons is consistent with the separate observation that the

droplets did not reform at the same locations after reapplication of the salt.

Under acidic conditions, at 1000°C the wetting had improved and the silica

was devitrifying under the drops. After 24 hours the cristobalite formed

characteristic patterns intermediate between spherulitic and globular (figure 2).

Devitrification occurred also under very small drops and no evidence of attack

was detected outside the drops. The weight loss of 0.3 mg/cm2, was smaller

than at 700°C. This can be explained by the greater stability of the cristobalite

compared to the vitreous silica which will result in the formation of less sodium

silicate to dissolve in the salt. The cristobalite layer which did not spall

probably offered some general protection to the vitreous silica but it was not

established if preferential attack occured at the boundaries between the

spherulate fibrils as reported for silicon carbide under similar conditions(l2).


Appendix B-Hot Corrosion of Silica 67

Figure I. Hot corrosion of fused silica under acidic conditions,


24 hoursat 700oC. SEM of sulfate drop area after
washing off sulfate showing wormy voids formed
preferentially at edge of drop.
68 High Temperature Corrosion of Ceramics

Figure 2. Fused silica exposed for 24 hours under acidic conditions


showing coarsened spherulites under sulfate drops (after
washing).
Appendix B-Hot Corrosion of Silica 69

Under basic conditions at 7OOoCthe salt tended to decompose with

extensive formation of cobalt oxide which complicated the interpretation of the

data, particularly weight change. Under the salt drops there was limited

formation of cristobalite spherulites mixed with some random globules after 24

hour exposure and some localized cracking (figure 3) but there was no evidence

of spalling or extensive attack of the substrate.

The most extensive degradation of vitreous silica occurred under basic

conditions at 1OOOoC. After 1 hour the salt wetted the coupon almost

completely and a layer of cristobalite spherulites had formed under the salt. As

shown in figure 4, the fibrils of the spherulites had already broken down

significantly into arrays of globules. The spherulites had grown to impingement

with radii of the order of 30 pm, underlining the high velocity of the surface

crystallization. After 24 hours nothing was left of the spherulitic surface

morphology as tridymite had formed at the silica-salt interface. Cristobalite

separates the tridymite from the vitreous silica. After 212 hours the tridymite

near the surface had coarsened into laths about 15 pm wide and over 60 pm long.

The crystalline layer spalled extensively. The washwater analysis indicated that

significant silicon was water soluble with the sulfate after 1 hour, however this

was no longer found after 10 or 100 hours. The sulfur was never depleted as the

Na/S ratios in the water remained of the order of 2 (t 0.4). This suggests that

less reaction occurred after a continuous crystalline layer was formed and that

at the longer times sodium silicates with low solubility in the sulfate at 1OOOoC

formed between the crystalline silica and the sulfate.

It became clear, as the previous results were obtained, that the major

mode of degradation of vitreous silica under hot corrosion conditions was due to

the crystallization and associated spailing. Qualitatively the extent of the

degradation at constant time increased in the order:


70 High Temperature Corrosion of Ceramics

Figure 3. Fused silica exposed under basic conditions at 700oC for 24


hours. Limited nucleation occured in small portions of drop
areas.
Appendix S-Hot Corrosion of Silica 71

Figure 4. Fused silica exposed to basic conditions at lOOOoC for 1 hour.


The fibrils of the spherulites have broken up into globular
arrays.
72 High Temperature Corrosion of Ceramics

Acidic 7000 < Basic 700° < Acidic 1000° < Basic 1000°C

The degradation increases with increasing Nag0 activity in the salt and decreasing

equilibrium PSO3 under these conditions. Since the hot corrosion results

emphasized the degradation associated with the devitrification, quantitative

data was obtained on the crystallization.

Kinetics of Crystallization

The thickness of crystallized silica was measured on fused silica under

basic hot corrosion conditions at 1000°C for times up to 300 hours. The

measurements were made on the unspalled region of fractured specimens. The

isothermal kinetics are shown in figure 5. They are for the propagation of the

cristobalite-glass interface. The plot is parabolic and the incubation time is

under an hour. A long term experiment was performed with exposure cycles of

46 hours. In between cycles, the salt was washed off and new salt was reapplied

as described previously. Since this was done at room temperature, the samples

were thermally cycled. Extensive spalling of the crystalline layer occured and

the thickness of the remaining silica glass was measured where the fracture had

propagated at the interface. The plot of the thickness crystallized for 6 cycles

is given in figure 6. The plot is linear instead of parabolic and the total thickness

crystallized (about 1000 urn) is over twice that without cycling. In a separate

experiment of 100 hours total duration, a sample was cycled without reapplication

of salt by cooling to room temperature after 1 hour and 10 hours of exposure. It

had the same layered structure as the other samples after the long exposures,

cycled or uncycled, (figure 7). One can identify a thin sodium silicate layer, a

tridymite layer and a cristobalite layer over the glass. The crystallized layer

had a thickness of the order of 400 urn which is also about twice the thickness
Appendix B-Hot Corrosion of Silica 73

I/2 l/2
TIME (Hours)

Figure 5. Kinetics of cr,stallization of silica glau during basic hot


corrosion at lODOW in Oxygen.
74 High Temperature Corrosion of Ceramics

900

800

600

500

400

300

200

IOC

2 3 Y 5 6
0 I

CYCLES (48 Hour, Resa It)

Figure 6. Kinetics of crystallization of silica glass exposed to basic hot


corrosion at 1OOOoC in Oxygen, with 45 hours cycles and
reapplication of sulfate. Note linear kinetics.
Appendix B-Hot Corrosion of Silica 75
+"..'0
01 ~ C
1n~0I
C tI ~
0 ~
.-In -
.-~
+" e
'0>.>.
Col'O .-
0-
tI ...
M-ln
tI In
.-~ 01
In~~-
.aoIC+"ba
.-~ ..
-tI -
C tI .-.c
.->.- +"
'0 tI .;n .c
~- e +"
In 01 .-
0 e ~ ~
0. .-
><..'0...
..~ 0 tI
-.Clnol
0I...~+"
tI C
.-.c ~ 0
-+"w
tI
1n~>.C ~ .-
'0
~~.Cln
In ~ E-o .-
~O .c
IntI
OO~:c
CO-~
O ba
c ~
... 0 ...
..+" .-
4i...0-
InUO~
1n0 eO
Inoln...
E o ~ .~
U °.C..
tI
~
GI
..
~
~
76 High Temperature Corrosion of Ceramics

crystallized for the same time under isothermal conditions. The larger rates of

crystallization and the change in the type of kinetics under cycling show the

strong influence of the salt on the crystallization but the increased damage under

the cycling experiments is due to the stresses generated during the cycling and

not to the replenishing of the salt.

General Discussion

The wetting was greater under basic than under acidic conditions. The

wetting angle decreased in the sequence:

Acidic 700 > Basic 700 > Acidic 1000 > Basic 1000°C

The increased wetting at higher temperature (lOOO°C) under acidic conditions is

due to a decreased acidity of the salt at higher temperature as predicted from

equation (1). The negative free energy of a reaction between a substrate and a

liquid enters in the interfacial energy balance of a sessile drop so as to increase

the wetting tendency of the solid by the liquid(ll). Therefore the wetting

behavior of the salt and in particular the influence of the atmosphere on it

indicate a tendency towards basic fluxing according to a reaction of the form:

xSi02 + yNa20 = yNa2OxSiO2 (2)

I(2 = a&/ (aSiO2)x. (aNa2O)y

Assuming that the silicate activity asil=l, y=l and x=2, the equilibrium

constant at 1000°C gives aNa20=9.2xlO- 11. Therefore reaction 2 should proceed

to the right when aNa is greater than this value. The experimental conditions

for equilibrium with 1.5x lo-5 atm So2 which correspond to aNa20=2xlO-15

should therefore not be .sufficient to form the silicate. A similar analysis was

performed by Jacobson et. al for the corrosion of silicon carbide(13). However


Appendix B-Hot Corrosion of Silica 77

the wetting behavior, the crystallization and the reaction of the sulfate with the

silica indicate that the equilibrium is shifted to much lower values of aNa so

that at least locally all conditions for this work might have been sufficiently

basic to permit for reaction 2 to proceed to the right. This can be explained at

least partially by aSiO2 > 1 since the glass is not an equilibrium phase and the

OH and Cl impurities decrease its stability. A discrete silicate phase was observed

under long term basic exposures at 1000°C. It is only expected to form under

strongly basic conditions where the silicate activity might get close to 1. In the

less basic experiments and early in the basic experiments the silicate formed is

dissolved in the sulfate and has activities much less than 1.

Some limited reaction occurs between the fused silica and the sodium sulfate,

even under the less basic conditions at 700°C (acidic 700°C in table I.) Under

these conditions, with cyclic replenishing of the salt, significant hot corrosion

accumulated over long term exposures. The quantitative interpretation of the

hot corrosion on the basis of reaction (2) was limited by the presence of cobalt in

the melt at 700°C and the overwhelming effect of the crystallization under all

conditions except the less basic (acidic 7OOOC). Neglecting these complications,

the major interactions between the salt and the silica occur at their interface.

They are the dissolution in the salt of the silicate formed via reaction (2) in the

salt and diffusion of sodium into the vitreous silica. Accordingly, under the more

basic conditions, the Si content of the washwater increases early during the

exposure but decreases later as the silica at the interface crystallizes; after a

long time a silicate layer is detected on top of the crystallized silica (after

washing). Under the less basic condition, this reaction is localized leading to the

observed ‘wormy’ texture. It is proposed that the reaction begins at impurity

clusters in the fused silica. The OH and Cl impurities are connected only to one
78 High Temperature Corrosion of Ceramics

silicon and they may generate microregions with more open structures more

susceptible to attack under reaction (2). Since for the less basic condition (acidic

7OOoC), the driving force for the reaction is small, the influence of these defects

is emphasized and it will be suppressed by increases in concentration of the gaseous

products in the salt. As a result it occurs more readily near the edge of the salt

drops (figure 1) where the SO3 produced is released more easily to the atmosphere.

Under more basic conditions these defects may play a role, but it is not as obvious

in the surface morphology observed after corrosion. Others have reported no

interaction of Na2SOq with fused silica under acidic conditions(14). This was not

the case in any of the present experiments which included lower temperatures.

Some of that difference in behavior may be due to the influence of the

impurities discussed above. Silica scales formed in combustion environment are

expected to contain OH impurities, so that the samples studied here are

expected to be more representative of industrial applications.

Sodium diffuses rapidly into fused silica with reported diffusion

coefficients (DNa) of the order of 10s6 cm2/s at 1000°C(15). In type III silica

glass such as used in the present experiments, the OH’s may slow down the

sodium. In a recent discussion it was pointed out that impurity and structure

seem to affect DNa( 13). Sodium can be incorporated in the glass as a network

modifier(17) through the reaction:

Na20 + Si-0-Si = 2Si-O- + 2Na+ (3)

(CSiO)2. (CNa)2
g3 =_________________
Cst. aNa

in which an oxygen bridge (Si-0-Si) has been broken into two single bonded

oxygens to incorporate the oxygen into the network. The sodium ions, as

network modifiers, are located in holes of the structure near the single bonded
Appendix B-Hot Corrosion of Silica 79

oxygens. The oxygen bridges most likely to be broken are the more strained in

the structure(l6), and their concentration is Cst. Although the sodium ions are

associated with single bonded oxygens they are mobile in the structure. These

considerations are important in the oxidation of silica formers and similar

concepts have been used by Schaeffer to discuss the influence of water on the

oxidation of silicon(lg). The penetration of sodium is expected to increase

rapidly with the activity of Nat0 in the melt as it increases the driving force

and probably the mobility of Na+ for the low concentrations of Na+. Diffusion of

sodium in silicate glass has been discussed in terms of a vacancy diffusion(20).

Initially it must enter the glass structure by reaction 3 which could be rate

controlling. If this reaction is slow it is likely that ion exchange between protons

of the hydroxyls present in the glass and sodium might play a role. However, the

situation is modified by the crystallization of the silica glass. Diffusion of

sodium through cristobalite is very difficult since it does not contain the

channels of vitreous silica or quartz. Diffusion is probably easier along the grain

boundaries.

Since it breaks up the network, reaction 3 is expected to promote the

crystallization of silica and the required rearrangement of the network to 6-

member rings. In this research, the silica did not crystallize under the lower

Na30 activity (acidic, 700°C) but it did at the same temperature under the more

basic condition. Qualitatively, the extent of the crystallization increased with

the Nag0 activity as expected from the previous discussion and reaction (3).

From reaction 3, the concentration of single bonded oxygens is proportional to

(aNaaO)*. The strong influence of sodium on the crystallization of silica is well

know in ceramics(21), here it has been shown that under basic hot corrosion

conditions at 1OOOoC the crystallization proceeds at a rate of the same order as

that observed above 1400°C on the same materials in oxygen or an enhancement


80 High Temperature Corrosion of Ceramics

of many orders of magnitude due to the sodium. This is in line with previous

results on the crystallization of sodium silicate glasses(21). Much greater rates

of crystallization, of the order of 0.1 mm/min. have been observed at 1000°C in

soda silica glasses with Na20 contents of the order of 1%. This shows that only

very low concentrations of sodium, in the ppm range are necessary to obtain the

crystallization rates of the order of 0.01 -0.1 um/min. observed here.

Under the more basic conditions, at 1000°C, the greater Na20 activities

lead to dissolution in the salt as well as sodium penetration in the vitreous silica,

however this is short lived because of the onset of devitrification. The

devitrification and the resulting spalling as the specimens are cooled through the

a-g cristobalite and the a-B-y tridymite inversions as well as the differential

contraction of the crystalline phases on cooling are responsible for most of the

degradation of fused silica under basic conditions. The increased degradation

under cycling is due to the cracking and spalling of the crystalline scale and

penetration of the salt through the cracks. In line with this conclusion, little

degradation was observed with no significant increase under cycling for the less

basic condition at 7OOoC (acidic 7OOoC) for which the silica did not crystallize.

Since it was responsible for most of the observed degradation the crystallization

of silica under hot corrosion conditions will be discussed extensively.

The crystalline phase formed initially was always cristobalite which forms

instead of tridymite since a lower energy path is available because of the greater

similarity between the structures of high cristobalite and vitreous silica(221. The

cristobalite was globular in morphology (figure 2-4). Under the more basic

conditions cristobalite globules were aligned in radial arrays related to their

spherulitic formation. Under less basic conditions the spherulites did not

nucleate under all the salt drops or over the whole areas covered by the drops
Appendix B-Hot Corrosion of Silica 81

and for less basic conditions yet (acidic 700°C) no crystallization was observed

even after nearly 500 hours. Under the intermediate conditions, cristobalite

globules formed randomly under the drops where no spherulites formed or in

between spherulites. This suggests that the crystallization was initiated by two

different mechanisms depending on the Na20 activity. In both cases the sulfate

reacts with the silica by basic fluxing as discussed earlier and shown in reaction

2. The relationship with SO3 pressure is obtained by combining reactions (1) and

(2). Under the greater Na20 activities the cristobalite nucleated at the sulfate-

glass interface and formed spherulites, with the well-known radial morphology of

their fibrils, however the previous results suggest that a threshold concentration

must be reached for the nucleation of the spherulites at the interface. These

fibrils coarsen and break down due to interfacial instability giving the radial

arrangements of globules. The spherulites grow in a thin surface layer in which

high Na20 activities promote the crystallization of silica by breaking down the

network, (reaction 3) and silica is transported rapidly through the liquid phase.

On the other hand under intermediate basic conditions, the surface

reaction of the sulfate with the silica is less extensive, the silicate ions dissolved

in the sulfate diffuse into the sulfate away from the interface on which no

crystallization is initiated. At the same time it is likely that the solubility of

silicon, probably as silicate, in the sulfate decreases as the activity of Na20 is

decreased as reported by Kim(22) in the same range of activities. Thus as the

reaction proceeds at the interface, gradients of Na20 activity and SO3 pressure

are set up across the salt. The solubility of the silicate decreases as it diffuses

away from the interface leading to supersaturation and homogeneous

precipitation of silica in the salt which generates the random globules of

cristobalite (figure 3).


82 High Temperature Corrosion of Ceramics

Discussion of Crystallization Kinetics

The crystallization of silica which has rapid parabolic kinetics under the

more basic isothermal conditions is not controlled by the transport of reactants

or products through the crystalline layer, since the crystal has the same

composition as the glass. As already discussed qualitatively, the very large rates

of crystallization are associated with high Na30 activities. The sodium breaks

oxygen bridges according to reaction (3). It is proposed that the crystallization

begins when a threshold Na30 activity is present at the silica glass-sulfate

interface generating sufficient concentrations of oxygen bridges to allow the

local rearrangement of the network to the crystalline form. This could occur by

a mechanism similar to that proposed by Fratello et. a&18) for the high pressure

transformation of fused silica to quartz. The sodium in Si-0-Na groups would

play a role similar to that of hydrogen in SiOH bonds, although the ONa bonds

are more ionic than the OH bonds and as a result they were considered ionized in

reaction (3). Since it is less tightly bound to the oxygen and because of size

considerations, the Na+ is more mobile than the protons(24). The active defects

sre still the non-bridging oxygens which are very mobile in combination with

either thermally created single bonded oxygens, SiOH or other SiO- Na+ groups.

The single bonded oxygens can attack oxygen bridges by simultaneous bond

formation and breaking, and reshape the network into six member rings. After

nucleation the crystallization front propagates at rates many orders of

magnitude greater than in the pure system. From the present results this is due

to the sodium which catalyses the crystallization. It has to be present at the

cristobalite-glass interface, although it could not be detected with the electron

microprobe, except in special cases. Only very small quantities of sodium are

required for the previous mechanism which can rapidly propagate a crystal ledge

into the glass.


Appendix B-Hot Corrosion of Silica 83

In such an interfacial reaction a sodium can rapidly rearrange many bonds and

the controlling step is the break up of a strained Si-0 bond(l@. Adapting this

kinetic model the rate of advance of the interface (u) should be a function of the

stoichiometry of the silica and directly proportional to the concentration of

sodium which take the place of the hydroxyls in the model. The hydroxyls have a

mobility 3 to 4 orders of magnitude smaller than Nalz41. It has been suggested

previously that most hydroxyls in synthetic silicas are strongly bound and nearly

sessile(z51. The sodiums are highly mobile although in small concentrations.

The concentration of sodium ions is not constant in the moving interface.

During the induction period some sodium diffuses into the glass. Soon this

diffusion is essentially stopped by the formation of a continuous cristobalite

layer, as discussed earlier. In the proposed model, the sodium at the

cristobalite-glass interface is supplied by this original diffusion through the glass

and essentially trapped once a significant crystalline layer has formed.

Neglecting any diffusion through the crystalline layer and the solubility in the

cristobalite, one can approximate the concentration of Na in the glass at the

interface with the cristobalite as a spike of constant area, representing the

sodium injected initially into the glass, which spreads as a function of time

(figure 8). This spread which is due to the sodium diffusion into the glass is

slowed by the rejection of the sodium ahead of the crystallization front. Based

on this model, the concentration of sodium CNa at the interface is expected to

decrease with time in line with the observed decrease of the rate of

crystallization. Quantitatively the constant sodium in the spike can be written

as:

cNa.(Dt)+ = s
84 High Temperature Corrosion of Ceramics

idym.

n
%I,0

Figure 8. Proposed mechanism of crystallization of silica during basic


hot corrosion. (1) Diffusion of Na into glass, nucleation of
cristobalite; (2) A continuous layer of cristobalite forms
rejecting Na; (3) with cristobalite growth the Na
concentration at the cristobalite-glass interface decreases by
diffusion into the glass.
Appendix B-Hot Corrosion of Silica 85

in which D is the diffusion coefficient of Na+ into the glass and t is the time.

And the rate of crystallization u is

u = dx/ dt = A CNa+ = A S / (Dt)*

which integrates to x = B tt / Dt and if D is

independent of time, x takes the form:

x=kt*

which predicts the observed parabolic behavior. In the previous discussion, x is

the thickness of the crystallized layer while S, A, B, and k are constants at

constant temperature.

One could still assume that the sodium at the interface is supplied mostly

be diffusion through the crystalline layer by grain boundary diffusion however

the cristobalite changes grain size by over an order of magnitude and it transforms

to tridymite at the salt interface side without significant departures from the

parabolic behavior. Furthermore Na is not consumed in the crystallization reaction,

it is rejected into the glass at the cristobalite-glass interface and since it

catalyzes the crystallization one cannot expect a falling rate of crystallization

such as the observed parabolic behavior. Therefore it is concluded that the

mechanism proposed above provides a better explanation.

Hot Corrosion of Silica Pormers

The present results are in general agreement with the protective properties

of silica scales on ceramics (silicon carbide and silicon nitride) and coatings on

superalloys. The silica was not affected by the SO3-SO3 containing environments

of this study, however sodium plays a dramatic role under hot corrosion conditions.
86 High Temperature Corrosion of Ceramics

The vitreous silica has a tendency to react with Na2S04 by basic fluxing,

particularly at low PSO3, but more importantly it is very sensitive to

crystallization which promotes spalling through temperature cycling. Under

these conditions silica scales should be strongly affected and fundamentally not

afford good protection to silica formers. The validity of these conclusions will

depend on the purity of the silica formed, the nature of the impurities associated

with the sealest and the rate of growth of the scale since thick scales are more

sensitive to spalling. It is likely that OH impurities promote the hot corrosion

and crystallization. Although only indirect indications are presented here, it is

expected from the role of OH in glasses( 18p2@. This is important for high

temperature structural applications since water is a major product of

combustion.

The results suggest also that some silicate scales might be more desirable

than vitreous silica if they could be used at high temperatures without

devitrification.

CONCLUSIONS

- No evidence of gaseous corrosion was obtained in atmospheres containing


initially up to 1% SO2 and balance oxygen at 700 and 1000°C for times up to 720
hours.

- No devitrification was observed under these conditions and similar


experiments in pure oxygen.

- The wetting of the silica by the sulfate varied significantly with the PSO3
and the Nap0 activity. It increased with aNa20. It was greater under basic than
under acidic conditions at both temperatures. Under basic conditions at 1000°C
the wetting was complete.

- In all cases there was a tendency towards basic fluxing which increased
with the activity of Na20.

- Some limited corrosion was observed under acidic conditions at 7000C.


At 1000°C under the same conditions, the silica devitrified and little reaction
was observed between the salt and the silica.
Appendix B-Hot Corrosion of Silica 87

- The most extensive degradation of vitreous silica occurs by crystallization


and the associated spalling during temperature cycling. In general it increased
with the activity of Na20 and was very severe under basic conditions (pure oxygen
atmosphere) at 1000°C.

- The sodium accelerates (catalyzes) dramatically the devitrification of


silica and the rate of crystallization at 1000°C under basic conditions is of the
same order as that observed by others at 1400°C in air (without sodium).

- The kinetics of crystallization at 1OOOoC under basic conditions were


parabolic with a short incubation time. Cycling increased the damage and changed
the kinetics to linear. The extra damage was associated with the strains due to
cycling not with additional salt applications.

- The parabolic behavior observed under isothermal conditions can be


explained with a model in which the crystallization is controlled by the sodium at
the crystal-glass interface which diffuses into the glass prior to crystallization.

- The need for the development of vitreous coatings for the protection of
ceramics and metallic alloys is stressed.

Acknowledgements

The authors gratefully acknowledge the support of Basic Energy Science Division

of the Department of Energy for their financial support.

REFERENCES

1. a. M. Taguchi, Adv. Ceram. Mat. 2 [4] 754-762 (1987).

b. Heat Engine Ceramics, Bull. Am. Ceram. Sec. 64 [Z] 268-294 (1985).

2. N. Birks and G.H. Meier, Introduction to High Temperature Oxidation of


Metals, Arnold 1984, Chap. 8.

3. P.J. Jorgensen et. al, J. Am. Ceram. Sot. 42 (12) 613-616 (1959).

4. J.R. Blachere and F.S. Pettit, High Temperature Corrosion of Ceramics,


DOE Basic Energy Sciences DEFG OZ-84-ER45117.

5. B. Draskovich, Hot Corrosion of Silicon Mitride and Silicon Carbide, M.S.


Thesis, University of Pittsburgh (1984).

6. J.I. Goldstein et. al, Scanning Electron Microscopy and X-ray Microanalysis,
Plenum, (1981), p. 378.

7. L.E. Murr, Interfacial Phenomena in Metals and Alloys, Addison-Wesley,


(1975), p. 69.

8. H.R. Kim, Gaseous Corrosion of Oxide Ceramics M.S. Thesis, University of


Pittsburgh (1983).
88 High Temperature Corrosion of Ceramics

9. M.G. Lawson, Hot Corrosion of Silica and Alumina, M.S. Thesis, University
of Pittsburgh (1987).

10. 0.0. Vander Biest et. al, J. Am. Ceram. Sot. 70 [7] 456-59 (1987).

11. I.A. Aksay et. ai, J. Phys. Chem. 78 [12] 1178-1183 (1978).

12. J.R. Blachere et. al, Paper #77-BEG-86P, Basic Science Div. Am. Ceram.
Sot., New Orleans (1986).

13. N.S. Jacobson, Workshop on Corrosion in Ceramics, Pennstate, Nov. 1987.

14. N.S. Jacobson et. al, Workshop on Corrosion in Ceramics, Pennstate, Nov.
1987.

15. G.H. Prischat, J. Am. Ceram. Sot 5_1[9] 528-530 (1968).

16. G.H. Prischat and W. Beier, J. of Non-CrystaLline Solids 7l, 77-85 (1985).

17. W.D. Kingery et. al, introduction to Ceramics, Wiley (1976) p. 103.

16. V.J. FrateIlo et. al, J. Appl. Phys. 51 [12] 6160-64 (1980).

19 H.A. Schaeffer, J. of Non-Crystalline Solids, 38 and 39 (1980), 545-550.

20. Z. Boksay: “Mass Transport in Non-Crystalline Solids”, in The Physics of


Non-Crystalline Solids, G.H. Frischat, Ed. Trans. Tech., (1977) p. 428.

21. H. Rawson, Inorganic Glass-Forming Systems, Academic Press (1967), p. 53.

22. Reference 17, p. 314.

23. G.M. Kim, M.S. Thesis, University of Pittsburgh, (198-).

24. R.H. Doremus, Glass Science, p. 168.

25. R.W. Lee and D.L. Fry, Phys. Chem. Glasses, 3 19 (1966).

26. Reference 24, p. 229.


Appendix C-Hot Corrosion of Alumina
M.G. Lawson, F.S. Pettit, and J.R. Blachere

INTRODUCTION

Failures due to corrosion limit the selection of materials in countless

operations. The corrosion of ceramic materials is a significant factor limiting

the design of new systems for coal gasification, coal liquification, energy

conversion, thermal storage, and the battery storage of power. (l) Advances in

the science and technology of processing, fabricating, and testing of brittle

materials have resulted in ceramic parts with mechanical properties acceptable

in a broad range of energy related applications.

Severe material degradation is common in environments where the combustion

of fuel occurs. As superalloys and complex cooling systems have been developed

the operating temperatures of turbine engines has been raised to increase their

efficiency. Many of the raw materials used in superalloys are expensive and can

be obtained from a limited number of foreign sources. Low heat rejection diesel

and gas turbine engines are in development which allow a substantial decrease in

the size of power source, primarily due to the elimination of bulky cooling

systems.(2T3) The continuing development of more efficient and compact engines

as well as the reduction of cost and dependence on strategic materials is dependent

on advances in materials technology. In addition to being inherently refractory

and resistant to corrosion, most ceramic materials are available from domestic

sources and are much less expensive than the elements used in superalloys.

Although ceramics tend to be refractory, the corrosion of ceramics occurs

at an appreciable rate in many systems. This is reflected in cost of refractoris

used by the steel and glass industries. The research resulting from the

89
90 High Temperature Corrosion of Ceramics

widespread use of ceramics as vessels to hold and direct the flow of molten metal

and glass has led to an understanding of the corrosion of ceramics in the presence

of a melt.

Alumina is an excellent bulk refractory which is used in impure and fairly

pure forms. It is also generated as a scale which is protective of coatings on

superalloys. Its corrosion resistance to gases at high temperatures, in presence

of deposits which may enhance the corrosion was studied in this research. Gaseous

corrosion of alumina was discussed in appendix A. The conditions of this study

are particularly relevant to turbine applications since the gases were generated

by SO2-02 mixtures and the deposits were Na2 S04.

In a turbine engine at 1000°C, gases have oxygen pressures of 0.2 atm or

greater and SO3 pressures of about 10e4 to lo- 5. f4) Deposits results from the

condensation of gases produced when burning fuels with sodium and sulfur impurities

in air-containing sodium. The composition of condensed Na2S04 at equilibrium

with gas mixtures will be governed by the decomposition reaction:

Na2S04 = Na20 + SO3 (11

K1 = [aNa [PSO3]
or
so42- = 02- + so3 (21

assuming that the salt is at unit activity. The Na20 and SO3 dissolved in the

melt are the basic and acidic species, respectively.

During hot corrosion the activity of Na20 and pressure of SO3 in the

deposit determine the type and extent of reaction. The composition of the

deposit in an engine can be determined by the gas phase or by the reaction of the

substrate material with components in the salt. If the activity of Na20 is high

and the pressure of SO3 is low or nil, solid oxides exhibit a basic solubilityin the

melt as in the reaction:


Appendix C-Hot Corrosion of Alumina 91

MO + 02- = M022- (3)

If the activity of Na20 is low and the pressure of SO3 is high, solid oxides exhibit

an acidic solubility in the melt as in the reaction:

MO = M2+ + 02- (4)

If the activity of Na20 (i.e. 02-) and pressure of SO3 in the deposit are within an

intermediate range then the oxide is stable and negligible solubility is observed.

More specifically, the degradation of an Al203 substrate by reaction with a

molten Na2S04 film may proceed by acidic or basic fluxing reactions:

3 Na2SO4(1) + Al203 = Al2(SO4)3 + Na20 (3)

Na2S04(1) + Al203 = s NaA102 + SO3(g) (6)

The regions where the hot corrosion of Al203 occurs by either mechanism are

shown on the stability diagrams in Figure l.@y6)

The solubilities of many oxides over a range of compositions of Na2S04

have been measured experimentally. Results for Al203 and SiO2 are shown in

Figure 2.(7-g) Plots of the solubilities of different oxides are displaced to the

left and right of each other depending on the relative acidity or basicity of those

oxides.

Alloys resistant to degradation at elevated temperatures are characterized

by the formation of continuous compact solid oxide scales with a low rate of

transport of metallic or oxidant species. In addition, the volatilization rate of

the oxides in the scale must be negligible. The extent of degradation by hot

corrosion is partially determined by the solubility of the oxide scale in the

molten salt.

Where the solubility of the oxide in the molten salt deposit is low or is not

a function of the acidity or basicity of the melt, corrosion may proceed at a

rapid rate until the liquid is saturated. When the solution reaction stops the

corrosion rate decreases.


92 High Temperature Corrosion of Ceramics

LOG
P BASIC
O2
FLUXING
-10

AIO;

-20 -10 0

LOG &03

Figure l(a). :Stability diagram showing phases of alumina that can


be stable in Na2SO4 at 700°C, and defining regions
where acidic or basic fluxing are possible. Dashed lines
are SO4 isobars (atm).
Appendix C-Hot Corrosion of Alumina 93

0
t
I/ 1
-16
IO
/
/
/
/
/ -10.6,
/ IO /
/ / 1
/ /
/ /
/
/ -s ’
BASIC / IO ’ ACIDIC
LOG / /
FLUXING / / FLUXING
/ t
P / /
/

-
02
/ /
/ /
I
AIO; /
/
A13+
-a
/

A’203 c

AG3
1
-8 -Y 0 Y

LOG Ps,
3

Figure l(b). Stability diagram showing phases of alumina that can be


stable in NaZSO4 at 1000°C, and defining regions where
acidic or basic fluxing are possible. Dashed lines are sulfur
isobars. Very high SO3 pressures are required for acidic
fluxing. Refractory metal oxides are believed to make acidic
fluxing favorable at lower SO3 pressures as indicated by the
displaced boundary (arrows).(ll)
94 High Temperature Corrosion of Ceramics

Figure 2. Solubilities of alumina in fused Na2S04 at 927OC.


Appendix C-Hot Corrosion of Alumina 95

In acidic or basic conditions sustained attack results if solution of oxide

and precipitation of the oxide in the salt away from the oxide-salt interface is by

a negative gradient of the solubility of the oxide across the salt film.(lO-lll The

gradient in the solubility of the oxide is established by the local variation in the

activity of Nat0 (i.e. 09-) across the film. The resultant solution and

precipitation of the oxide produces a porous scale. The growth of a non-

passivating oxide scale on a metallic engine component results in high corrosion

rates.

In acidic conditions sustained attack may arise when the acidic component

is derived from the gas phase.(5p12*131 Als o, acid fluxing can result when an

acidic component comes from the solid phase and reduces the Na90 activity in

the deposit. This causes the melt to become more acidic. Alloy induced acid

fluxing results from the solution of refractory metal components of superalloys

in Na$304. More specifically, the formation of tungstates, vanadates, and

molybdates reduces the oxide ion concentration in the melt.(121 In a similar

manner, the oxidation of the vanadium which is contained in many fuels can

promote acidic fluxing.

EXPERIMENTAL PROCEDURE

Materials

The materials chosen are single crystal alumina and three poiycrystalline

aluminas containing different levels of impurity and microstructures. They have

been discussed in previous reports. Their sources were given in Table I of the

first part of this report. The single crystal is 99.99% pure. The chemical

analysis of the polycrystalline aluminas is given in Table I.


96 High Temperature Corrosion of Ceramics

TABLE I

Chemical Analysis of Polycrystalline

Aluminas* (units: wt%)

High PP** Med PP Low PP

Fe203 O.Ol^ O.Ol^ 0.02

CaO 0.01,. 0.02 0.07

MgO 0.10 0.17 0.75

Ti02 O.Ol^ O.Ol^ 0.02

SiO2 0.11 0,17 1.65

K20 O.OOl^ 0.001- O.OOl^

Na20 O.OOl^ 0.02 0.09

** High PP = High Purity Polycrystalline Alumina

Med PP = Medium Purity Polycrystalline Alumina

Low PP = Low Purity Polycrystalline Alumina

Undetected, Limits of detection


Appendix C-Hot Corrosion of Alumina 97

Hot Corrosion Experiments

All specimen had areas about 1 x 1 cm which were polished down to 1 urn

diamond. After cleaning, they were usually coated with 5 mg/cm2 of Na2S04

and exposed for various times from 1 to 100 hours. Some samples were also

subjected to cyclic exposures with 45 hours exposure for cycle, and washing and

reapplication of Na2SO4 between cycles. Two temperatures 700°C and 1000°C

and two atmospheres, pure 02 and initially 1% SO2-balance oxygen at a total

pressure of 1 atm, were used in the experiments. The pure 02 tended to give

basic conditions and the SC2 containing atmospheres set up acidic environments

with pSO3 = 1.5 x 10e3 atm at 1000°C and pSO3 = 7 x 10N3 atm after passage of

the initial mixture over a platinized catalyst at 700°C. Since Na2S04 melts at

883OC, the sulfate used at 700°C was an equimolar mixture of Na2S04 and CoSO4

which was molten at that temperature.

The changes in morphology and products focused on the samples were

characterized mostly with the SEM, and its X-ray spectroscopy attachment (EDS

and WDS) for microanalysis. After exposures and observations of the samples,

they were washed and the weight changes were measured to iO.l mg. The wash

water was analysed as required using a semiquantitative microprobe method.

When sufficient products were formed, they were identified by X-ray diffraction.

These procedures were described in detail in previous reports.

RESULTS AND DISCUSSION

Wetting

The wetting angles for the 1 hour and 24 hours isothermal exposures were

measured and are given in Tables II, and III. In most cases the wetting morphology
98 High Temperature Corrosion of Ceramics

TABLE II

WETTING DATA Al203

1 & 24 Hour Isothermal Exposures

ACIDIC
1 HR 7oooc 24 HR 1000°/24

Low PP MS04 MS04 Na2S04


BM(S) 140 190
LD18O RP=50°
SD 40°

Med PP MS04 MS04 Na2S04


300 220 130
RP=48O RP 2O NC

High PP MS04 MS04 Na2SO4


8O 200 120

sxtaI MS0 MS04 Na2S04


140 220 180

KEY: MS04 (Na2,CO)S04 CWContinuous


Wetting NCNearly Continuous
Wetting PSPhase Separated
Drop Wetting Angle (Cobalt
Oxide+SaIt) BMBimodal (Two
Wetting Angles) BM(S)Angles -
f(Drop Size) LDLarger Drop
Wetting Angle SDSmaller Drop
Wetting Angle BM(H)Angles =
f(Drop Habit) RFReaction
Product Angle
Appendix C-Hot Corrosion of Alumina 99

TABLE III

WETTING DATA Al203

1 & 24 Hour Isothermal Exposures

BASIC
1 HR 1ooooc 24 HR 700°/24

Low PP Na2SO4 Na2SO4 MS04


30 BM(H1 PS
NC 40/21° =16O

Med PP Na2SO4 Na2S04 MS04


cw cw PS
=14o

High PP ;$2s04 Na2S04 MS04


40 PS
=14o

sxtsl ~o~so4 Na2S04 MS04


E” PS
200

KEY: MS04 WwWSO4


cw Continuous Wetting
NC Nearly Continuous Wetting
PS Phase Separated Drop Wetting
Angle (Cobalt Oxide+Salt)
BM Bimodal (Two Wetting Angles)
BM(S1 Angles - f(Drop Size)
LD Larger Drop Wetting Angle
SD Smaller Drop Wetting Angle
BM(H1 Angles = f(Drop Habit)
RP Reaction Product Angle
100 High Temperature Corrosion of Ceramics

which distinguished each set of exposure conditions, atmosphere, and temperature,

was nearly developed minutes after the coupon reached the furnace temperatures

except in a few eases.

The phase separation of the (Na2,Co)SO4 deposit used in the 700°C exposures

only occurred in the oxygen atmosphere as predicted by thermodynamics. After

exposure the deposits were composed of Na2S04 and masses of equiaxed cobalt

oxide crystals. The cobalt oxide formed a discontinuous layer on the substrate

surface. The cobalt oxide was preferentially wetted by the molten salt which

was prevented from wetting the alumina substrate.

In some cases two distinct wetting angles were measured on some of the

coupons. Bimodal wetting morphologies resulted from the phase separation of

the salt or from the formation a reaction product in the deposit. In other cases

the difference in contact angle was a function of drop size as reported in Tables

II and III.

The wetting of all the aluminas exposed in acidic conditions was limited.

In acidic conditions the wetting angle observed on the two higher purity aluminas

(High PP and single crystal) increased between 1 and 24 hours of exposure at

700°C. The wetting angle observed on the two lower purity aluminas (LowPP

and MedPP) decreased between 1 and 24 hours of exposure at 700°C. No trend in

the wetting with composition is apparent. In general, the wetting angle of the

salt on the aluminas in acidic conditions is lower at 1OOOoC than at 700°C. The

reaction of the more impure aluminas with the molten salt results in the formation

of multiple reaction products and the spreading of the droplets. There is less

reaction on the two higher purity substrates and the wetting decreases as the

angle of contact approaches the equilibrium wetting angle.


Appendix C-Hot Corrosion of Alumina 101

In basic conditions the wetting angle of the salt on the alumina is lower at

1000°C than at 700°C. However, the crystallization of cobalt oxide from the

salt at 700°C affected the results. At both temperatures, the wetting of the

alumina by the salt is more extensive in basic conditions than acidic conditions.

In generai, a slight decrease in the contact angle occured between 1 hour and 24

hours in basic conditions at 1000°C. In basic conditions the wetting generally

increases with the impurity content of the substrate. The lower purity aluminas

were wetted much more extensively by the molten salt than were the higher

purity aluminas. This is attributed to the fact that the wetting behavior is

affected by the reaction of impurity phases with the basic component of the

melt, sodium oxide, which is promoted by an increase in temperature.

Photographs taken during long term exposures where partial wetting of the

coupons occurred did not provide substantial evidence of the repeated wetting of

specific areas of a coupon when salt was reapplied before each cycle. This is

important since localized attack could be very damaging.

In those experiments where the wetting was poor and the resultant contact

angle was high, relatively thick droplets were observed. When the deposit is

thick the effect of the atmosphere on the composition of the salt, especially in

the vicinity of the substrate, may be significantly less than when the salt wet the

coupon in a thin continuous layer. On coupons where a low wetting angle is observed,

the area for reaction at both the oxide-salt and salt-gas interfaces is increased

and the distance over which the oxidant must be transported inward from the

salt-gas interface is decreased. Hence, the corrosion rate may be enhanced

when the contact angle is low. John observed that an increase in salt film thickness

results in a decrease in corrosion rate, for exposures of alloys with significant

solubility in the molten salt.(141


102 High Temperature Corrosion of Ceramics

Hot Corrosion of Al203

Even after long term exposure with thermal cycling and the reapplication

of the salt the total weight changes measured were very small, less than 1 mg/cm2,

even after 500 hours of cyclic exposure. The weight changes recorded for 1, 24,

405, and 495 hour experiments are listed in Tables IV,V, and VI.

A globular silica reaction product was observed on the washed polycrystalline

coupons from the exposures made in either atmosphere at 700°C. Well defined

crystalline silicate reaction products were observed on the washed polycrystalline

coupons from expures in either atmosphere at 1000°C. After exposure, the silica

or silicate reaction products were etched from some of the coupons using

concentrated hydrofluoric acid. The weight decreases are listed in Table VII.

The weight losses due to the etching of the coupons indicate that these reaction

products were present on the washed polycrystalline substrates in significant

quantities which were not unreasonably large compared to the impurity content

of the samples.

The weight changes must be affected by the smaller area of reaction when

the coupon surface was not wetted completely by the melt. In acidic conditions

at both temperatures all the materials lost weight. The formation of cobalt

oxide on the surfaces of the coupons exposed at 700°C in basic conditions increased

the weight of the samples. The weight changes recorded for any of the exposures,

particularly of the polycrystalline materials, are probably the result of two or

more contributions, possibly of opposite sign. Also, a combination of both gaseous

and hot corrosion occurred over fractions of the coupon surfaces during exposure.

Coupons of the same materials were exposed to gaseous corrosion in the same

atmospheres for 168 hours (appendix A) and no weight changes were detected

with a sensitivity of 0.1 mg.


Appendix C-Hot Corrosion of Alumina 103

TABLE IV

WT CHANGE ALUMINA [mg/cm2]


1 & 24 Hour Isothermal Exposures

BASIC 1OoOoC 1 HR 24 HR

LOWPP 0 - 0.2
MedPP 0 + 0.1
HighPP - 0.5 - 1.1
sxtal - 0.1 - 0.4

BASIC 7OOoC

LOWPP + 0.4
MedPP + 0.1
HighPP 0
sxtal + 0.1

ACIDIC 7OOoC 1HR 24 HR

LOWPP - 0.1 - 0.2


MedPP - 0.8 - 0.4
HighPP - 0.5 - 0.2
sxtal - 0.1 0

ACIDIC 1OOOoC 24HR

LOWPP - 0.3
MedPP -0.3
HighPP - 1.1
sxtel - 0.3

LowPP = Saxonsburg
MedPP = 3M
HighPP = Lucallox
sxtal = Single crystal
104 High Temperature Corrosion of Ceramics

TABLE V

WT CHANGE ALUMINA [mg/cm2]

495 HR Cyclic Exposure (45 HR Cycles + ResaIt)

ACIDIC 700°C

Material

Cycle LOWPP MedPP HighPP SXtal

-0.2 -0.3 -0.3 0


2 -0.1 -0.1 -0.1 0
3 0 0 0 0
4 0 0 0 0
5 0 -0.1 -0.1 0
6 -0.1 -0.1 0 -0.1
7 0 0 0 0
a 0 +0.1 0 0
9 -0.1 -0.1 0 0
10 -0.1 0 0 -0.1
11 0 0 0 0

-0.6 -0.6 -0.5 -0.2


Appendix C-Hot Corrosion of Alumina 105

TABLEVI

WTCHANGE ALUMINA[mg/cm2]

405 HR CyclicExposure(45
HR Cycles+ ResaIt)

BASIC 1000°C

Material

Cycle LOWPP MedPP HighPP SXtal

1 -0.2 0 0 0
2 +0.3 -0.1 0 0
3 0 +0.1 0 0
4 0 -0.1 0 0
5 0 0 0 0
6 0 0 0 0
7 0 0 0 -0.1
8 0 0 0 0
9 0 0 0 0

+0.1 -0.1 0 -0.1


106 High Temperature Corrosion of Ceramics

TABLE VII

RESULTS HP ETCH
WT CHANGE ALUMINA [mg/cm2]
Exposed Samples Etched in HP

BASIC 1000°C 24 HR 405 HR

LOWPP 0 - 0.6
MedPP - 0.7 - 0.3
HighPP 0

ACIDIC 700°C 495 HR

LOWPP - 0.3
MedPP - 0.1
HighPP - 0.1

ACIDIC 1000°C 24 HR

LOWPP 0
MedPP - 1.0
Appendix C-Hot Corrosion of Alumina 107

The data given in Table VIII were obtained from the semiquantitative

processing of the EDS spectra taken from the wash water. This analysis was

performed on 1, 10, and 100 hour exposures in acidic conditions at 700°C.

Results, Acidic Conditions

The 700°C exposures in acidic conditions were emphasized because sulfate

formation is favored at the low temperature for a given SO2 pressure.

Single Crystal Alumina

At 700°C limited reaction was detected after 1 hour of exposure.

The wetting was spotty and a wetting angle of 14O was measured. After 24 hours

of exposure the wetting angle was 22 o. The EDS spectra of many drops below

100 microns in diameter indicated that limited solution of Al had occurred.

Negligible reaction was detected on the washed substrate after 100 hours of

exposure. Aluminum ions were consistently present in the wash water at 1, 10,

100 hours. A limited amount of Si was detected after 10 and 100 hours of exposure.

Since negliglble silicon was present in the single crystal, it is assumed that

limited contamination of the samples occurred and that the levels of Si detected

for the single crystal alumina are not significant. The alumina furnace tube in

which the samples were heated may have been a source of Si. After 495 hours of

cyclic exposure, networks of shallow depressions were observed over portions of

the coupon.

The coupon exposed at 1000°C for 24 hours exhibited a faint etching at the

perimeters of drop areas. A significant amount of Al was present in some of the

EDS spectra of the salt at the drop edges. The substrate was wetted by drops up

to several millimeters across with a wetting angle of 180.


108 High Temperature Corrosion of Ceramics

TABLE VIII

Wash Water Analysis, Al203 Exposures


700°C Acidic Conditions

Material Element 1 Hour 10 Hour 100 Hour

LOWPP
Na 17.0 18.0 16.0
S 14.0 14.0 14.0
Si 0.99 0.96 1.4
Al 2.0 1.2 2.2
Mg 2.4 1.3 1.3
Ca 0.00 0.00 0.00
co 4.1 4.2 3.5
0** 60.0 60.0 61.0

MedPP
Na 18.0 19.0 16.0
S 13.0 11.0 13.0
Si 0.94 1.5 3.6
Al 2.2 1.5
Mb? 1.8 :::
Ci 0.00 0.00 I?“00
co 4.2 3.9 3.3
0 59.0 58.0 61.0

HighPP
Na 17.0 19.0 17.0
S 15.0 15.0 16.0
Si 0.46 0.00 0.40
Al 1.0 0.52 0.66
ME 1.4 1.0 0.43
Ca 0.00 0.00 0.00
co 3.0 3.9 3.9
0 62.0 61.0 62.0

sxtal
Na 19.0 18.0 16.0
S 15.0 15.0 16.0
Si 0.00 0.46 0.56
Al 0.60 1.3 0.16
I% 0.81 1.6 0.00
Ca 0.00 0.00 0.00
co 3.9 3.7 4.4
0 61.0 61.0 63.0

** by stoichiometry
Appendix C-Hot Corrosion of Alumina 109

High Purity Polycrystalline Alumina

At 700°C limited reaction was detected after 1 hour exposure. The

substrate was wetted by two large patches covering about half of the coupon

area. A wetting angle of 8Owas measured.

After 24 hours of exposure the wetting angle was approximately 20°. The

drops were patchy and typically under 1000 microns across. The solution of Al

was indicated by some of the EDS spectra of salt at the edges of drops. These

spectra also indicate the presence of significant amounts of Si. Washing the

substrate revealed the preferential solution of alumina grains and the presence

of a fine poorly defined silicate along the edges of the drop areas. A few

isolated patches of cobalt sulfate 10 to 20 microns across, containing Na and

significant amount of Si were observed.

The etching of alumina grains throughout the drop areas was evident after

100 hours of exposure. A thin discontinuous layer of silica on the edges of grains

in a drop area is shown in Figure 3 (top). Aluminum ions were consistently

present in the wash water after 1, 10, and 100 hour exposures.

After 495 hours of cyclic exposure small thin patches of silica covered

portions of the substrate and fractions of individual grains (Figure 3, bottom). In

these areas the etching and preferential solution of grains was apparent.

The coupon exposed at 1000°C for 24 hours was wetted by several discrete

drops about 2 millimeters across with a 12O wetting angle. Traces of Al and Si

were detected in EDS spectra of salt at the drop edges. Washing the substrate

revealed the presence of a thin poorly defined silica-containing layer with traces

of Na and Mg detected in some spectra. This discontinuous layer covers a large

fraction of the grains in the drop areas (Figure 4, top).


110 High Temperature Corrosion of Ceramics

Figure 3. High purity polycrstalline alumina exposed in acidic


conditions at 700oC or 100 hours {top) and 495 hours
{bottom). Etched grains in drop areas and thin patches of
silica are shown on washed substrates.
Appendix C-Hot Corrosion of Alumina 111

Figure 4. High, medium, and low purity polycrystalline aluminas


(top, middle, and bottom, respectively) exposed in
acidic conditions at lOOOoCfor 24 hours. Silicate
reaction products on washed substrates are shown.
112 High Temperature Corrosion of Ceramics

Medium Purity Polycrystalline Alumina

At 700°C after 1 hour of exposure a significant amount of Al and a

limited amount of Si were indicated by the EDS spectra of drops. Wetting was

spotty except for a single large patch of salt covering about a fifth of the coupon

area. The droplets were typically 10 to 200 microns across. A wetting angle of

30° was measured. A few isolated patches of cobalt sulfate 10 to 20 microns

across and containing Na and a significant amount of Si were observed.

After 24 hours exposure the wetting was spotty with some small patches.

Substantial solution of Al was indicated by the EDS spectra of salt in drops with

a measured wetting angle of 22O, and in droplets containing larger well defined

crystals richer in Al with a wetting angle of about 48O. Washing the substrate

revealed that fine poorly defined bands a few microns wide, containing Co, Si,

and traces of Mg and Ca, skirted large fractions of the perimeters of the large

drop areas. In the drop areas, some of the grain boundaries had been etched.

The etching of grains throughout the drop areas is evident after 100 hours

of exposure. Small globules rich in Si scattered on raised blocky areas 20 to 30

microns across were observed. Aluminum and Mg ions were consistenly present

in the wash water at 1, 10, and 100 hours in much greater concentrations than

for the single crystal or the high purity polycrystalline material. The

concentration of Mg detected after 10 hours of exposure was relatively large.

The concentration of Si ions in the wash water was also much greater, and increased

drastically between 10 and 100 hours.

After 495 hours of cyclic exposure patches of a continuous layer of well

defined globular silica covered portions of the substrate. The etching of grains

was apparent.
Appendix C-Hot Corrosion of Alumina 113

The coupon exposed at 1000°C for 24 hours was wetted nearly continuously

over some portions of the surface and by discrete droplets with a 13O wetting

angle in other areas. In both areas, substantial amounts of Ca, Mg, Si, Al, a

limited amount of Ba, and traces of K were present in EDS spectra of the salt.

Washing the substrate revealed the presence of numerous well defined sodium

aluminum silicate and sodium magnesium aluminum silicate crystals scattered

across the substrate (Figure 4, middle). Also, smaller sodium magnesium

aluminum silicate crystals had formed as platelets aligned in parcels on the

substrate. Some crystals were present in groups but in most cases they

protruded from grain boundaries where discontinuous etching had occurred.

Low Purity Polycrystalline Alumina

At 700°C after 1 hour of exposure the wetting was spotty with some

larger patches of salt over 500 microns across. Wetting angles of la0 for the

larger drops and 40° for the droplets were measured. A significant amount of Al

and a limited amount of Si were indicated by the EDS spectra of drops. Blocky

crystals about 20 microns across were visible below the salt in several droplets.

After 24 hours of exposure patches of salt nearly as large as 2000 microns

across wetted the substrate. Substantial solution of Al was indicated by the EDS

spectra of salt in drops with a measured wetting angle of 14O. Sulfate drops with

a wetting angle of about 500 contained well defined sulfate crystals rich in Al.

Some of the crystals contained Al and Mg as well as Si (Figure 5). Washing the

substrate revealed that a thin poorly defined layer as large as 100 microns wide

containing Co, Si, a significant amount of Mg, and traces of Ca skirted the

perimeters of the larger drop areas. In the drop areas a relatively even etching

between the grains had occurred. A few isolated patches of cobalt sulfate 10 to

30 microns across and containing Na and a significant amount of Si were

observed.
114 High Temperature Corrosion of Ceramics

Figure 5. Low purity polycrystalline alumina exposed in acidic


conditions at 700oC for 24 hours. Salt crystals containing Al
and Mg are shown (top and bottom).
Appendix C-Hot Corrosion of Alumina 115

The etching of grains throughout the drop areas is evident after 10 hours of

isothermal exposure. Distributions of silica globules about 1 or 2 microns across

have formed on fractions of the drop areas. After 100 hours of exposure many of

the globules coalesced. Aluminum and Mg ions were consistently present in the

wash water at 1, 10, and 100 hours in much greater concentrations than for the

single crystal or high purity polycrystalline materials. The concentration of Si

ions in the wash waters was also much greater, but the increase observed

between 10 and 100 hours was not as great as that observed for the medium

purity polycrystalline material.

After 495 hours of cyclic exposure a thin discontinuous layer of silica

covered portions of the substrate. The etching of grains was apparent.

The coupon exposed at 1000°C for 24 hours was wetted by drops up to 3000

microns across with a measured wetting angle of 19O. Spotty wetting was also

observed, particularly in the areas outside the perimeters of the larger drops. In

both areas, substantial amounts of Ca, Mg, Si, Al, and traces of Ba and K were

present in EDS spectra of the salt. Washing the substrate revealed the selective

solution of grains in the drop areas. Well defined aluminum silicate containing

Na and Ca formed in limited quantities (Figure 4, bottom). Minor amounts of Mg

were present along with the Ca in a few of the crystals.

Results, Basic Conditions

The 1OOOoC exposures in basic conditions were emphasized because more

extensive reaction occurred at the higher temperature when preliminary short

term exposures were made in oxygen. At 700°C a large fraction of the molten

salt did not remain in contact with the substrate throughout the exposure due to

the formation of cobalt oxide.


116 High Temperature Corrosion of Ceramics

Single Crystal

At 700°C phase separation of the salt occurred within 24 hours of

exposure. The wetting was patchy and irregular. A wetting angle of about 20°

was measured. The concentration of Al ions was limited to the EDS spectra of a

few small drops, where the molten salt had remained in contact with the substrate.

Cobalt oxide separated the sulfate from the substrate over most of the area of

the larger drops.

No reaction was detected after 1 hour of exposure at 1000°C. A wetting

angle of loo was measured and broad patchy wetting covered large areas of the

coupon. After 24 hours of exposure an So wetting angle was measured.

Some sodium aluminum silicate was observed after 405 hours of cyclic

testing at 1000°C. Negligible quantities of Si are present in the as-received

substrate material. The alumina furnace tube is a likely source of the Si.

High Purity Polycrystalline Alumina

At 700°C phase separation of the salt occurred within 24 hours of

exposure. The wetting was patchy and irregular. A wetting angle of about 14O

was measured. Traces of Si and limited concentrations of Al ions were indicated

by the EDS spectra of small drops where the molten salt had remained in contact

with the substrate. A network of globular silica associated with Co formed on

the substrate in the perimeter of the larger drop areas. From the morphology of

the washed substrate it appeared that preferential solution of grains and growth

of Si rich needles had occurred in the drop areas. The growth direction of the

needles was constant across the surface of each grain (Figure 6, top). A few

larger needles were observed at or near the triple points between grains and

contained significant amounts of Ca.


Appendix C-Hot Corrosion of Alumina 117

Figure 6. High purity polycrystalline alumina exposed in basic


conditions at 700oC for 24 hours (top) and 1000oC for
405 houl'S (bottom). Oriented silica rich needles (top)
and silicate reaction products at grain boundaries
(bottom) are shown on washed substrates.
118 High Temperature Corrosion of Ceramics

After 1 hour of exposure at 1000°C a distribution of well defined

hemispherical silicates formed across the substrate surface in the drop areas.

Crystals of sodium aluminum silicate between 5 and 10 microns across protruded

from beneath the salt near the centers of some of the drop. A wetting angle of

So was measured. After 24 hours of exposure a 4O wetting angle was measured.

A distribution of globular silicate had formed in the drop areas. Some

anomalously large globules contained Ca and Mg. A consistent pattern of

formation of sodium aluminum silicate along grain boundaries and sodium

magnesium aluminum silicate at the triple points between grains was apparent on

the washed substrate after 405 hours of cyclic exposure (Figure 6, bottom). The

preferential solution of the surfaces of certain grains is apparent.

Medium Purity Polycrystalline Alumina

At 700°C, phase separation of the salt occurred within 24 hours of

exposure. The wetting was patchy and irregular. A wetting angle of about 14O

was measured. The EDS spectra of small drops where the molten salt had

remained in contact with the substrate indicated that significant amounts of Ca,

Si, Al, and Mg were present in the salt. Washing the substrate revealed the

pronounced etching of the grains in drop areas accompanied by the deposition of

a poorly defined discontinuous layer of silica. A sparse network of globular silica

associated with Co formed on the substrate in the perimeter of the drop areas.

After 1 hour of exposure at 1OOOoCthe salt had formed a continuous film

over most of the coupon. Substantial amounts of Ca, Mg, Si, and Al were present

in EDS spectra of the salt. Washing the substrate revealed the presence of

numerous well defined magnesium aluminum silicate crystals scattered across

the substrate and a random distribution of vugs up to 20 microns across where


Appendix C-Hot Corrosion of Alumina 119

preferential solution of the substrate occurred. After 24 hours of exposure the

continuous wetting of the coupon resulted in a substrate surface which was

partially covered with poorly defined crystals. Some of the crystals, particularly

the one near vugs, contained substantial amounts of Si and Mg. Small patches of

a thin layer of silica were also observed on the substrate. On the unwashed

sample large peaks for Ca, Ba, Mg, Al, Si and traces of K were present in the

EDS spectra of the salt. Rosettes about 10 to 15 microns across were present in

some areas where substantial amounts of Be were present in the salt (Figure 7,

top). The pronounced etching of grains was apparent after 405 hours of cyclic

exposure. Numerous bunches of poorly defined sodium aluminum silicate crystals

were scattered over the substrate surface.

Low Purity Polycrystalline Alumina

At 700°C phase separation of the salt occurred within 24 hours of

exposure. The wetting was patchy and irregular. A wetting angle of about 16O

was measured. The EDS spectra of small drops where the molten salt had

remained in contact with the substrate indicated that significant amounts of Ca,

Al, Mg and Si were present in the salt (Figure 8). Washing the substrate revealed

a dense network of globular silica associated with Co on the substrate over

portions of the drop areas. A significant amount of Mg was present in many of

the globules suggesting that they were a magnesium silicate.

After 1 hour of exposure at 1OOOoC the salt had formed a nearly continuous

film on over half of the coupon surface. A wetting angle of 3O was measured.

Substantial amounts of Ca, Ba, Mg, Si, Al, and K were present in EDS spectra of

the salt, both in the salt film and in small groups of well defined salt crystals

rich in impurity ions. Washing the substrate revealed a fine globular silicate
120 High Temperature Corrosion of Ceramics

Figure 7. Medium purity polycrystalline alumina exposed in basic


conditions at lOOOoCfor 24 hours. A multi-phase
deposit with rosettes containing Ca, Ba, Al, Si, Mg, and
K (top) and a typical multi-phase deposit containing Ca,
Al, Si, Mg and K (bottom) are shown.
Appendix C-Hot Corrosion of Alumina 121

Figure 8. Low purity polycrystalline alumina exposed in basic


conditions at 700oC for 24 hours. Salt crystals
containing Ca, Mg, Al, and Si are shown (top and
bottom).
122 High Temperature Corrosion of Ceramics

containing a significant amount of Mg. An intermittent distribution of larger

globules up to

over 2 microns across contained Ca. A random distribution of vugs typically

below 10 microns across where preferential solution of the substrate had

occurred was also observed.

After 24 hours of exposure a bimodal wetting morphology had developed.

A wetting angle of 4O was measured on drops as large as 4000 microns across. A

thin discontinuous film of salt with a network of holes and branches was left

behind by the dewetting of the molten salt. A 21° wetting angle was measured

on the edges of the film. On the unwashed sample a substantial concentration of

Al ions and traces of Ca, Si, and Mg were indicated by the EDS spectra of the

salt. On the washed sample well defined tabular crystals of sodium magnesium

aluminum silicate and blocky sodium aluminum silicate crystals containing Ca

were observed (Figure 9, top).

The etching of poorly defined grains was apparent after 405 hours of cyclic

exposure. Blocky crystals of sodium magnesium silicate were predominantly free

of Ca and had become less well defined than after 24 hours of isothermal

exposure (Figure 9, bottom). A phase rich in Al and Mg formed on some sites on

the substrate, probably magnesium aluminate spineL

Discussion of Results

Impurities increase the attack of the polycrystalline aluminas by the

molten sulfate. This effect is enhanced by their pronounced segregation, at the

grain boundaries and at the triple points between grains in impurity second

phases. In each condition of exposure the hot corrosion of the lower purity

polycrystalline materials was dominated by the reactions of the impurity silicate

phases with the melt.


Appendix C-Hot Corrosion of Alumina 123

Figure 9. Low purity polycrystalline alumina exposed in basic


conditions at 1000oC for 24 hours (top) and 405 hours
(bottom). Aluminum silicate crystals containing Ca left
and Na right (top) and Mg and Na (bottom) are shown on
washed substrates.
124 High Temperature Corrosion of Ceramics

The data available on the solubility of alumina in Na2SO.t as a function of

the composition of the salt is not directly applicable when significant amounts of

Si are present in the melt. The formation of phases not predicted by thermodynamic

stability diagrams occurred at less than unity activity as in the previous work on

gaseous corrosion (Appendix A).

Activities of Na20 in the salt as great as 10-6 to 10e4 at temperatures

between 7OOoC and 1000°C in an oxygen atmosphere have been reported.

However, the actual activity of Na20 in the deposits during the exposures and

the profile of the activity across the thickness of the deposits is not known. In

the exposures of alumina the salt tended to form discrete droplets. Impurities

which result in more extensive wetting of the substrate increase the extent of

degradation by increasing the area of attack and promoting the formation of a

thinner molten salt film as discussed in the introduction. This is observed for the

polycrystalline aluminas, particularly under basic conditions.

At 400°C in acidic conditions all the coupons exposed were covered by

discrete droplets of salt. The data obtained from the analysis of the wash water

indicates that the atomic sodium to sulfur ratios were significantly below 2.0 for

all times of exposure (Table VIII). This is true for all four aluminas. Since no

insoluble products rich in Na were detected, it is likely that SO3 from the

atmosphere enriched the deposit as the composition of the melt shifted toward

being less acidic because aluminum sulfate and other sulfates formed in solution

in the melt.

The presence of Al in the wash waters and the weight losses recorded for

all exposures in these conditions are COnSiStent with the fOrtTIatiOn of A12(S0.+)3.

The wash water indicates that Mg was present as a soluble reaction product in

the salt, probably MgSO4, since it is stable at lower pressures of SO3 than
Appendix C-Hot Corrosion of Alumina 125

Al3(SO4)3. The well defined sulfate crystals in Figure 5, contain substantial

amounts of Al and significant amounts of Mg. An abundance of these crystals

was observed on the two least pure aluminas.

After 495 hour cyclic exposures silica was present on portions of all of the

substrates but the single crystal (Figure 10, top). The solution of Si rich impurity

phases at the grain boundaries of the two lower purity polycrystalline materials

resulted in a marked lack of connectivity of surface grains as in Figure 10 (bottom).

At 700°C in acidic conditions the principal reaction during the hot corrosion

of alumina is the formation of Al3(SO4)3, as was observed for gaseous corrosion

(Appendix A). A mechanism for the reactions which occurred during the long

term exposures of the less pure polycrystalline aluminas has been developed. It

is proposed that the acidic solution of alumina lowers the pressure of SO3 at the

substrate-salt interface below that at the salt-gas interface, resulting in a

negative solubility gradient for silica across the molten salt layer and satisfying

the Rapp-Gotto criteria for fluxing. (11) Impurity silicates are dissolved from the

grain boundary areas on the substrate. As the silicate ions are transported toward

the salt-gas interface they advance into a negative solubility gradient and the

precipitation of nearly pure SiO3 results. The coalesence, growth, and coarsening

of the SiO3 which is precipitating from the melt results in the formation of

globular silica patches on fractions of the substrate surface. Because the

precipitation of SiO3 in the melt occurs some distance away from the substrate-

salt interface the size and shape of the patches of globular silica were not dictated

by the microstructure of the substrate.

The morphology of the reaction products are shown in the schematic in

Figure 11. The solution of silicates present at the grain boundaries and as impurity

grains lowers the local Na30 activity of the melt in a manner analogous to the
126 High Temperature Corrosion of Ceramics

Figure 10. Medium purity polycrystalline alumina exposed in acidic


conditions at 700°C for 495 hours. Globular silica (top) and
etched grains (bottom) are shown on washed substrates.
Appendix C-Hot Corrosion of Alumina 127

S/G

Distance

O/S = Oxide-salt
Interface

S/G = Salt-gas
Interface

Q = Impurity Silicate Phases


R = Globular Silica Reaction Product
S = Sulfate Deposit Containing (Al, Mg, Si)

A = Alumina Grains at Substrate Surface


Solution Reaction at A:
3NagSO4 + Al303 = Al3(SO4)3 + 3Na30
B = Intergranular Attack
Solution Reaction at B:
Silicate + Nag0 = Na3O*xSiOg (in sulfate)
C = Away from Oxide-Salt Interface where Melt is More Acidic
Precipitation Reaction at C:
Na3O.xSiOg (in sulfate) = xSiO3 + Nag0 (in sulfate)

(Between grains and impurity phases at substrate


surface Reaction A + Reaction B:
3NagSOq + Al303 + xSilicate = Al3(SO4)3 + 3NagO-xSi0g)

Figure 11. Schematic diagram of the morphology of reaction


products on polycrystalline aluminas exposed in acidic
conditions at 7OOoC.
128 High Temperature Corrosion of Ceramics

way in which the solution of refractory oxides in molten Na3S04 lowers the

sodium oxide concentration of salt deposits during the alloy induced fluxing of

certain superalloys.(11~131 The formation of Al3(SO4)3 and MgSO4 occurs

concomitant to and in the vicinity of the localized solution of silicates, raising

the Nag0 activity and increasing the solubility of the silicate impurities in the

vicinity of the substrate surface. Intergranular corrosion results in a substantial

decrease in the connectivity of the grains of the substrate.

For the proposed model it is likely that the rate of reaction is controlled by

the inward transport of reactive species (i.e. SO3) from the salt-gas interface

across the thickness of the film or by the transport of soluble reaction products

away from the substrate-salt interface. This assumes that the transport of

reactant and product species proceeds more slowly than the dissolution of the

alumina grains or of impurity silicates at the substrate surface. The kinetics

could not be measured to check this statement, however, there is some evidence

that it is correct. As mentioned above, higher concentrations of Al were

detected in the drop edges than in the bulk of the salt on several of the samples

exposed in low temperature acidic conditions. The SO3 atmosphere has greater

effect on the chemistry of the salt at the drop edges because the distance for

transport of SO3 to interface or Na30 to surface is minimized.

At 1OOOoC in acidic conditions all of the coupons exposed were wetted by

discrete droplets of salt, except for the exposure of the medium purity

polycrystalline material where nearly continuous wetting was observed. Weight

losses were recorded for all the exposures in these conditions. After exposure,

well defined crystals with the form and chemistry of silicates were observed on

the washed polycrystalline substrates.


Appendix C-Hot Corrosion of Alumina 129

The silicates on the washed substrates of the polycrystalline materials

after 24 hours of exposure were shown in Figure 4. A fine poorly defined silicate

covers grains over large portions of the drop areas on the high impurity

substrate. Sodium aluminum silicate crystals containing substantial amounts of

Mg cover portions of the medium purity substrates. Sodium aluminum silicate

crystals containing substantial amounts of Ca cover portions of the low purity

substrate. In the drop areas on the washed substrates where the crystals are

present in substantial quantities, the substrates are etched more deeply than on

similar exposed at 700°C in acidic conditions. Within the etched drop areas the

intergranular corrosion which had been observed on the samples exposed at

700°C was absent (Figure 12).

The salt deposits on the two least pure materials contained significant

amounts of Al, Si, Mg, and Ca after exposure. The EDS spectra of most of the

smaller salt droplets do not contain Al or Si, so the Ca and Mg are in solution in

the salt as sulfates (Figure 13, top). Since the partial pressure of SO3 in the

atmosphere at 1000°C was less than at 7OOoC, the formation of Al2(SO4)3 would

not have been predicted by thermodynamics, but was observed for gaseous

corrosion (Appendix A). Under acidic conditions the formation of the CaS04

occurred at 1000°C but not at 700°C although it is stable at both temperatures.

This is believed to be the result of more severe degradation of impurity silicate

phases at the higher temperature. The uniform etching of the substrate, by the

formation of Al2(SO4)3, could result from local decreases in the basicity of the

melt at the oxide-salt interface in areas of copious silicate crystal growth.

The multi-phase salt deposit on the unwashed medium purity substrate

shown in Figure 13 (bottom) is characteristic of the two least pure materials.

The lighter blocky phase contains Ca, Si, Al, and Mg, and the remainder of the
130 High Temperature Corrosion of Ceramics

Figure 12. Medium purity polycrystalline alumina exposed in


acidic conditions at 1OOOoC for 24 hours. Sodium
aluminum silicate reaction products and etching of
drop area are shown on washed substrates.
Appendix C-Hot Corrosion of Alumina 131

Figure 13. Low and medium purity polycrystalline aluminas (top


and bottom, respectively) exposed in acidic conditions
at 1000°C for 24 hours. Sulfate crystals containing
Ca, Si, Al, and Mg (bottom) are shown.
132 High Temperature Corrosion of Ceramics

salt contains Al. The two phase salt droplet reaction product morphology must

be linked to the growth of aluminum silicates at surface heterogenieties in the

substrate.

As a result of the substrate microstructure unique reactions occurred

concomitantly over different areas of the substrate and produced local variations

in the composition of the melt. The formation of a network of well defined

aluminum silicate crystals occurred preferentially at the grain boundaries and

triple points of the microstructure. The results and the morphology of the

products described earlier and shown schematically in Figure 14 can be explained

as follows. As the alumina grains are dissolved by the sulfate, the local SO3

pressure decreases and the Nag0 activity increases accordingly near the alumina

grains. The increased Nap0 activity promotes the transport of Nag0 in the melt

to the grain boundaries and grains of silicate phases in the low purity aluminas,

where it reacts with these silicates, forming sodium silicates and alumina

silicates as well as calcium and magnesium sulfates. This maintains the acidity

of the sulfate melt thus promoting the dissolution of the alumina grains. As the

transport is rapid at 1000°C, the corrosion under the sulfate is quite uniform and

deeper than at ‘i’OO°C. At lower temperatures, 700°C under acidic conditions, a

similar mechanism operates but the transport is slower and the acid and basic

reactions cooperate mostly near the grain boundaries leading to strong

inter-granular attack (Figure 11).

The results of the exposures in basic conditions at 700°C are more difficult

to interpret due to the extensive formation of cobalt oxide crystals on the

substrate surface. Weight gains were recorded for the exposures.

Photomicrographs indicate that significant amounts of alumina were etched from

the substrates. The formation of nearly pure silica on the substrate did not

occur nearly as extensively as at 700°C in acidic conditions.


Appendix C-Hot Corrosion of Alumina 133

_ _----
r o 0-
,s
i
1
I
I0
I o 0 oI
I I

L-_-_-_-I

+ Transport of
Na20
-+ Transport of
so;
Q = Silicate Reaction Products
0 = Original Substrate Surface
Multi-phase Deposit
R = Bulk of Deposit
Sulfate Containing (Al, Mg, Ca, Si)
S = Other Phases
Complex Reaction Products Containing Silica

A = Alumina Grains at Substrate Surface


Solution Reaction at A:
3Na3SO4 + Al303 = Al3(SO4)3 + 3Na30
B = Intergranular Areas at Substrate Surface
Reaction at B:
Na3SO4 + Mg-, Ca-, Al- silicate =
NaSO*silicate + MgSO4 + CaSO4 + Al3604)3

Figure 14. Schematic diagram of the morphology of reaction


products on polycrystalline aluminas exposed in acidic
conditions at 1000°C.
134 High Temperature Corrosion of Ceramics

The presence of CaS04 in the salt on the two lower purity aiuminas, which

was not detected on samples exposed in acidic conditions at the same

temperature (700°C), may be due to a more severe attack of the impurity

silicate in the oxygen atmosphere. The conditions in the oxygen atmosphere are

not conducive to the type of sustained fluxing which occurred at 700°C in acidic

conditions. Significant amounts of both Si and Al were dissolved in the salt.

Because of the depletion of the basic component of the melt due to the

formation of cobalt oxide, it is likely that the initial dissolution products

generated at the aluminum oxide-salt interface were silica and aluminum

sulfate.

In basic conditions at 1000°C the most extensive wetting of all four of the

aluminas occurred. The lower purity materials were wetted more extensively

emphasizing the influence of the impurities. Vugs developed at some of the

triple points between grains on the surfaces of the two least pure substrates.

Even after 405 hours of cyclic exposure negligible reaction was detected on

the single crystal substrate. Aluminum silicates formed on the substrates of the

polycrystalline materials. They were sodium aluminum silicates at grain

boundaries and sodium magnesium aluminum silicates at triple points on the high

purity substrate. Sodium calcium aluminum silicates and sodium magnesium

aluminum silicate formed on the lower purity substrate. A variety of multi-

phase deposit morphologies formed on the two least pure materials after

exposure at IOOOoC in oxygen. These deposits consisted of mixtures of sulfate

with various aluminum silicates of sodium, potassium and the rare earth

elements contained as impurities in the substrates.


Appendix C-Hot Corrosion of Alumina 135

The formation of NaA102 as the primary reaction product in the deposits

would be predicted on the basis of solubility data published by Rapp (Figure 2).

While this may be the case for the single crystal material, the degradation of the

polycrystalline materials cannot be analyzed without consideration of the

formation of aluminum silicates which occurred on the substrates and in the

molten salt. This is obvious when one considers the multiple phase morphology

of the deposits on the two least pure materials.

Relatively small weight gains and losses were recorded for the exposures at

1000°C in basic conditions. Intergranular corrosion resulted in a decrease in the

connectivity of the surface grains of the two lower purity substrates. The

weight changes and photomicrographs indicate that the formation of alumina

silicates and the solution of alumina were not as extensive as that which

occurred in acidic conditions at the same temperature. It is proposed that the

silicate impurities shift the activities in the melt towards values intermediate

between the acidic and basic conditions promoted by the atmosphere, thus

making the corrosion behaviors at 1000°C under both atmospheres similar.

However, the greater acidicity of the melt under acidic conditions increases the

dissolution of alumina. Under basic conditions the cooperative reactions initiate

with the attack of the silicates impurities by the basic sulfate forming various

sodium alumina silicates including those of the rare earths. This raises the pSO2

of the melt providing for acidic solution of the alumina grains. This sulfate

formation is limited and the silicate formation is more extensive than under

acidic conditions, leading mostly to intergranular corrosion.


136 High Temperature Corrosion of Ceramics

REFERENCES

1. Wachtman, John B., et. al., “An Evaluation of Needs and Opportunities for
Fruitful Fundamental Research in Ceramics,” Ceramic Bulletin, Vol. 57,
No. 1 (1978), pp. 19-24.

2. Johnson, D.R., et.al., “Ceramic Technology for Advanced Heat Engines


Project,” Ceramic Bulletin, Vol. 64, No. 2 (1985), pp. 276-280.

3. T. Yonushonis, Cummins Engine Company, Private Communication.

4. Goebel, J.A. and F.S. Pettit, “Na2SO4 - Induced Accelerated Oxidation


(Hot Corrosion) of Nickel,” Metallurgical Transactions, Vol. 1 (1970), pp.
1943-1954.

5. Giggins, C.S. and F.S. Pettit, Hot Corrosion Degradation of Metals and
Alloys - A Unified Theory, Pratt and Whitney Aircraft, Report No. FR-
11545 (1979).

6. Goebel, J.A., P.S. Pettit and G.W. Goward, “Mechanisms for the Hot
Corrosion of Nickel-Base Alloys,” Metallurgical Transactions, Vol. 4 (1973),
pp. 261-278.

7. Jose, P.D., D.K. Gupta and R.A. Rapp, “Solubility of Alpha - Al203 in
Fused Na2S04 at lOOoK,” J. Electrochem. Sot., Vol. 132, No. 3 (1985), pp.
735-737.

8. Shi, D.Z. and R.A. Rapp, “The Solubility of SiO2 in Fused Na2S04 at
900°C,” J. Electrochem Sot., Vol. 133, No. 4 (1986), pp. 849-850.

9. Kim, G.M., “The Effect of Contaminants and Solubilities of SiO2 In Fused


Na2S04 at 1200°K,” Masters Thesis, University of Pittsburgh, 1982.

10. Rapp, R.A. and K.S. Goto, “The Hot Corrosion of Metals by Molten Salts,
“Symposium Fused Salts, Electrochemical Society Meeting (Pittsburgh,
1979).

11. Rapp, R.A. and K.S. Goto, Fused Salts, (J. Braunstein and J.R. Selman,
eds.), Electrochemical Society, 1978.

12. Birks, N. and G.H. Meier, Introduction to High Temperature Oxidation of


Metals (London: Edward Arnold, 1983), pp. 146-158.

13. Shores, D.A., “New Perspectives on Hot Corrosion Mechanisms,” H&$


Temperature Rapp, R.A. (ed.), (Houston: National Association
of Corrosion Eng., 1983), pp. 493-501.

14. John, R.C., “Corrosion of Metals by Liquid Na2C03,” Doctoral Thesis, Ohio
State University, 1979.
Appendix D-Hot Corrosion of Silicon Nitride
and Silicon Carbide
J.R. Blachere, D.F. Klimovich and F.S. Pettit

INTRODUCTION

Silicon nitride and silicon carbide are two ceramics materials considered

seriously for structural applications at high temperatures. They form a protective

silica scale in oxidizing atmosphere which is quite stable thermodynamically. In

highly corrosive environments such as those prevailing in incinerators or gas

turbines burning low grade fuels, SiOz may be attacked due to the pressure of

SOS, CO2 as well as oxides of metallic impurities (e.g. NagO, Na$O& It has

been shown in this program that Nap0 is the corrosive agent in the hot corrosion

of silica in contact with NagSO4 deposits (Appendix B). In particular the sodium

oxide promote devitrification and the crystalline layer formed tends to spa11

under temperature cycling. Using the results for silica as foundation, the

mechanisms and the extent of the hot corrosion of silicon nitride and silicon

carbide must be established.

EXPERIMENTAL PROCEDURE

The general procedures have been discussed in previous report&) and in

previous parts of this report. The materials are shown in Table I of the main

report. High purity materials (single crystal silicon carbide and CVD silicon

nitride) were studied in detail. Morphological studies of the corrosion were

performed on these materials and representative engineering materials usually

characterized by a significant level of impurities added during processing

particularly as sintering aids.

The two atmospheres used throughout the experiments were SOg-02

mixtures with a total pressure of 1 atmosphere. One contained 1% SO2 initially,

which generated a pressure of 1.5 x 1W3 atm of SO3 at 1000°C. The other was

pure oxygen. The gases flowed at the rate of 1 cm5/s. Usually the sodium

137
138 High Temperature Corrosion of Ceramics

sulfate was applied with a surface loading of 5 mg/cmz on polished substrates.

The surface loading as well as the temperature were varied in the studies on the

purer materials. The times of exposure varied from 1 hour to 168 hours, with

standard times of 24 and 168 hours for all materials. The usual characterization

techniques were used as described earlier. They depended strongly on the

scanning electron microscope (SEM) with microanalysis of salient features by

EDS and WDS. The thickness measurements based on X-ray spectroscopy,

developed for this part of the research(2), where used for scale of thicknesses

under 1 pm. Above 1 urn they were measured on cross sections in the SEM.

These experiments were supplemented with a number of other techniques, in

particular X-ray diffraction, weight change measurements, and many types of

surface analysis (ESCA, SIMS, ISS). The methods will be discussed as needed in

the text of this appendix. The morphology was characterized after exposure

before and after washing off the soluble materials in water. Analysis of the

wash water was performed as described earlier.

RESULTS AND DISCUSSION

The results of 24 and 168 hours exposures, in particular the product

morphologies, were described earlier for the various silicon nitrides and silicon

carbides(1~3~4). Scale thickness after 168 hours under gaseous, acidic and basic

conditions are compared in Fig. 1 and table I. For the purer materials (CVD and

single crystal) the thickness of oxide formed increased in order from gaseous

corrosion, which was essentially oxidation, acidic and then basic corrosion. This

trend is no longer clear for the more impure specimens whose behavior is

dominated by the impurities. Under basic conditions it must be emphasized that

the sulfate was not depleted on any samples which were not preoxidized before
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 139

168 hour:

sic-c Sic-Si Si3 IV,

Figure 1. Thickness of layers formed for the oxidation, acidic and basic hot
corrosion of C-side and S-side single crystal silicon carbide and
CVD silicon nitride after 168 hours at 1000°C (measured between
sulfate drops for acidic corrosion). Note that oxide thicknesses
increase as OxA&.
140 High Temperature Corrosion of Ceramics

Table I Thickness of oxide scales after 168 hours exposures (urn)

Gaseous Acidic Basic

SC SC - si 0.11 0.6
12-25

SC SC - c 0.5 1.1

CVD SC _- 1.2 (1.8)* 7.7

HP SC 0.78 0.61 -- 1.4 (24 hrs)

CVD SN 0.07** 0.3 (1.41) -6

HPSN 0.9 4.3 - 7.1 1.5 - 2.1

Sin SN 2.5 1.3 (2.1) 2.4 - 3.2

* Values in parenthesis are for spherulites under sulfate drops.


For acidic corrosion the values not parethesis were measured
between the droplets.
** 0.045 by WDS, 0.06 by ellipsometry, 0.09 SEM

SC SC -Si = single crystal Sic, silicon side

SC SC-C = single crystal Sic, carbon side

CVD SC = CVD silicon carbide

HP SC = Hot pressed silicon carbide

CVD SN = CVD silicon nitride

HP SN = Hot pressed silicon nitride

Sin SN = Sintered silicon nitride


Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 141

exposure to the corrosion. Experiments on preoxidized specimens with thick

scales (>lO um)(3) and on bulk fused silica(s) lead to the essentially complete

consumption of the sulfate on these materials as indicated by EDS of the surface

before washing and by the semiquantitative wash water analysis (table II). This

shows the fundamental tendency for the reaction

Nags04 + SiOq = Na silicate + SO3 (1)

In many cases, no sulfate could be detected in the washwater and the silicate

overall stoichiometry was about NagSiO3. This is probably indicative of a

gradient in the silicate composition as the silicates in equilibrium with

crystalline silicas are richer in silica than this average composition. Impurities

in large amounts slowed this reaction as surface phases were formed on

preoxidation (Mg silicates, yttrium silicates)(1p3).

For the purer materials not preoxidized it is clear from these results that

while reaction (1) occurs under basic hot corrosion, it is the rate of oxidation

which controls the amount of silicate formed and the transport through the

silicate and sulfate layers decreases the rate of oxidation as the scale thickens.

Therefore the scale is protective. This is indicated since the preoxidized

samples not only formed more silicates which was water soluble but they

oxidized more as indicated by greater weight gains (after washing) than the same

materials not preoxidized. Since the preoxidation scales were formed at 14OOoC,

they had crystallized and cracked extensively during cooling before application

of the salt and were not protective during the hot corosion.

Under basic conditions the sulfate wets completely the sample surface and

reacts with the scale as it forms leading to very thick silicate layers even on the

purer silicon nitride (table I). This has been documented by Mayer and Riley(G)

who reported a transient acceleration of oxidation with injection of NagCOS.


142 High Temperature Corrosion of Ceramics

Table II Wash Water Analysis From


Basic Hot Corrosion (168 Hours, 1000°C)

Material Si02 SO3 Na20


At% At% At%

SC SC 25 33 42
SC SC (PO)(l) 52 0 47

CVD SC 22 27 52
CVD SC (PO) 50 2 48

CVD SN 0 62 38
CVD SN (PO) 53 12 35

SN SN -_ -_ --
SN SN (PO) 59 13 27

SC SC = Silicon Carbide single crystal

CVD SC = CVD silicon carbide

CVD SN = CVD silicon nitride

Sin SN = Sintered silicon nitride

(1) PO = Preoxidized in 02 for 10 hours at 1400°C


Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 143

With Na2S04 the reaction is milder(3~4~71 as all sulfate was not consumed in 168

hours. Also the extent of the reaction is affected by impurities and the purer

materials formed thicker scales except possibly for the hot pressed silicon

carbide which contained alumina as a sintering aid. The alumina showed no

preference for segregation in the scale either on oxidation or hot corrosion. The

aluminum present in the scale tended to stabilize the glassy silicate phase. As

discussed in a previous report, magnesium and yttrium tend to segregate into the

scale on oxidation. They complicate also the hot corrosion conditions. MgC in

particular forms magnesium silicates and tends to promote basic conditions on

the surface under acidic environmental conditions. In general the acidic

corrosion on the impure samples formed continuous silicate layers, these layers

were quite thick for hot pressed silicon nitride and sintered silicon nitride and

were similar in morphology to those observed under basic conditions.

Early in the research it became apparent that some hot corrosion occured

under acidic conditions in addition to the expected corrosion under basic

conditions. Data on pure materials for shorter times were needed in order to

understand the fundamental mechanisms of hot corrosion under acidic conditions

and establish if it could lead to significant degradation. Therefore the single

crystal silicon carbide and the CVD silicon nitride were exposed for 1 to 24 hours

under standard acidic conditions at 1OOOoC. In other experiments the

temperature and Na2S04 loading were varied as necessary.

Under acidic conditions the sodium sulfate does not wet the silicon’nitride

and silicon carbide completely. The wetting angle was of the order of 40° for

both materials. The drop size distribution was bimodal (fig. 2). There was a

variation in the values of the wetting angles measured but it was not consistent

with drop size, exposure time or nature of the sample. However the values
144 High Temperature Corrosion of Ceramics
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 145

observed were greater than measured on fused silica under similar conditions.

The evolution of the deposit as a function of time is shown in figure 2 for single

crystal silicon carbide. The deposit which was sprayed as a layer around 120°C

breaks up as the sulfate melts and morphologies similar to those in fig. 2 were

observed after a few minutes of exposure at 1000°C. As shown in the figure the

smaller droplets tend to disappear with time, probably as a result of a coarsening

process. The large drops do not move on the surface and they are usually

surrounded after 10 to 24 hours by bands free of small droplets (fig. 2).

The kinetics of acidic corrosion for single crystal silicon carbide and CVD

silicon nitride are shown in figures 3-5 for times from 1 to 24 hours at 1000°C.

The plots of figures 3 and 4 were for thicknesses of vitreous oxide measured

between, and generally away from the sulfate droplets. All the thicknesses in

these two figures were obtained by the same method (X-ray spectroscopy(2)) in

order to minimize the influence of systematic errors in the interpretation. The

data for silicon carbide plots as good straight lines as function of the square root

of time suggesting a parabolic behavior. The kinetics for silicon nitride do not

show clearly this trend and are plotted as a function of time in figure 3a and as a

function of the square root of time in figure 3b. The lines for oxidation data

from previous workers and interpolated from previous experiments in this

research indicate that the hot corrosion, even under acidic conditions enhances

the formation of the oxide layers in all cases for the purer materials even

between the salt droplets.

The amount of sodium sulfate applied per unit area of surface was varied

from 0, 0.1 and 5 mg/cm2, the standard surface loading. While higher than that

for oxidation, the growth of the vitreous scale between the sulfate droplets is

independent of the surface loading as shown in fig. 4. The data for different
146 High Temperature Corrosion of Ceramics

1000- -0
CVDSN /
/
2 /
J; /
v) .

z / Hot Corrosion
5 /
i
b .
/

500. /
---Q
/
/

/ Oxidation

168
TIME(HR)

Figure 3. Thickness of scale formed in the acidic hot corrosion of CVD


silicon nitride at 1000°C (5mg/cm2 of NaZSOq).
(a) linear plot
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide
147

/
/
/
/
/

//

/’
b
/O
d
01

I--_-_-l I I L

I.0 5.0 9.0 13.0

VT (h OUTS ) ‘L

Fimre 3. (b) parabolic plot measured between the drops of


Figure 3(a).
148 High Temperature Corrosion of Ceramics

C-side

1.0 5.0

7rme
‘12 (/+%)

Figure 4. (a) Thickness of scale formed at 1OOOoC in the Acidic hot


corrosion of single crystal silicon carbide. Three different
loadings of NaZSO4 on the surface 5mg/cm2 (A), less than
0.1mg/cm2 (“) and 0.0mg/cm2 (0) fell on the same parabolic
plot. The data for silicon and carbon side fall on two
separate lines. (parabolic)
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 149

H(t)
-- --- -- -- -- A---
I

0 5.0

TIME “*H “*I

Figure 4. (b) Parabolic plots comparing of the data of figure 4(a) for single
crystal silicon carbide (solid lines) with the oxidation of
silicon (D + G )8 and single crystal silicon carbide (for
carbon side (“) and silicon side (A) (Harris) 22 at IOOOoC.
150 High Temperature Corrosion of Ceramics

0
SPHERULITES
UNDER DROPS

O-SC(C)
22
A- SC(Si) ____I_
O-CVDSN
/
0

0
I3
A i-.---.

l:o ’ 510
l/2 I/2
TIME (HR )

Figwe 5. Thickness of spherulites (oxides) formed under the sulfate droplets


for the C-side and Si-side of single crystal silicon carbide and
CVD silicon nitride (Acidic hot corrosion at 1000°C). The solid
line corresponds to the data of figure 4 for the acidic hot
corrosion of C-side silicon carbide.
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 151

surface loadings fall on the same lines. No salt was applied initially on the

samples with 0 mg/cm2 loading but they were exposed with the other samples,

and they picked salt apparently through vapor transport. The independence of

the data on the surface loading is consistent with a constant salt activity

maintained through the vapor phase and a fixed pSOS (1.5 x 10e3 atm) setting a

constant aNa for the experiments at 1000°C according to the reaction

Na2S04 = Na20 + SO3

Kg = aNa20.pS03/aNa$304

No sulfate could be detected on the vitreous scale by SIMS and ISS, but sodium

was found by ISS in the surface layers of the scale suggesting the formation of a

silicate layer about 10 A thick at the surface of the scale. This conclusion and

the role of sodium in the acidic hot corrosion will be discussed in detail below,

but it underlines that the role of sodium is different in acidic corrosion from that

in basic corrosion.

The rate of oxide build up under the sulfate droplets formed during acidic

corrosion on the purer silicon nitride and silicon carbides is much greater than

that outside the droplets except for C-side SIC as shown in fig. 5. The silica

crystallizes rapidly under the salt forming cristobalite spherulites. The evolution

of the average thickness of this spherulitic material at the center of the

droplets, measured after washing, follows approximate parabolic behavior which

is close to that calculated from Deal and Grove%(*) data for oxidation of silicon

at that temperature. However this is fortuitous since this part of the scale is

fine crystalline material which grows from the melt.


152 High Temperature Corrosion of Ceramics

Acidic Hot Corrosion of Silicon Nitride

The data of figure 3a, is fairly linear with an intercept with the vertical

axis at about 80 A. This suggests an initial oxide layer which is not consistent

with the present experiments. The samples were cleaned in HF prior to the hot

corrosion experiments so that an oxide layer close to 100 A is not expected

although some oxygen and water are always readsorbed, as established in the

electronic industry. The silicon carbide samples which were pretreated like the

silicon nitride do not have a significant y-intercept (fig. 4). The silicon carbide

data extrapolate essentially through the origin. The method of thickness

measurements by X-ray spectroscopy is also pushed to its limit with the very

thin oxide layers formed, but a systematic error in the measurements does not

appear responsible for this intercept. Another explanation is that oxynitride is

formed on oxidation as discussed below.

Laser ellipsometry measurements on the product layers formed on the

samples of table III were consistent with an index of refraction of 1.75 and the

thicknesses given in the table. This index of refraction is much higher than 1.46,

the one for vitreous silica. The corresponding thickness are larger than those of

figure 3 measured by X-ray spectrometry (Oku). IR measurements were not

conclusive since the spectrometer could not reach the major peak for the silicon

oxynitride. The method in general did not appear sensitive to the presence of

silicon oxynitride and was not pursued further. Tressler et al. recently reached

the same conclusion after extensive specular reflection FTIR spectroscopy@).

ESCA data is still expected. The thickness of the glassy product layers formed

on CVD silicon nitride were generally lower when measured by X-ray

spectrometry than measured on cross sections in the SEM. For the Sic single

crystal the thicknesses measured by X-ray were larger than measured in the
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 153

Table III

Scale Thickness of Selected Samples (A)

X-ray (4) Ellip (5)

A A

CVD SN (1) 450 605

CVD SN (2) 120 356

CVD SN (3) 130 326

(1) CVD silicon nitride exposed to gaseous corrosion 1% SO2-SO3


balance 02 for 168 hours at 1000°C.

(21 CVD silicon nitride exposed to acidic hot corrosion for 24 hours
at 91OOC.

(3) CVD silicon nitride exposed to acidic hot corrosion for 24 hours
at 955OC.

(4) X-ray spectrometry [2]

(51 Laser beam ellipsometry courtesy Terry O’Keefe Westinghouse


R & D.
154 High Temperature Corrosion of Ceramics

SEM(2). The of thicknesses in the SEM on very thin layers such as those formed

on the silicon nitride are difficult and subject to errors but the trend in the

thickness measurements is consistent with an oxynitride layer formed on the

silicon nitride. This is in qualitative agreement with the results obtained by

ellipsometry. Further analysis of the ellipsometry data indicated that it was

consistent with a graded film of oxynitride with possibly an overlayer of silica

50-100 A thick ( in contact with the atomsphere). Raider et al.(lO) reported

graded layers of oxynitride apparently without overlayer of silica.

The present results are in line with an initial formation of silica which

would be rapid and controlled by the surface reaction:

Si3N4 + 3 02 = 3 Si02 + 2 N2 (3)

giving very steep linear kinetics for short times in figure 3. As a result of

reaction 3, the pressure of nitrogen rises and the oxygen potential decreases at

the interface favoring the formation of oxynitride which grows between the

silicon nitride and the vitreous silica as proposed previously(ll). Diffusion

through the vitreous oxynitride is expected to be more difficult than through the

silica(12) since it is more tightly bound. The transport of oxygen through this

layered scale becomes rate controlling at steady state giving parabolic

kinetics(l3). The sequence of the three mechanisms and the distortion of the

thickness data due to the formation of oxynitride instead of the silica assumed in

the thickness calculation from X-ray spectrometry are responsible for the linear

shape of the plot of figure 3a. The scale grows more slowly after 24 hours. The

thickness after 168 hours was measured as 1 pm by X-ray spectroscopy and 1.2

urn by imaging of cross section in the SEM. However data for longer times yet

would be required to establish if the growth had reached steady state and

became parabolic after one week exposure. Some data was obtained at lower
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 155

temperatures and shorter times in an attempt to determinethe activation energy

of the linear process but this was not justified because of the complexity of the

early kinetics.

The formation of silicon oxynitride during the oxidation of silicon nitride

has been proposed by previous workers but often not demonstrated convincingly

since they usually used one technique which could not be interpreted

unambiguously. Hench used IR spectroscopy (141, others a combination of ESCA

and depth profiling and concluded that an oxynitride was formed during the

oxidation of silicon nitride. Others report only silica as the product of this

oxidation(l@. In the electronic industry silicon oxynitride is prepared readily

and used as passivating layer. Comparison of the vapor deposited oxynitride and

material formed by thermal oxidation of silicon nitride deposits on silicon

indicated that graded oxynitride layers were formed by thermal oxidation(lO).

ESCA and index of refraction measurements were used to reach these

conclusions. Tressler et a1.(gy13) performed extensive characterization of scales

formed on oxidation of silicon and silicon nitride using many different methods

such as etch back rates, SIMS, FTIR, ellipsometry and concluded that a thin layer

of silicon oxynitride formed under the silica.

The steady state kinetics for a layered scale growing by transport through

the layers are controlled by the mobility of a specie through the scale. This was

treated in detail by Yurek et al.( 17) for the diffusion of metal ions through two

layers of oxide scale to oxidize at the gas-scale interface. In the present case

oxygen must diffuse through two layers, the silica and then the oxynitride in

order to oxidize the silicon nitride. Part of the oxidation occurs at the

oxynitride-silica interface and the other at the silica-silicon nitride interface. It

is assumed that oxygen transport is controlling(g~l*) as generally accepted for


156 High Temperature Corrosion of Ceramics

dry oxidation of silicont lg). Since diffusion through the oxynitride is more

difficult than through silica, the oxynitride layer under steady state conditions

should be thinner than the oxide layer. Based on(17), it is expected that the

thickness of the layers under diffusion controlled steady state will be given by

y/x = Kp’oxy/ Kp’sil (4)

in which y, x, Kp’oxy, Kp’sil are the thicknesses and the rate constants for the

formation on silicon nitride of individual layers of oxynitride and silica,

respectively. Kp* for the total thickness y+x can be written

Kp* = (1 + Y/x)~ Kp’sil (5)

and if y is small Kp* = Kp’sil. Then the overall kinetics would be similar to those

of a silica layer forming directly on silicon nitride. This discussion assumes

distinct oxide layers. To a first approximation the formation of a graded layer

of vitreous oxynitride would not change this conclusion except that this layer

will spread over greater distance than a layer of specific composition. This is

expected from the previous kinetic arguments since the mobility of oxygen

through the oxynitride is expected to increase as the nitrogen content decreases.

The detailed shape of the nitrogen distribution would depend on the specific

concentration dependence of the diffusion coefficient through the layer. The

conclusion is consistent with the graded layer found by ellipsometry and the

large oxygen content of the scale found by X-rays. The index of refraction of

the scale appears high for the proposed graded structure of the scale, this may

be due to the fact that steady state was not achieved yet in the ellipsometry

samples. The conclusions of this analysis are in agreement with the extensive

work of Tressler et al.cg) on the oxidation of silicon nitride except that they

found a lower index of refraction to films grown under different conditions.

They find high actication energies for both the linear and parabolic constants of
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 157

a Deal and Grove analysis(*) for the oxidation of CVD silicon nitride between

1100 and 13OOOC. They conclude that the low rate of oxidation and

highactivation energies compared to silicon must be due to transport of oxygen

through the silicon oxynitride and it must be rate-controlling. The 110-120

kcal/mole for the parabolic activation energy seem to rule out the rate control

between oxygen molecular or ionic diffision through vitreous silica. While

oxygen transport, particularly molecular, through the oxynitride is expected to

be slower and with a higher activation energy than for fused silica the previous

analysis based on the formation of layered scale suggests that it cannot be rate

controlling.

Discussion of the Hot Corrosion of Silicon Nitride

The oxidation of silicon nitride is very slow. The larger oxidation rates(20)

observed in previous studies were due to impurities added for sintering as shown

in this research in the gaseous corrosion (and hot corrosion) of CVD, sintered and

hot pressed silicon nitride as shown in table I. The oxide layers formed on the

high purity silicon nitride (CVD) were thinner than observed on the high purity

single crystal or CVD silicon carbide under the same oxidation or hot corrosion

conditions. This may not establish that silicon nitride is intrinsically superior to

silicon carbide in this respect but it is certainly not inferior as stated in older

literature.

The oxidation of silicon nitride is much slower than that of silicon although

at 1000°C in both cases a vitreous oxide scale is formed and it appears oxygen

transport through the scale controls in both cases in that temperature range but

the detailed mechanisms are apparently different and not established for silicon

nitride. The general features of acidic corrosion already discussed are similar
158 High Temperature Corrosion of Ceramics

for silicon nitride and silicon carbide and in particular the growth of the scale

was increased under acidic corrosion. The presence of Na20 on the surface of

the sample, supplied as Na2S04 (or Na2C03 etc...) modifies this oxidation. The

behavior of the Na20 depends on its activity in the sulfate melt as determined

by reaction 2. From the equilibrium constant Kp, the activity of sodium oxide is

then determined by the SO3 pressure as described in previous reports. Under

acidic conditions (here at 1000°C, a pSO3 of 1.5~10~~ atm was selected) the

aNa is small and the sodium sulfate reacts little with the silica formed by

oxidation of the silicon nitride or silicon carbide as indicated by the high values

of the wetting angles on bulk silica(5) and on the scales formed on the purer

specimens as reported above. Some penetration of Na20 into the silica scale is

expected, loosening the network by the reaction

Na20 + -Si-O-Si- = 2 -Si-O- + 2 Naf (6)

Kg = [Csio12.[CNa+12/ aNa20.Csiosi

* [Csio14/aNa20

This reaction should be applicable also to the oxygen bridges of the oxynitride.

As a result of this opening of the network molecular diffusion of oxygen and

nitrogen is increased through the vitreous scales, thus generating a higher rate of

oxidation. It is likely that the sodium ion diffuses to the reaction interface

where it is reduced as proposed for silicon, it was proposed also that it would

help in the silicon-oxide structural transition at the interface(21). The sodium

may affect also the stability of the oxynitride phase. In the present experiments

sodium has been suggested in the scale with the electron microprobe and ISS, but

the evidence is not conclusive. Only very small amounts of sodium need to be

dissolved in the network according to reaction 6 in order to increase interstitial

diffusion in silica glass as discussed later for SIC. Under acidic conditions the
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 159

rate of oxidation of CVD silicon nitride is increased between the droplets

probably by the mechanism discussed above. A very thin (slOA) layer rich in

sodium was found by ISS on the scale between the droplets, it did not contain any

sulfur. This adsorbed layer fed by the droplets by surface diffusion or vapor

transport would supply the sodium into the glass.

The silicon nitride under the sulfate droplets oxidizes more rapidly than

outside and crystallization starts early (in the first hour). After 24 hours of

exposure and washing off the salt the region which were under the droplets stood

out(l). The corresponding thicknesses in Table I were measured outside the

droplets. The crystal-glass interfaces provide high diffusivity paths normal to

the surface of the sample. Some liquid is present in the intercrystalline regions

in which sodium and other impurities concentrated during the crystallization.

Thus the crystallization provides fast transport paths for oxygen leading to

thicker scale under the sulfate drops and greater penetration of the scale under

the intercrystalline regions such as the interfilamentary regions of the

spherulites shown in figure 6.

Under acidic conditions for pure materials, vitreous scales are formed

which crystallize very slowly outside the droplets. Some crystallization occurs

in between the droplets but it is sparse and the growth is extremely slow as

judged by a few very small spherulites disperse in the glass even after 168 hours

exposure. The difference in growth rates is illustrated in figure 6 where

nucleation started under the edge of a sulfate drop. The growth away from the

drops appear slower yet than in this example. As discussed in detail in appendix

A, the crystallization of vitreous silica depends on the Oxygen ion activity (or

aNa20) in the structure, and the presence of defects in the network. The sodium

oxide reaction with the network (reaction 6) provides those defects and tends to
160 High Temperature Corrosion of Ceramics

Acidic hot corrosion of CVD silicon nitride at lOOOoC. Under the


~6. sulfate drops (washed off in micrographs) the oxide crystallizes as
spherulites. After 24 hours the spherulites coarsen to globular
arrays.
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 161

occur preferentially at the interface between the salt and the sulfate where

aNa is highest. Under acidic conditions, in close proximity to the atmosphere

(outside the sulfate droplets) the gaseous potentials remain close to pSO3~1.5 x

1V3 atm and ~02~1 atm and the aNa is low. Under the sulfate droplets, the

oxygen and SO3 are more difficult to replenish as they are depleted by the

oxidation reaction at the silicon nitride-oxide or sulfate interfaces. Based on

equation (2) and

so3 = so2 + l/202 (7)

K7 = (p02)li2 .pSO2/pSO3

The activity of Na20 is raised under the sulfate droplets (it is higher than

outside) and the crystallization is promoted under the droplets.

Acidic Hot Corrosion Of Silicon Carbide

The morphologies of the samples of single crystal silicon carbide after

exposures were described previously (lp3) and for short times are generally

similar to those described above for the silicon nitrides (see fig. 2). They are

characterized by the formation of sulfate droplets separated by smooth vitreous

regions. The spherulitic growth of cristobalite occurs rapidly under the drops

and it is generally similar to the behavior observed on the silicon nitride (fig.6).

Preferential attack occurs also under regions between the cristobalite crystals as

shown in Figure 7 in which the silicon oxide has been etched away.

The kinetics of acidic hot corrosion for single crystal silicon carbide at

1OOOoCare plotted as a function of the square root of time in figure 4 and they

fall on two distinct lines both higher than the oxidation data of Harris(22)

sketched on the figure. The two lines have been associated with the carbon side

and the silicon side of the single crystal. In agreement with the oxidation results
162 High Temperature Corrosion of Ceramics

of Harris and Tressler et al.(g) the carbon-side is the fast or thick side and the

silicon-side is the thin or slow side. Hot corrosion data was obtained also at

955OC and 910 oC. The data at 910°C are plotted in figures 8 and 9 and both

sides still show parabolic behavior at that temperature. Activation energies for

the growth of oxide scale in the temperature range 910-1000°C were determined

from the parabolic constants B* calculated from x2 = Bt in which x is the scale

thickness at time t. The thickness measurements were all performed by one

method (X-ray spectroscopy) in order to minimize the influence of systematic

errors. The plots of figures 10 and 11 give activation energies of 34 and 118

kcai/mole for the carbon-side and the silicon-side, respectively.

Silicon carbide is a polar crystal and has different basal surfaces (0001) and

(0001)(23) which can be differentiated by a number of methods based on different

surface morphologies after high temperature wet oxidation or attack by fused

salts(24). The present hot corrosion experiments resulted in milder attack but it

was clear after extensive experience that slight differences in morphologies

existed between the two surfaces which allowed their identification. Namely

after hot corrosion for long times, the silicon side between the droplets appeared

as-polished except for some interference colors on the thicker scales and the

carbon-side was rougher, duller in texture. Different oxidation behaviors have

been reported(g*22) for the two sides. This difference in behavior was no longer

apparent at high temperatures (1400-15OOoC). Similar basic hot corrosion

behaviors were found for the two sides in this research at 10000C(1~3~4(. Under

acidic hot corrosion, the vitreous scale between the droplets grows faster on the

*B was determined also by the Deal Grove formulation but it gave the same
results since the data plotted as functions of t1/2 falls on good straight lines
extrapolating through the origin (fig. 3,7,8).
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 163

~7. Acidic Hot CorroSion 0( silicon carbide single crystal (lOOOOC).


The substrate was attacked under the spherulites as shown after
disSOlution 0( oxide in HF.
164 High Temperature Corrosion of Ceramics

31OC KINfllCS - C
5000

4000

3000
3
x
I?
5
0
5 2000

’ 000

0
0.00 2.00 4.00 6.00 a.00
SQRT(TIME(HR))

Figure 8. Kinetics of acidic hot corrosion of C-side single crystal silicon


carbide at 910°C (parabolic plot). Note reproducibility of data.
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide ‘I65

91OC KINETICS - SI
600

500

4oc

z
x
Y 300
$
I!
F

200

100

0
0 2.00 4.00 6.00 8.00
SQRT(TIME(HR))

Figure 9. Kinetics of acidic hot corrosion of Si-side single crystal silicon


carbide at 91OoC (parabolic plot).
166 High Temperature Corrosion of Ceramics

ACTIVATION ENERGY - C
1400

i 3.80

13.60

13.40

1320

13.00

i 2.80

12.60 1

0. 78 0.80 0.82 0.84


1000/T

Figure 10. Arrhenius plot for the acidic hot corrosion of C-side single crystal
silicon carbide.
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 167

ACTIVATION ENERGY - St
13.0

12.0

1 1 .O

10.0

8.0
3.78 0.80 0.82 0.84
1000/-T

Figure 1. Arrhenius plot for the acidic hot corrosion Si-side single crystal
silicon carbide.
168 High Temperature Corrosion of Ceramics

carbon side than on the silicon side as reported by Harris for oxidation. However

on figure 4 both sides show well defined parabolic behaviors while in Harris’

results the carbon-side was parabolic and the silicon side had linear kinetics over

the broad time scale of the experiments. For the hot corrosion the apparent

activation energy of the slow side (Si) is much larger than that of the fast

carbon-side.

Model for the Oxidation and Hot Corrosion of Silicon Carbide under Acidic
Conditions

In order to discuss the hot corrosion of silicon carbide under acidic

conditions in between the sulfate drops, it is important to understand its

oxidation. The formation of silica scales on silicon and on silica formers is

dependent on the transport of oxygen to the substrate for the oxidation to occur.

The molecular diffusion of oxygen through the scale might be rate controlling as

established for silicon dry oxidation. It has been proposed(g) that as the

temperature is increased contributions are made by the network oxygen

diffusion. Others have argued that the diffusion of CO produced in the oxidation

of silicon carbide is rate controlling and since the size of the two molecules is

about the same it is difficult to distinguish(z5). Considering recent

experiments(g) it is concluded that transport of oxygen is controlling the

oxidation of silicon carbide at low temperatures (below c 14OOoC). In earlier

work it was concluded that the diffusion of C-products was controlling at higher

temperatures(l*). They found two regimes in the oxidation of C-side silicon

carbide from 1200-1500°C, one with low activation energy (123kJ/mole) at low

temperatures and the other with high activation energy at high temperatures

(216kUmole) at 1 atm pressure of oxygen. However the oxidation of silicon-side

silicon carbide had only one regime associated with a high activation energy
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 169

(240kJ/mole in 1 atm of oxygen). The kinetics of the carbon-side oxidation

controlled by oxygen transport were modelled satisfactorily assuming that

several processes contribute to this transport; vacancy diffusion in the network

and molecular interstitial diffusion were considered specifically in the

analysis(g). The molecular diffusion dominates at low temperatures 1200-1350°C

while the vacancy diffusion dominates at high temperatures 1350-1500°C. In the

temperature range of the present experiments the oxidation of the carbon side is

similar to that of silicon, except slower with the same activation energy and

linear dependence of B on the oxygen pressure (gl. The molecular diffusion of

oxygen is dominating the kinetics. At low oxygen pressure a vacancy mechanism

appears to control the oxygen transport through the scale. The silicon side is

slower and has a high activation energy for the complete temperature range of

1200-1500°C(gl. Harris reported linear behavior at low temperatures. These

results must be considered in terms of the defect structure of the silica scale

and the surface reactions, then mechanisms will be proposed for the oxidation

and then for the hot corrosion.

Proposed Model

The following model is concerned with the lower temperature range

around 1000°C. It addresses:

il the different rates of oxidation and hot corrosion for Si-side


and C-side silicon carbide

ii) the enhancement of oxidation under acidic hot corrosion (the


role of Na20)

iii) the activation energies for hot corrosion

iv) the magnitude of B, the parabolic constant, relative to that


for silicon oxidation
170 High Temperature Corrosion of Ceramics

It is based on:

0 the variation in stoichiometry of vitreous silica with its


formation under different ~02 at the reaction interfaces.
This variation in stoichiometry and the associated defect
structure modify the transport of oxygen through the scale

ii) the oxidation results of others summarized above

iii) the oxygen transport model of Tressler and Speartg) applied


to lower temperatures

iv) the surface analysis results of Muehloff et a1.(26)

v) our acidic hot corrosion results.

Defect Structure and Stoichiometry of Silica

Vitreous silica is not often considered non-stoichiometric in the glass

literature although glassy structures which are great solvents allow greater

variation in composition than the corresponding crystals. It is known that the

silica formed under oxygen deficient environments crystallizes more slowly than

similar silica formed under more oxidizing conditions(27). Silicon rich silica has

been prepared in the electronic industry. Fratello et a1.(28) discussed the

influence of OH impurities on the crystallization of silica on the basis of the

defects and stoichiometry of vitreous silica. While the defect structure of

vitreous materials is relatively controversial for strpctural defects and mass

transport, the electronic defects have been studied in great detai1(2gy30). In

general it is accepted that the SiOg glass structure is built with SiO4 tetrahedra

connected through oxygen bridges to form a continuous random network. Thus

structural defects may extend to any kind of orderin$31). The major defects in

silica have been reviewed by Motttao); they are dangling bonds in particular the

well known single bonded oxygen, but also 3 bonded silicon. They include also

oxygen vacancies in oxygen bridges as well as Si-Si bonds. The single bonded
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 171

oxygens tend to form in silicate glasses to accommodate extra oxygens

associated with network modifying ions in the structure. It is reasonable to

expect that oxygen vacancies and Si-Si bonds are favored under oxygen deficient

conditions. It is anticipated also that compensating defects can be generated

thermally. Reaction 8 generates 2 single bonded oxygens and an oxygen vacancy

2 Si-0-Si = Si- -Si + ZSi-0


nil = Vo + 2SiO (8)

K8 = [Vo] [SiO12 = Cv [SiO12

in which [V,] = C, is the concentration of oxygen vacancies Si- -Si in the

network. Under low oxygen pressure oxygen vacancies could form by the

equation so familiar for crystalline oxides

00 = Vo + I/2 02 + 2e

K = [e12 [Vo] [~02]~/~ = 4Cv3 [~02]l/~

where the vacancies are formed at the oxide-gas interface and the oxygen is

supplied directly by the atmosphere and the electron concentration [e] is

assumed to be supplied only by the reaction so that [e] = 2 Cv. It is more

appropriate here to consider that the oxygen molecules are dissolved into the

glass at the scale-gas interface with the concentration of dissolved oxygen

proportional to the oxygen pressure through Henry’s law. Reaction 9 occurs

inside the scale under pO2 lower than at the interface and the molecular oxygen

is that dissolved in that environment (concentration Cm) so that

Si-0-Si = Vo + l/2 02 dis + 2 Si- (9)

Kg = 4Cv3 Cm112

in which the electronic charges are associated with silicon dangling bonds and Kg

includes the dissolution of 02 gas into the glass. The formation of electrons in

equation 9 may have appeared unrealistic since silica is an insulator. However,

they exist commonly in vitreous silica in trapped form such as dangling bonds on
172 High Temperature Corrosion of Ceramics

silicon shown in equation 9. It has been shown that thermally grown silica

contains positive and negative charge centers. Silica films support electronic

conduction during anodization(32).

The vitreous structure is built up on 5 and 6 member rings which are

randomly distributed on 3 dimensions. These rings provide tortuous paths for the

diffusion of gas molecules and impurity network-modifying ions smaller than the

size of the windows in the structure, a little over 3A(33). In the growth of a

scale the growth flux may introduce anisotropy to this structure by the

formation of more aligned channel in the flux direction(34). Larger channels less

than 50A in diameter have been postulated in order to explain silicon oxidation

and reported from observations in the TEM(35).

Oxidation of Silicon Carbide

The oxidation of silicon carbide at the reaction interface generates

very low oxygen pressures which cannot be calculated exactly except in presence

of free carbon, for the ternary equilibrium Sic-SiO2-C. The results of this

calculation which gives pO2 of the order of 10m30 atm at 1000°C can be

considered a lower bound for the oxidation of materials not containing excess

carbon. If active oxidation does not occur the silica formed at the reaction

interface is expected to be oxygen deficient with the formation of corresponding

defects: vacancies, 3-bonded silicons and possibly Si pairs. Thermal equilibrium

of the type of equation 8 will suppress the concentration of single bonded

oxygens

CJoxid = mNB/2x (10)

in which CJoxid is the sum of the oxidant fluxes which contribute to the

oxidation and N, B, x are the number of oxygen molecules incorporated in 1 cm3


Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 173

of scale, the parabolic constant and the scale thickness, respectively. The

constant m is determined by the mass balance between oxidants and products(g).

Two types of contribution are made to the transport of oxygen, molecular in

which molecules of oxygens are transported through channels, in the vitreous

silica and network contribution which can occur through the various network

defects. Assuming that the major mode of network transport is by a vacancy

mechanism equation 10 becomes

DvCv + DmCm = mNb/2 (11)

in which Dv, Dm, Cv and Cm are respectively the vacancy and molecular

diffusivities and concentrations, respectively. Therefore the kinetics of

oxidation of silica formers may depend on the relative contribution of the two

transport paths which are determined by the stoichiometry of the scale at a

given temperature. Different oxygen pressures on formation of the scale will

result in different stoichiometries through equations 8 and 9. Under low oxygen

pressure the quantity of dissolved oxygen (Cm) will be depressed and on the basis

of equation 9 the concentration of oxygen vacancies (Cv) will be increased

tending to favor transport by a vacancy mechanism over transport by molecular

diffusion. This in line with the proposal of a gradual change from molecular

mass transport at high pressures to ionic mass transport at low pressures for the

thermal oxidation of silicon(g). Based on equations 8,9 and 11 one can explain

the oxidation and hot corrosion of silicon carbide in the low temperature range in

which oxygen transport is assumed to be rate controlling.

First let us consider the oxidation of silicon in 1 atm of dry oxygen,

molecular diffusion is dominating the transport@) and equation 11 becomes

simply DmCm = NB/2. The parabolic constant B is proportional to pO2 and it has

an activation energy Q = 28 kcal/mole which is similar to that measured for the


174 High Temperature Corrosion of Ceramics

permeability of molecular oxygen through bulk vitreous silica(331. As discussed

later, if one assumes that the silica scale formed on C-side silicon carbide is

slightly oxygen deficient because it was formed under a low pO2 (Cm decreased),

then by equation 9, Cv increased. Although it is smaller than for the oxidation

of silicon, DmCm is still greater than DvCv. Molecular transport of oxygen

dominates but from equation 11, B is smaller than for silicon. This was observed

by Tressler et al.tgl, this oxidation has activation energy (27kcaVmole) and an

oxygen pressure dependence similar to that measured for silicon oxidation. On

the other hand if the scale on the silicon-side of silicon carbide was formed

under very low oxygen pressure, by the same reasoning Cm is greatly depressed

and Cv increased since Cv * Kg/(Cm) li6. So that now DvCv>>DmCm and the

network diffusion controls the oxidation resulting in a much lower B, a higher

activation energy (55-60 kcal/mole) and a low pO2 dependence(gl. Therefore

with the proposed model it is possible to derive a consistent picture for the

oxidation results obtained by others for silicon and silicon carbide. It postulates

different conditions of formation of the scales on silicon-side and carbon-side of

silicon carbide which are discussed below.

As discussed earlier the two sides of single crystal silicon carbide oxidize

at different rates for long periods of time at lower temperature below 1300° C.

This indicates that under those conditions the two sides maintain their specific

character as they oxidize. The two sides maintain their separate identity after

grinding and polishing as shown in the present experiments and after ion

bombardment (sputtering) to mix the surface. It was shown also by Muehloff et

a1.(261 that the two surfaces had different stabilities under low pressure

oxidation. Under ultra high vacuum the carbon side was observed to cover with

carbon above 900 K while the Si-side did not start to graphitize until 13000K. At
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 175

that temperature the graphitization of the carbon-side was extensive. The two

sides of silicon carbide maintained qualitatively their relative rates of oxidation

during the early oxidation for thicknesses of silica up to 20A, even after surface

mixing due to sputtering. This carbonization of the of the carbon side was

explained the evolution of silicon under the very low oxygen pressure at the site

of the reaction

Sic + l/2 02 = CO + Si (12)

Under 1 atm of oxygen, it is assumed that this difference in stability of the two

sides is maintained and that silicon still tends to volatilize from the C-side,

however both carbon and oxygen are oxidized in the reaction

SIC + 02 = SiO + CO (13)

In the pO2 range (s~O-~O atm) which is expected at the reacting interface SiO is

the prominant vapor specie in the Si-0 system(36). As shown in figure 12, it is

possible by this reaction to always maintain the order of the Si and carbon atoms

at the interface so that on the C-side carbon is always at the surface of the

silicon carbide as it oxidizes thus maintaining the C-side behavior. The SiO is

oxidized as it meets with higher oxygen pressure near the interface but not at

the interface and the silica forms as a rough porous scale and a rough interface

results from the active oxidation reaction. As the oxidation proceeds the SiO

oxidizes in the pores so that the scale porosity is filled. The scale is formed

under higher pO2 than is found at the interface and it will be more

stoichiometric.

On the Si-side, the silicons are at the surface and it is proposed that

oxygen is adsorbed first on the silicon (figure 13) and the carbon is oxidized by

penetration of oxygen through this oxide layer always maintaining silicon

between the oxide and the SIC in line with the Si character of that side. This in
176 High Temperature Corrosion of Ceramics

1) SiC

Figure 12. Model for oxidation of C-side silicon carbide. A rough interface
is formed and the C-side is maintained at the interface during
oxidation. The SO2 is formed away from the interface.
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 177

Sic

Sic

- Atmosphere
3) I02 I
01
Sib

sic

Figure 13. Model for oxidation of Si-side silicon carbide. The silicon side
nature of the interface is maintained.
178 High Temperature Corrosion of Ceramics

line with the results of Harris (linear kinetics of oxidation) as the CO would form

under the silica layer and desorption of CO from that position could control the

oxidation. At higher temperatures or in presence of sodium this desorption is no

longer rate controlling and parabolic kinetics are observed. The scale is formed

under very low oxygen pressure and it is expected to be very oxygen deficient as

discussed earlier. The scale formed under these conditions would be more

oxygen deficient than that formed on the C-side. In the oxidation of SiC the

network transport and the molecular transport add to supply the oxygen. This

interplay between the two types of transport processes is underlined as the

temperature of oxidation is increased and their relative contributions change for

the C-side and the Si-side. As the temperature is increased, transport by the

vacancy and other network mechanisms with high activation energies increases

rapidly and becomes a major contributor to the oxidation of the carbon side. The

oxidation of the silicon-side has been increasing also since it was dominated by

the network transport processes. As the temperature is increased the mobility

increases also in the SiC interfaces and they can no longer be differentiated by

the mechanisms of fig. 12 and 13 and around 1350oC the two sides oxidize at

about the same rate with the same activation energy as reported by Tressler(9).

Therefore the proposed model is qualitatively consistent with the data on the

oxidation of silicon and silicon carbide, now we shall apply it to our acidic hot

corrosion results.

Acidic Hot Corrosion

As discussed earlier acidic hot corrosion is oxidation enhanced by the

presence Na2Q at a low activity. Na2Q enters into the silica structure by

reaction 6 and generates single bonded oxygens thus modifying the defect
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 179

concentrations in the vitreous silica. Single bonded oxygens are oxygen excess

defects which lower the concentration of oxygen deficient defects through

reactions of the type

2 Si-0-Si = Si- -Si + 2 Si-O- (14)


K14 = [Vo] [Si-O12 = Cv [Si-012

Therefore as Na20 is introduced into the scale, it tends to increase the single

bonded oxygen concentration (eq. 6), this decreases the vacancy concentration

(eq. 14) which inturn increases the molecular oxygen concentration in the glass

(eq. S).This can be written as a single equation which is the algebraic sum of

equations 6, 14 and 9.

Na20 = 2 Na+ + l/2 02 gas + 2e- (15)

Kl6 = [Na+][02]1/2 [e12/aNa20

showing the Cm is proportional to aNa20*. Therefore in oxygen deficient scales

such as those formed on silicon carbide the sodium will decrease network

transport and increase molecular transport as it drives the scales towards more

stoichiometric compositions.

This result and the previous oxidation model will now be applied to the

acidic corrosion of C-side silicon carbide. According to the previous discussion,

the oxidation under 1 atmosphere dry oxygen at low temperatures (1OOOoC) is

controlled by molecular diffusion as DmCm >> DvCv. As shown in the previous

paragraph, in the hot corrosion Na20 increases Cm and decreases Cv, therefore

molecular transport is enhanced giving greater parabolic constant for the acidic

corrosion of C-side silicon carbide compared to that for oxidation under

*Equation 15 may be considered as the dissolution of sodium oxide in the glass in


an interstitial position (network modifier). Interpretation beyond the
relationship between aNa and [02dis] could be misleading since defects in the
glass structure are required to accommodate the Na+ ions and the electrons.
Therefore equations 6, 9 and 14 appear more representative of the reactions
occuring in the glass.
180 High Temperature Corrosion of Ceramics

corresponding conditions. In our experiments the parabolic constant B was

increased by a factor of about 4 at 1000°C. It was about 3x lo-l4 cm2js which

is similar to B=3.2xlO-I4 cm2js of Deal and Grove for the oxidation of

silicon(*). This result is in line with the scales of silicon carbide being more

stoichiometric under acidic corrosion than those formed in dry oxygen oxidation

and with lower oxygen pressures expected in the formation of scales by oxidation

of silicon carbidethan in the oxidation of silicon. From the previous analysis the

activation energy for acidic hot corrosion is expected to be that of the

permation of molecular oxygen through vitreous silica, the same as for the

oxidation of silicon carbide and silicon. Therefore it should be 26-28 kcal/mole.

The 34 kcal/mole measured is close but significantly higher because of the

nature of the experiment. The parabolic constants plotted in figures 11 and 12

were determined at constant sulfur content in the atmosphere and not constant

Na20 activity. The initial mixture of 1% S02-balance oxygen at one atmosphere

total pressure was used at all three temperatures. After passage over the

catalyst it generated decreasing pSO3 for increasing temperatures (equation 7).

From equilibrium relation 2 the activity of Na20 increases with decreasing pSO3

and therefore with increasing temperature. Since B increases with aNa20, an

apparent activation energy greater than the 26-28 kcal/mole for molecular

diffusion of oxygen and the oxidation of silicon and C-side SIC will be measured.

When the generation of 02 molecules by reaction 15 will be dominating the

activation energy for diffusion will be increased since B will be proportional to

DmCm and both terms will depend on temperature. It can be seen from reaction

15 that Cm o (K aNa20P in which both K and aNa have exponential

temperature dependences and n is a fractional exponent such as 2/9 obtained

from the defect equations considered in this report. However the enthalpy of

reaction 15 and in general the operating defect equations are not known so that
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 181

it is doubtful that an exact temperature dependence can be derived. It is clear

that the activation energy for hot corrosion should be significantly higher than

that for oxidation. The difference between the measured apparent activation

energy (34 kcal/mole) and that for molecular diffusion (27 kcal/mole) may not be

representative because the activity of sodium may be too low to completely

dominate the transport. Simple considerations suggest that the two mechanisms

of generation of the diffusing oxygen molecules were of the same order of

magnitude in the present experimental conditions. Experiments with higher

Na30 activities produced by initial gas mixtures such as 0.1% SOZ-balance

oxygen would answer this question.

For the acidic hot corrosion of Si-side silicon carbide, the same general

reasoning is applicable, however the parabolic oxidation was expected to be

dominated by network transport (vacancy mechanism) with a low parabolic

constant B. As for the C-side, the Na30 introduced by the hot corrosion

decreases the network contribution (lowers Cv) and increases the molecular

contribution (raises Cm) according to equations 6, 9, and 14 or 15. In the

reactions considered Cm increases rapidly as Cv decreases since Cm = k/Cv6 in

equation 9 and Dm >> Dv by about 4 orders of magnitude. Therefore the change

in defect concentration affect dramatically the relative contributions to the

transport of oxygen although it is not possible to state if and at what

temperature the molecular flux became dominant. B is increased dramatically

as observed experimentally. It is increased by a factor of about 16 compared to

B extrapolated from the oxidation data of Tressler et al.tgl. The large apparent

activation energy is due to the same effect as for the carbon-side, the Na30

activity increases with temperature as discussed above. A quantitative

correlation between the results for C-side and Si-side acidic hot corrosion is

being considered.
182 High Temperature Corrosion of Ceramics

REFERENCES

1. J.R. Blachere and P.S. Pettit “High Temperature Corrosion of Ceramics”


(a) DOE Report ER45117-2, March 1986
(b) DOE Report ER45117-1, June 1985
(c) DOE Report ER10915-4, June 1984

2. J.R. Blachere and D.F. Klimovich, *J. Am. Ceram. Sot. 70 [ll] C324-C326
(1987) paper appended in Appendix E.

3. B.S. Draskovich, MA thesis, University of Pittsburgh, 1985.

4, D.F. Klimovich, M.S. thesis, University of Pittsburgh, in preparation.

5. M.G. Lawson, M.S. thesis, University of Pittsburgh, 1987, (Appendix B).

6. M.I. Mayer and F.L. Riley, J. Mat. Sci., 13, (1978) p. 1319-1328.

7. (a) N.S. Jacobson, J. Am. Ceram. Sot., e, [l] 74-82 (1986).


(b) N.S. Jacobson and J.L. Smailek, J. Am. Ceram. Sot., a [8] (1985), p_
432-39.

8. B.E. Deal and A.S. Grove, J. Appl. Phys., 36 (1965), p. 3770.

9. R.E. Tressler and K.E. Spear, GRI Report GRI-87-0088, 1987.

10. S.I. Raider et. al., J. Electrochem. Sot., 123 [4] (1976), p. 560.

11. S.C. Singhal, Ceramurgia International, 2 [3] 123-130 (1976).

12. R.H. Doremus, Glass Science, Wiley 1973, p. 121.

13. R.E. Tressler and K.E. Spear, GRI Report, GRI-86/0066, 1985.

14. L.L. Hench et. al., Ceram. Eng. and Sci. Proc., 3 [9] (1982), p. 587-595.

16. I. Franz and W. Laugheinrich

17. G.J. Yurek et. al, Oxid. Metals, 4 265 (1974).

18. J.A. Costello and R.E. Tressler, J. Am. Ceram. Sot., 64, (1981) p. 327-331 .

19. F.P. Fehlner, Low Temperature Oxidation, Wiley 1986, p. 211.

20. D.C. Larsen et. al, Ceramic Materials for Advanced Heat Engines, Noyes
1985, p. 234.

91
-*. Reference 19
Appendix D-Hot Corrosion of Silicon Nitride and Silicon Carbide 183

22. (a) R.C.A. Harris and R.L. Call, in Silicon Carbide 1973, R.C. Marshall
et. al., eds., University of South Carolina Press, 1974, p. 329.
(b) R.C.A. Harris, J. Amer. Ceram. Sot., 68, 1975, p. 7-9.

23. J.W. Faust Jr., in Silicon Carbide, A High Temperature Semiconductor,


J.R. O’Connor and J. Smiltens, (Pergamon, 1960).

24. Ibid, pv 403.

25. D.M. Mieskowski et. al., J. Am. Ceram. sot., 67, (1984), C17-Cl8

26. (a) L. Muehloff et.al., J. Appl. Phys., fl [7] (1986), p. 2558-2563.


(b) Ibid, fl [8] (1986), p. 2842-2853.

27. H. Rawson, Inorganic Glass Forming Systems, Academic Press (1967), p. 53.

26.. V.J. Fratello et. al., J. Appl. Phys., 51 [12] (1980), p. 6160-6164.

29. D.L. Griscom, MRS Bulletin, June 16/ August 15, 1987.

30. N.F. Mott, Adv. Phys., 3, 1977, p. 363.

31. Reference 19 p. 67

32. Reference 19 p. 233

33. Reference 12 p. 133

34. A.G. Revesz and H.A. Schaeffer, J. Electrochem. Sot., 129, 1982, p0 357.

35. Reference 19 p. 230

36. U.S. Department of Energy, Thermochemical Stability Diagrams for


Condensed Phases and Volatility Diagrams - DOE/FE/13547-01, May 1980.
Appendix E-Publications

Many presentations have been made on this research. The most recent ones

are:

- Hot Corrosion of Non Oxide Ceramics [l]

- High Temperature Corrosion of Silicon Nitride and Silicon Carbide [2]

The research is being written up for publication. These publications are:

- Oxide Thickness Measurements in the Electron Probe Microanalyzer [3]

* Hot Corrosion of Silica (Appendix B) [4]

- Hot Corrosion of Alumina (Appendix C) [5]

- Hot Corrosion of Silicon Nitride and Silicon Carbide [6]

- Hot Corrosion of High Purity Silicon Carbide and Silicon Nitride [7]

- Corrosion of Ceramics [8]

The first paper is appended. Appendix B is ready for submission. The others

need further work. The material in Appendix C will be submitted next. Reprints will

be supplied as soon as they become available.

184
Appendix E-Publications 185

REFERENCES

1. J.R. Blachere, D.F. Klimovich and F.S. Pettit, “Hot Corrosion of Non Oxide
Ceramics,” paper # , Fail meeting, Basic Science Div. Am. Ceram. Sot.,
Nov. 5, 1966.

2. J.R. Blachere, D.F. Klimovich and F.S. Pettit, “High Temperature Corrosion
of Silicon Nitride and Silicon Carbide,” Invited paper, Workshop on Corrosion
of Ceramics, Penn State, Nov. 12-13, 1967.

3. J.R. Blachere and D.F. Klimovich, “Oxide Thickness Measurement in the


Electron Probe Microanalyzer,” J. Am. Ceram., 70 [ll], C324-C326 (1967).

4. M.G. Lawson et al., “Hot Corrosion of Silica,” ready for submission to J.


Am. Ceram. Sot.

5. M.G. Lawson et al., “Hot Corrosion of Alumina,” in preparation for J. Am.


Ceram Sot.

6. B.S. Draskovich et al., “Hot Corrosion of Silicon Nitride and Silicon Carbide,”
in preparation.

7. D.F. Klimovich et al., “Hot Corrosion of High Purity of Silicpn Nitride and
Silicon Carbide,” in preparation.

a. F.S. Pettit and J.R. Blachere, “Corrosion of Ceramics,” in preparation.


186 High Temperature Corrosion of Ceramics

,” wh,ch the/(X) funcoons prwdc the cor-


reel,““, for dx rbaorpo”” of Ihe chrrac-

Oxide Thickness Measurement in the Electron ,cr,s,,c


delccwr.
X-ray, along
,, 15 dcfmcd
[he” prthr
ar I,,/,,)
1” ,hc
csc + I”

Probe Microanalyzer wh,ch s/p I, the mar, abwrpr,“”


event for the hnc measured I; the titm. p
cocffi-

J. K. BLACHERE* AND D. E KL~MOV~CH* is the denrny “f the tilm. and $ IS the


L3ke”ff angle.
cepamle”~ “f Mawnrtr Sacncr and E.ng,nrennp. V”lw,#ty “f P,,,,burgh. P,,t,burph.
The ex~en~~vc. but slmplc. ealcuta-
Pcnniytunlr ,526,
,mns are dr,cr,bcd ,n dclrd ,n Ref, ?
and 3. They rcpon ~“4 rcwl,, tar ,mgte-
An X-ray method/or rhe meosuremenfs of the thickness of wpporred rhmjilms in
elemcnl tilmr For muloclemcnl lilms. It 15
Ihe micromobr is modified for silica tilms on silicon nrrride and silicon carbide.
powble I” obtam buth Ihc tilm rhrcknos
Thr inctvhrrirs o/rhe o.&rn Ka linrbre measured on bulk SIO- und on rhe/i/m.
nod Ihe c”mpo,mon by mcrrur,ng r( ~1,“s
The derivation of the colibrarion curve giving the rhicknru of the film from rhe for all co,noonems of Ihe film. In txxh
rario of rhesr inrensiries is ourlined. The method has been used /or srl~ca Jilnts CBYI, all Acrclc”lc”lrl effccr, (atomic
rhwwr rhrr~~ I um wirh a lawral resolurion of a few micromrrurs. number. rbrorpo”“. nnd lluoreicencct arc
a\rumrd nept~glblc il, mlfhr be cxpccwd I”
a first ~ppruxmu,~“” fur lhm f,lm,
0.
X,DE thxkness mcasuremen,s arc The m~enuty-thickness rclaoonsh~p is nor
nctded for mnny applications such as slmplc smcc the Xaya generdlcd are a
Ihe rczareh on rhc oxidation and hot corro- function of depth I” the sample. and they
SL”” of uhcon “node and sdxon carbblde. are abrorbed rccordmg Lo the prrh of their
The prcw”“, method was adrpled Lo
Many methods hzve been used LOassess the cx~l from rhc samolc
meas~remcnta of the ,hxknos of s,hca
ev”luo”n of Ihe oxidaoon process; bow- The mrcroan~lyrir of dun tilmr has
films on sd,c”n carb,de md s,hc”” mmde
ever. they arc often mdlrect or desuuctive. bee” revxwcd by Guldslem.’ In order 1”
wng lhc oxygen Ko hnc t,, appllulu””
Fur ~“rtimcc. wghc changes arc the nc, calculate rhe inw&ry of the X-rays gener-
rewll of simuluncous rcacu”“~ whxh g,ve nled in [hc aample. Yakowtz and Newbury’
LO light elcmcms tatamlc oumbcr Z< tI)
req&cd scvcr.d mod,f,cau”n, ,,ncc the
“lfvxing comnbuuona. More dwcct mca- “rooaed a” emomeal aooruach bdscd on
h”rrmcnts arc oficn dependent on etch- ii&g Ihe X-ray’depfh oi
~roducrion ewe
low-energy X-ray, p&wed
absortxd and Ihe ab,“rprw”
,,re wungly
c”rrec,,“n 13
mg or frxurc of the specimens Other +(pzt to rhe combmmoo of a parabola and
w_mlarge I” use Ihe approruna~c/(~) ral-
methud> dppbcrble to rhm films ax very an ripmrnt,~l. The mdss rhicknrss p: IS
dceur.w bu! do nor have spatial re)oIulion. ,hc pr”duct “f ,he d,,ww from Ihc rurfacc
“es. The mwn,mc,,~ crlsut,md for Eq CI)
The ctecrron pr”tw microanalyzer IEPMA) : by the drnrlry p “f Ihc sample m Ihrt
xe d,ffercm trum Ihe mc.,rured m,c”,meS
1. becau,e of ~brvp,,“” .l’he X-rry, SC
otfcrr a nundcslruct~vc means of mcaur- lhlchne,,. Cornoared t” Ihnr for a bulk a!“-
grncrad A! var,““, dcplh, I” ,hr san~ptc
img ttlc Thickness of suppancd films wnh a &ard of Ihe co&wlion of Ihe tilm. the
and I” calcul~!c I,. lhe,r m,cn,,,y must be
high lalcral resoluuon which we have uxd @(pi) ~“rvc for the lhm tilm IS tnmcrred at
mtcgraad over dw depth of prwtucuon:
for rihca films a” silicon nilnde and sili- [he mas, tblckncs, pz, of dw lilm and it
con carb,dc. mctudo P mo,hf~ed b;rckscattering y!eld of
The menrny of characteristz X-rays, the cIcc~r”“s wh,ch accoum, fur the sub-
yeneraed by a” ctec~ron barn for an ele- ,,ra,c c”nlr~bu!wn The mwn~~ty of the
ment conlrmcd m a thin titm and not in the charnclerwc X-rays emmcd frum the film
bubruale. 15 related 1” the thickness of the II compared to thrl en,,rled by a bulk For un4 ahsorou”” [he ~“lal mtcn,W
thm film and c”mp”w~“n of the sample. ,tand.ud of Ihe film ma!eriA The in!may
r.11,” I, of these I~ncr. which can be
mr~wed. IS pr”pon~onrt 1” tbr rat” of
Ihc mlcn&llcs of lhc X-raya gencrafrd
1” the lilm l/L) 1” [hat generated I” a bulk
standard /,:
Appendix E-Publications 187

r
November\ 1987

h,k

#(PZ)

00

IO loo 1000 I‘XOO


Thickness ,nm,

Parabolic

h
kh and 0.1425 were calcul.Acd rer~ctwely for
bulk SIO,. S,,N,. md SIC The>e value)
were “blamed for a pnmrry clec~ron barn
ener~v of IO keV bv m~erwlrtwn of the
elem;;l,al drta of He;“r,ch’;nd rddmon of
mars-weighted elcmcmal c”clf~c,en,~.
Therefore, the fdm and Ihe wb\,rrtcs .~e so
and iir IS calculr,sd from the prcdlcled in which simdar dxa, .I c”“a,a”, brckscartermp cc&
mca,ured ,“,cn,,,,eh as fic,en,. Iha, of the ondc. ua, urcd for dll
film thxknesse$ rouhmg I” an enor es,,-
mr,ed I” br <I% of ,bc film dnckne\ses
for xhca glass on s,bcon nitride ,ubw.oes
X[-@z-t&f (p:-h++ j] and slightly over 1% for sd~con crrbrde
wbsorlo. For other subsmale-tilm comb,-
“a,,““,. when ,h,r a,aump,,“” 1s no, \rhd.
The inlegrrl for Ihe parabolic and ex-
in which If 15 Ihe X-WV m,e”s,,v measured a film b;rcksca,ten”p cwflicwn, “wt be
ponenoal r.eg,on 8” a lhxkness p ,uch as
from Ihe f~irn md /Z ;s rhu lot& mca- calculawd as a funcoon “1 p:, d.,. I. aod h
pi,>p:>l.Sh 1s
wed fur the bulk aodard. The mwgration are different for lilm .md wb,trr,e.”
for /: ,s over ,hr mas lh,ckness of [he film Cabbra”“” curve, of 1, II film ihlck-
and Ihrl for Ihe bulk mtenwy is over Ihe ncss were calculr~ed u,,h a compu,er u,mg
uhole depth of X-ray producd”” p:, g,ven Eq. (4) and Ihe mclhod oullmcd above.
by Ihe X-ray praluc!,“” range of Hemnch.’ -(pz,- IShI S,l,ca glr,s u!,h .I den,,,y of 2 2 g/cm’
Equroon (4) ha, ken wed rucce,rfully by
x Q (8) was ured for the lilm I” ,dI ma rage.
?+K(p-_.-I 5hl
KelM,. ma,, Ih,ckne,,. and ab,“rp,,on COICUII-
Ttw mwosmc, of Eq (4) were integrated wnh u”“,. The cahbrat,“” cunc 1, ,houn m
I” clozd form u~,“e Ihe “arabolic-ex- Fig. 2 for sdx”” carbldc md ,dxo” m-
ponenwdexpresrm-for &p;) The ” Q=exP[P&&-X j] mde ,ubwa,es.
arrumrd 4(p:c) cwve 1s sketched I” Fig. I.

1
For [he prrabohc ,eg,on
1.Shx The inteosmes of Ihe “x)pc” Ku line
lip.-J=h~‘(~:-h)‘(a-k)+k (5) were measured 1” a ,ca”“,“~ elcc~“”
m,croscope* fmcd wlh IWO ud\slmglh
and for the rrponenwl rrg,o” I;=I, (for pi = I.Sh)
disperwe spec,r”mctcn equ~ptwd for hgh,
The panme,er, C&Bk. and h are a func- eleme”, .mrly~r. All wmple, uere coaled
don of [he backscattering clewon yield. w,lb &a~, 20 “m of carbon. Duphcnle
Using values for the bulk matenal, the I”- mmsurcment~ were made ,,muluncously

x exp(w ) (6)
te”,,,y for lhr bulk standard I, crlculaed: wth ,hr IWO spec,r”me,ers ill a” .,ccelcrd-
lion voltage of IO kV and a beam curre”,
,“.=I, ifor pr=pz,)
of 20 rA. The ra,a”s kr tie calculaled from
in which Ihe parameters 6, k. and h are cab Ihe measured intms,l~es corrected for back-
culzued followmg Keuwr’ and Goldstem.’ However, for thm films Ihe elrcrron back- ground. The oxide Ihicknes,es CA” be read
The mlegral for Ihe parabolic reg,““. for a xa,,er yield mcludes a conrnbution from from Fig. 2: ntedd. a am~ple m,erp&o”n
dwknes, ,x<l 5h. IP the subwale. Approx~ma,~“na of Ihe da, program and E.q. (4) were used LOgel them
of Cusslet were u,rd prev~ou,ly”’ dnd more exacdy. The meawreme”lS are very
wed m lh,s case Some var,~,,““s on Ihe reproducible. and a standard error of 1% or
electron ,rrnrm,ra~“n cwffiwn,’ of the less IP c;rlcula!ed’ for Le values corre,pond-
,ubauele were also med. The aweme val- me 10 th,cknesxs >IW nm for a se, 01
(7) “es of Ihe backsca,ter cwfticlen, of the lob-s co,m,~ on pea, md background. for
film ,h”uld be thrl of the ,ub,rrate for a film and srandsrd. I, i, ~mprowd by lhr
film mfmwly lhm md lbsl 01 Ihe bulk film comblnarlon of several of these mea-
mr,er,;ll for a IhIck film.’ I” tb,, case. suremen,,. They were uwlly repcrIed at
bilckscaltrr cocffic~ent?, of 0.1316. 0 I37 I, ,hr ,ame we and it, sevrral Icat~ons of
188 High Temperature Corrosion of Ceramics

Communicdonr of rhe American Ceramic Societv November 1987

Table I. Oxide Thickness Comparison* better SUN+ Oxan Eq (I) for tits La\k. As
Ilwkncncu shown in Rg. 2. httle addmonal lhrckness
(PI, can bc measured w,th dur higher electron
SImplc Typr WDS SEM energy beau% the sofr X-ray, produced
I SIC.SC 0.32 0.21 deep in tbe sample by hght elements such
a, “xygro are almost completely absorbed
: SIC, SC
SIC. 0.31
0.20 0.30
0.13 Tb,s rc~ults m Ihe shallow ,lopc of the
4 Sic. SC 0.17 0.11 curve for duckcr tilms and Ihe prormuty of
rhe two curves for low film ducknose,.
2 SI,N,
SI,N.. CVD 0.10
0.05 0.09
0.12 Imermedmte electron beam energies such
.sc=nng!x clyrul. WDS..X.R) mlcmuulvrir. ud sEM=lnugw “lr- as 20 kV did not provide signlftcanl
yzuuu. impmvemcn~~

The X-ray melhod ,u,r dexnbed has


similar momholoev a” Ihe surface Of the by as much as 30 to 50%. However. tbe bee” used consi~~rntly I” our rebesrch’ for
sample. It I; esu&red that tix rcprcduc~- repmducibday of the X-ray method 1s ex- the measoremeot of StOl films wrh thick-
bdtty of Ihe thicknesses measured by this cellent. and Ihe meahurements are srraight- nesses from IO “m I” 1.0 pm with a re-
procedure was of Ihe order of II. forward. Reuter’ clawned an accuracy of producibdtry of a few percent. Thicker
+10% for the thicknesses of lhin films of films were imaged m cros) secuon m the
REIWLTS AND Discuss& Al. Cu. N,. and Au on various substrates SEM. a procedure whtch is earbe, and more
The Ihicknenes of glassy regions of measured by a slmdill. method Application accurare than dzscussed earher for the than-
“xzde layers formed by oxidation and hot of the X-ray method I” hgh! elements I”- ncr films. The X-ray method I, non-
corrouon on GIlcon carbide and silicon ni- creases the uncenrmry smce Ihe physical do,,uc,,ve and ha a lswal rraoluuon oi a
tnde al IooO”C were measured bv lhls quanuties, such as roass absorpoon coeffi- few mrcrolnel~r,; II ;Ill”WS a correlauon of
mothal and by unagrng of cross sec& in cientb. used in Ihe caIcuIa11ons are not as tb~kneas rorasureznen~ with surface fca-
tbe SEM. The rcsulls are gwe” I” Table 1. well estabhshed. Also. X-ray absorplnn [ores such a, spheruhlrs or glassy regtom.
The mraruremen~s by SEM Imaging play, a mqor role for Lhe longer wave- Other techniques such as chemical ;maly-
proved difficult for films under I gm duck. ,eng,hs. Con~parison wnh Ihe SEM mra- sir. opr~cal mwferometry. ad elhp5”m-
The magnX~ar~“n was cahbraled wilh stao- sure,,,en,s ruaeerts that the accuracy euy. whrle roorc accur.~~c. do no! offer lhls
dud latex spheres. The samples were frac- ,sbe,rer dun 30-g. The shape of the callbra- lateral re,o1u~~“” whtch 1s highly dewable
tured and brhdv elched I” HBF. to reveal t,“” curve of F,g. 2 s”gge,t, Ihat thick- in roatmal5 sc~cnce.
the oxide idye;. They had been carbon nessts inIhc rang~l0nmlo0.8
to I pm can
coated pnor 1” fracrure to observe tie mor- be measured: these values correspond ap
phology of the wface. and I, was found “roxlmatelv to the linear part of tbe curve.
that dos coating helped mantain Ihe sur- ihe lower iimir
is se1 by dir
counring statis-
Iace edge of the oxide lilm during etching. Lies. For the thicker lilms, the sens~liwly
The samples were recoated to avold charg- decreass as the curve of Fig. 2 levels off.
mg I” Ihe electron beam and the coating It was clamxd’ lhrl the ongmal method
dwk”ea,es were large compared lo some \yas reasonably accumte up 1” 306 of dx
of the “ride layers of Table 1. Under tbr bulk inter~cf~on volume, correspondmg in
c”“d#i”ns of Ihe measurements. il is ex- this car to <0.5 ,.un. However. the more
pxred lhar Ihe resolution of the SEM did extensive absorprion correction performed
not reach 10 em. The experimenrai diffi- here should allow the range of the “lea-
culties in the SEM imagmg of these thin surements to be extended
Inyen resulled in large experimenlal scaLLer Increasing the pnmuy electron energy
and poor accuracy. II is estimated that the should also increase Ihe film thicknesses
uncexiaintv of tbe thicknesses measured dl- which can be measured by generaling X-
recrly I” Ihe SEM could be as high as 40 to rays deeper in!” the sample. This WPI
50% for the dunner oxide films. As shown shown’ for metal films with Eq. (I). A
I” Table I, rhe LW” methods gwc tie same calibruion cuwe for 30.kV prunary elec-
order of magmwde but their results differ moos was generated wrh Eq. (4). which 16

Das könnte Ihnen auch gefallen