Tải bản đầy đủ (.pdf) (30 trang)

Corrosion of Ceramic and Composite Materials Part 5 pps

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (165.58 KB, 30 trang )

106 Chapter 2
9. The spontaneity of a reaction depends upon more
than just the heat of reaction. To predict stability,
one must consider also the entropy.
10. If the reaction is spontaneous, the change in free
energy is negative, whereas if the reaction is in
equilibrium, the free energy change is equal to zero.
11. The real problem with predicting whether a reaction
may take place or not is in selecting the proper
reaction to evaluate. Care must be taken not to
overlook some possible reactions.
12. Since the corrosion of ceramics in service may never
reach an equilibrium state, thermodynamic calculations
cannot be strictly applied because these calculations are
for systems in equilibrium. Many reactions, however,
closely approach equilibrium, and thus the condition
of equilibrium should be considered only as a limitation,
not as a barrier to interpretation of the data.
13. There is a general tendency for oxides to be reduced
at higher temperatures at constant oxygen partial
pressures. One should be aware that any metal will
reduce any oxide above it in the Ellingham diagram.
14. Unit activity should be applied only to species in the
pure state.
15. The most important parameter of corrosion from the
engineering viewpoint is the reaction rate.
16. Diffusion coefficients depend upon the composition and
structure of the material through which diffusion occurs.
17. The rate of the reaction expressed as the rate of
change of concentration, dc/dt, depends upon the
concentration of the reactants.


18. The discrepancies between the experimental data and the
theoretical models are often due to nonspherical particles,
a range in sizes, poor contact between reactants,
formation of multiple products, and the dependency of
the diffusion coefficient upon composition.
19. Arnold et al. [2.141] concluded that dynamic
thermogravimetric studies provide insufficient data
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 107
for calculation of reaction kinetics, that the data are
influenced by the experimental procedures, and that
the results are uncertain.
20. The enthalpy of the reaction is often sufficient to raise
or lower the sample temperature by as much as 1000°C.
21. The flow of material by diffusion is proportional to the
concentration gradient and is directed from the region
of high concentration to one of low concentration.
22. In isometric crystals, the diffusion coefficient is
isotropic, as it is in polycrystalline materials as long
as no preferred orientation exists.
23. In most real cases, the diffusion coefficient can vary
with time, temperature, composition, or position
along the sample, or any combination of these.
24. Silica-forming reactions are the most desirable for
protection against oxygen diffusion.
25. A diffusion flux will set up a thermal gradient in an
isothermal system.
2.11 ADDITIONAL RELATED READING
Vetter, K.J. Electrochemical Kinetics; Academic Press, New York, 1967.
Bockris, O.M.; Reddy, A.K.N. Modern Electrochemistry; Plenum Press,

New York, 1970; Vol. 2.
Shaw, D.J. Charged Interfaces. Introduction to Colloid and Surface
Chemistry, 3rd Ed.; Butterworths, London, 1980; 148–182.
Chp. 7.
Marshall, C.E. The Physical Chemistry and Mineralogy of Soils: Soils
in Place; Wiley & Sons: New York, 1977; Vol. II.
Reviews in Mineralogy: Mineral-Water Interface Geochemistry;
Hochella, M.F. Jr., White A.F., Eds.; Mineral. Soc. Am.,
Washington, DC, 1990; Vol. 23.
Burns, R.G. Mineralogical Applications of Crystal Field Theory;
Cambridge University Press: Cambridge, 1970.
Shackelford J.F., Ed.; Bioceramics, Applications of Ceramic and Glass
Materials in Medicine; Trans Tech Publications: Switzerland, 1999.
Copyright © 2004 by Marcel Dekker, Inc.
108 Chapter 2
Reviews in Mineralogy: Health Effects of Mineral Dusts. Guthrie,
G.D. Jr.; Mossman B.T., Eds.; Mineral. Soc. Am., Washington,
DC, 1993; Vol. 28.
P.G.Shewmon. Diffusion in Solids. J.Williams Book Co., Jenks, OK, 1983.
2.12 EXERCISES, QUESTIONS, AND
PROBLEMS
1. Discuss the reaction products that may form and how
they may relate to any interfacial reaction layer formed.
2. If a “unified theory of corrosion of ceramics” were
to be developed, what structural characteristic would
be included and why?
3. Look up the vapor pressure of several materials to
confirm the concept that covalent materials vaporize
more quickly than ionic materials due to their higher
vapor pressure.

4. Why does the corrosion rate decrease when a thermal
gradient is present?
5. The Arrhenius equation has been used to represent
the temperature dependence of corrosion. Discuss
when this equation is most appropriate and why.
6. Discuss the difference between direct and indirect
dissolution. What other terms are used to describe
these types of dissolution?
7. What is the most predominant parameter in the
equation for corrosion rate under free convection?
Why is this parameter more predominant than the
others?
8. Discuss the various problems relating to the
experimental verification of the galvanic corrosion
of ceramics.
9. Describe how the cross-linking of silica tetrahedra
affect corrosion in silicates by aqueous solutions.
10. How does pH affect the corrosion of crystalline ceramics
and how does this relate to isoelectric point (IEP)?
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 109
11. Discuss the difference between electrochemical and
chemical dissolution. What material parameters are
important in each type?
12. Describe how one tells whether solid-solid corrosion
occurs by bulk, grain boundary, or surface diffusion.
REFERENCES
2.1. Phase Diagrams for Ceramists. Vol. I-XII, Am. Ceram. Soc.,
Westerville, Ohio.
2.2. Cooper, A.R. The use of phase diagrams in dissolution studies.

In Refractory Materials; Alper, A.M., Ed.; Academic Press:
New York, 1970; Vol. 6-III, 237–250.
2.3. Kramer, D.P.; Osborne, N.R. Effects of atmosphere and dew
point on the wetting characteristics of a glass-ceramic on two
nickel-based superalloys. In Ceramic Engineering and Science
Proceedings; Smothers, W.J., Ed.; Am. Ceram. Soc.
Westerville, Ohio, 1983; 4 (9–10), 740–750.
2.4. Noyes, A.A.; Whitney, W.R. Rate of solution of solid
materials in their own solutions. (Ger) Z.Physik. Chem.
1897, 23, 689–692.
2.5. Nernst, W. Theory of reaction velocities in heterogeneous
systems. (Ger) Z.Physik. Chem. 47, 52–55.
2.6. Berthoud, A. Formation of crystal faces. J.Chem. Phys. 10,
624–635, 1912.
2.7. Prandtl, L. NACE Tech. Memo. No. 452, 1928.
2.8. Levich, B.G. Physicochemical Hydrodynamics; Prentice Hall:
Englewood Cliffs, NJ, 1962.
2.9. Levich, E.G. Theory of concentration polarization. Discuss.
Faraday Soc. 1, 37–43, 1947.
2.10. Cooper, A.R., Jr.; Kingery, W.D. Dissolution in ceramic
systems: I, Molecular diffusion, natural convection, and
forced convection studies of sapphire dissolution in calcium
aluminum silicate. J. Am. Ceram. Soc. 1964, 47 (1), 37–3.
2.11. Samaddar, B.N.; Kingery, W.D.; Cooper, A.R., Jr. Dissolution
in ceramic systems: II, Dissolution of alumina, mullite,
Copyright © 2004 by Marcel Dekker, Inc.
110 Chapter 2
anorthite, and silica in a calcium-aluminum-silicate slag. J.
Am. Ceram. Soc. 1964, 47 (5), 249–254.
2.12. Oishi, Y.; Copper, A.R., Jr.; Kingery, W.D. Dissolution in

ceramic systems: III, Boundary layer concentration gradients.
J. Am. Ceram. Soc. 1965, 48 (2), 88–95.
2.13. Hrma, P. Contribution to the study of the function between
the rate of isothermal corrosion and glass composition. (Fr)
Verres Refract. 1970, 24 (4–5), 166–168.
2.14. Lakatos, T.; Simmingskold, B. Influence of constituents on
the corrosion of pot clays by molten glass. Glass Technol.
1967, 8 (2), 43–47.
2.15. Lakatos, T.; Simmingskold, B. Corrosion effect of glasses
containing Na
2
O-CaO-MgO-Al
2
O3-SiO
2
on tank blocks
Corhart ZAC and sillimanite. Glastek. Tidskr. 1967, 22 (5),
107–113.
2.16. Lakatos, T.; Simmingskold, B. Influence of viscosity and
chemical composition of glass on its corrosion of sintered
alumina and silica glass. Glastek. Tidskr. 1971, 26 (4),
58–68.
2.17. Chung, Y D.; Schlesinger, M.E. Interaction of CaO-FeO-
SiO
2
slags with partially stabilized zirconia. J. Am. Ceram.
Soc. 1994, 77 (3), 612.
2.18. Pons, A.; Parent, A. The activity of the oxygen ion in glasses
and its effect on the corrosion of refractories. (Fr) Verres
Refract. 1969, 23 (3), 324–333.

2.19. Blau, H.H.; Smith, C.D. Refractory problems in glass
manufacture. Bull. Am. Ceram. Soc. 1950, 29 (1), 6–9.
2.20. Woolley, F.E. Prediction of refractory corrosion rate from
glass viscosity and composition. In UNITECR ’89
Proceedings; Trostel, L.J., Jr., Ed.; Am. Ceram. Soc.
Westerville, OH, 1989, 768–779.
2.21. Fox, D.S.; Jacobson, N.S.; Smialek, J.L. Hot corrosion of
silicon carbide and nitride at 1000°C. In Ceramic
Transactions: Corrosion and Corrosive Degradation of
Ceramics; Tressler, R.E., McNallan, M., Eds.; Am. Ceram.
Soc. Westerville, OH, 1990; Vol. 10, 227–249.
2.22. Jacobson, N.S.; Stearns, C.A.; Smialek, J.L.Burner rig
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 111
corrosion of SiC at 1000°C. Adv. Ceram. Mater. 1986, 1 (2),
154–161.
2.23. Cook, L.P.; Bonnell, D.W.; Rathnamma, D.Model for molten
salt corrosion of ceramics. In Ceramic Transactions:
Corrosion and Corrosive Degradation of Ceramics; Tressler,
R.E., McNallan, M., Eds.; Am Ceram. Soc. Westerville, OH,
1990; Vol. 10, 251–275.
2.24. Gordon, S.; McBride, B.J. Computer Program for Calculation
of Complex Chemical Equilibrium Compositions, Rocket
Performance, Incident & Reflected Shocks, and Chapman-
Jongnet Detonations. NASA SP-273, US Printing Office:
Washington, DC, 1971.
2.25. Borom, M.P.; Arendt, R.H.; Cook, N.C. Dissolution of oxides
of Y, Al, Mg, and La by molten fluorides. Ceram. Bull. 1981,
60 (11), 1168–1174.
2.26. Le Clerc, P.; Peyches, I. Polarization of refractory oxides

immersed in molten glass. (Fr) Verres Refract. 1953, 7 (6),
339–345.
2.27. Godrin, Y. Review of the Literature on Electrochemical
Phenomena. International Commission on Glass: Paris, 1975.
2.28. Vetter, K.J. Electrochemical Kinetics; Academic Press: New
York, 1967.
2.29. Wall, F.D.; Taylor, S.R.; Cahen, G.L. The simulation and
detection of electrochemical damage in BMI/graphite fiber
composites using electrochemical impedance spectroscopy.
In High Temperature and Environmental Effects on Polymeric
Composites, STP 1174; Harris, C.E., Gates, T.S., Eds.; ASTM:
Philadelphia, PA, 1993; 95–113.
2.30. Lindsay, J.G.; Bakker, W.T.; Dewing, E.W. Chemical resistance
of refractories to Al and Al-Mg alloys. J. Am. Ceram. Soc.
1964, 47 (2), 90–94.
2.31. Busby, T.Hotter refractories increase the risk of downward
drilling. Glass Ind. 1992, 73 (1), 20, 24.
2.32. Lasaga, A.C. Atomic treatment of mineral-water surface
reactions. In Reviews in Mineralogy, Mineral-Water Interface
Geochemistry; Hochella, M.F., Jr., White, A.F., Eds.; Mineral.
Soc. Am. Washington, DC, 1990; Vol. 23, 17–85. Chp. 2.
Copyright © 2004 by Marcel Dekker, Inc.
112 Chapter 2
2.33. Marshall, C.E. The Physical Chemistry and Mineralogy of
Soils: Soils in Place; Wiley & Sons: New York, 1977; Vol. II.
2.34. Huang, P.M.Feldspars, olivines, pyroxenes, and amphiboles.
In Minerals in Soil Environments; Dinauer, R.C., Ed.; Soil
Sci. Soc. Am. Madison, WI, 1977, 553–602. Chp. 15.
2.35. Casey, W.H.; Bunker, B.Leaching of mineral and glass surfaces
during dissolution. In Reviews in Mineralogy, Vol. 23:

Mineral-Water Interface Geochemistry; Hochella, M.F., Jr.,
White, A.F., Eds.; Mineral Soc. Am. Washington, DC, 1990;
Vol. 23, 397–426. Chp. 10.
2.36. Borchardt, C.A. Montmorillonite and other smectite minerals.
In Minerals in Soil Environments; Dinauer, R.C., Ed.; Soil
Sci. Soc. Am. Madison, WI, 1977, 293–330. Chp. 9.
2.37. Schnitzer, M.; Kodama, H. Reactions of minerals with soil
humic substances. In Minerals in Soil Environments; Dinauer,
R.C., Ed.; Soil Sci. Soc. Am. Madison, WI, 1977, 741–770.
Chp. 21.
2.38. Jennings, H.M. Aqueous solubility relationships for two types
of calcium silicate hydrate. J. Am. Ceram. Soc. 1986, 69 (8),
614–618.
2.39. Marshall, C.E. The Physical Chemistry and Mineralogy of
Soils: Soil Materials; Krieger Publishing Company:
Huntington, NY, 1975; Vol. I.
2.40. Elmer, T.H. Role of acid concentration in leaching of cordierite
and alkali borosilicate glass. J.Am.Ceram. Soc. 1985, 68 (10),
C273-C274.
2.41. Burns, R.G. Mineralogical Applications of Crystal Field
Theory; Cambridge University Press: London, 1970; 162–
167.
2.42. Hawkins, D.B.; Roy, R. Distribution of trace elements
between clays and zeolites formed by hydrothermal alteration
of synthetic basalts. Geochim. Cosmochim. Acta 1963, 27
(165), 785–795.
2.43. Shaw, D.J. Charged interfaces. Introduction to Colloid and
Surface Chemistry, 3rd Ed.; Butterworths: London, 1980;
148–182. Chp. 7.
2.44. Brady, P.V.; House, W.A. Surface-controlled dissolution and

Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 113
growth of minerals. In Physics and Chemistry of Mineral
Surfaces; Brady, P.V., Ed.; CRC Press: New York; 1996, 225–
305. Chp. 4.
2.45. Parks, G.A. The isoelectric points of solid oxides, solid
hydroxides, and aqueous hydroxo complex systems. Chem.
Rev. 1965, 65 (2), 177–198.
2.46. Diggle, J.W. Dissolution of oxide phases. In Oxides and Oxide
Films; Diggle, J.W., Ed.; Marcel Dekker: New York, 1973;
Vol. 2, 281–386. Chp. 4.
2.47. Bright, E.; Readey, D.W. Dissolution kinetics of TiO
2
in HF-
HCl solutions. J. Am. Ceram. Soc. 1987, 70 (12), 900–906.
2.48. Hulbert, S.F.; Bokros, J.C.; Hench, L.L.; Wilson, J.; Heimke,
G. In High Tech Ceramics; Vincenzini, P., Ed.; Elsevier Science
Pub. B.V.: Amsterdam, 1987; 180–213.
2.49. Hench, L.L.; Wilson, J. Introduction. An Introduction to
Bioceramics, Advanced Series in Ceramics. World Scientific
Publishing Co. Ltd.: Singapore, 1993; Vol. 1, 1–24.
2.50. Reviews in Mineralogy, Health Effects of Mineral Dusts.;
Guthrie, G.D., Jr., Mossman, T., Eds.; Mineral. Soc. Am.
Washington, DC, 1993; Vol. 28.
2.51. Nolan, R.P.; Langer, A.M. Limitations of the Stanton
hypothesis. In Health Effects of Mineral Dusts; Guthrie, G.D.,
Jr., Mossman, T., Eds.; Reviews in Mineralogy. Min. Soc.
Am. Washington, DC, 1993; Vol. 28. Chp. 9.
2.52. Correns, C.W. Growth and dissolution of crystals under linear
pressure. Discuss. Faraday Soc. No. 5; 1949, 267–271.

2.53. Winkler, E.M. Salt action on stone in urban buildings. In
Application of Science in Examination of Works of Art; Joung,
W.J., Ed.; Museum of Fine Arts: Boston, 1973.
2.54. Skoulikidis, T.N. Atmospheric corrosion of concrete
reinforcements, limestones, and marbles. In Atmospheric
Corrosion; Ailor, W.H., Ed.; John Wiley & Sons: New York,
1982; 807–825.
2.55. Amoroso, G.G.; Fassina, V. Stone Decay and Conservation;
Elsevier: Amsterdam, 1983; 12.
2.56. Hoffmann, M.R.Fog and cloud water deposition. In Materials
Degradation Caused by Acid Rain; ACS Symposium Series
Copyright © 2004 by Marcel Dekker, Inc.
114 Chapter 2
318; Baboian, R., Ed.; Am. Chem. Soc. Washington, DC,
1986; 64–91.
2.57. Mulawa, P.A.; Cadle, S.H.; Lipari, F.; Ang, C.C.; Vandervennet,
R.T. Urban dew: Composition and influence on dry deposition
rates. In Materials Degradation Caused by Acid Rain, ACS
Symposium Series 318; Baboian, R., Ed.; Am. Chem. Soc.
Washington, DC, 1986; 61–91.
2.58. Semonin, R.G. Wet deposition chemistry. In Materials
Degradation Caused by Acid Rain, ACS Symposium Series
318; Baboian, R., Ed.; Am. Chem. Soc. Washington, DC,
1986; 23–41.
2.59. Neal, K.M.; Newnam, S.H.; Pokorney, L.M.; Rybarczyk, J.P.
Elemental analysis of simulated acid rain stripping of Indiana
limestone, marble, and bronze. In Materials Degradation
Caused by Acid Rain, ACS Symposium Series 318; Baboian,
R., Ed.; Am. Chem. Soc. Washington, DC, 1986; 285–300.
2.60. Reddy, M.M.; Sherwood, S.I.; Doe, B.R.Limestone and

marble dissolution by acid rain: An onsite weathering
experiment. In Materials Degradation Caused by Acid Rain,
ACS Symposium Series 318; Baboian, R., Ed.; Am. Chem.
Soc. Washington, DC, 1986; 226–238.
2.61. Kobussen, A.G. Corrosion in condensing gas-fired central
heating boilers. In Dewpoint Corrosion; Holmes, D.R., Ed.;
Ellis Horwood Ltd.: Chichester, UK, 1985; 179–190.
2.62. Penkett, S.A. Chemical changes in the air. SCI Sulfur
Symposium; May 1979; 109–122.
2.63. Cox, W.M.; Farrell, D.M; Dawson, J.L.Corrosion monitoring
for process control. In Dewpoint Corrosion; Holmes, D.R.,
Ed.; Ellis Horwood Ltd.: Chichester, UK, 1985; 191–217.
2.64. Cussler, E.L.; Featherstone, J.D.B. Demineralization of porous
solids. Science Aug. 1981, 213, 1018–1019.
2.65. Yoshimura, M.; Hiuga, T.; Somiya, S. Dissolution & reaction
of yttria-stabilized zirconia single crystals in hydrothermal
solutions. J. Am. Ceram. Soc. 1986, 69 (7), 583–584.
2.66. Sato, T.; Ohtaki, S.; Shimada, M. Transformation of yttria
partially stabilized zirconia by low temperature annealing in
air. J.Mater. Sci. 1985, 20 (4), 1466–1470.
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 115
2.67. Janowski, K.R.; Rossi, R.C. Mechanical degradation of MgO
by water vapor. J. Am. Ceram. Soc. 1968, 51 (8), 453–455.
2.68. White, W.B. Glass structure and glass durability. In Materials
Stability and Environmental Degradation, Materials Research
Society Symposium Proceedings; Barkatt, A., Verink, E.D.,
Jr., Smith, L.R., Eds.; Mater. Res. Soc. Pittsburgh, PA, 1988;
Vol. 125, 109–114.
2.69. Wald, J.W.; Messier, D.R.; DeGuire, E.J. Leaching behavior of

Si-Y-Al-O-N glasses. Int. J. High Technol. Ceram. 1986, 2
(1), 65–72.
2.70. Douglas, R.W.; El-Shamy, T.M.M. Reaction of glass with
aqueous solutions. J. Am. Ceram. Soc. 1967, 50 (1), 1–8.
2.71. Jantzen, C.M. Thermodynamic approach to glass corrosion.
In Corrosion of Glass, Ceramics, & Ceramic
Superconductors; Clark, D.E., Zoitos, B.K., Eds.; Noyes
Publications: Park Ridge, NJ, 1992; 153–217. Chp. 6.
2.72. Newton, R.G.; Paul, A. A new approach to predicting the
durability of glasses from their chemical compositions. Glass
Technol. 1980, 21 (6), 307–309.
2.73. Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous
Solution; NACE: Houston, TX, 1974. Eng Trans by
J.A.Franklin.
2.74. Garrels, R.M.; Christ, C.L. Solutions, Minerals, and
Equilibria; Harper and Row: New York, 1965.
2.75. Hench, L.L.; Clark, D.E. Physical chemistry of glass surfaces.
J. Non-Cryst. Solids 1978, 28, 83–105.
2.76. McVay, G.L.; Peterson, L.R. Effect of gamma radiation on
glass leaching. J. Am. Ceram. Soc. 1981, 64 (3), 154–158.
2.77. Hogenson, D.K.; Healy, J.H. Mathematical treatment of
glass corrosion data. J.Am. Ceram. Soc. 1962, 45 (4), 178–
181.
2.78. Budd, S.M. The mechanism of chemical reaction between
silicate glass and attacking agents; Part 1. Electrophilic and
nucleophilic mechanism of attack. Phys. Chem. Glasses 1961,
2 (4), 111–114.
2.79. Budd, S.M.; Frackiewicz, J. The mechanism of chemical
reaction between silicate glass and attacking agents; Part 2.
Copyright © 2004 by Marcel Dekker, Inc.

116 Chapter 2
Chemical equilibria at glass-solution interfaces. Phys. Chem.
Glasses 1961, 2 (4), 115–118.
2.80. Simmons, C.J.; Simmons, J.H. Chemical durability of fluoride
Glasses: I, Reaction of fluorozirconate glasses with water.
J.Am. Ceram. Soc. 1986, 69 (9), 661–669.
2.81. Thomas, W.F. An investigation of the factors likely to affect
the strength and properties of glass fibers. Phys. Chem. Glasses
1960, 1 (1), 4–18.
2.82. Wojnarovits, I. Behavior of glass fibers in strong acidic and
alkaline media. J. Am. Ceram. Soc. 1983, 66 (12), 896–898.
2.83. Hench, L.L. Bioactive glasses help heal, repair and build
human tissue. Glass Res. 2002–2003, 12 (1–2), 18.
2.84. White, J.E.; Day, D.E. Rare earth aluminosilicate glasses for
in vivo radiation delivery. In Rare Elements in Glasses; Key
Engineering Materials; Shelby, J.E., Ed.; Trans Tech Pub.,
Switzerland, 1994; Vols. 94–95, 181–208.
2.85. Oda, K.; Yoshio, T. Properties of Y
2
O
3
-Al
2
O
3
-SiO
2
glasses as
a model system of grain boundary phase of Si
3

N
4
ceramics,
Part 2: Leaching characteristics. J.Ceram. Soc. Jpn. 1991, 99
(11), 1150–1152.
2.86. Erbe, E.M.; Day, D.E. In Proceedings: Science & Tech of
New Glasses; Sakka, S., Soga, N., Eds.; Ceram. Soc. Japan:
Tokyo, 1991.
2.87. Conzone, C.D.; Brown, R.F.; Day, D.E.; Ehrhardt, G.J. In
vitro and in vivo dissolution behavior of a dysprosium lithium
borate glass designed for the radiation synovectomy treatment
of rheumatoid arthritis. J.Biomed. Mater. Res. 2002, 60 (2),
260–268.
2.88. Day, D.E. Reactions of bioactive borate glasses with
physiological liquids. Glass Res. 2002–2003, 12 (1–2),
21–22.
2.89. Bauer, J.F. Corrosion and surface effects of glass fiber in
biological fluids. Glass Res. 2000, 9 (2), 4–5.
2.90. Kubaschewski, O.; Hopkins, B.E. Oxidation of Metals and
Alloys; Butterworths: London, 1962.
2.91. Readey, D.W. Gaseous corrosion of ceramics. In Ceramic
Transactions, Corrosion and Corrosive Degradation of
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 117
Ceramics; Tressler, R.E., McNallan, M., Eds.; Am. Ceram.
Soc., Westerville, OH, 1990; Vol. 10, 53–80.
2.92. Yokokawa, H.; Kawada, T.; Dokiya, M. Construction of
chemical potential diagrams for metal-metal-nonmetal
systems: Applications to the decomposition of double oxides.
J. Am. Ceram. Soc. 1989, 72 (11), 2104–2110.

2.93. Grimley, R.T.; Burns, R.P.; Inghram, M.G. Thermodynamics
of vaporization of Cr
2
O
3
: Dissociation energies of CrO, CrO
2
and CrO
3
. J. Chem. Phys. 1961, 34 (2), 664–667.
2.94. Graham, H.C.; Davis, H.H. Oxidation/vaporization kinetics
of Cr
2
O
3
. J. Am. Ceram. Soc. 1971, 54 (2), 89–93.
2.95. Pilling, N.B.; Bedworth, R.E. The oxidation of metals at high
temperature. J. Inst. Met. 1923, 29, 529–591.
2.96. Jorgensen, P.J.; Wadsworth, M.E.; Cutler, I.B. Effects of oxygen
partial pressure on the oxidation of silicon carbide. J. Am.
Ceram. Soc. 1960, 43 (4), 209–212.
2.97. Engell, H.J.; Hauffe, K. Influence of adsorption phenomena
on oxidation of metals at high temperatures. Metall 1952, 6,
285–291.
2.98. Langmuir, I. Evaporation of small spheres. Phys. Rev. 1918,
12 (5), 368–370.
2.99. Tichane, R.M.; Carrier, G.B. The microstructure of a soda-
lime glass surface. J. Am. Ceram. Soc. 1961, 44 (12),
606–610.
2.100. Simpson, H.E. Study of surface structure of glass as related

to its durability. J. Am. Ceram. Soc. 1958, 41 (2), 43–9.
2.101. Tichane, R.M. Initial stages of the weathering process on a
soda-lime glass surface. Glass Technol. 1966, 7 (1), 26–29.
2.102. Burggraaf, A.J.; van Velzen, H.C. Glasses resistant to sodium
vapor at temperatures to 700°C. J. Am. Ceram. Soc. 1969,
52 (5), 238–242.
2.103. Johnston, W.D.; Chelko, A.J. Reduction of ions in glass by
hydrogen. J. Am. Ceram. Soc. 1970, 53 (6), 295–301.
2.104. Gibson, A.S.; LaFemina, J.P. Structure of mineral surfaces. In
Physics and Chemistry of Mineral Surfaces; Brady, P.V., Ed.;
CRC Press: New York, 1996; 1–62.
Copyright © 2004 by Marcel Dekker, Inc.
118 Chapter 2
2.105. Hochella, M.F. Jr.; White, A.F. Mineral-Water Interface
Geochemistry; Rev. Mineral; 23 pp.
2.106. Washburn, E.W. Note on a method of determining the
distribution of pore sizes in a porous material. Proc. Natl.
Acad. Sci. 1921, 7, 115–116.
2.107. Smithwick, R.W.; Fuller, E.L. A generalized analysis of
hysteresis in mercury porosimetry. Powder Technol. 1984,
38, 165–173.
2.108. Conner, W.C. Jr.; Blanco, C; Coyne, K.; Neil, J.; Mendioroz,
S.; Pajares, J. Measurement of morphology of high surface
area solids: Inferring pore shape characteristics. In
Characterization of Porous Solids; Unger, K.K., et al., Ed.;
Elsevier Science Publishers: Amsterdam, 1988; 273–281.
2.109. Moscou, L.; Lub, S. Practical use of mercury porosimetry
in the study of porous solids. Powder Technol. 1981, 29,
45–52.
2.110. Lapidus, G.R.; Lane, A.M.; Ng, K.M.; Conner, W.C.

Interpretation of mercury porosimetry data using a pore-throat
network model. Chem. Eng. Commun. 1985, 38, 33–56.
2.111. Conner, W.C.; Lane, A.M. Measurement of the morphology
of high surface area solids: Effect of network structure on
the simulation of porosimetry. J.Catal. 1984, 89, 217–225.
2.112. Van Brakel, J.; Modry, S.; Svata, M. Mercury porosimetry:
State of the art. Powder Technol. 1981, 29, 1–12.
2.113. Rootare, H.M.; Nyce, A.C. The use of porosimetry in the
measurement of pore size distribution in porous materials.
Int. J. Powder Metall. 1971, 7(1), 3–11.
2.114. Smith, C.S.Grains, phases, and interpretation of
microstructure. Trans. AIME 1948, 175 (1), 15–51.
2.115. White, J. Magnesia-based refractories. In High Temperature
Oxides, Part 1: Magnesia, Lime and Chrome Refractories.
Alper, A.M., Ed.; Refractory Materials: A Series of
Monographs; Margrave, J.L., Ed.; Academic Press: New York,
1970; Vol. 5–1, 77–141.
2.116. Inomata, Y. Oxidation resistant Si-impregnated surface
layer on reaction sintered articles. Yogyo Kyokaishi 1975,
83 (1), 1–3.
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 119
2.117. Messier, D.R. Use of Ti to enhance wetting of reaction-bonded
Si
3
N
4
by Si. In Ceramic Engineering and Science Proceedings;
Smothers, W.J., Ed.; Am. Ceram. Soc. Westerville, OH, 1980,
1 (7–8B), 624–633.

2.118. Puyane, R.; Trojer, F. Refractory wear and wettability by glass
at high temperatures. Glass 1980, 57 (12), 5–8.
2.119. Carre, A.; Roger, F.; Varinot, C. Study of acid/base properties
of oxide, oxide glass, and glass-ceramic surfaces. J.Colloid
Interface Sci. 1992, 154 (1), 174–183.
2.120. Gaskell, D.R. Introduction to Metallurgical Thermodynamics,
2nd Ed.; McGraw-Hill: New York, 1981.
2.121. Swalin, R.A. Thermodynamics of Solids. Wiley & Sons Inc.:
New York, 1962.
2.122. Bent, H.A. The Second Law. Oxford University Press: New
York, 1965.
2.123. Chase, M.W., Jr.; Davies, C.A.; Downey, J.R., Jr.; Frurip,
D.J.R.; McDonald, R.A.; Syverud, A.N. J. Phys. Chem.
Reference Data, Vol. 14, Suppl. No. 1, JANAF
Thermochemical Tables, 3rd Ed., Parts I & II, Am. Chem.
Soc. & Am. Inst. Phys., 1985.
2.124. Kubaschewski, O.; Evans, E.L.; Alcock, C.B. Metallurgical
Thermodynamics. Pergamon Press: Oxford, 1967.
2.125. Krupka, K.M.; Hemingway, B.S.; Robie, R.A.; Kerrick, D.M.
High temperature heat capacities and derived thermodynamic
properties of anthophyllite, diopside, dolomite, enstatite,
bronzite, talc, tremolite, and wollastonite. Am. Mineral. 1985,
70, 261–271.
2.126. Eriksson, G. Thermodynamic studies of high temperature
equilibria. XII. SOLGASMIX, a computer program for
calculation of equilibrium compositions in multiphase
systems. Chem. Scr. 1975, 8, 100–103.
2.127. Latimer, W.M. The Oxidation States of the Elements and
Their Potentials in Aqueous Solutions; Prentice-Hall:
Englewood Cliffs, NJ, 1952.

2.128. Brenner, A. The Gibbs-Helmoltz equation and the EMF of
galvanic cells, II. Precision of its application to concentration
cells. J. Electrochem. Soc. 1975, 122 (12), 1609–1615.
Copyright © 2004 by Marcel Dekker, Inc.
120 Chapter 2
2.129. Livey, D.T.; Murray, P. The stability of refractory materials. In
Physicochemical Measurements at High Temperatures;
Bockris, J. O’M., et al., Ed.; Butterworths Scientific
Publications: London, 1959; 87–116.
2.130. Luthra, K.L. Chemical interactions in ceramic and carbon-
carbon composites. In Materials Research Society
Symposium Proceedings: Materials Stability and
Environmental Degradation; Barkatt, A. Verink, E.D., Jr.,
Smith, L.R., Eds.; Mat. Res. Soc.: Pittsburgh, PA, 1988; Vol.
125, 53–60.
2.131. Ellingham, H.J.T. Reducibility of oxides and sulfides in
metallurgical processes. J. Soc. Chem. Ind. 1944, 63, 125.
2.132. Richardson, F.D.; Jeffes, J.H.E. The thermodynamics of
substances of interest in iron and steel making from 0°C to
2400°C; I Oxides. J. Iron Steel Inst. 1948, 160, 261.
2.133. Darken, L.S.; Gurry, R.W. Physical Chemistry of Metals;
McGraw-Hill: New York, 1953; 348–349.
2.134. Lou, V.L.K.; Mitchell, T.E.; Heuer, A.H. Review—Graphical
displays of the thermodynamics of high-temperature gas-
solid reactions and their application to oxidation of metals
and evaporation of oxides. J. Am. Ceram. Soc. 1985, 68
(2), 49–58.
2.135. Quets, J.M.; Dresher, W.H. Thermochemistry of the hot
corrosion of superalloys. J. Mater. 1969, 4 (3), 583–599.
2.136. Barret, P., Ed.; Reaction Kinetics in Heterogeneous Chemical

Systems; Elsevier: Amsterdam, 1975.
2.137. Sharp, J.H.; Brindley, G.W.; Narahari Achar, B.N. Numerical
data for some commonly used solid state reaction equations.
J. Am. Ceram. Soc. 1966, 49 (7), 379–382.
2.138. Frade, J.R.; Cable, M. Reexamination of the basic theoretical
model for the kinetics of solid-state reactions. J. Am. Ceram.
Soc. 1992, 75 (7), 1949–1957.
2.139. Freeman, E.S.; Carroll, B. The application of thermoanalytical
techniques to reaction kinetics. The thermogravimetric
evaluation of the kinetics of the decomposition of calcium
oxalate monohydrate. J. Phys. Chem. 1958, 62 (4), 394–397.
2.140. Sestak, J. Errors of kinetic data obtained from
Copyright © 2004 by Marcel Dekker, Inc.
Fundamentals 121
thermogravimetric curves at increasing temperature. Talanta
1966, 13 (4), 567–579.
2.141. Arnold, M.; Veress, G.E.; Paulik, J.; Paulik, F. The applicability
of the Arrhenius model in thermal analysis. Anal. Chim. Acta.
1981, 124 (2), 341–350.
2.142. Sestak, J. Thermophysical Properties of Solids. Thermal
Analysis; Wendlandt, W.W., Ed.; Part D; Svehla, G., Ed.;
Comprehensive Analytical Chemistry; Svehla, G., Ed.; Vol.
XII; Elsevier: Amsterdam, 1984.
2.143. Holman, J.P. Heat Transfer. McGraw-Hill: New York, 1963.
2.144. Courtright, E.L. Engineering limitations of ceramic
composites for high performance and high temperature
applications. In Proc. 1993 Conf. on Processing, Fabrication
and Applications of Advanced Composites, Long Beach, CA,
Aug 9–11; Upadhya, K., Ed.; ASM: Ohio, 1993; 21–32.
2.145. Tripp, W.C.; Davis, H.H.; Graham, H.C. Effects of SiC

additions on the oxidation of ZrB
2
. Ceram. Bull. 1973, 52
(8), 612–616.
2.146. Oishi, Y.; Kingery, W.D. Self-diffusion in single crystal and
polycrystalline aluminum oxide. J. Chem. Phys. 1960, 33,
480.
2.147. Sucov, E.W. Diffusion of oxygen in vitreous silica. J. Am.
Ceram. Soc. 1963, 46 (1), 14–20.
2.148. Kingery, W.D.; Pappis, J.; Doty, M.E.; Hill, D.C. Oxygen ion
mobility in cubic Zr
0.85
Ca
0.15
O
1.85
. J. Am. Ceram. Soc. 1959,
42 (8), 393–398.
2.149. Paladino, A.E.; Kingery, W.D. Aluminum ion diffusion in
aluminum oxide. J. Chem. Phys. 1962, 37 (5), 957–962.
2.150. Rhodes, W.H.; Carter, R.E. Ionic self-diffusion in calcia
stabilized zirconia, 64th Annual Mtg Abstracts. Am. Ceram.
Soc. Bull. 1962, 41 (4), 283.
2.151. Lindner, R.; Parfitt, G.D. Diffusion of radioactive magnesium
in magnesium oxide crystals. J. Chem. Phys. 1957, 26, 182.
2.152. Lindner, R.; Akerstrom, A. Self-diffusion and reaction in oxide
and spinel systems. Z. Phys. Chem. 1956, 6, 162.
2.153. Lindner, R.; Hassenteufel, W.; Kotera, Y. Diffusion of
Copyright © 2004 by Marcel Dekker, Inc.
122 Chapter 2

radioactive lead in lead metasilicate glass. Z. Phys. Chem.
1960, 23, 408.
2.154. Shewmon, P.G. Diffusion in Solids; J.Williams Book Co.:
Jenks, OK, 1983.
2.155. Crank, J. The Mathematics of Diffusion; Oxford University
Press: Fair Lawn, NJ, 1956.
Copyright © 2004 by Marcel Dekker, Inc.
123
3
Methods of Corrosion Analysis
One must learn by doing the thing; for though you think
you know it, you have no certainty until you try.
SOPHOCLES
3.1 INTRODUCTION
According to Weisser and Bange [3.1] it was Lavoisier who in
1770 first recorded the aqueous corrosion of a silicate glass
predominantly by use of an analytical balance. The analysis of
corrosion has been changing over the years with the greatest
changes probably taking place within the last 25 years. These
changes have been due mostly to the availability of
sophisticated computerized analytical tools. It has taken many
years for investigators to become familiar with the results
obtained and how to interpret them. In some cases, special
Copyright © 2004 by Marcel Dekker, Inc.
124 Chapter 3
sample preparation techniques had to be perfected. Although
one could conceivably employ all the various characterization
methods described below, in most cases, only a few are needed
to obtain sufficient information to solve a particular problem.
The determination of the overall mechanism of corrosion

requires a thorough detailed investigation using several
characterization methods. Many times, though, the investigator
has a limited amount of time and/or funds to obtain data and
thus must rely on a few well-chosen tools. It should be obvious
that considerable thought should be given to the selection of
samples, test conditions, characterization methods, and
interpretation of the results, especially if the data are to be
used for prediction of lifetimes in actual service conditions.
The reader is referred to the book by Wachtman [3.2] for a
review of the principles involved in the various characterization
techniques.
3.2 LABORATORY TEST VS. FIELD TRIALS
There are two general ways to approach a corrosion problem:
either to conduct some laboratory tests to obtain information
as to how a particular material will behave under certain
conditions, or to perform a postmortem examination of field
trial samples. It is best to perform the laboratory test first to
aid in making the proper selection of materials for a particular
environment and then perform the field trial. Laboratory
tests, however, do not always yield the most accurate
information since they rely on the investigator for proper
setup; however, they are easier to control. The investigator
must have a thorough understanding of the environment
where the ceramic is to be used and must select the portions of
the environment that may cause corrosion. For example, it is
not sufficient to know that a furnace for firing ceramicware is
heated by fuel oil to a temperature of 1200°C. One must also
know what grade fuel oil is used and the various impurities
contained in the oil and at what levels. In addition,
parameters such as partial pressure of oxygen, moisture

Copyright © 2004 by Marcel Dekker, Inc.
Methods of Corrosion Analysis 125
content, etc. may be important. Once all these various
parameters are known, the investigator can set up an
appropriate laboratory test.
One must also understand all the various things that cannot
be scaled down to a laboratory test, such as viscosity of liquids,
time, temperature, etc. Care must be exercised when
attempting to perform an accelerated laboratory test, which is
usually accomplished by raising the temperature or increasing
the concentration of the corrosive medium or both. Since the
mechanism of corrosion in the accelerated test may not be the
same (generally, it is not the same) as that under actual service
conditions, erroneous conclusions and inaccurate predictions
may be obtained. The mechanisms must be the same for
accurate application of laboratory test results to actual service
conditions. Sample size is one parameter that is easily scaled;
however, this can also cause problems. For example, when
testing the corrosion of a ceramic by a liquid, the ratio of liquid
volume present to the surface area of the exposed ceramic is
very important. The investigator must remember that
corrosion is controlled predominantly by thermodynamics and
kinetics. Assuming that the proper laboratory tests have been
conducted, the probability that any problems will arise is
minimal.
The only way to analyze corrosion accurately is to conduct
a field trial. This entails placing selected materials in actual
service conditions, generally for an abbreviated time, and then
collecting samples for analysis along with all the operational
data of the particular environment. The size and amount of

material or samples placed into actual service conditions for a
field trial can be as little as one small laboratory test bar, or,
for example, as large as a complete wall in a large industrial
furnace. The larger the installation for the field trial, the more
confidence one must have in the selection of materials. The
larger installations are generally preceded by several
laboratory tests and possibly a small-scale field trial.
Abbreviated times may be as long as several years or as short
as several days.
Copyright © 2004 by Marcel Dekker, Inc.
126 Chapter 3
Data such as temperature and time are the obvious ones to
collect, but there exists a large amount of other data that should
be examined. Many times, however, some of the more
important data do not exist for one reason or another. For
example, maybe the oxygen partial pressure was not determined
during the duration of the service life of the ceramic. In some
cases, it may be impossible to collect certain pieces of data
during the operation of the particular piece of equipment. At
these times, a knowledge of phase equilibria, thermodynamics,
and kinetics can help fill in the gaps or at least give an indication
as to what was present.
3.3 SAMPLE SELECTION AND
PREPARATION
It should be obvious that powders will present a greater surface
area to corrosion and thus will corrode more rapidly than a
solid sample. One may think this to be a good way to obtain a
rapid test, but saturation of the corroding solution may cause
corrosion to cease, or even cause a reverse reaction (i.e., crystal
growth), giving misleading results. This points to the extreme

importance of the surface area to volume ratio (SA/V) of the
ceramic to the corroding solution. Another factor related to
this is that during corrosion, the surface may change, altering
the SA/V ratio effect. Surface areas during dissolution have
been reported to increase presumably due to opening of etch
pits, microfissures, etc. [3.3].
Selecting samples for analysis provides another challenge
to the investigator. Foremost in the selection process is selecting
an area for analysis that is representative of the overall
corrosion process. If this cannot be done, then many samples
must be analyzed. Much of the modern analytical equipment
necessitates the analysis of very small samples, thus one must
be very sensitive to the selection of representative samples or
at least evaluate multiple samples.
Much care must be given to preparing samples that contain
an adherent reaction product surface layer. It is best to select a
Copyright © 2004 by Marcel Dekker, Inc.
Methods of Corrosion Analysis 127
sample that is many times larger than required by the final
technique and then mounting this in some metallurgical
mountant (e.g., epoxy). After the larger sample has been
encased, then smaller samples can be safely cut from the larger
piece.
Solid samples, when prepared for laboratory tests, should
be cleaned in a noncorrosive solution to remove any loose
particles adhering to the surface and any extraneous
contamination. Brady and House [3.3] have reported that
initially, an accelerated, nonlinear dissolution may occur from
high-energy sites caused by grinding and incomplete removal
of ultrafine particles. Best results are obtained if the cleaning

is done in an ultrasonic cleaner. These cleaning solutions can
be obtained from any of the metallographic supply companies.
If the sample is mounted into one of the epoxy-type
metallographic mountants, one must be aware that some
cleaning solutions will react with the mountant. It is best to
use supplies from one manufacturer to avoid these problems.
If as-manufactured samples are used for corrosion tests, one
should remove a thin surface layer by grinding and cleaning
before performing the corrosion tests. In this way, remnants
from such things as powder beds or encapsulation media used
in the production of the material can be eliminated and
therefore not interfere with the corrosion process.
Quite often, the as-manufactured surface of a ceramic will
have a different microstructure or even chemistry than the bulk.
This often manifests itself as a thin surface layer (as much as
several millimeters thick) that contains smaller grain sizes (more
grain boundaries) and possibly a lower porosity. If the corrosion
test corrodes only this thin surface layer, again, misleading
results will be obtained. One way to solve this problem is to
remove the surface layer by grinding. Grinding, however, must
be done with some thought to the final surface roughness since,
again, this will affect the SA/V ratio. Diamond-impregnated
metal grinding disks should be used rather than silicon carbide
paper disks or silicon carbide loose grit. Loose grit and the
grinding media from paper disks have a tendency to become
Copyright © 2004 by Marcel Dekker, Inc.
128 Chapter 3
lodged within the pores and cracks of the sample being prepared.
The final grinding media grit size should be no greater than 15
µm. It is best to clean samples after each grinding step in an

ultrasonic cleaner with the appropriate cleaning solution.
Surface roughness of solid samples is an item that is often
overlooked. Not only does a rough surface increase the area
exposed to corrosion, but it may also lead to problems with
some analytical techniques. For example, when the surface
roughness is on the order of the reaction layer thickness caused
by corrosion, errors will be present in the depth profiles
obtained by secondary ion mass spectroscopy (SIMS). In those
cases when surface analyses are planned, one should prepare
solid samples to at least a 5-µm finish.
Grinding and polishing of samples that contain a reaction
product surface layer should be done so that the reaction layer
is not damaged, or the interface obscured. If part of the sample
is metal, then polishing should be done in the direction ceramic
toward metal to eliminate smearing the metal over the ceramic.
If very thin reaction layers are present, one can prepare taper
sections to increase the area that is examined. Sample preparation
of composites presents some additional problems since materials
of very different characteristics will be presented at the surface
being polished. Chanat [3.4] of Buehler Ltd. has offered some
good advice for mounting, sectioning, and polishing. The most
effective method involves the use of diamond abrasives with
low nap cloths for FRCMC*. High nap cloths can induce
excessive relief at boundaries of different materials.
Many tips on how to prepare samples can be obtained by
reading the various technical journals published by the
manufacturers of consumable grinding and polishing supplies.
One particular article that offers some new ideas was that of
Damgaard and Geels [3.5]. They emphasized the importance
of polishing disk diameter and velocity, indicating that both

are directly proportional to material removal rates. Although
* FRCMC=fiber-reinforced ceramic matrix composite.
Copyright © 2004 by Marcel Dekker, Inc.
Methods of Corrosion Analysis 129
this may be true, one must be very aware of the amount of
lubricant used, the pressures applied, and the area of the sample
being polished. If the lubricant supply rate is constant, which
is generally the case, the material removal rate will peak at
about 300–350 rpm. Thus if one is thinking of purchasing
automatic grinding and polishing equipment, he should look
for something that has automatic lubrication flow rate control.
Although many advances have been made in the grinding and
polishing of ceramics, this area still is very much an art. Subtle
changes in the procedure can make a major difference in the
final finish of the sample*.
3.4 SELECTION OF TEST CONDITIONS
Although the selection of appropriate samples can be a major
problem, the selection of the appropriate test conditions is an
even more difficult task. The goal of the industrial corrosion
engineer in selecting test conditions is to simulate actual service
conditions. Selection of test conditions is much easier for the
scientist, who is attempting to determine mechanisms. The
major problem in attempting to simulate service conditions is
the lack of detailed documentation. This is caused by not
knowing the importance of such data in the corrosion of
ceramics, the cost of collecting the data, or both. Thus if one
wants to perform meaningful laboratory corrosion studies, it
is imperative that the industrial environment of interest be
accurately characterized.
When conducting laboratory oxidation studies, a convenient

way to obtain a range of oxygen partial pressures is desirable.
Very low partial pressures are never attained in practice by the
use of a vacuum system. Instead, a mixture of gases in which
oxygen is a component is used to establish the low partial
pressure. The most important mixtures that are used are
* One should also remember that a perfectly polished surface, although excellent
for reports, is not necessary to obtain sufficient data to solve a corrosion problem.
Copyright © 2004 by Marcel Dekker, Inc.
130 Chapter 3
CO
2
+CO and H
2
O+H
2
. Since the oxygen pressures are obtained
through the equilibrium reactions:
(3.1)
and
(3.2)
the partial pressure of oxygen is given by:
(3.3)
where k
1
and k
2
are the equilibrium reaction constants. For
constant ratios, the partial pressure of oxygen is independent of
the total pressure. Thus these gas mixtures provide a means to
obtain a range of oxygen pressures. Several techniques to mix

these gases are discussed by Macchesney and Rosenberg [3.6].
In the study of corrosion in coal gasification atmospheres,
gas mixtures such as CH
4
+H
2
and H
2
S+H
2
become important
along with the ones listed above. As the gas mixture becomes
more complex, the number of equations that must be solved
to obtain the equilibrium gas composition at elevated
temperatures and pressures also increases, making it convenient
to use a program such as SOLGASMIX [3.7] for the
calculations. One should not make the erroneous assumption
that gas mixtures are the same at all temperatures since the
equilibrium mixture is dependent upon the equilibrium
constant, which is temperature-dependent.
3.5 CHARACTERIZATION METHODS
3.5.1 Microstructure and Phase Analysis
Visual Observation
The most obvious method of analysis is that of visual
observation. The human eye is excellent at determining
differences between a used and an unused ceramic. Such things
Copyright © 2004 by Marcel Dekker, Inc.

×