throbber
Molten Glass Corrosion Resistance of Immersed Combustion-Heating
`Tube Materials in Soda-Lime-Silicate Glass
`
`S. Kamakshi Sundamn,* Jen-Yan Hsu,* and Robert F. Speyer*
`School of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332-0245
`
`The corrosion resistance of molybdenum, molybdenum
`disilicide, and a SiC,,,/AI,O, composite to molten soda-
`lime-silicate glass was studied. The ASTM-C621-84 cor-
`rosion test method was modified because of inherent
`inaccuracies in the method and Si attack of platinum cruci-
`bles. Specimen-glass interfacial regions were characterized
`using XRD, SEM, and EDS. After 48 h of exposure at
`1565"C, the half-down corrosion recessions of Mo, MoSi,,
`and SiC,,,/Al,O, were 0.11, 0.316, and 0.26 mm, respec-
`tively. Mo oxidized to form a MOO, surface scale which
`cracked, allowing glass seepage and further oxidation. Sili-
`con was leached out of MoSi, into the glass, leaving a Mo,Si,
`interface and particles of Mo near the interface. For the
`SiC,,,/AI,O, composite, bubbles observed at the interfacial
`regions formed from oxidation of Sic to form CO. Thermo-
`dynamic modeling corroborated
`these experimental
`observations.
`
`I. Introduction
`HE work discussed herein is part of a larger research thrust
`
`T to develop materials for immersed gas-fired radiant burner
`
`tubes for glass melters. In the evaluation of candidate materials,
`our research has adopted a parallel approach, separately evalu-
`ating molten glass corrosion, combustion gas corrosion, as well
`as using thermodynamic simulations for both processes. Gas
`corrosion investigations are reported elsewhere.'.' In the present
`paper, the molten glass corrosion behavior of candidate materi-
`als is presented.
`A selection of the candidate materials which were evaluated
`and the results of which are presented herein are the following:
`molybdenum, molybdenum disilicide, and SiC,,,/AI,O, com-
`posite (p refers to particulate). Molybdenum has been used as
`the electrode material for electric melting and boosting in glass
`melting tanks for many years.'., Molybdenum disilicide has
`demonstrated a self-protective mechanism of forming a protec-
`tive SiO, layer at the surface in an oxidizing atmosphere.' For
`temperatures below 600°C. oxidation of molybdenum disilicide
`forms solid MOO, and SiO, products on the surface. The forma-
`tion of MOO,,,, in microcracks extends them and ultimately the
`solid fragments into powder, referred to as MoSi, pest.6 In the
`case of SiC,,,/AI,O, composite, the combination of high ther-
`mal conductivity of SIC and the refractory properties of A1,0,
`were expected to make it potentially suitable for the combustion
`product side of the present application. The composite is fabri-
`cated by the Lanxide process:' directed oxidation of molten alu-
`minum infiltrating into a particulate Sic matrix. Fused-cast
`AZS (50.8 wt% alumina, 32.6 wt% zirconia, 14.9 wt% silica,
`plus impurity phases), used as molten glass container material
`in this investigation, is well known for its corrosion resistance
`
`to molten glass"o and is a common commercial glass tank
`refractory.
`Bockris and co-workers" used molybdenum crucibles to
`study the interaction of Mo with molten binary silicates and
`detected a 0.15 wt% molybdenum absorption as MOO, by mol-
`ten Na,O-SiO, at 1750°C after 3 h in a H,-N,
`atmosphere.
`Minute reaction between pure silica and Mo under vacuum at
`1600°C has also been observed." Soda-lime-silicate glass, free
`of iron and sulfur, has been shown to be only very slightly cor-
`rosive toward molybdenum in the temperature range 1OOO" to
`1300°C in an oxidizing atrno~phere.'~ The slight observed cor-
`rosion was due to reaction with dissolved oxygen, postulated
`(though not shown experimentally) to form MoO,.I4
`The oxygen solubility of a glass melt is not governed entirely
`by the oxygen partial pressure of the furnace atmosphere; bas-
`i ~ i t y , ' ~ e.g., the concentration of alkali, and the concentration
`and type of polyvalent species present in the melti6." are also
`controlling factors. Based on solubility measurements of CO,,,,
`in glass, Pearce18 showed that oxygen ion activity increased
`dramatically with increasing alkali content in sodium silicate
`glasses. Because of its open structure, however, oxygen is pre-
`dominantly absorbed molecularly rather than atomically.
`Oxidation of molybdenum by dissolved oxygen in pure
`soda-lime-silicate glass melts was minute as compared to
`melts containing 0.1 wt% Fe and 0.3 wt% SO,. On reacting
`with Na,SO,, MoS, formed by
`~Mo,,, + 2Na,SO,,,, = MoS,,,, + 2Na,MoO,,,,
`The addition of 0.10% Fe,O, only (no sulfur compounds) to the
`soda-lime-silicate melt did not show any appreciable corro-
`sion.', Ooka" studied the effect of the amount of As,O,,
`Na,SO,, and Sb,O, present in soda-lime-silicate glass on the
`weight loss of molybdenum at 1500°C for 50 h. As,O, additions
`were found to be highly corrosive to Mo, Na,SO, was found to
`be slightly corrosive (forming MoS,), and Sb,O, had no effect.
`Lead-containing glasses were highly corrosive to Mo. Pecoraro
`el uLZo observed that the extent of corrosion by the mechanism
`of direct elemental solubility was negligibly small; corrosion
`proceeded primarily through the oxidation of the metal caused
`by oxidizing agents, dissolved oxygen and water, or a melt con-
`stituent which could accept electrons from the metal. Literature
`studies do not clearly indicate the form of corrosion product nor
`were they able to accurately measure corrosion rates.
`Muan and S p e d ' evaluated thermodynamic calculations for
`the systems Mo-0, Mo-Si-O, and Na,O-CaO-SiO, and sug-
`gested MoSi, as a potentially suitable material for gas boosters
`immersed in glass melting tanks. No experimental evidence of
`the corrosion rate or mechanism has been reported. The corro-
`sion resistance of SiC,,,/AI,O, composite in molten glass has
`also not been reported. However, the corrosion resistance of
`pure alumina has been studied extensivelyzz-26 and shown to be
`less than optimum.
`
`1. Smialek+ontributing editor
`
`Manuscript No. 194696. Received April 12,1993; approved February 16.1994.
`Supported by the Gas Research Institute under Contract No. 5090-298-2073.
`'Member. American Ceramic Society.
`
`11. Experimental Procedure
`
`( I ) Corrosion Testing
`The ASTM standard method C621-84 for isothermal corro-
`sion resistance of refractories to molten glass was initially
`adopted for the present investigation." This method involves
`1613
`
`GE-1029.001
`
`

`
`Journal of the American Cerurnic Society-Sundaram et al.
`
`Vol. 77, No. 6
`
`1614
`partial immersion of a test specimen of rectangular or circular
`cross section in molten glass contained in a Pt crucible. The
`recession from original material dimensions at the glass line
`and at an immersed position are measured after high-tempera-
`ture exposure for various periods of time. Modifications of the
`standard were made based on the following encumbrances: (i)
`molybdenum oxidizes and volatilizes above the glass line at
`temperatures exceeding -600°C, and (ii) silicon dissolves into
`the melt from the MoSi, and SiC,,,/AI,O, test coupons, and in
`turn alloys with the platinum crucible to form low-temperature
`eutectic phases, destroying the crucible.
`The Pt crucible as specified in ASTM C62 1-84 was replaced
`by an AZS (UNICORSOI, Corhart Refractories Corp.) crucible
`with a containment volume of 3.81-cm diameter by 3.81-cm
`depth to contain molten glass. The choice of AZS was made
`after static corrosion testing of fusion-cast AZS, MgO-partially-
`stabilized zirconia (PSZ) (Coors Ceramics Co., Golden, CO),
`fusion-cast chromia (Carborundum Co., Monofrax Refractories
`Division, Falconer, NY), and bonded chromia (Corhart Refrac-
`tories Corp., Buchannon, WV) by following the ASTM
`C62 1-84 standard. The results of corrosion testing, presented in
`Figs. I(a) and (b), indicate that AZS was the least corroded
`material at both the glass-line and half-down regions.
`
`Chromia (F)
`
`The soda-lime-silicate glass (Owens-Illinois, Inc.) used in
`the present investigation is a common container-glass composi-
`tion. A chemical analysis of the as-received glass in weight per-
`cent shows 73.3% SiO,, 13.3% Na,O, 11% CaO, 1.5% AI,O,
`0.3% K 2 0 , 0.2% MgO, 0.2% SO,, 0.2% ZrO,, and 0.07%
`FezO,. Molybdenum (Johnson Matthey/AESAR) specimens
`were of 12.5-mm diameter by 12.5-mm length for the purpose
`of complete immersion. Chemical analysis for trace elements in
`as-received molybdenum showed less than 1 ppm of Al, Ca, Cr,
`Cu, Mg, Mn, Ni, Pb, Si, Sn, and Ti, less than 2 ppm of C and 0,
`and less than 14 ppm of Fe. Cylindrical specimens of MoSi,
`(Kanthal Super 33, Kanthal Corp.) and SiC,,,/AI,O, (Du Pont
`Lanxide Composites) of 12.5-mm diameter by 60-mm length
`were used. The as-received MoSi, contained 1.7 vol% Mo,Si,
`grains, generally in contact with amorphous aluminosilicate
`glassy pockets comprising 18.6 vol% of the specimen, as deter-
`mined by scanning electron microscopy image analysis.
`The test configuration is sketched in Fig. 2. The quantity of
`glass powder put into the AZS crucible was adjusted so that a
`molten glass level of 18.8 mm was maintained. The as-received
`glass was crushed to -40 mesh, and the AZS crucible with
`48.5 g of glass was placed inside an electrically heated furnace.
`The glass was heated to 1565°C and allowed to equilibrate at
`
`0.4
`
`T
`
`0
`
`10
`
`40
`
`50
`
`30
`20
`30
`20
`Time (hrs)
`Time (hrs)
`(a)
`(b)
`Fig. 1. Corrosion rate of refractory containment materials at (a) glass-line and (b) half-down regions. The error bars indicate the standard deviation
`in the measurement of specimen diameters before the corrosion tests. These out-of-roundness deviations represent the largest errors of all contribu-
`tions to error in the measurement.
`
`40
`
`50
`
`Fig. 2. Fully immersed static corrosion test (unit: mm).
`
`GE-1029.002
`
`

`
`June 1994
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials
`
`1615
`
`i
`
`Fig. 3. Correlation of measured diameters 20 and 2b to the true diameter 2r of the corroded specimen, for a known blade thickness T. Left: case
`where blade cuts through the center. Right: case where blade misses the center.
`
`that temperature for 2 h to obtain bubble-free glass. Test speci-
`mens were then loaded into the crucible.
`The diameters of the pre-exposed specimens at the glass-line
`as well as the half-down positions were measured using digital
`vernier calipers (Mitutoyo Corp., Model CD-8") with a preci-
`sion of 20.01 mm. Each specimen was rotated and the mea-
`surement was performed at 10 points of contact at the two
`positions. The standard deviation of measured diameters ranged
`from 0.0052 to 0.0255 mm. The planeness and parallelness of
`the surfaces of the specimens were checked using a machinist's
`square. If the faces were not truly flat and perpendicular to the
`cylindrical axis, they were ground using 600-p,m' polishing
`paper and checked again until they were flat. In the case of
`molybdenum specimens, the center was marked using a lathe
`(Rivett Lathe and Grinders Inc., Boston, MA) and a self-center-
`ing collet before the corrosion test (precision: 2 0.05 1 mm).
`MoSi, and SiC,,,/AI,O, specimens were affixed to zir-
`con support wafers with zircon cement and allowed to dry
`overnight. They were preheated to 1OOO"C and maintained at
`that temperature for 10 min, to avoid specimen thermal shock
`when transferred to the equilibrated molten glass. Each molyb-
`denum specimen in a Pt wire basket was suspended by Pt wire
`into the molten glass so that the specimen was completely
`immersed. These specimens were quickly immersed in the mol-
`ten glass at the test temperature; they were not preheated so as
`to avoid atmospheric oxidation. The durations of corrosion tests
`were 12, 24, and 48 h at 1565°C in static air. The test tempera-
`ture of 1565°C corresponds to the hot-spot temperature in a
`commercial glass tank. One specimen was tested for each
`experimental condition. After exposure, the specimens were
`removed from the melt and allowed to cool in ambient.
`The corroded specimen rods were cut lengthwise along the
`rod axis using a diamond wafering blade (Buehler, Lake Bluff,
`IL) with a thickness of 0.34 2 0.1 mm ( 2 0.1 mm positioning
`precision). Several initial steps were taken to ensure that the
`wafering blade propagated through the center of the specimen:
`For Mo after exposure, the top Mo surface was ground until a
`flat surface was exposed, while for other specimens, the uncor-
`roded top surface was already exposed. The specimens were
`then mounted on this surface using thermal glue (Buehler). A
`cut running parallel to the axis of each specimen was made. The
`edges along this side face were checked as to whether they were
`parallel to each other using a digital vernier calipers. If they
`were not parallel, the side face was repeatedly ground with
`uneven pressure, and remeasured, until they were parallel; this
`indicated that the side face was parallel with the cylinder axis.
`For MoSi, and SiC,,,/AI,O, specimens, the center of the top
`surface was marked by the intersections of drawn diameters,
`
`while for Mo, the center was previously marked (see above).
`The specimens were then mounted on the side face, and another
`cut was made through the marked center, down the center axes
`of the specimens.
`The diameters at the glass-line and half-down positions on
`the cut surfaces were measured using the platform and micro-
`scope of a micro-hardness tester (Model HMV 2000, Shimadzu
`Co., Kyoto, Japan) where the precision of diameter measure-
`ment was 20.01 mm. An oil of refractive index of 1.51 was
`used to cause the glass to vanish from view so that the interface
`could be clearly seen through the microscope.
`The ASTM C621-84 procedure does not account for the
`thickness of the cutting blade when evaluating cylindrical spec-
`imens; it specifies only to determine the average of the mea-
`sured diameters of the specimen halves. Figure 3 illustrates the
`effect of blade thickness on the corrosion measurement. If 2u
`and 2b are the measured diameters, and 2r the genuine diameter
`of the specimen, then r can be calculated:
`. , ]
`
`[(" + ;; - UZf +
`
`r = ?
`
`where T is the thickness of the blade. The calculated propagated
`error by using the equation of r ranged from 0.0003 to 0.007
`mm. The computational procedure is elaborated upon else-
`where.2R Since the calculated errors corresponded to less than
`the y-axis dimension of data markers, error bars for radius
`recession plots are not shown.
`The reaction of the AZS container with molten glass was sep-
`arately evaluated. A 48.5-g sample of glass was placed in an
`AZS container and maintained at 1565°C. After equilibration, a
`small quantity of the melt was removed from the center of the
`exposed melt surface using a Pt boat, and quenched after 3.6,
`12,24, 36, and 48 h of reaction with the container. These sam-
`ples were then chemically analyzed for SO,, A1,0,, and ZrO,.
`(2) Characterization
`The glass-specimen interface was characterized using scan-
`ning electron microscopy (SEM) (Model 18 10, AMRAY, Bed-
`ford, MA). Cross sections of specimens were mounted in epoxy
`and the surfaces were ground smooth using 120-, 240-, 320-,
`400-, and 600-grit abrasive papers and then polished with dia-
`mond paste (30, 15,6, and l p,m). The polished samples were
`etched in 1% HF for -30 s, rinsed with distilled water, and then
`cleaned in an ultrasonic cleaner using distilled water, and dried.
`The etched specimens were coated with conductive Au-Pd
`alloy. Point and line scan elemental compositions were deter-
`mined using energy dispersive spectroscopy (EDS) (Model
`
`GE-1029.003
`
`

`
`1616
`
`Journal of the American Ceramic Society-Sundaram et al.
`
`Vol. 77, No. 6
`
`1.41
`
`0.8
`
`0.6
`
`0.4
`
`0.2
`
`0
`
`10
`
`40
`
`50
`
`30 30
`
`20 20
`
`30
`20
`Time (brs)
`Time (hrs)
`(a)
`(b)
`Fig. 4. Specimen recessions at (a) glass-line and (b) half-down regions when exposed to soda-lime-silicate glass at 1565°C.
`
`
`
`0 0
`
`
`
`10 10
`
`
`
`40 40
`
`
`
`50 50
`
`TS 16-501 8, Princeton Gamma Technology, Princeton, NJ) to
`establish the chemistry at and across the interfacial regions.
`A 48-h corrosion test was repeated using a specimen with a
`,emicircular cross section where the flat interface was for the
`convenience of XRD analysis. Diffraction patterns were taken
`on the glass-coated surface and then repeated after grinding the
`flat surface with 600-pm abrasive paper for 10 min and clean-
`ing. This process was continued until the interface was exposed
`and then penetrated. The specimen was analyzed using a com-
`puter-interfaced (Model VAX- I 1/730, Digital Equipment Co.,
`Northboro, MA) X-ray (CuKa) powder diffractometer (Model
`12045 X-ray diffraction unit, Philips Electronic Instruments
`Co., Mount Vernon, NY) with a I-s time constant and 0.02” step
`4ze over a 28 range of 20-80”. Using a computer search pro-
`gram, the X-ray diffraction patterns were matched to JCPDS
`data and corresponding phases were identified. In the phase
`identification procedure (D5000, Alfred University X-ray Lab-
`oratory, Alfred, NY), two types of searches, LIST and SUB-
`TRACT, were used. The LIST search performed a forced match
`of the input data with the entire JCPDS powder diffraction file
`and produced a list of potential matches. The SUBTRACT
`search performed a Hanawalt search and subtracted a matching
`compound that met a figure-of-merit criterion.
`(3) Thermodynamic Analysis
`Thermodynamic analysis of Mo-O, M-Si-0,
`Si-C-0, and
`A 1 4 systems were performed to predict solid-molten glass
`interfacial chemistry. The standard-state free energies of reac-
`tion, forming MOO,, MOO ?, ”.”’ Mo,Si, and Mo,Si,, and
`CaMoO,” were computed and plotted against temperature
`(Ellingham diagrams). Thermodynamic data were obtained
`from the Facility for the Analysis of Chemical Thermodynam-
`ics (F*A*C*T).+ These thermodynamic plots represent equilib-
`ria among solids and molecular oxygen in their standard states,
`that is, pure oxygen at 1 atm and pure solids.
`Since oxygen dissolves predominantly molecularly
`in
`glass,i9 the chemical potential of 0, in molten glass is equal to
`that in the furnace (electrically heated) atmosphere, where
`po. = 0.2. The activities of oxygen in the glass and in the
`atmosphere, however, are not equal, since soluble gas is gener-
`ally considered as a dilute solution which is non-Raoultian.
`Norton” has determined that the ratio of the concentration of
`molecular oxygen in the glass relative to that in the furnace
`atmosphere at 1078°C is 0.01. The “coefficient of solubility”
`does not change appreciably with temperature.”.” Thus, the
`Henrian activity of molecular oxygen in glass was taken to be
`2 X lo-’. and Richardson lines corresponding to this oxygen
`activity are drawn on the Ellingham diagrams.
`
`_ _ _ _ _ _ _ ~
`’Accessed through time-sharing with McGill University. Montreal. Canada
`
`111. Results and Discussion
`The chemical analyses of the glass (removed from the top-
`center of the melt) after interaction with AZS for varying times
`are shown in Table 1. No significant changes in these constit-
`uents were observed up to 12 h of exposure. Beyond 12 h, ALO,
`and ZrO, concentration gradually increased. After 48 h of expo-
`sure, the A120, content increased to 2.8%. and the ZrO, content
`to 0.15%. The effect of this compositional change on the speci-
`men corrosion rate is not known. Replenishing glass melt after
`every 12 h of exposure was considered and rejected since it
`changes the present static testing to a semidynamic corrosion
`test.2u
`Figures 1 (a) and (b) show the glass-line and half-down reces-
`sions, respectively, of potential containment materials. In the
`case of the glass-line corrosion, AZS was the least corroded
`material. In the case of the half-down region, initially fusion-
`cast chromia showed a corrosion rate less than that of AZS, but
`increased steeply beyond 12 h of exposure. Chromia also
`severely discolored the glass. The half-down corrosion rates of
`bonded chromia and PSZ could not be measured because of
`specimen cracking.
`Figures 4(a) and (b) show the glass-line and half-down reces-
`sions of cylindrical specimens (in turn contained in AZS cruci-
`bles), respectively, by molten glass. In the case of glass-line
`corrosion, SiC,,,/AI,O, showed a corrosion rate initially less
`than that of MoSiz but showed greater recession rates after 12 h
`of exposure, and was the most corroded specimen after 48 h of
`exposure. With regard to half-down corrosion, Mo was the least
`corroded material. MoSiz and SiC,,,/AI,O, showed similar cor-
`rosion behavior. The indicated downward trend for SiC,,,/AI,O,
`could not be corroborated by repeated testing because of speci-
`men availability.
`( I ) Molybdenum
`Figure 5 shows XRD patterns starting from the glass side,
`penetrating through the interface. On the glass side, only one
`crystalline peak is apparent, which matches with CaMoO,. This
`
`Table I. Chemical Analysis of Glass after Interactions with
`AZS Crucibles at 1565°C
`Composition ( ~ 1 % )
`A I J h
`I .5
`1.4
`1.4
`1.4
`2.6
`3.2
`4.3
`
`~~
`
`~
`
`Corrosion lime ( h )
`0
`3
`6
`12
`24
`36
`48
`
`SiO,
`73.3
`72.6
`72.8
`73.5
`73.3
`73.2
`73.1
`
`Zr02
`0
`0
`0
`0
`0.06
`0.1
`0.15
`
`GE-1029.004
`
`

`
`June 1994
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials
`
`1617
`
`500 I
`
`1
`
`I
`
`I
`
`I
`
`I
`
`Y 20
`
`40
`
`60
`
`Two Theta
`
`80
`
`XRD patterns penetrating from the glass, through the interface. into the molybdenum bulk of a specimen after an exposure of48 h at 1565°C:
`Fig. 5.
`a. MOO,: c. CaMoO,; m. Mo; p. Pt: u. unidentified phase.
`
`peak gradually reduces, then increases in intensity with penetra-
`tion. Low-intensity peaks corresponding to this phase are still
`apparent within the interfacial region. The trace corresponding
`to deepest penetration indicates the presence of MOO, and Mo,
`with minor quantities of CaMoO,. The increase and decrease in
`the intensity CaMoO, peak are attributed to the discontinuous
`composition of the debns in the glass near the interfacial layer.
`A few small peaks (marked “u” in Fig. 5) could not be matched
`with any known phases in the JCPDS database. Pt XRD peaks
`correspond to the Pt wire used in suspending the specimen in
`the glass melt.
`Figure 6 shows the Mo-glass
`interfacial morphology. An
`interfacial reaction layer is apparent. From the backscattered
`image (right-hand side), atomic number contrast of the interfa-
`cial layer implies that it was composed of (on average) atomi-
`cally lighter elements. This was corroborated by the line EDS
`scan shown in Fig. 7 in which a Mo-depleted (relative to Mo
`bulk) interfacial region is shown. This corresponds to XRD
`results showing MOO? at penetration depths corresponding to
`the highest trace in Fig. 5. It is therefore clear that the interfacial
`layer is Moo2.
`
`The sharp increase in Si intensity in Fig. 7 near the Mo bulk/
`interfacial layer interface implies the penetration of glass into
`the interfacial region. It is interpreted that oxidation of molyb-
`denum caused local expansion and associated compressive
`stress. These layers thus cracked, allowing penetration of the
`molten glass.
`Our XRD results showed that no MOO,,,, or MoS2,\, was
`detected within the interfacial region. The latter is in spite of the
`fact that the glass used in this investigation contained 0.2 wt%
`of SO,. Figure 8 shows the Ellingham diagram of various
`Mo-O,-Ca equilibria. MOO, is clearly favored over MOO,
`since the equilibria line falls above the Richardson line at the
`testing temperature. This is in agreement with XRD results at
`the interfacial region. Only near an atmospheric activity of sol-
`uble molecular oxygen would MOO, be expected to form.
`XRD results identified the formation of CaMoO, at the
`interfacial region and floating out into the glass. From Fig. 8,
`reaction between formed MOO,, calcium oxide, and dissolved
`oxygen is more favorable than reaction between elemental
`molybdenum, calcium oxide, and dissolved oxygen. This is
`
`Molybdenum-glass interfacial region after 48 h of corrosion (B = bulk, I = interface). Left: secondary electron image. Right: backscat-
`Fig. 6.
`tered image.
`
`GE-1029.005
`
`

`
`1618
`
`Journal of the American Ceramic Society-Sundaram et al.
`
`Vol. 77. No. 6
`
`Line EDS of the molybdenum-glass
`
`interfacial region in
`
`Fig.7.
`Fig. 6 .
`
`consistent with XRD results showing the presence of both Mo
`and CaMoO,. If CaMoO, formation using elemental Mo had
`been more favorable, then the formation of MOO? would not
`have been expected.
`(2) Molybdenum Disilicide
`The material used in this investigation is a cermet principally
`containing MoSi, with a minor aluminosilicate amorphous
`phase, and a fused silica surface layer of - 10 pm, as shown in
`phase, regions of Mo,Si, in contact with the ahminosilicate
`Fig. 9. After pre-heat treatment at 1000°C for 10 min, the sur-
`face layer became somewhat thinner and discontinuous,
`implying the onset of minute pesting. The presence of this sur-
`face layer on all specimens is expected to offset the absolute
`values of recession by 0.01 mm, but not expected to affect the
`trend of recession of MoSi2. This thickness was subtracted from
`all calculated specimen recessions.
`
`Fig. 9. Secondary electron SEM micrograph of a cross section of the
`surface of as-received MoSi, showing a -7-km fused silica coating:
`s, silica-rich vitreous protective coating; m, MoSi,; f. Mo,Si,; a, alumi-
`nosilicate glass.
`
`Figure 10 shows successive XRD patterns starting from the
`glass, penetrating through the interface. On the glass side, sev-
`eral peaks were identified as a-quartz. This phase was not
`detected in the interfacial region. On approaching the interfacial
`region, first molybdenum was identified. The interfacial region
`was found to be a mixture of primarily Mo,Si, with minor quan-
`tities of Mo,Si and Mo,O?,.
`
`100 I
`
`I
`
`-700 '
`
`0
`
`I
`I
`2000
`1000
`Temperature (K)
`
`I
`3000
`
`Fig. 8.
`Ellingham diagram of oxidation equilibria. A Richardson line of 2 X 10.' represents the molecular oxygen activity in soda-lime-silicate
`glass. The dotted lines indicate extrapolations. The dot-dashed line indicates the testing temperature ( I838 K).
`
`GE-1029.006
`
`

`
`June 1994
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials
`
`1619
`
`20
`
`40
`
`60
`
`80
`
`Two Theta
`
`Fig. 10. XRD patterns as a function of penetration, analyzing glass,
`through the interfacial region. to the molybdenum disilicide bulk: m.
`molybdenum; q, a-quartz; 1, MoSi,; 2, Mo,Si,; 3, Mo,02,; and u,
`unidentified phase.
`
`.,
`0
`
`1
`
`3
`2
`Energy (KeV)
`
`4
`
`5
`
`Fig. 11. EDS patterns of flat-faced MoSi, specimen corresponding to
`the XRD pattern second from the top in Fig. 10. The spectra marked
`“bulk” is from an as-received specimen.
`
`Molybdenum disilicide-glass interfacial region at (a) glass-line and (b) half-down regions after 48 h of corrosion: B, bulk; I, interface; and
`Fig. 12.
`G, glass. Left: secondary electron image. Right: backscattered image.
`
`GE-1029.007
`
`

`
`1620
`
`Journal of the American Ceramic Society-Sundaram et al.
`
`Vol. 77, No. 6
`
`(a)
`(b)
`Fig. 13. Line EDS of molybdenum disilicide-glass interfacial region at (a) glass-line, (b) half-down regions.
`
`100
`
`0
`
`-1 00
`
`-200
`
`.....2
`
`p
`a3 2 -300
`:$
`es
`2
`; -600
`P
`8
`
`-400
`
`-500
`
`-700
`
`-800
`
`-900
`0
`
`2000
`1000
`Temperature (K)
`
`3000
`
`Fig. 14. Ellingham diagram of compounds of interest in MoSi+xygen
`
`interaction.
`
`Figure 1 I shows an SEM EDS pattern of the outer interface
`region from which the XRD second from the highest in Fig. 10
`was taken. This figure shows appreciable molybdenum
`enhancement relative to silicon, corresponding to an XRD pat-
`tern showing predominantly Mo,Si,.
`interfacial fea-
`Figures 12(a) and (b) show the MoSi,-glass
`tures at the glass-line and half-down regions, respectively.
`From the backscattered images, the interfacial region was richer
`in high atomic weight elements (relative to the bulk). Since
`oxygen is lighter than both metals, the interfacial region is not
`an oxide such as MOO,, or SO,. Mo being heavier than Si, the
`interfacial region is interpreted to be Mo-rich. Further corrobo-
`ration is provided by EDS line scans (Figs. 13(a) and (b)). At
`the interface, the Si concentration decreased sharply. The fluc-
`tuations in Si and Mo concentrations in the MoSi, bulk were
`attributed to the amorphous aluminosilicate phase (dark second
`phase in Fig. 12). The interfacial layer had a thickness of -35
`pm at the half-down region, and -25 pm at the glass-line
`region. A small Mo EDS peak suggests a limited molybdenum
`debris stream into the glass.
`
`XRD, SEM, and EDS results provide irrefutable evidence
`that the mechanism of MoSi, corrosion is removal of silicon
`from MoSi, to the glass, leaving a silicon-deficient molybde-
`num silicide. This can be visualized as an oxidation reaction
`where oxygen was provided from dissolved molecular oxygen:
`SMoSi, + 70, = Mo,Si, + 7Si0,.
`Figure 14 is an Ellingham diagram for equilibria of interest
`among Mo, Si, and 0,. The formation of silica from MoSi, via
`the most oxygen conservative mechanism is favored. Thus sil-
`ica is formed by removing silicon from MoSi, to form Mo,Si,,
`rather than complete oxidation of the silicon component, the
`molybdenum component, or both. In addition, the most favored
`reaction is that most conservative in silicon removal from the
`compound; e.g., Mo,Si, forms in favor of Mo,Si or Mo. If no
`MoSi, was locally available, then it would be thermodynami-
`cally favorable for Mo,Si, to convert to Mo,Si, and by the same
`argument, for Mo,Si to convert to Mo.
`These predictions are consistent with our experimental
`results where at the interface, Mo,Si, forms along with a minor
`amount of Mo,Si. The Mo,Si, at the interface would not be
`
`GE-1029.008
`
`

`
`June 1994
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials
`
`1621
`
`expected to convert to Mo,Si or Mo since MoSiz is locally
`available which has a greater affinity for dissolved oxygen. A
`volume change would be associated with silicon removal from
`MoSi,. Thus Mo,Si, surface layers may dislodge from the inter-
`face and float away as debris. No longer near regions of MoSi,,
`local dissolved oxygen converts Mo,Si, to Mo,Si, and Mo,Si to
`Mo. The latter was observed by XRD in regions in the glass
`extended away from the interface.
`After an appreciable silicon-deficient interfacial layer has
`formed on the specimen, the diffusion distance for silicon from
`layer, to the surface
`the MoSi, bulk, through the Mo,Si,
`becomes arduous, and the rate of further silica formation is
`decreased. This appears to be the justification for the decrease
`in corrosion rate for the half-down tests, and the apparent termi-
`nation of recession after I2 h for the glass-line tests. This flat-
`tening of the corrosion rate may not be found in a commercial
`glass tank where greater convective flow forces are expected,
`which may periodically break loose and sweep away portions of
`the interfacial layer.
`(3) SiC,,,IA~,O,
`Figures I S(a) and (b) show the SiC,,,/Al,O,-glass interfacial
`features at the glass-line and half-down regions, respectively.
`Bubbles formed within the glass, near the composite surface at
`both regions. Near the glass line, the bubbles are seen in contact
`
`with the bulk of the specimen. CO or CO, can be the product of
`reaction between S i c and oxygen.”
`Competing oxidation reactions are considered in Fig. 16. The
`Sic portion of the composite oxidizes most favorably to form
`silica and carbon monoxide. If the carbon monoxide diffuses
`away from the interface, it is favorable for it to react with dis-
`solved oxygen to form CO,. Thus, the observed bubbles are
`anticipated to be a mixture of the two gases, of unknown ratio.
`The presence of bubbles in contact with a refractory in mol-
`ten glass has been demonstrated to be a mechanism of highly
`accelerated corrosion referred to as upward or downward
`“drilling,”34 where convection driven by surface tension gradi-
`ents sweeps fresh glass to the interface and corrosion products
`away. Thus, fresh glass with a higher oxygen activity would
`replace more viscous silica-rich glass. This corresponds well to
`the observed high corrosion rates for these specimens.
`(4) Remarks
`It is interesting to note that the MOO, interfacial layer formed
`in Mo specimens (-IS km) was much less

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket