Sie sind auf Seite 1von 8

Materials Science and Engineering A261 (1999) 16 23

Boron-doped molybdenum silicides for structural applications


Mut Akinc *, Mitchell K. Meyer 1, Matthew J. Kramer, Andrew J. Thom, Jesse J. Huebsch 2, Bruce Cook
Ames Laboratory and Department of Materials Science and Engineering, Iowa State Uni6ersity, Ames, IA 50011, USA

Abstract The addition of as little as 1 wt.% ( =3 at.%) boron improved the oxidation resistance of Mo5Si3 by as much as ve orders of magnitude over a temperature range of 8001500C. The mechanism of oxidation protection is the formation of a borosilicate glass scale that ows to form a passivating layer over the base intermetallic. The compositional homogeneity range for T1 (Mo5Si3Bx) phase was determined to be much smaller than that reported previously by Nowotny. Compressive creep measurements show that materials based on the phase assemblage of T1-T2 (Mo5SiB2) Mo3Si have high creep strengths similar to single phase Mo5Si3. Electrical resistivity of selected compositions was also measured and varied from :0.06 mV-cm at room temperature to 0.14 mV-cm at 1500C. Temperature coefcient of resistivity (TCR) was estimated to be on the order of 1 10 4 C 1 for most compositions. 1999 Elsevier Science S.A. All rights reserved.
Keywords: Molybdenum silicides; Oxidation protection; Creep strength

1. Introduction In order to meet an ever-increasing demand on the operating temperature capability of structural components, a considerable amount of research has been focused on developing new materials. In particular, ceramic matrix composites and intermetallics have received considerable attention. Recently, a large body of research has centered on nickel aluminide and titanium aluminide. These materials exhibit signicant room temperature ductility compared to other intermetallics. Unfortunately, their operating temperature is typically limited to less than 1000C. The end result is that these materials provide little material property advantage over existing nickelbased superalloys. MoSi2 has also been studied extensively and has been found to have several signicant applications. As a number of other articles in this volume allude to, MoSi2 possesses a number of unique properties, which makes

* Corresponding author. 1 Present address: Argonne National Laboratory West, Idaho Falls, ID 83408, USA. 2 Present address: Seagate Technology, Bloomington, MN 55435, USA.

it an attractive high temperature material. In particular, excellent oxidation resistance up to 1700C and relatively easy processibility makes this material an attractive candidate for structural applications. On the other hand, low creep strength at T\ 1000C, pest oxidation at moderate temperatures, and poor low temperature fracture toughness are severe limitations and are the subject of signicant current research activity. Development of single-phase ceramic or intermetallic systems capable of meeting the demanding requirements for ultra-high temperature applications doesnt seem likely in the near future. Additionally, compositebased materials typically pose a large number of processing challenges which again hinder their near future development. Multiphase (in-situ composite) intermetallics based on MoSiB compositions around the T1 phase present some unique solutions to these problems. The present paper reviews the work conducted by the authors in this system. Most of the material discussed in the present paper is based on articles previously written by the authors. Experimental details have mostly been omitted, and readers should consult these articles for specic details. Where available, new data has been incorporated in the present paper to bring up to date the state of understanding of the MoSiB research effort at Ames Laboratory.

0921-5093/99/$ - see front matter 1999 Elsevier Science S.A. All rights reserved. PII: S 0 9 2 1 - 5 0 9 3 ( 9 8 ) 0 1 0 4 5 - 4

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

17

2. Mo5Si3-a structural intermetallic material Mo5Si3 is the highest melting compound in the MoSi binary phase diagram with a melting point of 2180C. In addition, it exhibits a small compositional homogeneity range. It crystallizes in tetragonal W5Si3type structure with no phase transformation from room temperature to its melting point, and it melts congruently. Compared to molybdenum disilicide, Mo5Si3 has a more complex structure with 4 formula units per unit cell, hence it is expected to have better creep resistance. In fact, compressive creep rate for this material was measured to be an order of magnitude lower than that of MoSi2 [1]. Mo5Si3 has been previously considered for high temperature structural applications. It was quickly abandoned for mainly two reasons, low oxidative stability at moderate to high temperatures, and microcracking presumably due to thermal expansion anisotropy. Bartlett, et al. [2], modeled the oxidation of MoSi2 and Mo5Si3 using a kinetic model based on the rate of dissociation of respective silicides to generate silicon and its consumption by the available oxygen at the reaction interface. Though the model ignores the microstructural aspects of the SiO2 scale formation, it predicts that Mo5Si3 will oxidize actively in air atmosphere unless the temperature is above 1650C. Thus, Mo5Si3 was considered intrinsically unstable during oxidation in air at temperatures below 1650C. These predictions were corroborated by experimental data at various oxygen partial pressures and temperatures by several studies including ones carried out in our laboratory. Cyclic oxidation by Anton and Shah [3], at 1149C also showed catastrophic failure in 20 1 h cycles, indicating that an adherent, passivating layer does not form at these temperatures.

Common features of oxidation plots for both Bdoped and pure Mo5Si3 samples include a small mass gain followed by an abrupt mass loss during initial heating to the test temperature. Catastrophic mass loss occurs for undoped samples whereas borondoped samples passivate and show little to no mass change over several hundred hours. Undoped samples exhibit active oxidation due to the formation of an initial porous oxide scale and its inability to subsequently form a protective scale. Small additions of boron to Mo5Si3 promote the formation of a non-porous, protective scale as shown in Fig. 2. The micrographs indicate that the oxide scale is several hundred micrometers thick for undoped Mo5Si3 (Fig. 2A), whereas the boron-doped Mo5Si3 (Fig. 2B) forms a continuous, non-porous scale that is B 10 mm thick. Obviously, the addition of boron modies the scale microstructure dramatically. Oxidation resistance provided by the boron addition of : 1.2 wt.% as previously mentioned is not limited to a narrow composition range. Several other MoSiB compositions around Mo5Si3 were also tested for oxidation resistance. These materials also exhibit excellent oxidation resistance as shown in Fig. 3. The magnitude of the initial weight loss is dependent on the molybdenum content of the sample and the B/Si ratio. Initial mass loss ranged from about 2 mg cm 2 for B/Si = 0.02 to about 11 mg cm 2 for B/Si =0.24. Again, all compositions show the similar behavior of a rapid mass loss followed by a region of near zero mass change. The isothermal portions of the oxidation curves at long times (several hundred hours) show some clear differences in oxidation rates. Compositions with low B/Si ratio or low (B + Si) content show a negative linear oxidation rate (kl = 3.6

3. MoSiB intermetallics as oxidation resistant structural materials In order to improve the oxidation behavior of Mo5Si3 by scale modication, the Mo Si B system was chosen for investigation [4,5]. Compositions around the Mo5Si3 compound requiring small boron additions were selected. These materials were exposed to a owing air atmosphere at elevated temperatures, and Fig. 1 shows the oxidation behavior of a typical boron-doped Mo5Si3 composition (16.1 Si, 1.24 B by wt.%). Mass change was measured as a function of time at temperatures ranging from 800 to 1500C. For comparison, the oxidation behavior of undoped Mo5Si3 over a temperature range of 800 1200C is also illustrated.

Fig. 1. Isothermal oxidation of undoped Mo5Si3 and boron-doped Mo5Si3 (Mo 16.1 Si 1.2 B) from 800 to 1500C in owing air. Dashed curves denote undoped Mo5Si3 while solid line curves denote boron-doped Mo5Si3.

18

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

4. Oxidation protection mechanism The mass gain followed by large mass loss during the initial heating to the test temperature can be attributed to oxidation of silicon and molybdenum, followed by sublimation of MoO3 A clearer understanding of transient state oxidation and the oxidation protection mechanism provided by boron addition was determined by characterizing the initial stages of oxidation for several boron-doped materials using SEM (scanning electron microscopy) and XRD (X-ray diffraction) [6]. Fig. 4 shows SEM surface micrographs of the development of the oxide scale on a specimen with a high B/Si ratio. The material is designated as V, and the composition is given in Fig. 3. The starting material consists of an Mo3Si matrix with both T1 and T2 as distributed phases. At 600C, the scale is composed of small pockets of borosilicate glass, molybdenum oxide grains, and a matrix consisting of mixed molybdenum and silicon oxides. There are no crystalline phases in the scale since the XRD pattern at 600C in Fig. 5 only reveals the intermetallic base material. At 633C, MoO3 crystals begin to grow over the scale surface, and growth continues up to 725C, as shown in Fig. 4. The vapor pressure of MoO3 increases signicantly at 750C, and MoO3 begins to sublime, evident in Fig. 4(B). By 775C, a signicant fraction of the MoO3 crystals has evaporated, leaving behind a surface containing borosilicate glass and a sizable fraction of porosity (Fig. 4C). At 1000C, Mo presence is suggested by the appearance of the Mo (100) diffraction maxima. Viscous ow and closure of submicron scale porosity occurs after holding isothermally for 20 min at 1000C. The scale looks like a glass surface after 40 min at 1000C (Fig. 4D), and there are no longer any signs of MoO3 formation. The enrichment of Mo at the oxidation interface is indicated by the Mo (100) peak that grows in intensity with time at 1000C relative to Mo5Si3, T2, and Mo3Si intermetallic peaks near 41 2U (Fig. 5). These observations suggest the selective oxidation of silicon over molybdenum during passivating scale growth. For all compositions tested, oxidation occurs in two stages, an initial transient period followed by steady state oxidation. For some compositions, a protective scale forms rapidly so that transient mass loss is relatively small and essentially completes by 850C. On the other hand, transient mass loss continues well into the isothermal temperature regime and a protective scale forms more slowly where Si+ B content is relatively low or B/Si ratio is low. There is a transition to a slower, steady state oxidation regime, regardless of the magnitude of the initial mass loss. This transition is the crucial step in

Fig. 2. Micrographs of scale cross-sections of samples oxidized in air. (A) Undoped Mo5Si3 oxidized at 1100C for 8 h. Note only small center fraction of the original silicide remains unoxidized. (B) B Mo5Si3 oxidized at 1300C for 120 h. External scale is B10 mm. For both micrographs, U, underlying base alloy, S, scale.

10 3 mg cm 2 h). Compositions with a higher B/Si ratio or higher (B+Si) content exhibit a near zero to positive linear oxidation rate (kl = + 7.3 10 4 mg cm 2 h).

Fig. 3. Isothermal oxidation of MoSiB compositions I V at 1000C. Silicon and boron content for each composition is indicated by its respective curve.

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

19

Fig. 4. SEM micrographs of the surface of composition V to demonstrate formation of the initial oxide scale. (A) 725C, (B) 750C, (C) 775C, and (D) 1000C for 40 min. Samples in AC were immediately quenched to room temperature after exposure to the test temperature.

formation of a protective surface layer. If the substrate surface is sealed by a continuous scale, oxygen transport through the scale to the reaction interface occurs by atomic scale diffusion. On the other hand, a microporous scale will force oxygen and oxidation products to diffuse through a torturous path to the substrate surface. The resulting passivating scale will be rate-limited by diffusion of oxygen and oxidation products through the porous scale. If oxidation is limited by atomic oxygen diffusion and the silicon activity at the oxidation interface is sufcient, oxygen partial pressure will be xed by the Si/SiO2 equilibrium [7]. Silicon dioxide has a much lower free energy of formation than any of the molybdenum oxide species and hence molybdenum will not oxidize. The dominant oxidation reaction may be expressed as: Mo5Si3 +3O2 5Mo +3SiO2 (1)

will thus form. It has been shown that SiO2 can exist in equilibrium with Mo and Mo5Si3 so that formation of a Mo3Si layer according to MoSi phase diagram is not thermodynamically necessary [8]. Eq. (1) predicts a net mass gain on oxidation. Fig. 4 describes a composition that exhibited an isothermal mass gain and formation of a molybdenum interlayer after heating to 1000C. For this composition, oxidation at 1000C is controlled by oxygen diffusion through a completely passivating scale. If oxidation is limited by diffusion of oxygen through a microporous layer, the partial pressure of oxygen at the interface is high enough to oxidize molybdenum. Oxidation will proceed according to: 2Mo5Si3 + 21O2 10MoO3 (volatile)+ 6SiO2 (2)

A silicon depleted and molybdenum rich interlayer

A mass loss is predicted by Eq. (2) since molybdenum oxide is volatile above : 750C. Also noteworthy is that a metallic molybdenum interlayer will not form. A slow steady state mass loss was observed for compo-

20

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

sitions with low Si+ B content or low B/Si ratio. No molybdenum interlayer was detected by XRD, and therefore Eq. (2) is the dominant oxidation reaction for these compositions at 1000C. A transition from mass loss to near zero mass change was observed for intermediate compositions after about 30 h of oxidation. In this case, Eq. (1) and Eq. (2) must be competing to yield an essentially zero net mass change. Another possibility is that micropores close off with time so that the extent of Eq. (2) decreases with time to become negligible. For the sake of simplicity, the preceding discussion neglects the presence of boron and/or other phases usually present such as T2 and Mo3Si. This assumption however, does not change the qualitative assessment of the oxidation mechanism proposed. It is important to consider the scale porosity and scale viscosity when describing the transition to steady state oxidation. If viscous ow does not occur to close pores on a reasonable time scale, initial scale porosity and pore size will affect the steady state oxidation rate. The steady state oxidation rate will be controlled by the diffusion rate of oxygen through the scale if viscous ow of the scale to form a coherent passivating layer does occur. Fig. 6 shows the variation of isothermal oxidation rate at 1000C with the estimated viscosity of the scale. The steady state mass loss rate does increase with increasing scale viscosity. The scale viscosity was calculated assuming that the B/Si ratio in the scale is identical to the initial B/Si ratio of the substrate [9].

Fig. 6. Variation of isothermal mass loss rate at 1000C with the calculated viscosity of the scale. The atomic B/Si ratio for each composition is given.

5. MoSiB ternary All of our oxidation tests were conducted on three phase samples. T1 was usually the matrix phase with T2, Mo3Si, MoSi2, or MoB as the distributed phases, depending on the overall composition point. Therefore, it was not possible to accurately assess the role that individual phases play in the oxidation behavior. Realizing the need to assess the oxidation behavior of single-phase T1, we began a concerted effort to synthesize phase pure T1 [10]. A number of compositions were prepared using the 1600C isothermal cut of the ternary phase diagram according to Nowotny shown in Fig. 7. Numerous preparations with careful consideration of the starting materials failed to produce single phase T1. The resulting multi-phase assemblage materials were analyzed by XRD and EMPA (electron micro probe analysis). These data suggested that either Nowotnys phase diagram was not accurate or that equilibrium was not achieved during annealing of the chosen points. However, careful heat treatment periods at 1800C and chemical analysis indicated that the single-phase homogeneity region for T1 is much smaller than that reported by Nowotny. Fig. 8 shows the narrow homogeneity range measured for T1 phase. During these studies, arc-cast samples were originally heat treated for 2 h at 1800C in an attempt to reach equilibrium. XRD patterns indicated that even this high temperature heat treatment was not sufcient for thermodynamic equilibrium. In order to reach the equilibrium

Fig. 5. XRD patterns from 600 to 1000C for composition V shown in Fig. 4. Dashed lines indicate MoO3. Arrows at top of plot give the 100% intensity peak for both molybdenum and cristobalite.

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

21

Fig. 9. Arrhenius plot of steady state creep rate with varying compressive loads for undoped and boron-doped Mo5Si3.

Fig. 7. Isothermal cut at 1600C of the MoSiB ternary phase diagram [14].

phase distribution, arc-melted samples were ground into powders, compacted, and sintered in an argon atmosphere for 48 h at 1800C. Although this heat treatment promoted the equilibrium condition, the vapor pressure of silicon at this temperature is signicant and shifts the initial composition to the molybdenum rich side. Nevertheless, the homogeneity range for the T1 phase was determined to be 61.5 62.9% Mo (at.%) and 1 to 2% B (at.%). The maximum solubility for B occurs at about 62.0% molybdenum content.

Preliminary oxidation testing of single phase T1 samples suggests that the material also has good oxidation resistance [11]. This is consistent with the previous oxidation data for which T1 containing materials exhibit excellent oxidation resistance. Obviously, the oxidation protection afforded to Mo5Si3 by boron doping is independent of the resulting phase assemblage in which it resides. Very small quantities of boron (B3 at.%) are sufcient to render the Mo5Si3 oxidatively stable.

6. Creep behavior As mentioned earlier, Mo5Si3 and boron-doped compositions around the T1 phase eld are expected to exhibit higher creep strength than other silicides due to its complex unit cell structure. For creep testing, both Mo5Si3 and a typical MoSiB sample were investigated [12]. Material was prepared by powder compaction and sintering. The undoped Mo5Si3 material is designated A, and the boron-doped sample of composition 13.0 Si1.3 B (wt.%) is designated B. Sample B possessed a three-phase microstructure with T1 as the matrix phase and T2 and cubic Mo3Si as the distributed phases. The sintered specimens were about 97% of theoretical density as determined by Archimedes method. The average grain size for all phases was about 4 mm. Compressive creep rates were determined from 1220 to 1320C and stress levels of 140180 MPa. At temperatures below 1200C, and loads less than 100 MPa, no measurable deformation was detected. Fig. 9 shows a reciprocal temperature plot of creep rate for each applied load for both compositions A and B. From the data in Fig. 9, apparent activation energies were obtained from a least squares t. The calculated activation energy for sample A was 399 kJ mol 1, and the activation energies for sample B ranged from 377 to

Fig. 8. Experimentally determined MoSiB ternary phase diagram at 1800C in the region of T1 phase.

22

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

412 kJ mol 1. Since the values for sample B and between samples A and B were similar, the same or similarly activated creep processes dominate the creep behavior of both A and B. Fig. 10 shows calculated creep stress exponents for B, along with the standard deviation associated with the calculation of each stress exponent. The stress exponent appears to decrease with increasing temperature from a value of 5.0 at 1242C to a value of 3.7 at 1302C. The errors associated with the calculation of the stress exponent are in the order of 0.50.7. Since the differences between stress exponents are in the order of this error, no conclusions about the change in creep mechanism with temperature can be drawn from this data. The average stress exponent for all conditions tested is 4.3. Stress exponents in this range often indicate creep by dislocation related processes. The creep rate of the undoped and boron-doped Mo5Si3 samples was about the same. A very low dislocation density was observed in the T1 phase during TEM examination of the crept material. Only a few dislocations were noted in the T2 phase, while a high dislocation density was observed for Mo3Si. At 5% strain levels, Mo3Si actively undergoes polygonization and shows dislocation climb. Cracks were seen in the T1 phase at higher strain levels (: 13%), although the specimen retained its integrity. The density of cracks was much higher in the T1 phase than in Mo3Si or T2. It is believed that cracking and sliding of the T1 phase is the mechanism by which deformation is accommodated at higher temperatures and strain rates. Some toughness enhancement for the composite microstructure may be afforded by Mo3Si because cracks appeared to be arrested at the Mo3Si phase.

Fig. 11. Temperature dependence of electrical resistivity for borondoped Mo5Si3 with commercial Kanthal Super MoSi2 heating element material for comparison.

7. Electrical resistivity One potential application for these materials is electrical furnace heating elements. This application is targeted for providing improved performance over existing MoSi2 heating elements in use today. In order to assess the utility of the material as potential heating elements, we measured the electrical resistivity of select compositions using the four point modied Van der Pauw method [13]. Electrical resistivity was measured in air atmosphere up to 1100C. Measurements up to 1500C, were conducted in an inert atmosphere. Thin oxide scales that formed in air above 1100C contributed to high resistance electrode contacts and erroneous measurements. Data taken in an inert atmosphere and in air atmosphere compare very well up to 1100C. Therefore, the inert atmosphere data up to 1500C are considered to be reliable and accurate. Fig. 11 shows the variation of electrical resistivity with temperature for select compositions. Superkanthal heating elements were also measured and are compared to the manufacturers published data (solid line). The data obtained in our laboratory for MoSi2 agrees well with the published data, and this lends strong support to the validity of the electrical resistivity measurements. The materials developed in our lab have a room temperature electrical resistivity of 0.060.08 mV-cm, depending on the specic composition. The electrical resistivity increases slightly up to 1500C. Even after several heating/cooling cycles, no hysteresis was observed. No drift in resistivity was observed as evidenced by heating a sample up to 1500C and holding at the temperature for 6 h. Mo5Si3Bx compositions have a relatively low temperature coefcient of resistivity (TCR), indicating that these materials are excellent candidates for electric fur-

Fig. 10. Creep rate versus stress plot for boron-doped Mo5Si3 showing the temperature dependence of the stress exponent. Uncertainty value for each estimate is given in parentheses.

M. Akinc et al. / Materials Science and Engineering A261 (1999) 1623

23

nace heating elements. Rigorous current control to maintain certain power levels upon heating may not be required.

8. Conclusions The oxidation behavior of Mo Si B compositions near the T1 (Mo5Si3Bx) phase eld was studied from 800 to 1500C. Steady state oxidation in a owing air atmosphere indicates that the addition of small amounts of boron dramatically increases the oxidation resistance of these materials. Steady state linear oxidation rates on the order of kl = + 10 4 mg cm 2 h were measured. Scale formation proceeds with an initial oxidation of the surface to form borosilicate glass and crystalline MoO3. Sublimation of MoO3 and reow of the porous borosilicate glass promotes formation of a passivating scale. Further oxidation is limited to diffusion of oxygen through this scale to the reaction interface. Boron doping of Mo5Si3 does not appear to decrease the creep strength of Mo5Si3. Compressive creep data at several temperature and stress levels yields a creep stress exponent in the range of n =3.8 5.0. The average activation energy was calculated to be Q =400 kJ mol 1. No dislocation motion was evident in the T1 phase. Some dislocation motion was observed in T2, and a much higher density of dislocations was observed in Mo3Si. Electrical resistivity measurements of several compositions indicated that these materials show electrical conductivity much like metals. Electrical resistivities on the order of 0.1 mV-cm were measured up to 1500C with a TCR in the order of 10 4C 1. It appears that boron-doped Mo5Si3 compositions offer promise for use in high temperature oxidative environments. Oxidative stability, compressive creep strength, and electrical resistivity characteristics suggest that these materials have signicant potential for structural applications. Several other engineering properties have not yet been studied including cyclic oxidation

and mechanical properties such as exural strength and fracture toughness. Different phase assemblages available around the T1 phase offer the possibility of designing microstructures to provide a combination of chemical, physical, and mechanical properties for a given application.

Acknowledgements Ames Laboratory is operated for the US Department of Energy by Iowa State University under contract number W-7405-ENG-82. This research was supported by the Ofce of Basic Energy Science, Materials Science Division. This work was also supported by the Ofce of Energy Research, Ofce of Computational and Technology Research, Advanced Energy Projects Division.

References
[1] D.L. Anton, D.M. Shah, Mater. Res. Soc. Symp. Proc. 213 (1991) 733. [2] R.W. Bartlett, J.W. McCamont, P.R. Gage, J. Am. Ceram. Soc. 4811 (2) (1965) 551. [3] D.L. Anton, D.M. Shah, Mater. Res. Soc. Symp. Proc. 213 (1991) 733. [4] M.K. Meyer, M. Akinc, J. Am. Ceram. Soc. 794 (4) (1996) 938. [5] M.K. Meyer, M. Akinc, J. Am. Ceram. Soc. 7910 (10) (1996) 2763. [6] M.K. Meyer, A.J. Thom, M. Akinc, Intermetallics, in press. [7] E. Fitzer, in: R.B. Tressler, M. McNallan (Eds.), Ceramic Transactions, Corrosive and Erosive Degradation of Ceramics, vol. 10, American Ceramic Society, Westerville, OH, 1989, p. 19. [8] R. Beyers, J. Appl. Phys. 56 (1) (1984) 147. [9] M.F. Yan, J.B. MacChesney, S.R. Nagel, W.W. Rhodes, J. Mater. Sci. 15 (1980) 1371. [10] J.J. Huebsch, M.J. Kramer, M. Akinc, Intermetallics (1998), submitted for publication. [11] J.J. Huebsch, M. Akinc, Intermetallics (1998), submitted for publication. [12] M.K. Meyer, M.J. Kramer, M. Akinc, Intermetallics 4 (1996) 273. [13] B. Cook, S. Zelle, M. Akinc, Ceram. Eng. Sci. Proc., in press. [14] H. Nowotny, E. Dimakipoulu, H. Kudiekla, Monatsh. Chem. 88 (1957) 180.

Das könnte Ihnen auch gefallen