Você está na página 1de 7

Materials Science and Engineering A 485 (2008) 415421

Processing and properties of ultra-high temperature


ceramics for space applications
F. Monteverde a , A. Bellosi a , Luigi Scatteia b,
a

ISTEC-CNR-Institute of Science and Technology for Ceramics, National Research Council,


Via Granarolo 64, 48018 Faenza, Italy
b CIRA-Centro Italiano Ricerche Aerospaziali, Via Maiorise, 81043 Capua, CE, Italy

Received 27 February 2007; received in revised form 1 August 2007; accepted 6 August 2007

Abstract
The processing and the properties of two ultra-high temperature ceramics (UHTCs) designed for the manufacturing of aerospace sharp-shaped
hot-structures are presented, along with the results obtained in the electrical discharge machining (EDM) of these UHTCs into sharp hot-structure
components.
The powder mixtures in the (ZrB2 SiC)-based systems were brought to full density by hot-pressing. The hot-pressed bodies were characterized
by fine and uniform microstructures (typical grain size < 5 m), along with controlled thermo-mechanical properties. Average flexure strength and

fracture toughness at room temperature (RT) up to about 900 MPa and 4.1 MPa m, respectively, were measured. Elastic moduli at RT approaching
500 GPa were also measured. At 1500 C in air, flexure strength decreased to 250 MPa.
Beyond basic mechanical properties, the machinability of UHTC blocks into a more complex shape by means of the EDM was also assessed.
Sharp-shaped hot-structures in the form of a nose-cone were produced from 12 cm 10 cm hot-pressed cylindrical blocks. The machined surfaces
showed limited roughness Ra < 1 m. With the exception of the typical size of critical flaws induced by EDM seeming slightly larger than those of
specimens machined with conventional diamond-loaded tools, the EDM technique proved to be as effective in machining UHTC sintered pieces
into more complex components.
2007 Elsevier B.V. All rights reserved.
Keywords: UHTC; ZrB2 ; HfB2 ; Hot-pressing; Electrical discharge machining

1. Introduction
Diborides of transition metals such as ZrB2 and HfB2
are commonly referred to as ultra-high temperature ceramics
(UHTCs) for their extremely high melting temperatures (ZrB2
3040 C, HfB2 3250 C), solid-state stability, good thermochemical and thermo-mechanical properties. These extremely
promising high performance materials are also characterized by
hardness above 20 GPa, high wear resistance, high emissivity,
high electrical conductivity (
=106 S/cm), excellent corrosion
resistance against molten iron and slags, and good thermal shock
resistance [1,2]. UHTCs are eligible for high temperature applications in several industrial sectors like foundries or refractory
industries. However, leading applications are currently found in

Corresponding author. Tel.: +39 0823 623576; fax: +39 0823 623515.
E-mail address: l.scatteia@cira.it (L. Scatteia).

0921-5093/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.08.054

aerospace, more specifically in the possibility to employ them


to realize sharp-shaped hot-structures like wing leading edges
and nose-caps able to withstand the severe thermal requirements
of next generation of hypersonic re-entry vehicles. The highly
thermal demanding trajectories foreseen for future spaceplanelike winged re-entry vehicles dictate the need for base materials
able to sustain operating temperatures approaching 2200 C, to
resist evaporation, erosion, and oxidation in the harsh re-entry
environment. When combined with SiC, ZrB2 , and HfB2 -based
composites exhibit indeed excellent refractoriness, high oxidation resistance [1,3,4], and are as such good potential candidates
for the above-mentioned application.
Nonetheless, the production of dense UHTCs pieces having
good thermo-mechanical properties has shown some difficulties.
Densification of UHTCs, primarily the case of metal carbide
like HfC or TaC, needs for instance sintering temperatures
even higher than 2300 C, due to their extremely high melting
point. Such conditions generally induce coarsening of the final

416

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

microstructures, and lead to non-negligible amounts of residual


porosity and decreased strength in the obtained compacts.
Different routes have been investigated in order to optimize
fabrication procedures and performances of these ceramics. One
approach relies on the use of sintering aids through conventional
techniques like hot-pressing (HP). A second foresees the use of
novel production methods such as the spark plasma sintering
(SPS), or the self-sustaining high temperature synthesis (SHS):
both can be used as reactive processing routes that allow for
the synthesis and further densification of UHTC compacts by
simultaneously applying an external pressure.
As for the approach dealing with sintering additives, some
studies on liquid phase sintering and on the incorporation of reinforcing phases are worth of mention. Indeed, recent works have
shown that high-density ZrB2 -based materials can be obtained
through liquid phase sintering at temperatures lower (about
1800 C) than those necessary for undoped compositions [58].
However, type and amount of grain boundary phases deriving
from the sintering aids often deteriorate the high temperature
properties, particularly when metal sintering aids are used [5].
Along the same lines, another processing variant to improve the
properties of monolithic diborides ceramics was the introduction
of SiC particulates, which proved to limit grain growth during
heat treatment and to enhance oxidation resistance [3,4,913].
Concerning the novel production methods, SPS is likely the
most promising technique, although the underlying mechanisms
that allow for such rapid processing times are still matter of open
discussions [1417]. However, shaping and maximum size of the
components, as well as costs and reproducibility of properties at
scales larger than that achieved in a laboratory, represent significant technological barriers. These challenges presently prevent
the acceptance of the SPS technique for industrial purposes.
An important parameter such as the upper limit of the operational temperature is strongly related to the characteristics of
grain boundary phases (that are in turn dependent upon the sintering aids if any) and of second phases. For example, above
1200 C the oxide scale formed in air on the surface of pure ZrB2
is unstable and non-protective due to intensive volatilization of
B2 O3 [1], while the SiC-containing MB2 , M = Zr or Hf, showed
enhanced resistance to oxidation up to 1600 C compared to the

pure MB2 . For temperature higher than 1300 C, the addition of


SiC promotes, on the exposed surface, the formation of borosilicate glass, which gives much more oxidation protection than
B2 O3 alone. Several studies have dealt with the thermal stability
and physical properties of ultra-refractory MB2 -based ceramics
in oxidizing environments [1830], and highlighted the role of
composition and microstructure on the mechanisms governing
the materials response to hostile environments.
In the present contribution, two ultra-high temperature
ceramics based upon ZrB2 SiC were produced and characterized in their microstructure and mechanical properties.
Moreover surface-modified microstructure and flexure strength
of specimens machined with EDM were analyzed and discussed.
Sharp-shaped hot-structures in the form of a nose-cone were then
manufactured.
2. Experimental
Two compositions (vol%) were selected.
Composition A: ZrB2 + 15 SiC
Composition B: ZrB2 + 15 SiC + 10 HfB2
On the basis of the rule-of-mixture, the expected bulk densities are 5.67 g/cm3 for composition A, and 6.17 g/cm3 for
composition B. The powder mixtures were wet-mixed in a nonaqueous medium using SiC milling balls, and were further dried
with a rotary evaporator under a continuous stream of inert gas.
Table 1 summarizes some characteristics of the raw powders
used. As sintering aid, 2 vol% MoSi2 was batched in both the
initial powder mixtures.
Hot-pressing was performed in vacuum (1 mbar) using an
induction-heated graphite die at the conditions shown in Table 2,
and about 20 C/min of heating rate. The shrinkage of the green
powder compacts was measured during the hot-press runs by
recording the displacement of the rams. The temperature was
measured by means of an optical pyrometer focused on the
graphite die.
The microstructures of both the materials were analyzed
with a scanning electron microscopy (SEM, Leica Cambridge

Table 1
Some characteristics of the raw powders: specific surface area (s.s.a, by BET) and particle size range (psr, by SEM)

ZrB2
HfB2
SiC
a

Company

Type

s.s.a. (m2 /g)

psr (m)

Main impuritiesa (wt%)

H.C. Starck, Germany


Cerac Inc., USA
H.C. Starck, Germany

Grade B
325 mesh
BF12

11.6

0.210
0.55
0.052

O 1, Hf 0.2
Zr 0.5
O 1.65

From suppliers.

Table 2
Base compositions, onset temperature TON , peak temperature TPEAK , applied pressure P, soaking time t, bulk density dB , and relative density rd
Label

Base composition (vol%)

Ta ON ( C)

TPEAK ( C)

P (MPa)

t (min)

dB g/cm3

rdb (%)

A
B

ZrB2 + 15SiC
ZrB2 + 15SiC+10HfB2

1350
1575

1820
19001940

35
40

7
45

5.61
6.06

99
98

a
b

TON : temperature at which measurable shrinkage takes place.


Based on the rule-of-mixture.

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

S360, UK) equipped with an energy dispersive analyzer (EDS,


INCA Energy 300, Oxford Instruments, UK), and an X-ray
diffractometer (Siemens D500, Germany). For microstructural
investigations, some specimens were polished with diamond
pastes down to 0.25 m.
Vickers microhardness (HV1.0) was measured on polished surfaces, with a load of 9.81 N, using a Zwick 3212
tester. Youngs modulus (E) was measured by the resonance
frequency method on 28 mm 8 mm 0.8 mm specimens
using gain-phase analyzer (mod. 4194A, Hewlett-Packard,
Yokogama, Japan). Fracture toughness (KIc ) was evaluated
using the chevron-notched beam (CNB) in flexure. Five bar,
25 mm 2 mm 2.5 mm (length width thickness, respectively), were notched with a 0.08 mm diamond saw; the
chevron-notch tip depth, and average side length were about
0.12 and 0.80 of the bar thickness, respectively. The slice
model equation of Munz et al. was used for the calculation
of KIc [31]. Flexure strength () at room temperature (5 bar)
and at 1500 C in air (3 bar) was measured on chamfered bars
25 mm 2.5 mm 2 mm (length width thickness, respectively), using a cross-head speed of 0.5 mm/min. For the
high-temperature tests, a soaking time of 18 min was
set to reach thermal equilibrium. This series of flexure
bars was machined with conventional diamond-loaded tools
(DLTs).
Another series of 5 bars (machined with DLTs) from both
the compositions was aged in stagnant air at 1600 C for four
consecutive heatingcooling cycles (5 min each) using a bottomloading furnace box. Once the bars have been slotted into the
hot furnace, the temperature of 1600 C was reached within
1 min. The retained flexure strength ( RET ) of the aged bars
was measured at room temperature.
Roughness was measured upon surfaces processed by DLTs
and EDM using a Taylor Hobson Talysurf Plus profiler (cut-off
0.8 mm, gaussian filter, band width 300). Typical parameters
such as Ra (arithmetric mean deviation of the assessed profile) and Rt (total height of the profile) were evaluated. Flexure
strength was also measured on 25 mm 2.5 mm 2 mm bar
prepared by EDM (3 bar) and further compared to that machined
with DLTs.

417

3. Results and discussion


3.1. Densification behaviour
The curves in Fig. 1 display different densification behaviours
associated to compositions A and B. Composition A densified
at a lower temperature with respect to composition B. In addition, the temperature TON (Table 2) indicates that the measurable
shrinkage of the powder mixture A started at about 1350 C,
while an excess of about 220 C was necessary before composition B begun shrinking appreciably. The SEM examination of
the microstructure did not give evidence of residual porosity in
any appreciable amount.
The need of a higher peak temperature (Fig. 1a) as well as
a longer isothermal hold (Fig. 1b) for composition B derived
from the presence in the starting mixture of a metal diboride like
HfB2 that is more refractory than ZrB2 . Once set the parameters
(associated to the isothermal hold of composition B) for the
excitation of the graphite die at 1900 C, a temperature drift to
about 1940 C was allowed in order to provide a controlled extra
heating to such (more refractory) powder compact.
The selection of MoSi2 as a ceramic additive for ZrB2 SiC
and ZrB2 HfB2 SiC compositions follows previous attempts on
MB2 -based composites containing MoSi2 as second phase, M
being Zr [3233], Hf [17], or Ti [34]: in all cases the addition
of MoSi2 proved to be effective for the densification of metal
diborides. MoSi2 powder particles are externally coated by a
silica layer [35], which can react with other oxides, for example B2 O3 , forming liquid phases. According to the B2 O3 SiO2
phase diagram [36], a liquid borosilicate phase forms at temperatures well below the sintering temperatures. Some hypotheses
can be put forward about the role of liquid phases for these systems. One possibility is that such a liquid phase facilitates the
removal of oxide species by wetting the surface of MB2 particles,
and thus reacting with them. These oxides, most likely B2 O3 and
MO2 [37], behave as obstacles to densification, because they
promote an anticipated coarsening of the diboride grains before
effective densification might take place.
The exact sequence through which such a liquid phase
would interact with the oxide impurities is, at the moment,

Fig. 1. (a) Heating-up stage: relative density (rd) vs. temperature (T) and (b) isothermal stage: densification rate vs. time (t).

418

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

Fig. 2. Polished cross-sections of the hot-pressed materials A and B (SEM micrographs): shell (S), core (C), and residual glass (G).

unclear. However, Fig. 2 shows evidence of discrete shells


around ZrB2 cores, enriched of Mo (composition A) or of
Hf/Mo (composition B). This result reinforces the existence of
an active role of MoSi2 through a dissolutionreprecipitation
mechanism that effectively promoted densification during hotpressing.
3.2. Microstructure
The fracture surfaces examined by SEM revealed regular
diborides grains, and a typical grain size range of 15 m for
material A, and 38 m for material B. In both the materials
fracture propagated with a predominantly intra-granular path.
The XRD analyses of the as-sintered material A identified as
crystalline phases, other than ZrB2 and SiC, only traces of monoclinic ZrO2 and MoSi2 . On the contrary, the comparison in Fig. 3
of the XRD patterns before and after hot-pressing of composition
B verified considerable alterations of the starting powder mixture: ZrB2 and HfB2 strongly interacted during hot-pressing and
gave rise to new ZrB2 /HfB2 solid solution(s), hereafter indicated

Fig. 3. XRD patterns from the as-sintered material B, and its initial powder
mixture.

as (Zr, Hf)B2 . Such a solid solution kept the same crystalline


structure of the original diborides, but incorporated at the same
time Zr and Hf atoms.
The examination by SEM-EDS of the polished surfaces
(Fig. 2) revealed that in composite A the SiC particulates are
located intergranularly within the ZrB2 matrix, or in agglomerates up to 10 m in size. Such agglomerates sometimes include
silica-based glassy residues. Thin shells enriched of Mo, i.e
(Zr, Mo)B2 solid solutions, around ZrB2 cores were also seen
(Fig. 4). Some zirconia particles, originated during hot-pressing,
are also present. In composite B, the SEM-EDS verified the formation of (Zr, Hf)B2 solid solutions: beside SiC (the dark round
features in Fig. 2), a core-shell configuration of the diboride
matrix clearly appears. The core consists of ZrB2 , whilst the
shell is constituted by (Zr, Hf)B2 solid solution(s). Silica-based
glassy pockets were sometimes found in proximity of SiC
agglomerates. Isolated HfB2 particles were rarely found. Such a
complex configuration of the basic microstructure accounts for
the deviation of the relative density evaluated using the rule-ofmixture (97.7%) and that stated by SEM (100%). The formation
of (Zr, Hf)B2 solid solution(s) indicates that the two starting
diborides are mutually soluble at high temperatures, as already

Fig. 4. Polished cross-section of material A (SEM micrograph): shell (S) and


core (C) are indicated.

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

419

Table 3
Grain size range (gsr), Vickers micro-hardness HV1.0, Youngs modulus E, Poisson ratio , fracture toughness KIc , flexure strength , and retained strength after
aging RET
Sample

A
B

gsr (m)

15
38

a Repeated

HV1.0 (GPa)

17.7 0.4
18.2 0.5

E (GPa)

480 4
506 4

0.13
0.128

(MPa)

KIc (MPam1/2 )

RET a (MPa)

RT

1500 C

RT

1500 C

4.07 0.03
4.08 0.75

2.53 0.23
3.43 0.02

887 125
763 73

255 25
240 20

564 67
631 69

exposures at 1600 C in air, 5 min each.

reported [38]. A similar behaviour has also been observed for


the TiB2 ZrB2 [39,40].
3.3. Mechanical properties
The values of the mechanical properties are summarized
in Table 3. Micro-hardness of the two composites is similar
(both about 18 GPa) and close to values reported in literature
for ZrB2 and HfB2 -based materials [1,7,912,15,16,18]. The
Youngs modulus values which range from 480 MPa (material
A) to 506 MPa (material B) are quite high: this result, given the
Youngs modulus high dependence upon residual porosity, further confirms the quality of the densification process realized in
this research work.
Both the compositions A and B have the same room temperature mean fracture toughness (KIc ) 4.1 MPa m1/2 : this value is
close to those reported in literature for MB2 -based composites
containing SiC as second phase [3,7,912]. At 1500 C under
air, the composite B showed a value of KIc higher than that of
the composite A (Table 3).
The room temperature flexure strengths, 887 125 MPa
(composition A) and 763 73 MPa (composition B), lie
within the range of values reported for similar highstrength ceramics in the MB2 SiC system, i.e 7001000 MPa
[1,7,912,18,26]. These flexure bars, machined with diamondloaded tools (DLTs), have the following roughness parameters:

Ra = 0.140 0.001 m and Rt = 1.33 0.10 m. The mean


flexure strength values measured at 1500 C in air, 255 25 MPa
(composition A), and 240 20 MPa (composition B), showed
an appreciable decrease compared to the room temperature values. Considering however the standard deviations, data obtained
at 1500 C for the two composites are not statistically different,
having at the same time reduced scattering.
The strength retention at 1500 C benefited to some extent by
the repair of surface defects from silica glass forming during the
test. This was confirmed by the fractographic studies on the (fractured) specimens tested in air [11,32]. At this temperature, the
formation of silica is consequence of the SiC oxidation according to the reaction SiC + (3/2)O2 SiO2 + CO. The presence of
SiC has proven to limit the surface degradation due to oxidation,
as after all previously demonstrated [1,3,4,13,1719,2126,29].
The other oxidation products of the composites A and B are
MO2 , M = Zr or Zr/Hf, respectively, that originated from the
reaction of MB2 with oxygen. Thanks to the protective action
of the surface borosilicate layer, the inner diffusion of oxygen
is greatly hindered.
Concerning the flexure strength values RET after cycled
treatments at 1600 C in air, 564 67 MPa (composite A),
and 631 69 MPa (composite B), the second retained its initial strength (i.e 82.7%) better than composite A (i.e 63.6%),
thanks to the formation of more compact external oxide scale
(Fig. 5).

Fig. 5. Polished cross-section of materials A and B after aging (SEM micrographs): external silica glass (1), oxide sub-scale (2), and un-changed volume (3).

420

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

3.4. Production of nose-cap prototypes


Nose-cap ceramic prototypes made of compositions A and
B were manufactured from the hot-pressed cylinders (10 cm
wide and 12 cm high). The final shape (Fig. 6) was obtained
by EDM. Some characteristic parameters of the surface roughness are reported in Table 4. On the average, the surface of the
prototype B has roughness slightly higher than that of the prototype A: Ra is about 0.6 m for material A and 1 m for material
B, whereas Rt is about 4.9 m, and 7.5 m for material A and
B, respectively. Examples of the surfaces machined with EDM
are shown in Fig. 7. The discrepancies are related to differing
surface damage introduced by the EDM process on the surfaces
of the two materials. The surface modifications seem more pronounced in material B due to its higher refractoriness, which
tends to retard the removal of matter through local melting phenomena. The SEM-EDS analyses upon the electrical discharged
surfaces confirmed the formation of a partially oxidized thin
layer, which in turn includes contaminants like Cu and Zn from
the electrically active tools used for EDM (Fig. 8).
Flexure strength measured upon 25 mm 2.5 mm
2.0 mm bar machined with EDM showed a decreased mean
value of about 26.4%, compared to that obtained from specimens machined with DLTs (Table 4). The major reason of such
a decrease lies in the micro-cracked surface layer induced by the
action of the electrical discharges. This surface micro-cracking
(Fig. 8) acts as a new population of critical defects that most
likely are characterized by sizes larger than those of specimens
prepared by DLTs. To account for such anincrease of the
critical flaw size (c), the Irwing relationship c = KIc /Y was

Fig. 6. Ceramic prototype A of a nose-cone obtained by hot-pressing and further


shaped through the EDM technique.

Table 4
Roughness parameters (Ra and Rt), flexure strength () of specimens machined with diamond-loaded tools (DLTs) or electrical discharge (EDM)
Sample

Composition

Machining

Raa (m)

Flexure bar
Flexure bar
Flexure bar
Prototype
Prototype

A
A
B
A
B

DLT
EDM
DLT
EDM
EDM

0.140
0.58
0.140
0.58
0.98

0.001
0.03
0.001
0.03
0.03

Rta (m)
1.33
1.33
1.33
4.93
7.44

0.10
0.10
0.10
0.46
0.46

(MPa)

887 125
653 80
763 73

Mean 1S.D.

Fig. 7. Electrical discharge machined surfaces of materials A and B (SEM micrographs): micro-cracking (white arrows) and re-solidified features (black arrows)
upon melting are indicated.

F. Monteverde et al. / Materials Science and Engineering A 485 (2008) 415421

Fig. 8. Flexure bar of material A machined with EDM (polished section, SEM
micrograph): micro-cracking (black arrows) and surface oxidized thin layer
(white arrows) are indicated.

used, being Y = 1.3. Values of 16 m (DLTs) and 29 m (EDM)


were estimated for the critical flaw size that induced failure at
room temperature during loading.
In spite of those minor hindrances, whose effects can be taken
into account and controlled in the design of a thermal structure,
the employment of the EDM technique has proven to be effective for shaping a sintered ceramic piece into a more complex
medium-scaled component.
4. Conclusions
Two ultra-high temperature ceramics-based on the
ZrB2 15 vol% SiC system were densified to full density
by hot-pressing. The peak temperature during hot-pressing
for the more refractory composition (that included 10 vol%
HfB2 ) did not exceed 1940 C. The two materials showed
uniform microstructures, controlled size dispersion and average
grain sizes below 5 m. MoSi2 (2 vol%) added as sintering
aid facilitated densification through dissolution-reprecipitation
mechanism.
Fine mean grain size and controlled population of defects
were the causes of flexure strength at room temperature (RT)
ranging from 760 MPa to about 900 MPa. At 1500 C (in air),
flexure strength decreased
to 250 MPa in both the compositions.
A mean value of 4.1 MPa
m at RT in
both the compositions

decreased to 3.4 MPa m and 2.5 MPa m at 1500 C in the


composition with or without HfB2 , respectively. The electric
discharge machining was successfully applied for the realization
of medium-sized massive nose-caps with a sharp profile.
Acknowledgements
The authors wish to thank the S. Guicciardi and C. Melandri
(ISTEC-CNR) for mechanical testing.
References
[1] W.G. Fahrenholtz, G.E. Hilmas, I.G. Talmy, J.A. Zaykoski, J. Am. Ceram.
Soc. 90 (5) (2007) 13471364.
[2] C. Mroz, Am. Ceram. Soc. Bull. 73 (6) (1994) 141142.
[3] S.R. Levine, E.J. Opila, M.C. Halbig, J.D. Kiser, M. Singh, J.A. Salem, J.
Eur. Ceram. Soc. 22 (2002) 27572767.

421

[4] K. Upadhya, J.-M. Yang, W.P. Hoffman, Am. Ceram. Soc. Bull. 58 (1997)
5156.
[5] A. Bellosi, F. Monteverde, D. Dalle Fabbriche, C. Melandri, J. Mater. Proc.
Manuf. Sci. 9 (2000) 156170.
[6] F. Monteverde, A. Bellosi, Scripta Mater. 46 (2002) 223228.
[7] F. Monteverde, S. Guicciardi, A. Bellosi, Mater. Sci. Eng. A346 (2003)
310319.
[8] F. Monteverde, A. Bellosi, Adv. Eng. Mater. 5 (2003) 508512.
[9] F. Monteverde, A. Bellosi, Adv. Eng. Mater. 6 (2004) 331336.
[10] F. Monteverde, A. Bellosi, J. Mater. Res. 19 (12) (2004) 35763585.
[11] F. Monteverde, A. Bellosi, Solid State Sci. 7 (2005) 622630.
[12] A.L. Chamberlain, W.G. Fahrenholtz, G.E. Hilmas, J. Am. Ceram. Soc. 87
(2004) 11701172.
[13] W.C. Tripp, H.H. Davis, H.C. Graham, Ceram. Bull. 52 (1973) 612616.
[14] M. Nygren, Z. Shen, Silic. Ind. 69 (2004) 211218.
[15] V. Medri, F. Monteverde, A. Balbo, A. Bellosi, Adv. Eng. Mater. 7 (3)
(2005) 159163.
[16] A. Bellosi, F. Monteverde, D. Sciti, Int. J. Appl. Ceram. Technol. 3 (1)
(2006) 3240.
[17] D. Sciti, L. Silvestroni, A. Bellosi, J. Mater. Res. 21 (6) (2006) 14601466.
[18] M. Gasch, D. Ellerby, E. Irby, S. Beckman, M. Gusman, S. Johnson, J.
Mater. Sci. 39 (2004) 59255937.
[19] J. Bull, M.J. White, L. Kaufman, Ablation resistant zirconium and hafnium
ceramics, US Patent 5,750,450 (1998).
[20] D. Sciti, M. Brach, A. Bellosi, J. Mater. Res. 20 (4) (2005) 922930.
[21] F. Monteverde, Corros. Sci. 47 (2005) 20202030.
[22] J.D. Bull, D.J. Rasky, J.C. Karika, Stability characterization of diboride
composites under high velocity atmospheric flight conditions, in: 24th
International SAMPE Conference, 1992, pp. T1092T1106.
[23] M.M. Opeka, I.G. Talmy, E.J. Wuchina, J.A. Zaykoski, S.J. Causey, J. Eur.
Ceram. Soc. 19 (1999) 24052414.
[24] C.R. Wang, J.-M. Yang, W. Hoffman, Mater. Chem. Phys. 74 (2002)
272281.
[25] M.M. Opeka, I.G. Talmy, J.A. Zaykoski, J. Mater. Sci. 39 (2004)
58875904.
[26] F. Monteverde, J. Alloy Comp. 428 (12) (2007) 197205.
[27] A. Chamberlain, W. Fahrenholtz, G. Hilmas, D. Ellerby, Refract. Appl.
Trans. 1 (2) (2005) 18.
[28] A.G. Metcalfe, N.B. Elsner, D.T. Allen, E. Wuchina, M. Opeka, E. Opila,
Oxidation of hafnium diboride, in: Electrochemical Society Proceedings,
vol. 99-38, 1999, pp. 489501.
[29] F. Monteverde, A. Bellosi, J. Electr. Soc. 150 (2003), B552-B55930.
[30] S.S. Hwang, A.L. Vasiliev, N.P. Padture, Mater. Sci. Eng. A 464 (2007)
216224.
[31] D.G. Munz, J.L. Shannon Jr., R.T. Bubsey, Int. J. Fract. 16 (1980)
R137R141.
[32] D. Sciti, F. Monteverde, S. Guicciardi, G. Pezzotti, A. Bellosi, Mater. Sci.
Eng. A 434 (2006) 303309.
[33] A. Chamberlain, W.G. Fahrenholtz, G.E. Hilmas, Characterization of zirconium diboridemolybdenum disilicide ceramics, in: N.P. Bansal, J.P. Singh,
W.M. Kriven, H. Scheider (Eds.), Advances in Ceramic Matrix Composites IX, Ceramic Transactions, vol. 153, The American Ceramic Society,
Westerville, OH, 2003, pp. 299308.
[34] T.S.R.Ch. Murthy, B. Basu, R. Balasubramaniam, J. Am. Ceram. Soc. 89
(1) (2006) 131138.
[35] Y.-L. Jeng, E.J. Lavernia, J. Mater. Sci. 29 (1994) 25572560.
[36] E.M. Levin, C.R. Robbins, H.F. Mc Murdie, Phase diagram for
Ceramists, Suppl., The American Ceramic Society, Columbus, OH, 1969,
p. 98.
[37] S. Torizuka, K. sato, H. Nishio, T. Kishi, J. Am. Ceram. Soc. 78 (6) (1995)
16061610.
[38] A.E. McHale (Ed.), Phase Equilibria Diagrams, Phase Diagrams for
Ceramists, vol. X, National Institute of Standards and Technology/The
American Ceramic Society, 1994, pp. 163166.
[39] F. Monteverde, A. Bellosi, S. Guicciardi, J. Eur. Ceram. Soc. 22 (2002)
279288.
[40] M. Moriyama, H. Aoki, Y. Kobayashi, J. Ceram. Soc. Jpn. 106 (12) (1998)
11961200.

Você também pode gostar