Sie sind auf Seite 1von 7

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/227851774

Formation of Core/Rim Structures in Ti(C, N)


WCNi Cermets Via a Dissolution and
Precipitation Process
ARTICLE in JOURNAL OF THE AMERICAN CERAMIC SOCIETY MAY 2000
Impact Factor: 2.61 DOI: 10.1111/j.1151-2916.2000.tb01415.x

CITATIONS

DOWNLOADS

VIEWS

72

53

282

2 AUTHORS:
Sun-Yong Ahn

Shinhoo Kang

Korloy, Korea

Seoul National University

11 PUBLICATIONS 204 CITATIONS

128 PUBLICATIONS 1,349 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Sun-Yong Ahn


Retrieved on: 10 September 2015

J. Am. Ceram. Soc., 83 [6] 1489 94 (2000)

journal

Formation of Core/Rim Structures in Ti(C,N)-WC-Ni Cermets via a


Dissolution and Precipitation Process
Sun-Yong Ahn and Shinhoo Kang*
School of Materials Science and Engineering, Seoul National University, Seoul 151742, Korea
The core/rim structures of compounds in the Ti(C,N)-xWC20Ni system have been investigated to determine the effect of
WC and nitrogen content on the microstructure of the system.
In addition, the relative dissolution rate of WC to Ti(C,N) in
the system was studied by analyzing the rim compositions.
Variations in the WC content had a much-lesser influence, as
an additive, on the general microstructure than that of other
carbides that have been used in previous studies. However, the
nitrogen content in Ti(C1XNX) had a significant effect on its
microstructure. The composition of the rim structure was
determined by the ratio of solutes that were dissolved in the
liquid binder, under the given processing conditions. The
dissolution rate of WC was 2 and 5 times faster than that of
Ti(C,N) in the system at temperatures of 1300 and 1510C,
respectively. The results have been interpreted in terms of
phase stability and precipitation phenomena.
I.

microstructural changes of a Ti(C,N)-WC-Ni system as a function


of the nitrogen and WC content. Particular attention was directed
to the composition of the rim structure, which, to a large extent,
reflects the equilibrium reaction during liquid-phase sintering. The
results are related to the dissolutionprecipitation phenomena.

II.

Experimental Procedure

The particle sizes, stoichiometry, and manufacturers of the


powders used in this study are listed in Table I. The carbon,
nitrogen, and oxygen contents of the powders were determined
using a carbon/nitrogen analyzer (Leco, St. Joseph, MI). The
samples were prepared using a conventional powder metallurgy
(P/M) technique. After weighing, the powder was mixed in
acetone via ball milling with WC-Co balls for 24 h. A change in
WC content, within 0.4 wt%, was observed after ball milling.
Then, the mixed powder was dried for 12 h and sieved to remove
agglomerates, which might contribute to poor sinterability. Disktype compacts were prepared under a pressure of 50 MPa. The
compacts, with a composition of Ti(C0.7N0.3)-xWC-20Ni (where
x 525 wt%), were sintered at a temperature of 1780 K for 1 h
under vacuum, and the resulting sintered specimens were sliced
with a low-speed saw and polished with 1 6 m diamond pastes.
Polished sections were observed via SEM in the backscattered
electron (BSE) mode. X-ray diffractometry (XRD) was conducted
to identify the phases that formed during sintering. In addition,
after sliced specimens were mechanically ground to a thickness of
100 m, ion milling was performed at an acceleration voltage of
6 kV. Then, the compositions of the rim structure were analyzed
using transmission electron microscopy (TEM) in conjunction
with energy-dispersive X-ray analysis (EDXA).

Introduction

ECAUSE of their superior mechanical properties and chemical


stability, Ti(C,N)-based cermets frequently are used as materials for cutting tools, in parallel with relatively modern protective
coatings.1 Currently, Ti(C,N)-based cermet cutting tools typically
are used for high-speed milling, semifinishing, and finishing work
of both carbon steel and stainless steel.2
Ti(C,N)-based cermets possess a fine and stable microstructure.
Such a microstructure, when formed during liquid-phase sintering,
shows a typical core/rim structure, where the cores are partially
dissolved raw-material particles on which the rim structure has
grown through a dissolutionprecipitation process.3,4 Figure 1 is a
schematic of a microstructure of a commercial cermet, imaged via
scanning electron microscopy (SEM). In general, most of the
particles in the microstructure consist of black Ti(C,N) cores and
gray (Ti,W,..)(C,N) solid-solution rims.
WC, which is a secondary carbide, is a necessary carbide
component in a cermet system; WC serves to improve the
mechanical and functional properties. The general role of added
WC in a cermet system is to enhance the density by improving the
wetting and sinterability.5,6 The addition of WC to a Ti(C,N)-Ni
system also has been reported to improve the toughness and
decrease the particle growth rate.6,7 Finally, WC, which forms a
solid solution with Ti(C,N), has been noted to remain as an
independent phase as the nitrogen content of a Ti(C,N)-Ni system
increases. However, to date, few studies of the dissolution and
precipitation behavior of Ti(C,N) and WC in the Ti(C,N)-Ni
system have been reported.8,9 The present study investigates the

T. M. Besmanncontributing editor

Manuscript No. 189464. Received March 29, 1999; approved December 10, 1999.
This research was sponsored by the Korean Institute of S&T Evaluation and
Planning (KISTEP) Research Fund for Advanced Materials through the Highly
Advanced National (HAN) project (1998).
*Member, American Ceramic Society.

Fig. 1. Schematic representation of an SEM image of a Ti(C,N)-based


cermet.

1489

1490

Journal of the American Ceramic SocietyAhn and Kang

Table I. Characteristics of Initial Powders Used for


Preparation of Cermets

Ti(C0.7N0.3)
Ti(C0.5N0.5)
Ti(C0.3N0.7)
WC
Nickel

Particle size (m)

Stoichiometry

Manufacturer

35
35
35
1.88
4.2

Ti(C0.68N0.28O0.01)
Ti(C0.50N0.46O0.01)
Ti(C0.31N0.64O0.01)

Kennametal
Kennametal
Kennametal
H. C. Starck
Novamet

III.

Vol. 83, No. 6

Results and Discussion

(1) Effect of Composition on Microstructure


(A) Effect of WC: Figure 2 shows the SEM microstructures
of compounds in the Ti(C0.7N0.3)-xWC-20Ni system (where x
525 wt%). All the systems exhibited a core/rim structure, as has
been shown for other carbides.10,11 The core/rim boundary, which
appears rather irregular, can be explained by interface instability
that is due to the presence of strain energy.12 In addition, the

Fig. 2. SEM/BSE micrographs of the Ti(C0.7N0.3)-xWC-20Ni systems (for x (A) 5, (B) 10, (C) 15, (D) 20, and (E) 25 wt%), sintered at 1510C under
vacuum for 1 h.

June 2000

Formation of Core/Rim Structures in Ti(C,N)-WC-Ni Cermets via Dissolution and Precipitation

Fig. 3. Change in particle and core size, with respect to WC content in


Ti(C0.7N0.3)-xWC-20Ni systems sintered at 1510C under vacuum for 1 h.

boundary morphology between the rim structure and the binder


phase is less faceted, which indicates that the surface energy at the
boundary becomes more isotropic. This observation is in contrast
to the distinct faceted morphology of Ti(C,N)-Ni systems that
contain NbC or VC.10,11
The change in WC content in the Ti(C,N)-20Ni system has less
effect on the microstructure than those of other carbide additives in
similar systems.11,13 That is, the thickness of the rim remains
relatively thin as the WC content increases. As will be discussed
later in this paper, this result is due to (i) the low dissolution rate
of Ti(C,N), in comparison to that of WC in the nickel melt, and (ii)
limited tungsten solubility in the solid-solution rim phase. The
particle and core sizes were measured by a standard intercept
method and plotted in Fig. 3. The term particle implies a
separate Ti(C,N) entity with its rim structure. The results are the
average of 200 particle measurements. Generally, the changes in
the size of the particles and cores are not significant. The known
effect of particle refinement, as the result of added WC14 to
Ti(C,N)-Ni cermets, was not evident in this study.
It has been reported that the dissolution of WC in the nickel
binder phase can be limited if the total nitrogen concentration is
high in a cermet.9 This phenomenon can be attributed to the low
chemical affinity between tungsten and nitrogen. However, undissolved WC particles are not found in the microstructure of the
present systems. The results of the XRD analyses shown in Fig. 4

Fig. 4.

1491

indicate the absence of a WC peak after sintering at 1510C. Given


the detection limit of X-rays, the undissolved WC content cannot
be more than 5 wt%. Despite the radii of Ti and W atoms (1.47 and
1.41 , respectively), because of the low affinity between tungsten
and nitrogen, the Ti(C,N) peak shifts to the low angles with the
formation of the (Ti,W)(C,N) phase. With an increase in WC
content, a double peakwhich is indicated by arrows in Fig.
4 becomes obvious. This result corresponds to the rim structure
of the (Ti,W)(C,N) solid solution and to the Ti(C,N) cores. The rim
structure is observed to be the major peak.
(B) Effect of Nitrogen Content in Ti(C1XNX): To understand the effect of nitrogen in Ti(C1XNX) on the microstructure,
the nitrogen content (X) was varied over a range of X 0.3 0.7 in
Ti(C1XNX)-10WC-20Ni. Figure 5 shows the microstructures of
the systems that were composed of the Ti(C0.7N0.3), Ti(C0.5N0.5),
and Ti(C0.3N0.7) phases, respectively. For other systems,10,11,14,15
the microstructures have been reported to be strongly affected by
the nitrogen content of Ti(C,N). That is, the morphology
especially the size of the core/rim structure is dependent on the
nitrogen content. In this system, the particle size and rim thickness
decreases as the nitrogen content in Ti(C,N) increases. The
(Ti,W)(C,N) rim that surrounds the Ti(C,N) core was observed in
the system with a low nitrogen content. However, the system that
is composed of the Ti(C0.3N0.7) phase exhibits an almost-rimless
structure and a small particle size. This observation indicates that
the tendency of Ti(C0.3N0.7) to dissolve in the liquid-binder phase
(i.e., the dissolution rate) is the lowest among the various
Ti(C1XNX) phases at 1510C, assuming that the dissolution rate
of WC in the binder is not affected by different Ti(C,N) phases.
Somewhat different results were reported by Fukuhara and
Mitani14 for a Ti(C,N)-Mo-Ni system via the P/M technique and
by Qi and Kang10 for compounds in a Ti(C,N)-NbC-Ni system
obtained via an infiltration technique. Both investigations showed
that, when molybdenum or NbC was added to the Ti(C,N)-Ni
compound, the Ti(C0.5N0.5) phase exhibited the slowest dissolution rate among the various Ti(C1XNX) phases. However, we
determined that the Ti(C0.3N0.7) phase often exhibits the slowest
dissolution rate among various Ti(C,N) compounds,15 including
the result of this study.
Based on thermodynamic calculations for Ti(C,N) stability,16 it
has been reported that Ti(C0.3N0.7) is the most stable compound in
the 14001600C temperature range. However, the small difference in the formation energy of Ti(C0.5N0.5) and Ti(C0.3N0.7)
(12 kJ/mol) might partially explain the discrepancy. The low
dissolution tendency of Ti(C0.3N0.7) is caused primarily by difficulties in the dissociation of the phase. It would appear that the
dissolution rates of the Ti(C,N) phases are still affected, to some

XRD profiles of Ti(C0.7N0.3)-xWC-20Ni systems before and after sintering at 1510C under vacuum for 1 h.

1492

Journal of the American Ceramic SocietyAhn and Kang

Vol. 83, No. 6

Ti(C,N) particles.17 Thus, the equilibrium reactions in the dissolution and precipitation of the hard phase during sintering can be
inferred from the composition analysis of the core/rim structure.10,13 The compositional analyses are made, using TEM/
EDXA, across several particles for Ti(C0.7N0.3)-xWC-20Ni, where
x 525 wt% in the system. The black spots shown in Fig. 6(A)
are some of the positions for the point detection. The results show
that the core regions contain titanium, whereas the rim regions
contain titanium and tungsten. Carbon and nitrogen in the system
were not analyzed, because of the detection limit of the instrument.
The rim phase, which is (Ti,W)(C,N), consists of two regions,
as shown in Fig. 6(B): (i) an inner rim, which is close to the
core/rim boundary and richer in tungsten, and (ii) an outer rim,
which contains less tungsten than the inner rim. The boundary
between the inner and outer rims is not clear in Fig. 6(A); however,
the SEM micrographs in Fig. 2 show inner-rim regions as a bright
area around the cores. The bright appearance of the inner rim is
due to a high concentration of heavy elements in the BSE mode.
No significant difference is observed in the signal levels of Ti and
W in several positions throughout the outer rim; however, the Ti
and W contents clearly differ between the inner- and outer-rim
regions.
The inner rim exhibits a uniform composition with an average
composition of 74 at.% titanium and 26 at.% tungsten in the range
of 10 25 wt% of added WC. The titanium and tungsten concentrations in the outer rim are 84 and 16 at.%, respectively, for x
1525 wt% WC. This result indicates that the compositions of the
inner and outer rims are (Ti0.74W0.26)(C1XNX) and
(Ti0.84W0.16)(C1XNX), respectively, in Ti(C,N)-(1525)WC20Ni. The inner rim is believed to form during the heating stage
from the onset temperature for liquid formation (1300C),
whereas the outer rim forms at the sintering temperature
(1510C).13,17 Thus, the inner-rim composition represents the
average of many different layers that form at different times and
temperatures.
(B) Relative Dissolution Rates of Constituent Particles: The
Ti(C,N)-15WC-20Ni system (in units of weight percent), which is
equivalent to Ti(C,N)-5WC-23Ni (in units of atomic percent),
resulted in WC contents of 26 and 16 at.% for the inner and outer
rims, respectively. This observation indicates that the dissolution
rate of Ti(C0.7N0.3) is slower than that of WC and increases more
rapidly as the temperature increases from 1300C to 1510C. As
shown by the different dissolution tendencies of various
Ti(C1XNX) compounds, this difference seems to be mainly due to
the fact that Ti(C,N) has a higher phase stability than WC in this
temperature range: the entropy change for Ti(C,N) (fGTi(C,N)) is
161 kJ/mol, whereas that for WC (fGWC) is 33 kJ/mol at
1510C.
The relative dissolution rate of WC into Ti(C,N) in a system can
be estimated using the following equation:
Q i S i i m t it i

Fig. 5. SEM/BSE micrographs of Ti(C1XNX)-10WC-20Ni systems (for


X (A) 0.3, (B) 0.5, and (C) 0.7) sintered at 1510C under vacuum for 1 h.

extent, by the composition of the liquid binder in the system and


are strongly affected by the sintering temperature and time.11,15

(2) Dissolution and Precipitation of Constituent Carbides


(A) Composition of Rim Structure: The cores of undissolved
Ti(C,N) particles act as nucleation sites for the rim structure. The
rims form when oversaturated solutes precipitate out onto the

(1)

where Qi represents the average concentration of dissolved carbide


i (in units of mol/cm3). Si is the average surface area of carbide i
per unit weight (in units of cm2/g), i the concentration of carbide
i particles in the system (in units of g/cm3), mti the average
dissolution rate of carbide i (in units of mol(cm2s)1), and ti
the dissolution time (given in seconds). If the ratio of the average
dissolution rate of WC to that of Ti(C,N) remains approximately
constant when the rim phase forms during the sintering process,
and if the ratio of elements dissolved in the binder determines the
final composition of the rim structure, the relative dissolution rate
(i.e., the ratio of (mt)WC to (mt)Ti(C,N)) can be calculated from Eq.
(1). For well-dispersed spherical particles, the following equation
is valid:18
S i im ti

Qi
d

t i dt

V ij i fV ij ,t j

(2)

where Vij is the volume of the jth particle of carbide i and i is the
density of the particle. The term f(Vij,tj) is a particle-distribution

June 2000

Formation of Core/Rim Structures in Ti(C,N)-WC-Ni Cermets via Dissolution and Precipitation

Fig. 6.

1493

(A) TEM micrograph of a Ti(C0.7N0.3)-WC-Ni system; (B) concentrations of titanium and tungsten in the rim.

function that shows the number of particles of Vij size at time tj. The
size change of particles, per unit time, can be expressed as18
DX r ij X a
dr ij

dt
r ij

(3)

Here, D represents the diffusivity, Xrij is the solubility of the jth


particle (of rij size), and Xa is the average concentration; the latter
two parameters are given in units of mole fraction per cubic
centimeter. Equations (2) and (3) and the GibbsThompson
relationship18 suggest that the amount of dissolved carbide per unit
of processing time (Qi/ti) is constant, which guarantees a constant
ratio between the two different carbide solutes in the liquid binder.
Furthermore, the ratio between the total surface area of each
carbide can be regarded as a constant if the ratio between the
average dissolution rates of the carbides is constant.
The measured values for the surface area per unit weight of
Ti(C,N) and WC used in this study are 8.96 105 and 4.17 105
cm2/g, respectively. Using these initial values and the composition
of the outer rim in the Ti(C,N)-15WC-20Ni compound, the

dissolution rate of WC is calculated to be 1.8 times higher than


that of Ti(C,N) at a temperature of 1510C. Similarly, from the
Ti(C,N)-(10 25)WC-20Ni compound, the dissolution rate of WC
was calculated to be 5.3 times higher than that of Ti(C,N) at
1300C. The Ti(C,N)-10WC-20Ni and Ti(C,N)-15WC-20Ni cermets were considered to be reference compositions for the purpose
of calculating the relative dissolution rate. The compositions are
considered just to reach, for a given processing time, the solubility
limits of WC at 1300 and 1510C, respectively.
(C) Precipitation of Rim Structure: In the similar manner,
the compositions of the inner rim and the outer rim were calculated
for the Ti(C,N)-5WC-20Ni system to be (Ti0.86W0.14)(C1XNX)
and (Ti0.95W0.05)(C1XNX), respectively. These values are well
within experimental error, in comparison with the corresponding
(Ti0.83W0.17)(C1XNX) and (Ti0.96W0.04)(C1XNX). Furthermore, the
predicted composition of the outer rim for the Ti(C,N)-10WC-20Ni
system(Ti0.90W0.10)(C1XNX)also is similar to the measured
composition: (Ti0.88W0.12)(C1XNX). All these results were obtained
based on the fact that the Ti:W ratio in the rim phase is similar to that

1494

Journal of the American Ceramic SocietyAhn and Kang

in the binder phase at the sintering temperature. From this observation, it can be concluded that the final compositions of the rim are
determined by the ratio among solute amounts up to the solubility
limits in the binder phase; in addition, this observation means that the
relative dissolution rate is a function of temperature, rather than the
(initial) carbide content.
Given this information, one might surmise that the relative
stability of the solid-solution phases, at high temperatures, increases as the tungsten content in (Ti1YWY)(C1XNX) increases.
The carbon and nitrogen contents in the solid solution can be
estimated by the rule of mass conservation. If a negligible amount
of free carbon and oxygen is present and the process for nitrogen
gasification ([N]Ni 0.5N2) is considerably difficult, then the
compositions of the inner and outer rims are approximately
(Ti0.74W0.26)(C0.78N0.22) and (Ti0.84W0.16)(C0.75N0.25), respectively, in the case of the Ti(C,N)-(1525)WC-20Ni material. This
approach is valid in a separate experiment with TiC-TiN-Ni
systems.
The final composition could be determined by the solute
contents when at least one of the constituent elements, such as
titanium, tungsten, carbon, and nitrogen, attains its maximum
solubility in the melt. According to the binary phase diagrams for
Ni-X systems (where X is the constituent element),19 the solubility
limits of titanium, tungsten, carbon, and nitrogen in nickel are ,
20, 9, and 0.025 at.%, respectively, at 1510C whereas they are
5, 12.5, 0, and 0 at.%, respectively, at room temperature. When
these elements form a nickel-based multicomponent system, the
maximum solubility of each element changes. The effect of
secondary solutes on the solubility of the primary solute can be
quantified using this interaction coefficient.20
Determination of the maximum solubility of each element in
nickel will become a more-complex issue if other reactions are
involved in the liquid solution, in addition to the precipitation of
the rim phase. Other possible reactions could be oxidation of
carbon (C O ^ CO) and the gasification of nitrogen ([N]Ni
0.5N2) in the melt. Based on the data from a binary system,
assuming that there is no significant influence of interaction
coefficients on solubility, the saturation of nitrogen in the melt
would be expected to initiate the precipitation of the rim phase.
IV.

Conclusions

The effect of WC and nitrogen content on the microstructure of


the Ti(C,N)-Ni system was investigated. More importantly, attention was given to the process by which the core/rim structure is
formed and to the determination of the rim composition in the
systems. The following conclusions can be drawn from this study.
(1) The WC content in the Ti(C,N)-Ni system exerts much
less influence, as a parameter, on the microstructure than other
carbides. This phenomenon is due to the fact that the dissolution
rate of Ti(C,N) is significantly slower than that of WC, which
causes the formation of a solid-solution rim that is less sensitive to
sintering time and the amount of WC present.
(2) The Ti(C0.3N0.7) phases exhibits an almost-rimless structure in the Ti(C,N)-WC-Ni system, which suggests that the
dissolution rate of Ti(C0.3N0.7), among the various Ti(C1XNX)

Vol. 83, No. 6

compounds, is the slowest. The dissolution rates of the Ti(C,N)


phases in the liquid-nickel binder seem to be controlled primarily
by the stability of the phases.
(3) Based on the compositions of the outer rim in Ti(C,N)xWC-20Ni systems, the dissolution rate of WC is 2 times higher
than that of Ti(C,N) at 1510C. The average dissolution rate of
WC is 5 times higher than that of Ti(C,N) at 1300C.
(4) The compositions of the inner and outer rims are
(Ti0.74W0.26)(C0.78N0.22) and (Ti0.84W0.16)(C0.75N0.25), respectively, in Ti(C,N)-(1525)WC-20Ni. All the rim compositions are
determined by the amount of solutes available in the binder at
1300 and 1510C. The phases that are obtained from transient
compositions, i.e., Ti(C,N)-(510)WC-20Ni, also are formed with
the same dissolution rates, which means that the relative dissolution rate is a function of temperature, rather than carbide content.
References
1

S. Zang, Titanium Carbonitride-Based Cermets: Process and Properties, Mater.


Sci. Eng. A, A163, 14148 (1993).
2
E. B. Clark and B. Roebuck, Extending the Application Areas for Titanium
Carbonitride Cermets, Refract. Met. Hard Mater., 11, 2333 (1992).
3
P. Ettmayer, H. Kolaska, W. Lengauer, and K. Dreyer, Ti(C,N) Cermets
Metallurgy and Properties, Int. J. Refract. Met. Hard Mater., 13, 34351 (1995).
4
P. Ettmayer and W. Lengauer, The Story of Cermets, Powder Metall. Int., 21,
3738 (1989).
5
H. Matsubara, S. Shin, and T. Sakuma, Grain Growth of TiC and Ti(C,N) Base
Cermet During Liquid Phase Sintering, Solid State Phenom., 25&26, 55158 (1992).
6
H. Suzuki, H. Matsubara, and T. Saitoh, The Microstructures of Ti(C,N)Mo2C-Ni Cermet Affected by WC Addition, Jpn. Soc. Powder Powder Metall., 31
[7] 236 40 (1983).
7
H. Suzuki and H. Matsubara, Some Properties of Ti(C,N)-WC-Ni Alloy, Jpn.
Soc. Powder Powder Metall., 33 [4] 199 203 (1986).
8
M. Rynemark, Investigation of Equilibria in the Ti-W-C-N System at 1750C,
Refract. Met. Hard Mater., 10, 18593 (1991).
9
A. Doi, T. Nomura, M. Tobioka, K. Takahashi, and A. Hara, Thermodynamic
Evaluation of Equilibrium Nitrogen Pressure and WC Separation in Ti-W-C-N
System Carbonitride; pp. 825 43 in Proceedings of the 11th International Plansee
Seminar 85, Vol. 1. Edited by H. Bildstein and H. M. Ortner. Reuttre, Tirol, Austria,
1985.
10
F. Qi and S. Kang, A Study on Microstructural Changes in Ti(C,N)-NbC-Ni
Cermets, Mater. Sci. Eng. A, A251, 276 85 (1998).
11
S. Ahn and S. Kang, Effect of VC Addition on the Microstructure and
Mechanical Properties of Ti(C,N)-Based Cermet, J. Korean Ceram. Soc., 35 [12]
1316 22 (1998).
12
K.-W. Chae, D.-I. Chun, D.-Y. Kim, Y.-J. Baik, and K.-Y. Eun, Microstructural
Evolution During the Infiltration Treatment of Titanium CarbideIron Composite,
J. Am. Ceram. Soc., 73 [7] 1979 82 (1990).
13
J. K. Yang and H.-C. Lee, Microstructural Evaluation during the Sintering of
Ti(C,N)-Mo2C-Ni Alloy, Mater. Sci. Eng. A, A209, 21317 (1996).
14
M. Fukuhara and H. Mitani, Effect of Nitrogen Content on Grain Growth in
Ti(C,N)-Ni-Mo Sintered Alloy, Powder Metall. Int., 14 [4] 196 200 (1982).
15
S. Mun and S. Kang, The Effect of HfC Addition on the Microstructure of
Ti(C,N)-Ni Cermet System, Powder Metall., 42 [3] 25156 (1999).
16
I. J. Jung, S. Kang, S. H. Jhi, and J. Ihm, A Study of the Formation of Ti(C,N)
Solid Solutions, Acta Mater., 47 [11] 3241 45 (1999).
17
M. G. Gee, M. J. Reece, and B. Roebuck, High Resolution Electron Microscopy
of Ti(C,N) Cermets, J. Hard Mater., 3 [2] 119 43 (1992).
18
G. W. Greenwood, The Growth of Dispersed Precipitates in Solutions, Acta
Metall., 4, 243 48 (1956).
19
A. E. McHale, H. F. MacMurdie, and H. M. Ondik, Phase Equilibria Diagrams,
Vol. X, Borides, Carbides, and Nitrides; 1st Ed. American Ceramic Society,
Westerville, OH, 1994.
20
R. D. Pehlke and J. F. Elliott, Solubility of Iron Alloys: 1, Thermodynamics,
Trans. Metall. Soc. AIME, 218, 1088 101 (1960).

Das könnte Ihnen auch gefallen