Вы находитесь на странице: 1из 11

Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Journal of the European Ceramic Society


journal homepage: www.elsevier.com/locate/jeurceramsoc

Original Article

Analysis of pressureless sintering of titanium diboride ceramics with nickel,


cobalt, and tungsten carbide additives
Rainer Telle
Institute für Gesteinshüttenkunde (Mineral Engineering), Chair of Ceramics and Refractories, RWTH Aachen University, Mauerstrasse 5, D-52064 Aachen, Germany

ARTICLE INFO ABSTRACT

Keywords: The mechanisms during pressureless sintering of TiB2 with 0–50 wt.% Ni were analysed by differential dilato-
TiB2 metry after mechanical alloying with hard metal beads. Three main stages of sintering could be identified by
Ni means of the changing dilatation rate. The first interval (200–600 °C) comprises reactions with oxygen and
Co moisture (H3BO3, B2O3 formation, melting, evaporation). The second stage (600–700 °C) is governed by solid
Mechanical alloying
state sintering and reactions of Ni and/or Co with TiB2 yielding Ni3B, Ni2B, the corresponding Co-compounds
Sintering reactions
and the ternary ω-phase of W2NiB2-type. The third main shrinkage interval is due to liquid phase sintering
Grain growth
(987–1640 °C) and is accompanied by a significant weight-loss. Both precipitation of ω-phase as well as a (Ti,W)
B2 solid solution in form of a core-shell structure are responsible for a grain growth retardation and the
homogeneous microstructure. The results explain well all features reported in the corresponding literature while
encountering the role of WC, too.

1. Introduction grain boundary energy, i.e. by use of oxide-layer reducing agents such
as carbon and aluminium.
In recent times, the increasing demand for novel materials used in Due to its high melting point of 3250 °C TiB2 requires sintering
armour, hypersonic aerospace vehicles, and electronically conductive temperatures in the range of 1800–2300 °C to activate grain boundary
ceramics has once again attracted interest to transition metal diboride- and volume diffusion to obtain > 95% of the theoretical density (th.d.).
based materials such as TiB2 because of their high melting point and At lower temperatures, evaporation of boron and boron suboxides en-
corrosion resistance, exceptional hardness and stiffness combined with hance grain growth without shrinkage by re-condensation and surface
a remarkably high fracture strength, fracture toughness, and electronic diffusion. These sintering mechanisms give rise for a particle growth in
and thermal conductivity [1–4]. Many publications deal with non-re- a well-facetted crystal shape together with the formation of wide pore
active, reactive or spark-plasma assisted hot pressing procedures of channels as well as pore trapping inside large grain clusters
conventional TiB2 powders aiming at the perfection of mechanical [1,9,12–15]. Therefore, pressureless densification of TiB2 powders
properties [5,6]. Today, only few publications [e.g. 7, 8] treat, how- conventionally produced by the reduction of TiO2 with B4C and C
ever, pressureless sintering although much fundamental work was al- usually fails even at temperatures between 1900–2500 °C.
ready done between 1960 and 1985 which is quite often disregarded In 1984, Baumgartner and Steiger [10] accomplished pressureless
although presenting valuable hints for pressureless sintering systems of solid-state sintering to very high density by means of ultra-pure TiB2
TiB2-based composites [e.g. 9–11]. nano powders produced by a vapor phase reaction of TiCl4 with BCl3.
As TiB2 possesses a pronounced anisotropy of thermal expansion its After sintering at 2000–2100 °C in inert gas densities between 98.4 and
major disadvantage for fabrication is the sensitivity to spontaneous 99.4% th.d. were achieved together with a relatively fine micro-
mechanical degradation by internal stress-induced cracking during structure of 1–18 μm average grain size. Baumgartner and Steiger
cooling from sintering temperature if grain growth has generated a identified inclusions of impurities such as TiC and TiO responsible for
microstructure with particle size exceeding 5–10 μm. This fact limits the preventing grain growth according to the particle drag mechanism.
use of very high sintering or hot-pressing temperatures to gain full Another approach to gain high density is liquid phase sintering
density. To overcome this problem, the sinter activity has to be im- where capillary forces generate a fast particle re-arrangement and allow
proved either by increasing the total surface energy, i.e. by using sub- for further densification by the well-known mechanisms such as shape
micron powders, or by increasing both the specific surface energy and accommodation, dissolution/re-precipitation and skeletal grain growth.

E-mail address: telle@ghi.rwth-aachen.de.

https://doi.org/10.1016/j.jeurceramsoc.2019.02.036
Received 15 November 2018; Received in revised form 14 February 2019; Accepted 18 February 2019
0955-2219/ © 2019 Elsevier Ltd. All rights reserved.

Please cite this article as: Rainer Telle, Journal of the European Ceramic Society, https://doi.org/10.1016/j.jeurceramsoc.2019.02.036
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Liquid phase sintering, however, also fosters rapid mass transport


through the melt and significantly accelerates grain growth in parti-
cular by Ostwald ripening. This fact requires further consideration to
prevent grain growth such as the use of a minimum volume of liquid
phase, the implementation of grain growth-retarding additives, or the
formation of lattice strain by structural misfit during growth [16]. Ac-
cordingly, there are many publications on the use of iron-group metals,
i.e. Fe, Ni, Co, Fe-Ti or Ni-Al alloys as optional metallic liquid phases,
e.g. [17–26], dealing with the wetting behaviour and phase diagrams
including ternary phases. The formation of a cermet-type composite of a
hard but brittle boride phase bonded and reinforced by a ductile metal
matrix, is, however, not that easy to achieve as the wetting is sensitively
controlled by traces of contaminants as well as by the precipitation of
Fig. 1. TiB2 powder untreated.
brittle binary or ternary metal borides. Moreover, all composites had to
be hot-pressed or hot-isostatically pressed after pressureless pre-sin-
tering for final densification. As the sintering of cermets is not the topic
of this paper, it is referred to [2], which provides a brief survey on the
achievements in the field of cemented borides.
Recently, Fu and Koc reported on liquid phase sintering of TiB2
which was prepared by reduction of a carbon-precursor coated TiO2
powder and metallic cobalt [27,28]. Since the powder had a particle
size of 0.3-0.8 μm the driving force for sintering was almost as high as
in the case of Baumgartner´s starting material (0.2-0.5 μm). The addi-
tion of 1 wt.% Co was found to be insufficient yielding 89.5% and
91.5% after sintering at 1500 °C and 1600 °C for 1 h, respectively, while
the addition of 3 wt.% was considered optimum leading to 98% density
at 1600 °C. Fu and Koc obtained an average grain size of 1.75 μm
whereas a higher volume fraction of liquid phase initiated coarsening of Fig. 2. Nickel powder untreated.
TiB2 up to 10 μm particle size. After sintering, the matrix phase con-
sisted of residual Co and Co2B phase while the TiB2 was unaffected.
nickel powder was HCST 495H grade by H.C. Starck, Goslar, Germany,
If no special TiB2 powders of extremely fine grain size and high
(in wt.%: Ni: 68.8, B: 30.4, C: 0.21, N 0.19, O: 0.4) with an average
purity are available in larger quantities rather than produced in a la-
grain size of 1–1.5 μm (Fig. 2). Powder batches of TiB2 + 0, 5, 7.5, 10,
boratory-scale, commercial raw materials of larger grain size have to be
and 12.5 wt.% Ni (100 g in total) were milled together with 3500 g of
ground by high-energy milling. If the milling bead materials are che-
Co-bonded hard metal beads of 1.1–1.2 mm diameter (Netzsch-Ger-
mically compatible, the debris from the milling beads can be used as an
ätebau GmbH, Selb, Germany; composition in wt.%: W 44.6; Ti 43.7; C
additive, and the procedure is called “mechanical alloying”. Ağaoğullan
6.0; Co 7.6 equal to 92.4% (W,Ti)C + 7.6% Co). The blend with
et al. [8] reported on the pressureless sintering of TiB2 with 0–20 wt.%
12.5 wt.% Ni also contained 2.0 wt.% amorphous boron (Sigma Aldrich,
Co addition ground with WC balls for 3–9 h. Sintering at 1550 °C for 1 h
Steinheim, Germany, in wt.%: B: 95.0, C: 0.0, N: 0.08, O: 2.2, Si: 0.2,
in Ar yielded Co and Co2B together with metallic Ti as the matrix
Mg: 1.0, Fe: 0.2) to prevent boron losses from the TiB2-powder. An
phases. Unfortunately, the authors did not argue how the WC-debris
additional composition with 50 wt.% Ni was produced in order to make
from the milling balls interacted with the liquid Co or the TiB2 particles
the liquid phase reactions more evident and to allow for grain growth
and how metallic Ti could be obtained. Furthermore, the relative
and, therefore, for a better chemical analysis by EDS. The attrition mill
density calculations were based on the unreacted TiB2-Co composition.
with hard metal stirrer (Moulinette, Netzsch-Gerätebau GmbH, Selb,
In this work, a pressureless sintering system for TiB2 ceramics is
Germany) and stainless steel container was run at 1000 rpm. The sus-
presented and analysed in respect to the temperature-dependent phase
pension medium was ethanol to prevent hydroxylation. The wear of the
reactions, which makes use of the general rules of liquid phase sintering
stainless steel container and the stirrer was negligible while the pick-up
and grain growth prevention. Nickel was introduced for liquid phase
from the milling balls relative to the powder input increased from
formation while tungsten carbide was used to provide in-situ-formed
5.8 wt.% after 0,5 h milling to 22.5 wt.% after 4 h milling time
grain growth retarding particles. Since WC was added by mechanical
(Table 1). The wear debris from milling balls was nanometre-sized and
alloying from hard metal milling beads, also cobalt was present as a
generated a coating onto the TiB2 particles after drying in a rotary dryer
minor constituent of the balls.
(Rotavapor, Büchi, Switzerland). Fig. 3 shows the as-dried powder
The influence of milling debris on the sintering of boride-based
composite.
materials is of general importance, as powder processing by high-en-
Shaping of the samples for sintering studies was carried out by cold
ergy milling with hard metal beads has become a quite common pro-
isostatic pressing of cylindrical specimens in silicon rubber moulds
cedure which is quite often applied by powder suppliers to enhance the
under 630 MPa for 1 min (tapping density 30–45% th.d.; green density
sintering activity even if the powder will be used for hot pressing.
50–60% th.d.). The final green body size was 14 mm in diameter and
Moreover, boronization of hard metal cutting tools has also become a
21 mm in height, approximately. Since both green body strength and
popular method to harden surfaces subjected to wear [23].
shape accuracy were excellent, no organic binder was added.
Sintering was carried out in a resistant-heated tungsten-tube va-
2. Material and methods
cuum furnace (Heraeus Combivac, Germany) under 0.04 Pa pressure
after three times flushing with 99.98% pure N2. The samples were
Starting material was commercial TiB2 grade 2936C by H.C. Starck
coated by hexagonal boron nitride powder to prevent reaction with the
GmbH, Goslar, Germany, (in wt.%: Ti: 68.8, B: 30.4, C: 0.21, N: 0.19, O:
alumina crucible. The crucible provided space to sinter several samples
0.4) produced according to the boron carbide method. The particles
simultaneously together, which guaranteed same heating conditions for
show a well crystallised hexagonal-prismatic habit, an average particle
each test series. Temperatures was recorded by both PtRh-Pt and W-Re
size (d50) of 3.0 μm and a total size range of 0.1–20 μm (Fig. 1). The

2
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Table 1
Starting material composition before and after 4 h attrition milling (boron analysed by wavelength-dispersive X-ray spectroscopy).
Input wt.% Output wt.% Output vol.%
Mix TiB2 Ni B TiB2 Ni Co WC B TiB2 Ni Co WC B

0.0 100.0 0.0 0 75.9 0.0 1.6 22.5 0 91.1 0.0 1.0 7.7 0
5.0 95.0 5.0 0 65.8 3.5 2.0 28.7 0 85.8 2.3 1.3 10.6 0
7.5 92.5 7.5 0 71.6 5.8 1.5 34.7 0 84.1 3.4 0.9 11.5 0
10.0 90.0 10.0 0 56.6 6.3 2.4 34.7 0 79.9 4.5 1.7 13.9 0
12.5 85.8 12.2 2.0 64.3 9.2 1.6 23.4 1.5 81.3 5.9 1.0 8.4 3.4

Fig. 3. TiB2 + 5 wt.% Ni after 4 h attrition milling with (W,Ti)C-Co beads. Note
the fine wear debris attached to the surface of TiB2 particles. Fig. 4. Density dependence from temperature after 2 h isothermal sintering
time. The continuous increase of absolute density results from evaporation of
Ni-compounds and the relative enrichment in W-compounds.
thermal couple which was calibrated by the melting point of high-
purity Co, Ni, and Cu. Sintering was carried out at 800, 1000, 1100,
1200, 1350, 1500, 1600, 1700, and 1900 °C with isothermal annealing
for 1–2 h, respectively, in some cases for 5 min to 4 h. Heating rate was
40 K/min, cooling rate 100 K/min. After removal of the hBN coating,
the density was measured according to Archimedes principle. In case of
open porosity, the geometric density was measured. Weight-loss and
dilatational changes were recorded, too. Continuous shrinkage was
measured in an argon atmosphere up to 1700 °C by a graphite-lined
differential dilatometer (L75VD2400C, Linseis Messgeräte GmbH, Selb,
Germany) with a heating and cooling rate of 10 K/min. Calibration runs
with a graphite sample for the compensation of the expansion of the
entire device were carried out once before and another time right after
every experiment. DTA-measurements (DSC 404, Netzsch-Gerätebau
Fig. 5. Time-dependent absolute density and pore volume of a 7.5 wt.% Ni
Gmbh, Selb, Germany) were carried out in He atmosphere with a composition sintered at 1500, 1600 and 1700 °C. The curves belong to both
heating rate of 10 K/min up to 1400 °C with a tantalum reference axes, respectively.
sample. The phase composition was analysed by x-ray powder diffrac-
tion (Bruker D8 Advance DaVinci, BrukerAXS, Karlsruhe, Germany),
and by SEM together with EDX-analysis (LEO 440i with Oxford EDX
system Link Isis). Grain size distribution and porosity were determined
from high-magnification SEM images by manually recording the peri-
meters of at least 350–400 particles or pores followed by computer-
aided classification and statistical evaluation.

3. Results

3.1. Vacuum furnace sintering

The densification behaviour depending from the Ni content is


shown in Fig. 4 for 2 h isothermal annealing at the respective tem-
perature. Obviously, the onset of shrinkage at 990–1090 °C is in- Fig. 6. Time-dependent weight-loss of a 7.5 wt.% Ni composition after iso-
dependent from the sample composition while a maximum density was thermal sintering at 1500 °C in a vacuum.
obtained in case of 10 wt.% Ni. As the theoretical density could not be
calculated due to phase reactions, the pore volume was determined by
ultimate temperatures. At a density of 5.0 gcm−3, approximately,
evaluation of SEM-micrographs of polished sections. The lower limit of
closed porosity is obtained which can be taken from the microstructural
2.8% porosity is due to ongoing evaporation of Ni-compounds, while
analysis. This corresponds to the isothermal weight-loss curve (Fig. 6)
the continuous increase in density at high temperature is attributed to
changing from a parabolic to a linear rule.
evaporation of volatiles and the relative enrichment in tungsten.
Looking at the temperature-dependent weight-loss recorded above
The time-dependent densification is illustrated in Fig. 5 for the
800 °C after 2 h of annealing it becomes obvious that evaporation of
7.5 wt.% Ni composition fired at 1500, 1600 and 1700 °C as the

3
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Fig. 7. Temperature-dependent weight-loss of a 7.5 wt.% Ni composition after


2 h sintering in a vacuum. Note the steep increase at > 1000 °C. Fig. 9. Dilatation rate dl/lodt of a 7.5 wt.% Ni composition, 10 K/min heating
rate between 250 and 850 °C.
volatile species starts between 1000 and 1100 °C which is associated
with the onset of shrinkage followed by the maximum densification
rate. At temperatures above 1350 °C the weight-loss is constant (Fig. 7).

3.2. Dilatometry

As recorded by the dilatometric measurement of the 7.5 wt.% Ni


containing greenbody, the sample first expands up to 325 °C and tends
to shrink stepwise up to 708 °C. Following, there is a temperature in-
terval up to 1150–1170 °C with an almost unchanged elongation. At
1160–1170 °C, the main shrinkage occurs up to 1600 °C (Fig. 8). This
behavior can even be better resolved by recording the first derivate of
the dilatation curve, i.e. the shrinkage rate dl/lo⋅dt in Figs. 9 and 10.
Dilatometer curves for Ni concentrations of 5, 10, and 12.5 wt.% do
Fig. 10. Continuation of Fig. 9 in the range of 1000–1600 °C showing the main
not significantly differ from the measurements for the 7.5% composi-
sintering interval.
tion. The intensities of the reactions increase slightly with increasing
Ni-content, and the temperatures vary in the range of ± 7 K. All sin-
tering curves imply that the main shrinkage of 22% is accomplished at nickel addition, the sintering curve principally shows the same beha-
1700 °C so that pressureless sintering should be possible between 1500 viour: Main densification starts after a short period of expansion and an
and 1700 °C if an isothermal sintering time is applied. intermediate shrinkage tendency (1040 °C) at around 1300–1320 °C
As the dilatation rate is, similar to DTA measurements, very sensi- (Fig. 11). At the highest temperature of 1700 °C the shrinkage rate is,
tive for both sintering and phase reactions, it becomes obvious that however, still not decreasing. A linear shrinkage of 12% is obtained at
heating up is accompanied by distinct effects between 325–460 °C, that point and isothermal sintering finally results in 16% shrinkage
455–510 °C, 605–710 °C, while the main sintering takes place between after 3 h. This implies that mechanical alloying with WC/TiC-Co beads
1160 and 1600 °C, interrupted by another shrinkage-accelerating re- should be a sufficient treatment for pressureless sintering if a longer
action between 1230 and 1303 °C (Fig. 10). All of these little changes in isothermal annealing at 1700 °C follows.
the dilatation rate have been well reproduced by several dilatometer Comparing the sintering kinetics of green bodies with and without
runs and are therefore attributed to true phase reactions or sintering nickel addition, the plots of the dilatation rates exhibit similar effects
phenomena, respectively. (Figs. 12 and 13). Some of these are, however, shifted to lower and
In order to separate the influence of nickel from that of the me- some to higher temperatures if nickel is present. Thus, the reaction
chanical alloying, a pure TiB2 powder batch has been prepared by at- series from the maximum at 432 °C to the minimum at 672 °C for 0% Ni
trition milling under exactly the same conditions. In this case of no has moved to 455 °C and 708 °C for 7.5% Ni, respectively, while the
pronounced minimum at 1134 °C (0% Ni) is missing in case of a nickel
content. In contrast, the following onset of the main shrinkage at
1304–1320 °C is the same for all sintering systems. Another result is
that any nickel addition yields by more than ten times higher shrinkage
rates compared to the blend without nickel.
The DTA-measurement of the initial powder mixture with 7.5 wt.%
Ni reveals first a very broad exothermic range between 600 and 1050 °C
followed by a smooth endothermic effect starting at 1113 °C with a
maximum at 1123 °C. The reaction is completed at 1131 °C but im-
mediately followed by another weak endothermic reaction at 1163 °C
which changes to a long-lasting exothermic reaction at 1235 °C up to
1400 °C/maximum temperature. During cooling down an exothermic
effect is observed at 1038 °C (maximum) followed by a very steep
Fig. 8. Dilatometer diagram of a 7.5 wt.% Ni composition, 10 K/min heating
exothermic peak between 1035 and 1029 °C. Repeated heating shows
rate and 20 min isothermal sintering at 1700 °C. The numbers give the tem- only one endothermic reaction starting at 1317 °C and terminating at
peratures of changes in the shrinkage rate. Note the onset of the main sintering 1398 °C while repeated cooling yields an exothermic effect again at
phase at 1168 °C. 1035 °C. To correlate the DTA and dilatometer results, a

4
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Fig. 11. Dilatometer diagram of mechanically alloyed TiB2 powder without, 10 K/min heating rate and 3 h isothermal sintering at 1700 °C. Note the onset of the main
sintering phase at 1304 °C. At 1700 °C the shrinkage still accelerates.

Fig. 12. Dilatation rate dl/lodt of TiB2 without Ni-addition, 10 K/min heating
rate between 250 and 850 °C.

Fig. 14. Superimposition of DTA and differential dilatometer curves of a TiB2


+ 7.5 wt.% Ni sample during the first heating. Note the long-term exothermic
range and the significant endothermic peaks due to melting which are followed
by incipient shrinkage.

Fig. 13. Continuation of Fig. 12 in the range of 1000–1600 °C showing the


onset of the main sintering inverall at > 1304 °C.

superimposition of both curves is presented in Fig. 14.


Fig. 15. Micrograph of a TiB2 + 7.5 wt.% Ni sample, sintered at 1930 °C for
30 min in Ar atmosphere. All grains, in particular the large one in the centre as a
3.3. Phases and microstructure residual unmilled powder particle, exhibit the typical core-rim structure with a
TiB2 center and a (Ti,W)B2 solid solution around.
The microstructure of the samples mainly consists of three groups of
phases. The first is a series of (Ti,W)B2 solid solutions containing up to lattice. Also, no β-WB, β-(W,Ti)B or (W,Ti)B2 of W2B4-structure have
14 mol% WB2, in average 6.9 ± 3.8 mol%. After grain growth in an been found as proposed by other authors [5,6,29].
intentionally higher volume fraction of liquid phase, i.e. in the case of The second group of phases comprises ternary borides with a gen-
the 50 wt.% Ni series, or in case of unmilled larger particles it becomes eral formula WCoB, the so-called φ-phase (orthorhombic, Pnma, # 62),
obvious that all grains exhibit a typical core-rim structure with an un- W2NiB2, the so-called ω-phase (orthorhombic, Immm, #71), and
reacted dark TiB2-core inside surrounded by the (Ti,W)B2 solid solution W3CoB3, the so-called γ-phase (orthorhombic, Cmcm, #63) [30–32]. In
appearing light due to the atomic number contrast (Figs. 15 and 16). our system these formulas contain Ti and Co, too, and may be written as
There is no indication that (Ti,W)B2 may contain Ni or Co in the crystal

5
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

(W,Ti)C is partially already liquefied, at least reacted and sintered. The


matrix phase contains small residues of undissolved (W,Ti)C debris
from the milling beads as well as facetted precipitates of (W,Ti)2(Ni,Co)
B2 (frame in Fig. 18). Thus, the liquid is W-richer compared to sintering
at 1600 °C when all the debris is dissolved and the tungsten is in-
corporated into the (Ti,W)B2 and the grown (W,Ti)2(Ni,Co)B2 solid
solutions.

3.4. Exudation

The significant weight-loss above 1000 °C in vacuum implies that


pore channels are kept open while the surrounding microstructure
shrinks during sintering. Accordingly, there are stages of sintering in
Fig. 16. Micrograph of a TiB2+50 wt.% Ni specimen, sintered at 1500 °C for which the samples exhibit strongly differentiated cross sections. While a
2 h in vacuum. Dark: Euhedral TiB2 crystals with core-rim structure. Bright: heat treatment up to 1000 °C still reveals a homogeneous, open-porous
W2(Ni,Co)B2 with TiB2-inclusions. Matrix: Ni/Co-B-Ti/W-liquid phase. and green-body-like appearance, rising temperature to 1100 °C already
indicates concentric zones of different grinding and polishing hardness.
Table 2 Using 5 wt.% Ni, this graded structure is not visible any more after
Chemical composition of ternary borides in at.%. Note, that the boron content sintering at 1700 °C. With higher Ni-content, the temperature required
has not been determined. for homogenisation decreases to 1500 °C A close-up of a corresponding
zonal structure in 50% Ni containing sample sintered at 1200 °C is
W Ti Ni Co (W + Ti)/(Ni + Co)
shown in Fig. 19. The outer surface layer of the sample is created by a
ω-Phase 60.2 8.0 22.0 10.8 2.1 strong exudation of liquid phase leaving degassing channels behind.
Std. Dev. 3.2 1.5 11.0 8.2
γ-Phase 52.8 21.6 12.6 13.3 2.9 3.5. Grain growth
Std. Dev. 3.4 6.2 2.5 0.5

In homogeneous areas of the samples, all phases exhibit an equiaxial


shape and an average grain size (d50) of 0.5-0.7 μm. Typical micro-
structures of 7.5 wt.% Ni samples (1000, 1350 and 1500 °C, 2 h, re-
spectively) are shown in Fig. 20a-c.With increasing temperature, higher
Ni-content or prolonged sintering time the grain growth of the (Ti,W)B2
solid solution is not significant. The maximum grain size is between 2
and 7 μm (Fig. 21a–c). Except for the temperature dependence
(Fig. 21a), the variations are not systematic, which may be due to the
effect of evaporation as well as due to errors in monitoring the micro-
structural data.
The correlation between the various phases can be taken from the
kind of intergrowth. The first visible interaction between the newly
generated liquid phase is the partial dissolution of (W,Ti)C and TiB2
particles. The next step is the epitaxial precipitation of a (Ti,W)B2 solid
solution onto the TiB2 residues acting as nucleation sites (Figs. 15–17).
The composition of the solid solution ranges from 4 to 14 mol-% WB2.
Fig. 17. Large facetted (W,Ti)2(Ni,Co)B2 crystal intergrown with TiB2-particles This can also be proven by TEM-studies. Due to the misfit of lattice
in Ni/Co-B-Ti/W melt. TiB2 + 50 wt.% Ni, sintered at 1700 °C for 30 min in parameters the boundary between core and shell is marked by dis-
argon. locations and stresses (Fig. 22). This phenomenon has been observed
quite often and regarded as a hardening and strengthening mechanism
(W,Ti)(Co,Ni)B, (W,Ti)2(Ni,Co)B2 and (W,Ti)3(Ni,Co)B3, respectively. [16,33,34]. Thus, the sintering system presented here provokes active
The average chemical compositions depending from the local avail- dissolution/re-precipitation as sintering mechanism and therefore ex-
ability of surrounding liquid phase are given in Table 2. Both phases hibits the well-known “core-shell” structure known from many hard
can easily be recognised in SEM-images due to their bright atomic metals and cermets. The (Ti,W)B2 solid solution may have grown
number contrast. They always appear in a euhedral shape and even may during further heating as well as during the initial steps of cooling
include smaller particles of the (Ti,W)B2 phase (Fig. 17). From the down. It is obvious from the idiomorphous crystal shape that the grain
microstructural analyses it is, therefore, evident that the ternary borides growth occurred unhindered into a liquid phase.
are growing from the supersaturated liquid phase, most probably
during further heating or cooling down. 4. Discussion
The last group of phases comprises the matrix, i.e. the liquid phase,
which solidifies as Ni3B and Ni2B containing some Co, respectively, 4.1. Sintering reactions
together with the ternary borides mentioned before. In rare cases also
Ni4B3 is found by XRD. XRD-measurements reveal that in case of 7.5 wt. For an understanding of the sintering effects occurring in the TiB2-
% Ni addition, all metallic Ni has already reacted to Ni-borides at Ni-Co-WC system both the well-known solid-state and liquid phase
1000 °C, as shown by the broad range of exothermic reactions in sintering phenomena as well as the chemical interactions between the
Fig. 14. Only in case of 50 wt.% Ni residual metallic Ni is stable up to phases have to be considered.
1300 °C together with Ni-borides. Table 3 presents EDS-analyses of First of all, oxygen impurities must be taken into account even if
larger homogeneous matrix areas which are not available if the Ni- high-purity raw materials have been used since mechanical treatment,
content is less than 50 wt.%. At 1200 °C the blend of Ni, Co, and drying and handling of powders and green bodies in air will necessarily
result in the formation of an oxide layer at the surface of both metal and

6
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Table 3
Composition of the matrix phase in at.-%. Note, that the boron and carbon content could not be determined.
Element/at.% / Ti W Ni Co Ti + W Ni + Co (Ti + W)/
Conditions (Ni + Co)

50% Ni, 1200C, 2 h 14,7 ± 0,8 2,6 ± 0,3 81,6 ± 2,8 1,1 ± 0,4 17,3 82,7 0,2
50% Ni, 1600C, 1 h 8,4 ± 0,4 0,5 ± 0,05 90,1 ± 0,5 1,0 ± 0,2 8,9 91,1 0,1

Before melting occurs, these metals undergo solid state sintering first,
which happens between 600 (grain boundary diffusion) and 750 °C
(volume diffusion) in case of Ni [39]. This explains the first significant
shrinkage peak between 511 and 708 °C in case of Ni-addition as well as
the shift to 532 and 733 °C without nickel which is due to the higher
melting point of Co and the accordingly higher activation energy for
diffusion.
At 900 °C, Márquez-Herrera et al. [28] reported two solid state re-
actions: 2 Co + B → Co2B and Co + 2 WC + 2 B → W2CoB2 + 2 C after
4 h annealing of a B4C-coated WC-Co cutting tool. The first reaction, i.e.
the formation of Co2B, has to be modified in our case because of the
presence of Ni and the different source of boron from TiB2. X-ray dif-
fraction analysis reveals, that metallic nickel does not exist anymore at
temperatures as low as 1000 °C if the initial Ni-content is less than
Fig. 18. Ni3B/Ni2B matrix with bright (W,Ti)2(Ni,Co)B2 precipitates and area of 50 wt.%. It can be concluded that Ni and Co first undergo solid state
integral EDS-analysis in TiB2+50 wt.% Ni, 1200 °C 2 h, vacuum. sintering between 600 and 700 °C (Fig. 9), while TiB2 releases boron at
800 °C and higher for the formation of Ni and Co borides. Yet, there is
no indication that (W,Ti)2(Ni,Co)B2 already crystallised, while WC and
TiC are still detectable.
Calculations in the ternary B-Ni-W system by Morishita et al. [32]
proved a ternary eutectic (E2 in Fig. 23) with L = Ni + Ni3B + W2NiB2
at Te = 987 °C which is closely located to the tie-line between WB2 and
Ni. Unfortunately, there is no information on the liquidus surface of the
isotypic B-Co-W system available, except the isothermal section at
1000 °C published by Haschke et al. [30] showing an area around the τ-
phase in which melting was observed between 950 and 1100 °C. In the
ternary B-Ni-Ti system, an ω-phase of W2NiB2-type does not exist but
the ternary τ-phase Ni21Ti2B6, which is in equilibrium with elemental
Ni but does not form a quasi-binary section together with TiB2 [40–42].
According to these data the tie line crosses the ternary equilibrium TiB2
Fig. 19. SEM-micrograph of a 50 wt.% Ni sample sintered at 1200 °C/2 h re- + Ni3B + τ from room temperature to 998 °C. At higher temperatures,
vealing open degassing channels underneath the surface. TiB2 is stable together with a liquid phase and both Ni3B and Ni2B. A
more recent study by Ajao [26] treated the formation and decomposi-
tion of the τ-phase in the Ni-rich corner by DTA and set the ternary
TiB2 powders. In laboratory atmosphere, even the presence of a small
eutectic L = τ + Ni + Ni3B to Te = 952 °C. During sintering or hot-
amount of humidity will generate hydroboric acid as next. Therefore,
pressing under technical conditions, the τ-phase is generally not ob-
during the first steps of heating, both H3BO3 and B2O3 will liquefy
tained. Its formation may be suppressed either by the presence of TiO or
(Tm = 170.9 and 475 °C, respectively), while at 400 °C and higher TiB2
TiO2 [43–45] or for kinetic reasons due to fast firing and quenching.
will actively oxidise to TiBO3 and further to TiO2 and B2O3 [35–38] as
Hence the residual matrix phase mainly consists of Ti-containing Ni3B
long as traces of oxygen are available at surfaces or in the atmosphere.
solid solution [3,44,45]. Since other Ni-borides such as Ni2B and Ni4B3
Both H3BO3 and B2O3 tend to evaporate which does not significantly
and even metallic Ni are found after hot-pressing at 1600 °C [46],
contribute to recordable weight-losses [37,38]. These reactions cer-
equilibrium conditions are obviously not easy to obtain. It is therefore
tainly will give rise for the changing slopes of the dilatation rate up to
not surprising that, due to the presence of tungsten and carbon, the τ-
500 °C (Figs. 9 and 12).
phase was not found in our sintering system. Nevertheless, following
As next we consider, that metallic Ni and Co were intentionally
the calculations by Morishita and taking Co, Ti and C into account as
added or picked up as a debris from the hard metal milling beads.

Fig. 20. a -c: SEM-micrographs of 7.5 wt.% Ni samples sintered at (a) 1000, (b) 1350 and (c) 1500 °C for 2 h.

7
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Fig. 21. a -c: Cumulative frequency plots of the grain size. (a) Temperature-dependence of a 5 wt.% Ni composition after 2 h sintering; (b) Time-dependence of a
7.5 wt.% Ni composition at 1500 °C; (c) Dependence from the Ni-content after sintering at 1700 °C for 2 h.

shrinkage starts at 991 °C. As recorded by DTA measurement (Fig. 14),


in Ar atmosphere the maximum of liquid phase formation occurs at
1123 °C, i.e. at slightly higher temperature.
At 1000 °C we recorded the onset of a significant weight-loss (Fig. 7)
as well as the first time the appearance of the ω-phase in the matrix by
XRD after cooling which is a proof that both boron and tungsten have
been dissolved in a liquid. In the ternary B-Ni-W subsystem Morishita
et al. [32] presented a second ternary eutectic E3 with L =
Ni2B + Ni3B + W2NiB2 at Te = 1018 °C which is simultaneously the
eutectic temperature of the L = Ni3B + Ni4B3 binary subsystem
(Fig. 23). At 1031 °C the melting point of Ni4B3 is reached so that the
volume of liquid phase is steadily increasing upon further heating.
Other liquid phase-promoting reactions are: 1093 °C: L = Ni + Ni3B;
Fig. 22. TEM micrograph of the core-rim structure separated by dislocations 1111 °C: L = Ni2B + Ni3B; 1125 °C: Tm of Ni2B; 1127 °C: Tie line WB2-
and stress contours. Image courtesy B. Freitag, Inst. Inorganic Chemistry, Ni crosses the melting groove; 1158 °C: Tm of Ni3B. In conclusion, all Ni-
University of Bonn, Germany. borides, if not dissolved earlier, have now completely been liquefied. In
accordance, we observed the onset of the major endothermic DTA-re-
action at 1130 °C, its maximum speed at 1155 °C and the end at 1200 °C
while the Ni-containing samples with < 50 wt.% Ni now exhibited the
onset of the main shrinkage by liquid phase sintering at 1168 °C. After
sintering at 1100 °C for 20 min solidified Ni3B, Ni4B3, NiB liquid phase
besides TiB2, φ, ω and residual WC and TiC have been identified by X-
ray diffraction (Fig. 24). Please, note that the patterns of the starting
powder contains reflexions of (W,Ti)C after attrition milling which are
quite broad due to the mechanical treatment and the lattice distortion
of the crystal lattices.
Samples sintered at 1200 °C do not show any further weight-loss but
an increasing amount of ω- and γ-phase. This coincides with the end of
the endothermic DTA-reaction. Only in the case of 50 wt.% Ni metallic
Ni and Ni3B can be detected. Hexagonal WC and cubic TiC are still
available but their ongoing dissolution in the melt contributes to the
formation of the ternary phases. At 1216, another weak but long lasting
endothermic reaction was recorded which superimposes the slight in-
crease of the shrinkage rate at 1222 °C for 0% Ni and 1230 °C for 7.5 wt.
% Ni samples, respectively. Ajao [26] reported on a weak endothermic
reaction at 1221 °C during sintering of B + Ni + Ti powder mixes, too.
Going to higher temperatures, the partial dissolution of WC and TiC
gives a significant contribution to the melt volume. According to Zhou
et al. [47], who thermodynamically treated the C–Co–Fe–Ni–W system,
Fig. 23. Liquidus surface of the B-Ni-W system after Morishita et al. [32] in [K],
the equilibrium plane L + WC = (Co, Ni) + WC + C is reached at
completed along the B–W edge. Note that the tie-line WB2-Ni approaches the
1313 °C at which a 50 at.% Co/50 at.% Ni melt is able to partially
low-melting eutectics E2 and E3.
dissolve WC. If the Co/Ni ratio approaches 15 at.% Co/85 at.% Ni, the
temperature increases to 1332 °C (Fig. 25). As the latter temperature is
additional constituents, it is obvious that the first liquid phase appears higher than the binary eutectics Co-C with 1310 °C and Ni-C with
between 950 and 980 °C which now gives rise for liquid phase sintering. 1321 °C [48], all carbon is now either bonded as residual TiC (Fig. 24)
This is the temperature range in which another intermediate tendency or dissolved in the liquid phase. Therefore, graphite was never observed
for shrinkage appears in the dilatometer diagrams of samples con- in all compositions. As the solubility of WC and TiC in the liquid in-
taining up to 12.5 wt.% Ni, while in the 50 wt.% Ni-blends the main creases with higher temperature, the liquid phase volume increases,

8
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

Fig. 26. Ni-B-compounds condensed at the alumina cover of the crucible after
vacuum sintering of the TiB2-Ni/Co-(W,Ti)C composites.

losses by evaporation of Ni- and Co-compounds, respectively. XRD


pattern taken from 7.5 wt.% Ni samples sintered at 1400 °C for 20 min
show φ- and ω-phase besides TiB2 and TiC while only traces of Ni3B are
recognised (Fig. 24). This effect drives, unfortunately, the permanent
formation of new pores. If, for instance, a body with an initial Ni-
content of 7.5 wt.% looses 5 wt.% during sintering for 2 h at 1700 °C
(Figs. 6 and 7), there are not much more than 2.5 wt.% Ni left in the
binder phase. A composition of 12.5 wt.% Ni treated at 1500 °C for 2 h
Fig. 24. XRD pattern of a TiB2-7.5 wt.% Ni sample sintered at 1100, 1400 and contains only 3.7 wt.% Ni. Even in 50 wt.% Ni compositions, there is
1700 °C for 20 min, respectively, in comparison to the as-milled raw material only little Ni3B or Ni2B found after cooling down. An excellent proof for
blend. the Ni-B-exudation is the condensation material collected from the
alumina cover of the crucible which contains Ni and all Ni-borides as
well (Fig. 26). Thus, the samples “dry out” during prolonged sintering
or at higher temperatures. Accordingly, the microstructure consolidates
as TiB2, TiC and φ-/ω-/γ-phase, which arrests almost the entire Ni-/Co-
content in solid state as can be seen from the XRD pattern of 7.5 wt.%
Ni samples sintered at 1700 °C for 20 min (Fig. 24). Thus, depending
from the Ni + Co-content, liquid phase sintering turns into solid state
sintering again, i.e. it is some kind of transient liquid phase sintering. As
a result, the shrinkage rate decreases from -0.75%/min at 1472 °C to
-0.1%/min at 1573 °C. Going to higher temperatures results in a further
increase in relative density, but even sintering at 1900 °C for 2 h yields
the same phase composition and, fortunately, almost the same grain
size.
In conclusion, the addition of Ni to WC-Co-mechanically alloyed
TiB2 powder fosters first solid state sintering and then liquid phase
sintering by lowering the temperatures and by accelerating the
shrinkage. Fu and Koc [7] discussed the minimum amount of Ni or Co
and found that 1–3 wt.% are not sufficient to get more than 89.5–91.5%
th.d. if sintered at 1500 or 1600 °C. In contrast, additions > 3 wt.% Co
result in 98–99% th.d. even if sintered at 1500 °C, while additions of
20 wt.% binder phase did not significantly improve densification. This
is in a general agreement with our results. We found, however, that our
Fig. 25. Isopleth of the W-C-(Ni/Co) system across the quasi-binary equilibrium much coarser powder requests for 5–7.5 wt.% Ni to get sintered while
WC-Liquid at 20 wt.% (Co + Ni) as calculated by Zhou et al. [47]. Note the higher volumes of liquid phase are automatically lowered by the self-
temperature shift of the melting groove as a function of the Co/Ni ratio.
regulation mechanism of exudation.

too. This results in the onset and continuous acceleration of the main
4.2. Microstructural evolution
shrinkage for 0% Ni material at 1304–1323 °C, while for Ni-containing
samples the main shrinkage is already in progress.
As shown in Figs. 21a–c, the grain size of all phases does not sig-
The microstructures of samples sintered between 1200 and 1350 °C
nificantly increase neither by temperature nor by annealing time. In the
are characterised by dense local clusters of liquid (Fig. 18) and evidence
case of solid state sintering it is known that grain growth occurs via
for capillary flow (Fig. 19). At higher temperatures, the weight-loss by
evaporation/re-condensation, surface diffusion, grain boundary diffu-
exudation of volatile Ni and Ni-borides increases again and results in
sion and volume diffusion. As long as open pore channels are available,
channel pore and sinter skin formation. At 1420–1430 °C the maximum
shrinkage rate of -0.75%/min is obtained in the case of Ni-containing evaporation/re-condensation may be active but the temperature is too
low for the refractory TiB2 or WC to coarsen. Thus, this mechanism only
mixes, while samples without Ni-addition show a continuously accel-
erating shrinkage up to 1700 °C and, maybe, higher temperatures. In drives the loss of nickel compounds. Furthermore, the solid state sin-
tering interval is too short (600–700 °C) in order to initiate grain
both cases, the volume shrinkage is, however, not only a liquid phase
sintering phenomenon but has to be attributed to the tremendous mass growth of TiB2 by the other mechanisms as a liquid phase is formed at
rather low temperatures of 950–980 °C. Thus, liquid phase sintering

9
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

changes entirely.
Differential dilatometry was demonstrated to be a powerful tool for
revealing both chemical and sintering reactions by recording dilatation
rate versus temperature. Three main periods of sintering TiB2-Ni/Co-
(W,Ti)C powder blends could be separated. The first interval ranges from
200 to 600 °C, approximately, and encounters reactions with chemically
adsorbed oxygen and moisture. H3BO3, B2O3 and later TiBO3 formation,
melting and evaporation give rise for several shrinkage and swelling
tendencies.
The second interval starts at 600 °C and is governed by the solid state
Fig. 27. Scheme of the liquid phase sintering mechanisms upon heating. sintering of Ni and/or Co, respectively, leading to a maximum densifi-
cation rate at 650–670 °C. From ˜ 800–900 °C on both Ni and Co un-
dergo solid state reactions together with TiB2 which releases boron so
may trigger coarsening by dissolution/re-precipitation mechanisms.
that Ni3B and Ni2B and the corresponding Co-compounds appear as
This happens in fact as can be recognised by the core-rim structure of
new phases. Ni and Co may also react with WC and B to the ternary φ-
(Ti,W)B2 solid solution around unreacted TiB2 particles. The misfit of
phase of WCoB-type and the ω-phase of W2NiB2-type by solid state
lattice constants results, however, in the formation of stresses which
reaction up to 1220 °C.
limit the epitaxial growth of thick layers (Fig. 22). Moreover, the well-
The third reaction and main shrinkage interval comprises liquid phase
known mechanism of “solute-drag” [49] provides a thermodynamic
sintering. Liquid phase formation starts at the ternary eutectic E2: L ↔
reason for preventing grain growth as the (Ti,W)B2 solid solution is
Ni + Ni3B + ω at 987 °C and continues faster as all melt-generating
more stable than the pure phase. Also in case of the ternary ω-phase
quasi-binary and other ternary equilibria are met. As WC/TiC starts to
grain growth is limited to the contact and thickness of the surrounding
dissolve in the liquid phase at ˜ 1300 °C and higher, the melt volume is
liquid phase. In case of a 50 wt.% Ni composition, (W,Ti)2(Ni,Co)B2
steadily increasing. Thus, for Ni-containing compositions liquid phase
grows for several hundreds of micrometres through the continuous
sintering starts at 1163 °C with a maximum at 1420–1470 °C and is
matrix phase incorporating TiB2 particles (Figs. 16 and 17). This kind of
completed at ˜ 1640 °C A transient acceleration of shrinkage between
microstructure implies anyway, that grain growth preferentially occurs
1222 and 1303 °C is attributed to wetting and particle re-arrangement
during cooling down if the solubility of W, Ti and B in the liquid phase
by capillary forces.
becomes continuously smaller or if undercooling a eutectic temperature
The formation of liquid phase is accompanied by a strong weight-loss
forces the solidification of all phases. This can be proven by the strong
from 1000 to 1100 °C on due to the evaporation of Ni- and Co-species.
exothermic reaction during first cooling at 1053 °C (Ni-containing
Therefore, all literature data on that sintering system reveal 2–3% re-
samples) and at 1115 °C for repeated re-heating and cooling.
sidual porosity even, except those using very reactive nano powders.
The individual steps of liquid phase sintering are summarised in
Thus, the optimum amount of Ni or Co for initiating liquid phase sin-
Fig. 27. First, the TiB2 particles are wetted by liquid which infiltrates
tering is between 3 and 7.5 wt.%.
the meniscuses at the grain contacts. Particle re-arrangements occurs by
The residual liquid phase increases in W, Ti, and B concentration
capillary forces. Due to the anisotropic nature of the TiB2 surface en-
which causes precipitation of ω-phase and γ-phase of W3NiB3-stoichio-
ergy and depending from the local amount of liquid, a skeleton may be
metry as well as a (Ti,W)B2 solid solution in form of a core-shell structure.
formed. Next step is the partial dissolution of TiB2, WC and TiC in the
Both effects are responsible for the retarded grain growth of (Ti,W)B2 and
melt and the precipitation of a facetted (Ti,W)B2 rim onto the residual
thus for the homogeneous microstructure found in all corresponding
TiB2 cores which is known from the liquid phase sintering of hard
publications, even at higher liquid phase concentrations or prolonged
metals, SiAlONs and SiCAlONs as well. Also, ω-(W,Ti)2(Ni,Co)B2 or γ-
sintering time.
(W,Ti)3(Ni,Co)B3 are precipitated. At the end, ω-/γ-phase crystals grow
Accordingly, conventional TiB2 and Ni/Co powder mixes can be
further either by dissolution and re-precipitation for lowering the in-
pressureless sintered in the range of 1500–1700 °C. Even without any Ni
terfacial energy or by the reduction of the W-B solubility in the liquid
additive, mechanical alloying by hard metal beads has proven sufficient
due to the evaporation of Ni- and Co-compounds or by temperature
to activate sintering to almost full density if annealed at 1700–1750 °C
decrease during cooling.
for 60–120 min as only a small amount of volatile Co-compounds is
Unfortunately, all liquid phase sintering systems with TiB2 con-
generated.
taining WC, intentionally or not, will suffer from either brittle ternary
Last, former less-readily accessible literature on boron containing
phases or Ni/Co-boride residues. Without any W-compound, typical
materials dating back to ˜1985 and earlier may reveal helpful in-
metal contents required for a successful liquid phase hot-pressing of
formation for the interpretation of phenomena supposed to be novel.
TiB2 are 5–25 wt.% Ni or Co. In order to avoid reactions consuming
TiB2, the borides of Ni or Co were also quite often used [11,18,50–52].
Acknowledgements
By these methods, the sintering temperatures have been decreased from
2100 °C to 1400 °C. In these cases, the liquid phase intensifies the mass
This research was not funded by any external source. Instead, the
transport but causes an accelerated grain growth, too. The micro-
author likes to gratefully acknowledge the strong technical support by
structures of composites prepared by liquid phase sintering are similar
P. König (powder processing), V. Schmitz (dilatometry), P. Schott (X-
to those of hard metals. The TiB2 grain size usually exceeds 20 μm and
ray diffraction), and R. Coenen (ceramography and SEM).
the particles form a rigid skeleton of facetted crystals whereas the
binder, e.g., Ni3B, Ni2B, Ni4B3, or comparable compounds of Fe, Cr, or
References
Co, is the matrix phase.
[1] E.V. Clougherty, R.L. Pober, Physical and mechanical properties of transition metal
5. Conclusion diborides, Nucl. Metall. 10 (1964) 423–443.
[2] R. Telle, L.S. Sigl, K. Takagi, R. Riedel (Ed.), Boron-Based Hard Materials. In: Hard
Materials, Part III, Chpt. 4, Wiley-VCH, Weinheim, 2000, pp. 802–943.
Liquid phase sintering of TiB2 with Ni or Co as additives usually [3] R. Telle, G. Petzow, Strengthening and toughening of boride and carbide hard
results in rather coarse microstructures similar to WC-Co hard metals. material composites. Proc. 3rd Int. Conf. Hard. Mat., Nassau (Bahamas) 1987,
Mater. Sci. Eng. A 105/196 (1988) 97–104.
The grain sizes between 5 and 50 μm have been proven sensitive for [4] J. Gild, Y. Zhang, T. Harrington, S. Jiang, et al., high-entropy metal diborides : a
spontaneous microcracking. With the presence of WC, the situation

10
R. Telle Journal of the European Ceramic Society xxx (xxxx) xxx–xxx

new class of high-entropy materials and a new type of ultrahigh temperature ma- [29] Z.T. Zakhariev, M.S. Ivanova, Hard materials based on cemented TiB2-WB-Co-al-
terials, Nat. Sci. Rep. (2016) Doc. No. 37946 ; DOI : 10.1038/srep37946, 10. loys. Proc. 11th int. Symp. boron, Borides and related compounds, Tsukuba 1993,
[5] M. Sibuya, T. Yoneda, Y. Yamamoto, M. Ohyanagi, Z.A. Munir, Effect of Ni and Co JJAP Series 10 (1994) 230–231.
additives on phase decomposition in TiB2-WB2 solid solutions formed by induction [30] H. Haschke, H. Nowotny, F. Benesovsky, Untersuchungen in den Dreistoffen:
field activated combustion synthesis, J. Am. Ceram. Soc. 86 (2) (2003) 354–356. (Mo,W)-(Fe,Co,Ni)-B, Monatshefte Chemie 97 (5) (1966) 157–168.
[6] M. Sibuya, M. Ohyagi, Effect of nickel additive on simultaneous densification and [31] H. Jedlicka, F. Benesovsky, Nowotny, H.: Die Kristallstruktur des W3CoB3 und der
phase decomposition in TiB2-WB2 solid solutions by pressureless sintering using dazu isotypen Phasen Mo3CoB3, Mo3NiB3 und W3NiB3, Monatshefte Chemie 100
induction heating, J. Eur. Ceram. Soc. 27 (2007) 301–306. (11) (1969) 844–850.
[7] Z. Fu, R. Koc, Pressureless sintering of TiB2 with low concentration of Co binder to [32] M. Morishita, K. Koyama, K. Maeda, G. Zhang, Calculated phase diagram of the Ni-
achieve enhanced mechanical properties, Mater. Sci. Eng. A 721 (2018) 22–27. W-B ternary system, Mater. Trans. JIM 40 (7) (1999) 600–608.
[8] D. Ağaoğullan, I. Gökçe Duman, M.L. Öveçoğlu, Influences of metallic Co and [33] H.-B. Ma, J. Zou, J.-T. Zhu, L.-F. Liu, G.-J. Zhang, Segregation of tungsten atoms at
mechanical alloying on the microstructural and mechanical properties of TiB2 ZrB2 grain boundaries in strong ZrB2-SiC-WC ceramics, Scripta Mater. 157 (2018)
ceramics prepared via pressureless sintering, J. Eur. Ceram. Soc. 32 (2012) 76–80.
1949–1956. [34] L. Silvestroni, S. Failla, V. Vinokurov, I. Neshpor, O. Grigoriev, Core-shell structure:
[9] H. Pastor, Metallic borides: preparation of solid bodies - sintering methods and an effective feature for strengthening ZrB2 ceramics, Scripta Mater. 160 (2019) 1–4.
properties of solid bodies, in: V.I. Matkovich (Ed.), Boron and Refractory Borides, [35] A. Kulpa, T. Troczynski, Oxidation of TiB2 powders below 900°C, J. Am. Ceram.
Springer, Heidelberg, Berlin, New York, 1977, pp. 457–493. Soc. 79 (2) (1996) 518–520.
[10] H.R. Baumgartner, R.A. Steiger, Sintering and properties of titanium diboride made [36] R.A. Andrievskii, Yu.M. Shul’ga, L.S. Volkova, I.I. Korobov, N.N. Dremova,
from powder synthesized in a plasma arc heater, J. Am. Ceram. Soc. 67 (3) (1984) E.N. Kabachkow, G.V. Kalinnikov, S.P. Shilkin, Oxidation behavior of TiB2 Micro-
207–212. and nanoparticles, Inorg. Mater. 52 (7) (2016) 686–693.
[11] Y. Murata, H.P. Julien, E.D. Whitney, Densification and wear resistance of ceramic [37] Y.-H. Koh, S.-Y. Lee, H.-E. Kim, Oxidation behavior of titanium boride at elevated
systems. Part 1. Titanium diboride, Am. Ceram. Soc. Bull 46 (7) (1967) 643–648. temperatures, J. Am. Ceram. Soc. 84 (1) (2001) 239–241.
[12] F. Binder, Refraktäre metallische hartstoffe, RadexRundschau (1975) 531–557. [38] B. Shahbahrami, H. Bastami, N. Shahbahrami, Studies on oxidation behavior of TiB2
[13] R.L. Coble, H.A. Hobbs, Investigation of boride compounds for very high tem- powder, Mater. Res. Innov. 14 (1) (2010) 107–109.
perature applications, in: L. Kaufman, E.V. Clougherty (Eds.), N.T.I.S. Report AD [39] V.V. Skorokhod, G.O. Ranneva, Study of the sintering of nickel powders produced
428006, Clearinghouse for Federal Scientific and Technical Information, by different methods, Poroshkovaya Metallurgiya 3 (15) (1963) 25–29 translated to
Springfield, VA, 197382 120. English in: Sov. Powder Met. and Met. Ceram. 2 [3] (1963) 194-198.
[14] P.S. Kislyi, O.V. Zaverukha, Vacuum sintering of titanium diboride, Poroshkovaya [40] J.D. Schöbel, H.H. Stadelmaier, Die nickelecke im dreistoffsystem nickel-titan-Bor,
Metallurgiya 91 (7) (1970) 32–35 Engl. Translation in: Sov. Powder Metall. Met. Metallwissenschaft und Technik 19 (7) (1965) 715–717.
Ceram. 7 (1970) 549-551. [41] Y.B. Kuz’ma, M.V. Chepiga, An X-Ray diffraction investigation systems Ti-Ni-B, Mo-
[15] S. Baik, P.F. Becher, Effect of oxygen contamination on densification of TiB2, J. Am. Ni-B, and W-Ni-B, Poroshkovaya Metallurgiya 8 (10) (1969) 71–75 (in Russian);
Ceram. Soc. 70 (1987) 527–530. translated to English: Sov. Powder Metall. Met. Ceram. 8 [10] (1969) 832-835.
[16] T. Watanabe, Lattice strain in TiB2 related to initial-stage sintering of TiB2 powder [42] E. Lugscheider, H. Reimann, E. Pankert, Ternary tau-borides in nickel-boron-Alloys
and TiB2-Ni mixed powder under high pressure, J. Am. Ceram. Soc. 60 (3) (1977) stabilized by refractory metals of the 4a and 5a groups, Metall 36 (3) (1982)
176–177. 247–251.
[17] V.F. Funke, S.I. Yudkovskii, G.V. Samsonov, Alloys in the system B-Ti-Fe. Zhur. [43] P. Angelini, P.F. Becher, J. Bentley, et al., Processing and microstructural devel-
Prikladn. Khim. 34 (1961) 1013-1020. Engl. translation in, Russ. J. Appl. Chem. 34 opment in liquid phase sintered and pure TiB2 ceramics. science of hard materials,
(1961) 973–978. Proc. 2ndInt. Conf. Science of Hard Materials, Rhodes, 1985; Inst. Phys. Conf. Ser.
[18] T. Watanabe, S. Kouno, Mechanical properties of TiB2-CoB metal borides alloys, 75, (1986), pp. 1019–1032.
Ceram. Bull 61 (1981) 970–973. [44] P.S. Sklad, C.S. Yust, Characterization of TiB2-Ni ceramics by transmission and
[19] L. Ottavi, J.M. Chaix, C. Allibert, H. Pastor, Thermodynamic guidelines for the li- analytical Electron microscopy, Proc. Int. Conf. Sci. Hard Mat., Bordeaux, (1981).
quid phase sintering of titanium diboride cermets, Solid State Phenom. 25/26 [45] P. Angelini, P.F. Becher, J. Bentley, C.B. Finch, Processing and microstructural
(1992) 543–550. characterization of TiB2 liquid phase sintered with Ni and Ni3Al, MRS Proc 24
[20] L.S. Sigl, Th Jüngling, Effects of carbon and oxygen species on the sintering of (1983) 299–303.
titanium diboride-iron cermets, J. Hard Mater. 3 (1992) 39–51. [46] G. Petzow, R. Telle, S. Somiya (Ed.), New Developments in the Field of Refractory
[21] Th. Jüngling, R. Oberacker, F. Thümmler, L.S. Sigl, Sintering and properties of TiB2- Hard Metals Based on Cemented Borides. Advances in Ceramics, Terra Scientific
Hardmetals, Proc. Int. Conf. Advances in Hard Materials Production, (1992), pp. Publ. Comp., Tokyo, 1987, pp. 131–143.
15/1–15/14. [47] P. Zhou, Y. Peng, Ch Buchegger, Y. Dua, W. Lengauer, Experimental investigation
[22] V. Ghetta, N. Gayraud, N. Eustathopoulos, Wetting of iron on sintered titanium and thermodynamic assessment of the C–Co–Fe–Ni–W system, Int. J. Refractory
diboride, Solid State Phenom. 25/26 (1992) 105–114. Metals Hard Mat. 54 (2016) 60–69.
[23] J.M. Sánchez, M.G. Barandika, J. Gil-Sevillano, F. Castro, Consolidation, micro- [48] H. Ohtani, M. Hasebe, T. Nishizawa, Calculation of Fe-C, Co-C and Ni-C phase
structure and mechanical properties of newly developed titanium diboride-based diagrams, Trans. ISIJ 24 (1984) 857–864.
materials, Scripta Metall. Mater. 26 (1992) 957–962. [49] S.J. Bennison, M.P. Harmer, Effect of MgO solute on the kinetics of grain growth in
[24] I. Azkona, J.M. Sánchez, F. Castro, Abnormal growth of TiB2 crystals during sin- Al2O3, J. Am. Ceram. Soc. 66 (5) (1983) C-90–C92.
tering of TiB2-Ni3(Al,Ti) cemented borides, Metall. Mater. Trans. 36A (2) (2005) [50] S. Takatsu, E. Ishimatsu, Sintering and properties of TiC, TiN and TiB2 alloys with
459–466. refractory metal binders. In: trends in refractory metals, hard metals and special
[25] L. Xi, I. Kaban, R. Nowak, G. Bruzda, N. Sobcak, J. Eckert, Wetting, reactivity, and materials and their technology vol.2, Proc.10th Plansee Seminar (1981) 535–548.
phase formation at interfaces between Ni-Al melts and TiB2 ultrahigh-temperature [51] V.J. Tennery, P.F. Becher, G.W. Clark, C.B. Finch, C.S. Yust, P.S. Sklad, Task 1:
ceramic, J. Am. Ceram. Soc. 101 (2018) 911–918. boride ceramics, in: C.J. McHargue, S. Peterson (Eds.), Metals and Ceramics
[26] J.A. Ajao, Phase transformation in some nickel-rich nickel-boron-titanium hard Division Materials Science Program Annual Progress Report W-7405-Eng-26, Oak
alloys, J. Alloys Comp. 493 (2010) 314–321. Ridge Natl. Lab. Publ., 1981, pp. 32–46. TM-7970, Sept..
[27] Z. Fu, R. Koc, Synthesis of TiB2 from a carbon coated precursors method, J. Am. [52] R. Telle, Ceramics for high-tech applications, in: T. Valente (Ed.), Material Science,
Ceram. Soc. 100 (6) (2017) 2471–2481. European Concerted Action COST 503 Powder Metallurgy - Powder Based
[28] A. Márquez-Herrera, G. Bermúdez-Rodríguez, E.N. Hernández-Rodríguez, Materials, vol. 5, European Community Directorate-General, Science, Research and
M. Melendez-Lira, M. Zapata-Torres, Boride coating on the surface of WC–Co-based Development, Brussels, 1997296pp.
cemented carbide, Int. J. Mater. Res. 107 (7) (2016) 676–679.

11

Вам также может понравиться