Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Powder Metallurgy

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/ypom20

Microstructural observations of high temperature


creep processes in hardmetals

K. Mingard , S. Moseley , S. Norgren , H. Zakaria , D. Jones & B. Roebuck

To cite this article: K. Mingard , S. Moseley , S. Norgren , H. Zakaria , D. Jones & B. Roebuck
(2021): Microstructural observations of high temperature creep processes in hardmetals, Powder
Metallurgy, DOI: 10.1080/00325899.2021.1877866

To link to this article: https://doi.org/10.1080/00325899.2021.1877866

Published online: 29 Jan 2021.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ypom20
POWDER METALLURGY
https://doi.org/10.1080/00325899.2021.1877866

RESEARCH ARTICLE

Microstructural observations of high temperature creep processes in


hardmetals
K. Mingarda, S. Moseleyb, S. Norgren c
, H. Zakariaa, D. Jonesa and B. Roebucka
a
National Physical Laboratory, Teddington, UK; bHilti AG, Schaan, Liechtenstein; cSandvik & Coromant R&D, Stockholm, Sweden

ABSTRACT ARTICLE HISTORY


High-temperature properties of hardmetals are critical to their use in many applications but a Received 8 December 2020
challenge to measure accurately. Creep behaviour is not well understood so this work has Revised 14 January 2021
studied uniaxial tensile testing of small simple geometry samples to look at how Accepted 14 January 2021
modifications to the microstructure can affect creep behaviour at temperatures between
KEYWORDS
800 and 900°C. In particular, a carbon-ladder series with high, medium and low carbon Hardmetals; creep;
contents in the 10wt-%Co binder has been investigated. Significant differences between the microstructure; WC-Co;
stress–strain curves of the different carbon contents have been observed, but the carbon-ladder; miniaturised
underlying microstructural mechanisms appear to be similar in detailed large area testing
examination of samples after failure. Penetration of Co along WC-WC boundaries with
‘precipitation’ of discrete islands is seen as well as the formation of continuous thin lamellae
while void formation tends to occur at WC-Co boundaries. EBSD mapping suggests Co
penetration varies as a function of WC-WC misorientation.

1. Introduction conditions as creep, high-temperature deformation


The high-temperature properties of hardmetals or has also been studied by examination of the micro-
cemented carbides are often critical for their successful structure of cutting tools after use [4,5]. The exact
use in conditions such as drilling in the mining and temperatures experienced and the stress state were
construction industries, where high loads and abrasive not easily defined although the latter was probably
conditions and temperatures in excess of 800°C can be generally compressive. A significant number of binder
encountered. However, measurement of properties phase lamellae were observed (more than in [2]) in
such as creep under load at high temperature is SEM and TEM studies and WC was seen to have
difficult, as is obtaining an understanding of the mech- deformed with extensive slip in one case [4] but not
anisms by which the materials fail to enable selection in the other [5]. Additional understanding was sought
of hardmetal grades with longer or more effective using three point bending and mechanical spec-
operational life. troscopy [6] from which it was suggested that ≈
Most reported work has concentrated on analysis of 800°C was a transition temperature between brittle
the mechanical properties measured between 750°C fracture of the entire WC-Co skeleton and plastic
and 1350°C. Direct microstructural investigation of deformation of Co and the WC skeleton enhanced
the creep processes has however largely looked at by the presence of Co.
temperatures above 1000°C. Lay et al. [1] used Trans- Useldinger and Schleinkofer [7] investigated com-
mission Electron Microscopy (TEM) on 3 point bend pressive creep at slightly lower temperatures, 700–
test samples between 1050°C and 1350°C as well as 950°C, and deduced from measurements of strain
measuring stress–strain curves and observed dislo- rates that diffusion control was a more significant
cation behaviour in WC grains. Coupling observations creep mechanism in high binder content (>8 wt-%)
with measured activation energies, they concluded materials, with grain boundary sliding more signifi-
creep was controlled by diffusion of W and thus dislo- cant at lower contents. However, WC grain size
cation movement in the WC phase with grain bound- measurements before and after indicated no changes
ary sliding only occurring during stage II of the creep as the result of creep. The effect of binder phase com-
curve. Yousfi et al. [2,3] studied compressive creep at position became more significant at higher tempera-
1000–1100°C and reported WC grain growth perpen- tures and at high binder phase contents with a
dicular to the load axis as well as observing a small significant reduction in creep in a 30wt-% CoNiCr
number of narrow binder phase lamellae (20 nm and binder. Buchegger and Lengauer [8] measured creep
greater) forming between WC grains and break-up in a 3 point bend test system at lower temperatures
of the large Co grains. While not defining the stress for a wide range of binder phase compositions to

CONTACT K. Mingard ken.mingard@npl.co.uk National Physical Laboratory, Hampton Road, Teddington, Middlesex, TW11 0LW, UK
© 2021 Institute of Materials, Minerals and Mining Published by Taylor & Francis on behalf of the Institute
2 K. MINGARD ET AL.

determine creep activation energies and found similar cross sectional views perpendicular to the polished
creep rates in for Co, FeCoNi and austenitic FeNi at surface.
725°C, but significantly reduced elongation to failure Table 1 lists the different grades strained in the
in the latter two binders. The effects of different stress ETMT and examined by SEM, naming the grades by
states and strain rates, as well as the effect of tempera- use of the approximate weight % binder phase and
ture and load could account for differences in reported the grain size range descriptions given in ISO4499.
behaviour, as discussed by Roebuck and Moseley [9]. The 10 M grade was produced with 3 different carbon
The current work has sought to investigate the concentrations, described here simply as L, M and H
microstructural processes occurring in the lower to indicate low, medium and high carbon concen-
temperature regime (800–900°C) using high resol- trations in the binder phase. The low and high carbon
ution scanning electron microscopy (SEM) on well- concentrations were aimed to achieve binder concen-
polished samples of a range of hardmetal grades. trations close to the eta-phase and graphite phase
The creep testing itself was carried out on small boundaries but still within the WC-Co phase region;
samples of each grade by resistive heating of the no evidence of either eta phase or graphite was
samples in a thermomechanical testing system, with observed in the polished microstructures.
most samples strained in tension to failure. Results of the microstructural examination are pre-
sented first to show the general features observed,
before relating these observations to variation in
stress–strain curves measured on different samples.
2. Experimental details
Samples for creep testing were cut by electro-discharge
machining from larger rectangular blocks before being 3. Results
ground along the four long faces to produce testpiece
3.1. Microstructural examination
bars 40 × 2×1 mm. Creep testing was carried out in
an NPL designed electro-thermomechancial test system The effect of creep on the microstructure of a 6wt-%
(ETMT). This system strains the sample between Co F grade hardmetal is shown in Figure 2 after it
water-cooled grips while heating it with a direct electric had been subject to a final tensile step creep load of
current, all held within a reducing Ar/5%Hydrogen 720 N at 800°C (which resulted in a strain of 1.7%).
atmosphere. Plastic strain is measured by changes in The structure typical of the material before creep
electrical resistance between contacts approximately can be seen from the example in Figure 2(a,b), taken
3 mm apart within the central heated zone (uniform 2 mm outside of the high-temperature zone, 4 mm
within ≈ +/−5°C over 4 mm). Further details of this away from the creep fracture face at the centre of the
equipment are given in [9]. sample. Boundaries between the WC grains and
Figure 1(a) shows a typical testing cycle, in which between WC and the Co binder are all completely con-
the sample is heated rapidly (5 K/s) from room temp- tinuous, as expected.
erature to the test temperature. A constant tempera- At a distance of 1 mm from the creep fracture
ture is then maintained while the tensile load on the (Figure 2(c,d)), which would be within 5°C of the
sample is incremented in 50 N steps, maintaining a peak temperature, there is a clear change at a number
constant load at each increment and measuring the of the WC-WC boundaries. Figure 2(d) shows
strain from changes in the electrical resistance (Figure examples at high magnification, at top left and bottom
1(b)). A stress–strain curve can then be derived from right, of the formation of a series of small rounded fea-
these measurements (Figure 1(c)). Stresses at failure tures in a line along the WC-WC boundaries. Less fre-
were of the order of 500–700 MPa with typical strain quent were continuous gaps, an example of which is
rates of between 1 × 10−4 and 1 × 10−5 s−1. Electrical arrowed in Figure 2(c). Although it cannot be deter-
noise during the hold at each load produces the mined from these images, it will be shown later that
short horizontal lines in the vertical part of the both the small individual features (described in the
stress–strain curve before the measurable deformation following as precipitates) and continuous gaps (lamel-
is observed. lae) are not voids but are filled by the Co binder phase.
After testing to failure, one half of each sample was These features were also visible very close to the
mounted and polished by conventional mechanical fracture (Figure 2(e,f)), perhaps with an increase in
metallographic processes, finishing with a broad the thickness of the Co precipitates and lamellae.
argon ion polish [Hitachi IM4000]. The as-polished What is more noticeable is the increase in number
microstructure was examined in a Zeiss Supra 40 of what are clearly voids, between WC grains and at
SEM and mapped by electron backscatter diffraction WC-Co boundaries, and now extending between mul-
(EBSD). The fracture face of the other sample half tiple WC grains. A few small voids are visible in Figure
was also examined in the SEM. A Zeiss Auriga 2(c), confined close to triple points of WC-WC and
Focused Ion Beam (FIB) SEM was used to prepare binder, but in Figure 2(e) these become more frequent
POWDER METALLURGY 3

Figure 1. (a) time base profile of temperature (upper flat line) and load control (stepped line) with consequent increase in grip
displacement (lower curving line), (b) resulting measure of plastic strain calculated from resistivity and (c) stress strain curve.

and much larger, extending between multiple WC progress further into the material. One side of the
grains. boundary always appears straight, while the other side
Figure 3 shows that the formation of precipitates, is irregular from the Co features extending into the
lamellae and voids are common across a range of WC grain.
WC grain sizes and Co binder phase fractions. The It may be noted that in the above example, as at
previous example was from a 6wt-% Co F grade: other similar sites, the Co features are not associated
Figure 3 shows examples in 10 or 11wt-% Co binders with voids. The voids are often seen as small narrow
from both coarser (10M, Figure 3(a,b)) and finer grade gaps occurring at WC-Co boundaries, as for example
(11F and 10UF), Figure 3(c,d) materials. the feature ringed in Figure 4(a). This void increases in
The previous views are clearly 2D sections but by size, apparently at the expense of the Co binder phase,
use of the FIB a section perpendicular to the polished until it extends between a number of WC grains
surface of the 6F sample was milled close to the frac- (Figure 4(d)) before shrinking again to lie again
ture face. By milling of 100 nm slices and imaging mostly on a WC-Co boundary.
each new face revealed, a 3D understanding could be Figure 5 shows images of the creep failure fracture
built up of the structure. The FIB section also pro- face of the 6F grade which add to the three-dimen-
duced a very planar surface and, coupled with high sional understanding of the Co distribution between
magnification imaging, showed clearly that the fea- WC grains, even though only one WC grain is visible.
tures at the WC boundaries were Co. This is seen in In addition to the expected necking of Co typical of
Figure 4, in which representative images are shown monotonic fracture of WC-Co, in Figure 5(a), small
after removal of approximately 500 nm (or every 5 features appear dotted across several WC grain facets.
slices) of material which enables tracking of features In some cases these features are densely packed (to the
at a WC-WC boundary across the width and depth left of ‘1’), and in others more widely separated (at ‘2’).
of the boundary. An example WC-WC grain bound- To the right of ‘1’ and at ‘3’ the WC facets are not com-
ary towards the top of Figure 4(a–e) is arrowed and pletely covered, and while it is difficult to interpret the
is best described by working backwards from Figure contrast with certainty, the features appear in recesses
4(e). Initially (Figure 4(e)) a thin continuous band of (unlike at ‘2’ where they sit on a flat surface. The fea-
Co (lamella) extends across the small contact between tures in Figure 5(a) are all randomly distributed, but
the grains. This continuous band is still present when the large facet shown in Figure 5(b) clearly has
there is much greater contact between grains (Figure 4 elongated features in ordered rows.
(d)) but is of variable thickness. Finally, Figure 6 shows that a 10% Ni binder phase
After even more milling, further still along the shows very similar behaviour to the Co samples. In the
boundary, the Co only exists in isolated pockets or pre- polished face of Figure 6(a), taken close to the creep
cipitates, and these become smaller and clustered nearer fracture face, binder phase penetration of WC-WC
one end of the WC-WC boundary as the sections boundaries is clearly visible, as are voids, although in
the area imaged these are present at possibly more of
a mix of WC-WC and WC-Ni boundaries. Figure 6
Table 1. Sample microstructure details. (b) shows the creep fracture face of this alloy and
Hardmetal Binder volume Mean WC grain size (linear again one WC facet is clearly covered in discrete fea-
grade fraction (%) intercept) (µm) tures formed by the binder phase.
6F 10.1 1.1
10M (L/M/H) 16.1 1.6/1.7/1.8
10UF 16.3 0.4 3.2. EBSD analysis
11F 17.8 0.9
11Ni 20.0 2.8 Figure 7 shows a high-resolution SEM image of the
10SM 16.3 0.5
10M grade after creep testing which, when inspected
4 K. MINGARD ET AL.

Figure 2. Typical examples of microstructure after creep failure in a 6 wt-% Co F grade, at reducing distances from the final creep
fracture. The tensile direction is horizontal. (a, b) 4 mm from fracture (and thus effectively outside the high temperature zone,
unaffected by creep), (c, d) 1 mm from fracture, within the high temperature zone and (e, f) 0.25 mm from fracture face. Micron
marker bar in 2 (a, c and e) is 1 µm; the location of the enlargements in 2 (b, d and f) is shown by the box in the corresponding
lower magnification image. Arrows indicate features discussed in the text.

at higher magnification, enabled the identification of


3.3. Stress–strain curves
all the grain boundaries showing penetration by Co.
It was observed that the smallest angle between the Having established some of the basic microstructural
penetrated boundaries and the tensile direction (hori- features of creep which appear common to a wide
zontal in the image) was 38° so there was a clear tendency range of hardmetal grades, the following section
for these boundaries to tend towards perpendicular to makes comparisons between similar grades to see if
the applied stress. However, some of these perpendicular small variations can account for the difference in
boundaries did not show Co penetration. These are observed mechanical behaviour. Figure 9 shows
identified by a white ring in Figure 7, and matching stress–strain creep curves for the three carbon ladder
these to the grain orientations revealed by the EBSD samples of the 10 M grade, acquired at both 800 and
map of the same area in Figure 8, shows that several of 900°C. At the higher temperature there is an obvious
these boundaries had prism-prism, basal-basal or trend of increasing resistance to creep with decreasing
prism-basal plane orientation relationships. carbon content. The medium carbon grade sample did
POWDER METALLURGY 5

Figure 3. Voids and Co precipitates or lamellae observed in three Co binder hardmetal grades (a, b) 10M, (c) 11F and (d) 10UF.
Arrows indicate grain boundary penetration by the binder phase.

not then follow this trend at 800°C but the difference most form between WC grains, particularly at triple
between low and high is repeated with the high carbon points between multiple WC grains. By contrast
sample showing, at both temperatures, a greater ten- about 25% of the voids in the high carbon sample
dency to deform at lower stresses during stage II form along WC-Co boundaries. Whether these differ-
creep as well as failing at a lower overall stress in ences are an effect of testing the high carbon samples
stage III. Higher carbon contents reduce the liquidus at a higher homologous temperature or other changes
temperature of the binder phase, so it is possible that to properties of the binder phase or interface strengths
the reduction in strength of the high carbon samples needs further investigation.
is at least in part due to being tested at a higher hom-
ologous temperature.
4. Discussion
Figures 10 and 11 show high-resolution images of
large areas of the low and high carbon 800°C samples, In relating the microstructural behaviour observed
respectively. These were acquired by stitching together here to possible mechanisms for creep, it must be
smaller images with a 5 nm pixel size to give total areas noted that the fracture face and regions close to it
of ≈ 60 × 60 µm and several thousand grain bound- will show features of both secondary and tertiary
aries to inspect at resolutions similar to that shown creep under tensile loading. The key features observed
in Figure 2. While not visible at the resolution repro- in this work are the penetration of Co on WC-WC
duced in the figures, close inspection shows that, as grain boundaries and the formation of porosity,
might be expected from the previous results, large often at WC-Co boundaries. Both features have been
numbers of boundaries show Co penetration as lamel- observed before; Co penetration [2–5] and porosity
lae or discrete islands in both samples, the majority or grain boundary decohesion in [2,3], the latter
perpendicular to the stress axis. Where there is a sig- clearly under compression and secondary creep but
nificant difference between the two samples is, how- in three point bending and turning studies [4–6] the
ever, in the voids formed at WC boundaries. In the stress state and creep stage is less clear cut. Most
high carbon sample, at least 47 examples of voids reports of property measurement concentrate on the
can be found, but more than 60% fewer are seen in steady state secondary creep and discussion of mech-
the low carbon sample; in the low carbon sample, anisms also assumes this stage.
6 K. MINGARD ET AL.

Figure 4. (a–f) Series of cross sections milled by FIB in the 6F grade from the top surface of the sample shown here on the left of
the figures (tensile direction is vertical, along the long edge of each image).

However, the following examples from a 10SM which was initially strained at 400 MPa before increas-
grade, in which one sample was strained to failure (ter- ing to 450 MPa and allowing continued strain into the
tiary creep) and a second halted after secondary creep tertiary creep regime until failure at about 1% strain.
before failure, suggests that valid comparisons can be The images in Figure 13 show that voids (Figure 13
made with earlier work: both samples, tested at 900 (a,b)) and Co penetration (Figure 13(c,d)) are seen in
°
C show the same features of voids and grain boundary both (the fractured and interrupted test) samples with
penetration already described even though the first a similar appearance to these features shown in the
had only been tested to the late secondary stage. The previous figures from other fractured samples. The
similarity of the strain applied to these two samples density of voids in the interrupted test is lower than
before the tertiary creep stage is seen in Figure 12. in the fractured sample, just as it is at a distance of a
The lower curve shows strain against time for the few millimetres from the fracture of the failed speci-
interrupted test sample for which loading was men. From the similarity of the features seen after sec-
removed after ≈0.65% strain produced by a stress of ondary and tertiary creep, it would seem reasonable to
450 MPa. The upper curve shows an identical sample conclude that comments on the mechanisms
POWDER METALLURGY 7

Figure 5. SEM images of two sites on the creep fracture face of the 6F sample seen in Figure 2. (a) small islands or precipitates of
Co seen at random positions on faces of WC grains (see text for explanation of numbers), (b) aligned islands of Co on one WC grain
face.

Figure 6. Ni binder phase sample. (a) polished section close to fracture face (tension horizontal) with lamealle (black arrows),
precipitate (white arrows) and cracks/voids visible at WC-Ni and along WC-WC boundaries. (b) fracture face showing densely
packed precipitates, aligned, on one facet.

Figure 7. 10M grade with grain boundaries penetrated by Co ringed in black. Example boundaries close to perpendicular to stress
with no Co penetration ringed in white.

suggested by the failed samples described in the results Co-WC boundaries (e.g. top right of Figure 2(d),
are still relevant to a general discussion of creep in the Figure 3(b,a)) while other more crack-like voids at
context of earlier reported observations. WC-WC boundaries or triple points. The latter
Void or porosity formation leading to failure is a would be consistent with the W-type wedge-shaped
common feature in tertiary stage creep of metals. crack formation mechanism when sliding on some
From the 2D images shown in this paper, it might grain boundaries interacts with a non-deforming or
be possible to conclude that some voids form at static grain [10]. However, the 3D interpretation
8 K. MINGARD ET AL.

Figure 8. EBSD map of area in Figure 7. (a) shows a cluster of slightly misoriented grains, with Co penetration between grains. (b–
d) show example boundaries with no Co penetration and with orientations close to basal-prism, prism-prism and basal-basal.

Figure 9. Creep curves for low, medium and high carbon binder phase samples (L, M, H) 10 M grade samples, for (a) 800°C and (b)
900°C. (lower case letters in graph legend refer to specific sample.

Figure 11. High carbon 10M grade with 47 voids ringed. 12


Figure 10. Low carbon 10M grade with 21 voids ringed. 3 WC-Co WC-Co boundary voids ringed in blue. Red lines show Co
boundary voids ringed in blue. Red lines show Co penetration. penetration.

possible from the sequence in Figure 4 suggests that and morphology of the voids at the WC-Co bound-
narrow cracks at WC-WC boundaries are linked to aries would seem to rule out the suggestion [3] of
larger voids formed at WC-Co boundaries. The size intergranular cavities arising from unaccommodated
POWDER METALLURGY 9

Figure 12. 10SM samples strained to failure (black, upper Figure 14. Contrast variation in Co binder (cross hatched
curve) and interrupted after only secondary creep (green bright and dark bands) of 10M grade showing hcp formation
lower curve). in what was uniform contrast at room temperature from
nearly fully fcc binder, resulting from deformation at 800°C
during creep.
grain boundary sliding. The alternative explanation is
the significant diffusion of vacancies, along grain
the predominant observation in the current case is of
boundaries, and possibly generated by grain boundary
separate pockets or precipitates of Co along the length
sliding. Vacancy diffusion implies the mass transport
of a boundary, with a tendency to bulge into one grain
of Co in the opposite direction.
only while leaving a flat facet on the other, which
Indeed the penetration of Co between WC grains
matches with the TEM observation of [5]. The fracture
must clearly be by diffusion, with an additional driving
face observations suggest these precipitates can be
force of the applied stress since penetration was only
either randomly arranged in small islands or elongated
observed on boundaries closer to perpendicular to the
and aligned with one another.
tensile axis and thus more likely to be ‘opened’ by the
The implication of the precipitate morphology is
applied stress. Uniform thickness lamellae of Co have
that the grain boundary habit planes are probably
been seen in both SEM and TEM samples from com-
important in controlling the pattern of Co
pression [2,3] and machining samples [4,5]. However,

Figure 13. Secondary electron image (a) of tertiary creep 10SM sample strained to failure and (b) of secondary creep 10SM sample
at an equivalent (central) position to that of (a). (c, d) Backscatter electron images showing Co precipitation between WC grains.
10 K. MINGARD ET AL.

Figure 15. Examples of Co penetration in tensile creep at 800°C between irregular WC-WC boundaries which suggest some grain
boundary sliding may occur.

precipitates. Ostberg and Andren [5] show an enabling creep in the current work. The temperature
example where the flat face was a prism plane and of the tests is only 800°C which is much lower than
in one case the development of new facets on the that in [1–3] and based on the summary by Mari
bulge faces. The EBSD mapping in Figure 8 was [11], this is below the temperature at which grain
not carried out with the resolution that would be boundary sliding (and certainly any WC deformation)
needed to look at the facet formation of the bulges, would be possible.
but analysis of the orientation of the flat face is At this temperature, Mari suggests deformation of
possible. Of the 41 boundaries identified with lamel- the binder phase will be responsible. The image con-
lae in Figure 7, 13 could be positively identified as trast seen in the binder phase in Figure 14 (an enlar-
having a flat face with bulges into the other grain gement from Figure 7) supports this mechanism,
and of these, 7 had a prism plane parallel to the with the binder phase that was largely FCC before
flat facet, 3 a basal plane and 3 no obvious align- creep now showing multiple slip bands visible,
ment with any plane. Of the remaining infiltrated indexed in many cases as the HCP phase by
boundaries, most were not resolved clearly enough EBSD. Some caution is needed with the interpret-
in the SEM image to determine if one face ation as binder phase transformation does occur
remained flat. A few lamellae were neither paral- on cooling from creep, and comparative studies on
lel-sided or flat/bulging and indeed were curved. non-crept samples would be needed to see if the
The observations from the EBSD map of some creep stress induces more HCP phase or HCP with
boundaries that have not been penetrated with a a different morphology than simple cooling of an
possible CSL2 tilt or twist type orientation between unstressed sample.
the grains is in agreement with several other obser- Some of the creep could be accounted for by void
vations [1,4]. formation, but the voids make up <0.5% by area so
Ostberg and Andren suggest boundary infiltration substantially more of typically 1–1.5% strain must be
is stress assisted dissolution. In the current case, the accounted for. The lamellae formation could just
majority of the infiltrated boundaries were suitably result from dissolution of the WC grains, but the
oriented, nearer to perpendicular than parallel to examples in Figure 15 suggest some grain boundary
the tensile axis, but the discontinuous Co suggests sliding with the lamellae then accounting for some
that both diffusion between WC-WC boundaries of the strain. While the void formation cannot account
with some dissolution at specific points on the for the full strain it is probably significant that the
boundaries. Diffusion of Co along WC-WC bound- number of voids increases with the increasing carbon
aries away from the main Co skeleton (and thus content of the binder phase.
vacancy diffusion in the opposite direction) might
be expected to account for some of the void for-
mation seen, but it will be noted that precipitates
5. Conclusions
arrowed in Figure 4 are not associated with voids
at adjacent interfaces. Tensile creep behaviour of a range of hardmetal grades
The Co penetration has been suggested to be a has been studied between 800 and 900°C. This temp-
necessary precursor to grain boundary sliding, but in erature is low relative to many previous microstruc-
this stress state, only a few of these penetrated bound- tural studies carried out at 1000°C or more. Despite
aries have much alignment with the tensile axis so this this, similar characteristic features have been observed
does pose the question of what mechanisms are in the final microstructure:
POWDER METALLURGY 11

(1) Co penetration along grain boundaries, most Funding


commonly as multiple discrete precipitates rather This work was supported by Department for Business,
than the straight lamellae previously reported Energy and Industrial Strategy [NMS].
(2) Voids which, when considered as 3D features, are
generally largest at WC-Co interfaces, narrowing
to smaller cracks at WC-WC interfaces ORCID
S. Norgren http://orcid.org/0000-0003-1471-8307
The Co precipitate structure is frequently
related to the boundary orientations of the two
WC grains between which they form. They appear References
to grow into grains with random boundary orien-
[1] Lay S, Vicens J, Osterstock F. High temperature creep
tations and not into those with basal or prism
of WC-Co alloys. J Mater Sci. 1987;22:1310–1322.
habit planes. [2] Yousfi MA, Nordgren A, Falk LKL, et al. Evolution of
Void formation is difficult to explain as (a) they the microstructure during creep testing of WC-Co
appear large relative to the likely flux of vacancies based cemented carbide, Proceedings of the 18th
from WC-WC boundaries (and are not always near Plansee Seminar 2013, Reutte; Austria.
WC-∼WC boundaries with precipitates) and (b) do [3] Yousfi MA, Weidow J, Nordgren A, et al. Deformation
mechanisms in a WC–Co based cemented carbide
not appear to form by wedging open by rigid WC during creep. Int J Refract Met Hard Mater.
skeletons. 2015;49:81.
The same features are observed in grain sizes ran- [4] Östberg G, Buss K, Christensen M, et al. Mechanisms
ging from 0.4 to 2.8 µm and binder phase fractions of plastic deformation of WC–Co and Ti(C, N)–WC–
of 6–11wt-%, and also in samples with a Ni binder Co. Int J Refract Met Hard Mater. 2006;24: 135–144.
[5] Östberg GK, Andrén H-O. Microstructural changes
phase. Evidence of binder phase deformation has
during wear by plastic deformation of cemented car-
been observed in a Co containing material, but bides and cermet cutting inserts. MetallMater Trans.
void formation and grain boundary sliding may 2006;37A:1495–1506.
also be occurring despite the relatively low creep [6] Mari D. Understanding the mechanical properties of
temperature. Despite the similar mechanisms hardmetals through mechanical spectroscopy. Mater.
observed, the carbon content of the Co binder Sci. Eng. 2009;A521-22:322–328.
[7] Useldinger R, Schleinkofer U. Creep behaviour of
phase was shown to have a significant effect on the cemented carbides – Influence of binder content, bin-
creep strength, with high carbon contents failing at der composition and WC grain size. Int J Refract
lower stresses. MetHard Mater. 2017;62:170–175.
[8] Buchegger C, Lengauer W. ‘Creep behaviour of hard-
metals with alternative binder alloys at elevated temp-
Acknowledgements eratures’; Proceedings of the 18th Plansee Seminar
2013, Reutte; Austria.
Funding by Hilti, Sandvik and Ceratizit and from the
[9] Roebuck B, Moseley S. Tensile and compressive asym-
National Measurement System of the UK Government
metry in creep and monotonic deformation of WC/Co
Department for Business, Energy and Industrial Strategy is
hardmetals at high temperature. Int J Refract Met
gratefully acknowledged. Members of the British Hardme-
Hard Mater. 2015;48:126–133.
tals Research Group are thanked for the provision of
[10] Chang HC, Grant NJ. Mechanism of intercrystalline
samples.
fracture. Trans AIME J Met. 1956;8:544–551.
[11] Mari D. ‘Mechanical behaviour of hardmetals at high
Disclosure statement temperature’, p. 405–421, in Comprehensive Hard
Materials. Vol. 1 ed. D. Mari, L. Llanes and C.E.
No potential conflict of interest was reported by the author(s). Nebel, Elsevier.

You might also like