0% found this document useful (0 votes)
18 views13 pages

Thermal Treatment Effects on Cobalt Behavior

This study investigates the effects of thermal treatment on the microstructure and mechanical behavior of polycrystalline cobalt, focusing on the fcc↔hcp phase transformation. Isothermal annealing and thermal cycling were performed, revealing that grain size increases and fcc phase content decreases with higher temperatures and cycles, affecting yield strength and fracture strain. The findings suggest that controlling the fcc phase content can optimize the mechanical performance of cobalt for various applications.

Uploaded by

manoranjan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Topics covered

  • conclusions,
  • research findings,
  • deformation dynamics,
  • results interpretation,
  • mechanical properties,
  • ductility,
  • thermal hysteresis,
  • recrystallization,
  • yield strength,
  • phase transformation
0% found this document useful (0 votes)
18 views13 pages

Thermal Treatment Effects on Cobalt Behavior

This study investigates the effects of thermal treatment on the microstructure and mechanical behavior of polycrystalline cobalt, focusing on the fcc↔hcp phase transformation. Isothermal annealing and thermal cycling were performed, revealing that grain size increases and fcc phase content decreases with higher temperatures and cycles, affecting yield strength and fracture strain. The findings suggest that controlling the fcc phase content can optimize the mechanical performance of cobalt for various applications.

Uploaded by

manoranjan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Topics covered

  • conclusions,
  • research findings,
  • deformation dynamics,
  • results interpretation,
  • mechanical properties,
  • ductility,
  • thermal hysteresis,
  • recrystallization,
  • yield strength,
  • phase transformation

materials

Article
The Effect of Different Thermal Treatment on the
Allotropic fcc↔hcp Transformation and Compression
Behavior of Polycrystalline Cobalt
Michal Knapek 1,2, * , Peter Minárik 1 , Patrik Dobroň 1,2 , Jana Šmilauerová 1 ,
Mayerling Martinez Celis 1,3 , Eric Hug 3 and František Chmelík 1
1 Department of Physics of Materials, Faculty of Mathematics and Physics, Charles University,
Ke Karlovu 5, 12116 Prague, Czech Republic; [email protected] (P.M.);
[email protected] (P.D.); [email protected] (J.Š.);
[email protected] (M.M.C.); [email protected] (F.C.)
2 Nuclear Physics Institute of the Czech Academy of Sciences, Husinec—Řež 130, 25068 Řež, Czech Republic
3 Laboratoire de Cristallographie et Sciences des Matériaux, Normandie Université, CNRS UMR 6508,
6 Boulevard Maréchal Juin, 14050 Caen, France; [email protected]
* Correspondence: [email protected]

Received: 28 November; Accepted: 13 December 2020 ; Published: 17 December 2020 

Abstract: Pure polycrystalline cobalt is systematically thermally treated in order to assess the effect
of the microstructure on the compression behavior. Isothermal annealing of the as-drawn material
leads to recrystallization and grain growth dependent on the annealing temperature (600–1100 ◦ C).
Consequently, the yield strength decreases and the fracture strain increases as a function of rising
grain size; the content of the residual fcc phase is ∼6–11%. Subsequent thermal cycling around the
transition temperature is applied to further modify the microstructure, especially in terms of the fcc
phase content. With the increasing number of cycles, the grain size further increases and the fraction
of the fcc phase significantly drops. At the same time, the values of both the yield strength and
fracture strain somewhat decrease. An atypical decrease in the fracture strain as a function of grain
size is explained in terms of decreasing fcc phase content; the stress-induced fcc→hcp transformation
can accommodate a significant amount of plastic strain. Besides controlling basic material parameters
(e.g., grain size and texture), adjusting the content of the fcc phase can thus provide an effective
means of mechanical performance optimization with respect to particular applications.

Keywords: cobalt; phase transformation; microstructure; deformation

1. Introduction
Metals with a hexagonal close-packed (hcp) crystal lattice belong to the prominent engineering
materials used in the transport, aerospace, environmental, and information technology industries.
For example, in the production of vehicles and airplanes, magnesium- and titanium-based materials
deliver an excellent strength-to-weight ratio [1]. Furthermore, in the memory and data storage
industries, miniaturized devices are being developed, which utilize cobalt and other hcp metals [2,3].
The design and reliability of such devices are essentially related to the understanding of their
mechanical behavior, especially of the dynamics of the deformation mechanisms involved with respect
to the material microstructure.
Pure cobalt features an almost ideal ratio of the lattice parameters (c/a), very low stacking fault
energy, and a high melting point compared to other metals with an hcp structure [4]. hcp metals
typically exhibit complex deformation dynamics due to only a limited number of slip systems, unlike
face-centered cubic (fcc) metals [5]. Systematic studies of mechanical properties in relation to the

Materials 2020, 13, 5775; doi:10.3390/ma13245775 www.mdpi.com/journal/materials


Materials 2020, 13, 5775 2 of 13

microstructure of polycrystalline cobalt are, however, very scarce. Several existing studies mostly
date back to the 1950s–1970s when the microstructural observations were still rather limited [6,7].
Other studies focused on cobalt thin films [8], micro-/nano-crystalline samples [9–11], or single
crystals [6,12,13] or investigated the activity of various deformation modes including mechanical
twinning mostly in tension [14–18].
Investigations of the strength-microstructure relation in polycrystalline cobalt are of particular
interest because cobalt exhibits an allotropic hcp→fcc transition. The transition belongs to the class
of martensitic phase transformations [19,20] and is rather sluggish as it is associated with a very low
free-energy change of ∼500 J/mol [4,21]. Martensitic transformation is a concept used currently in
a relatively broad context, which describes diffusionless crystallographic transformations that are
(almost) reversible. They are brought about by a movement of the interface between two phases as
the atoms in the parent lattice realign into the more energetically favorable structure. During this
first-order transformation, cooperative atomic shift results in crystal shape and symmetry changes [22].
Besides cobalt, martensitic transformations have been long known to occur in many metals and
metallic alloys, such as Ti, steels, Ni-Ti alloys, and entropy alloys, frequently investigated using
electron microscopy techniques [23–25].
In cobalt, the martensitic hcp→fcc transition takes place at a relatively low temperature of
∼420 ◦ C. Upon cooling through this temperature, a significant fraction of the high-temperature
fcc phase is typically retained in the microstructure (i.e., the reverse fcc→hcp transformation is
incomplete). The aforementioned temperature is an average value of the martensite start and austenite
start temperatures, which can exhibit a thermal hysteresis of up to 40 ◦ C in the case of polycrystalline
cobalt [21]. It was shown that the fraction of the retained fcc phase in cobalt could be reduced by
repeated thermal cycling around the transition temperature, thus providing an additional routine of
microstructure tailoring [21,26–28]. Moreover, the residual fcc phase transforms to hcp modification
upon loading at room temperature, thus making the deformation dynamics of cobalt even more
complex [7,29]. Hence, the residual fcc phase stands as another microstructural parameter affecting
the mechanical performance of the material, in addition to typical parameters such as grain size,
dislocation density, crystallographic texture, etc.
Along these lines, the objective of this study is two-fold. Firstly, we systematically examine
microstructure modifications and evolution in relation to the applied thermal treatments: isothermal
annealing at different temperatures and thermal cycling in the vicinity of the hcp↔fcc transition.
In particular, the recovery/recrystallization dynamics, grain growth, and fcc phase content
evolution are examined in detail with the help of scanning electron microscopy and thermal
analyses: thermodilatometry and differential scanning calorimetry. Secondly, we aim at assessing the
microstructure-deformation behavior dependence, primarily focusing on the effect of the fcc phase
content, which has been seldom investigated up until now. It is indeed shown that the fraction of the
high-temperature fcc phase plays a significant role in the room-temperature deformation behavior of
polycrystalline cobalt.

2. Materials and Methods


The polycrystalline cobalt was purchased from Goodfellow Cambridge Ltd. (Huntingdon,
England) in the form of as-drawn rods (length of 200 mm, diameter of 6.35 mm). The purity of
the material was 99.9%, having the following quoted content of impurities (in ppm): Fe—180, Ni—800,
C—30, S—150. The rods were cut to obtain cylindrical samples with a height of 9 mm. The first set of
samples was annealed using the vertical furnace Nabertherm RHTV 120-600 (Lilienthal, Germany)
under vacuum for 1 h at the temperatures of 600–1100 ◦ C (100 ◦ C step) and water-quenched. The second
and third set of samples were subjected to the same annealing procedure and subsequently thermally
cycled (second set—10 cycles, third set—20 cycles). One thermal cycle consisted of linear heating
from 300 to 550 ◦ C and subsequent cooling down to 300 ◦ C, both at a rate of 5 ◦ C/min. Thermal
cycling was performed using the Linseis L75 PT vertical thermodilatometer (Selb, Germany) under
Materials 2020, 13, 5775 3 of 13

an Ar atmosphere (for the thermodilatometric (TD) data used in the paper, correction curves were
subtracted). Differential scanning calorimetry (DSC) was carried out using the Netzsch DSC 404 C
Pegasus apparatus (Selb, Germany) under vacuum with the same thermal cycling program. For DSC
measurements, small discs were cut from the samples and mechanically polished using 4000 grit paper
to obtain a mass of ∼30 mg.
Microstructural changes in the investigated samples were studied by electron backscatter
diffraction (EBSD). The samples for EBSD were prepared by mechanical grinding by emery papers
and polishing using diamond suspensions with particle size decreasing down to 1 µm. Subsequently,
the final ion-etching was performed by Leica EM RES102 (Leica Mikrosysteme, Wetzlar, Germany).
The measurements were performed in the scanning electron microscope (SEM) ZEISS Auriga Compact
FIB-SEM (Jena, Germany) equipped with the EDAXEBSD camera (Berwyn, PA, USA). The EBSD study
was focused (i) on the volume fraction of the residual fcc phase and (ii) on the grain size and texture
of the hcp phase. The observations regarding (i) were performed on the maps showing an area of
400 × 400 µm2 collected with a step size of 0.4 µm, while the analyses concerning (ii) were performed
on EBSD maps with an area of 1000 × 2000 µm2 and with a step size of 1.5 µm. Such a variation in
the scan size was crucial because of the significant difference in the sizes of fcc and hcp phase grains.
Subsequently, the measured data were partially cleaned using the EDAX OIM TSL 7 software by one
step of confidence index (CI) standardization, one step of phase neighbor correlation, and one iteration
of grain dilatation. Only points having a CI higher than 0.1 were used for further analysis. The average
grain size was calculated for areas separated by grain boundaries with a misorientation higher than
15◦ [30], as a weighted average with the area fraction as the weight. Inverse pole figures (IPF) were
generated from the EBSD data by harmonic series expansion up to the rank of 16.
The compression tests on the samples (height of 9 mm, diameter of 6.35 mm) of each set were
performed using the Instron 5582 (Norwood, MA, USA) universal testing machine at room-temperature
and the initial strain rate of 10−3 ·s−1 . The machine stiffness was subtracted from the deformation
data before further processing. Apiezon M grease was used as a lubricant to suppress barreling of the
samples. The systematic experimental error was determined to be 3%. Several tests were repeated in
order to verify the consistency of the results.
The following naming convention is used throughout the paper: AD—initial (as-drawn) samples,
(600–1100)—samples annealed for 1 h at 600–1100 ◦ C, and (600-xxc–1100-xxc)—samples annealed for
1 h at 600–1100 ◦ C and subjected to the thermal cycling of xx cycles.

3. Results
The microstructural features of the initial as-drawn samples and the sample sets subjected to
different thermal treatments were examined in detail. The EBSD micrographs of the selected samples,
showing the orientation maps, as well as the phase maps, are presented in Figure 1. The AD sample
featured a deformed (see the color variations within a single grain) and relatively fine microstructure
with a mean grain size of ∼9 µm and less than 2% of the residual fcc phase (see also Table 1). Isothermal
annealing of the initial AD material for 1 h led to an increase in the grain size to 22–46 µm with
apparent dependence on the annealing temperature, as observed in Figure 1c–e, demonstrating the
occurrence of recovery, recrystallization, and grain growth (for the values, see also Table 1). At the
same time, the amount of the residual fcc phase increased to ∼6–11%, with no clear dependence
on the annealing temperature. The samples annealed at different temperatures were subsequently
subjected to thermal cycling in the temperature range of 300–550 ◦ C, i.e., the range within which the
hcp↔fcc allotropic phase transformation takes place. Ten such thermal cycles resulted in further grain
growth (to ∼30–60 µm), still reflecting the differences in the microstructure after isothermal annealing
(Figure 1f–h). Concurrently, the fraction of the fcc phase decreased to ∼1% for the sample 600 and
below 0.5% for the samples 700–1100. Finally, after 20 thermal cycles, the grain growth was further
promoted, achieving a grain size of ∼60 µm, and the fraction of the fcc phase was reduced below 0.5%
for all the samples (Figure 1i–k and Table 1).
Materials 2020, 13, 5775 4 of 13

Figure 1. SEM EBSD micrographs of the selected samples before the thermal treatment: (a) AD—orientation
map, (b) AD—phase map; and after the thermal treatment: (c–e) samples 600, 800, and 1000, respectively;
(f–h) samples 600-10c, 800-10c, and 1000-10c, respectively; (i–k) samples 600-20c, 800-10c, and 1000-20c,
respectively. The parent images of the samples (c–k) represent orientation maps and the insets represent
phase maps of the respective part of each image. The processing direction (i.e., the rod lengthwise direction
during drawing) is perpendicular to the paper plane.
Materials 2020, 13, 5775 5 of 13

Table 1. Microstructural and mechanical data of all the tested sample sets.

Sample Grain Size (µm) fcc Fraction (%) εmax σmax (MPa) σ0.2 (MPa)
AD 9±1 ∼2 0.25 ± 0.01 1047 ± 31 486 ± 15
600 22 ± 3 ∼6 0.32 ± 0.01 1091 ± 33 422 ± 13
700 26 ± 1 ∼10 0.33 ± 0.01 961 ± 29 312 ± 9
800 32 ± 2 ∼8 0.36 ± 0.01 984 ± 30 290 ± 9
900 32 ± 2 ∼11 0.39 ± 0.01 1024 ± 31 283 ± 8
1000 39 ± 8 ∼11 0.38 ± 0.01 1014 ± 30 274 ± 8
1100 47 ± 4 ∼6 0.38 ± 0.01 1008 ± 30 269 ± 8
600-10c 31 ± 2 ∼1 0.29 ± 0.01 918 ± 28 287 ± 9
700-10c 50 ± 1 <0.5 0.32 ± 0.01 926 ± 28 277 ± 8
800-10c 51 ± 3 <0.5 0.29 ± 0.01 895 ± 27 257 ± 8
900-10c 51 ± 3 <0.5 0.27 ± 0.01 887 ± 27 261 ± 8
1000-10c 57 ± 1 <0.5 0.31 ± 0.01 885 ± 27 267 ± 8
1100-10c 70 ± 2 <0.5 0.32 ± 0.01 894 ± 27 276 ± 8
600-20c 65 ± 2 <0.5 0.25 ± 0.01 879 ± 26 276 ± 8
700-20c 59 ± 5 <0.5 0.28 ± 0.01 885 ± 27 268 ± 8
800-20c 59 ± 1 <0.5 0.26 ± 0.01 878 ± 26 268 ± 8
900-20c 59 ± 1 <0.5 0.28 ± 0.01 879 ± 26 259 ± 8
1000-20c 68 ± 2 <0.5 0.26 ± 0.01 863 ± 26 267 ± 8
1100-20c 67 ± 4 <0.5 0.27 ± 0.01 864 ± 26 254 ± 8

Figure 2 shows the crystallographic texture by means of inverse pole figures of the selected
samples: AD, 600, 600-10c, and 600-20c. A certain texture can be observed for the AD sample,
comprising the components typical of drawn or extruded rods of hcp metals [31]. This rather weak
texture, however, vanished already during isothermal annealing at the lowest temperature of 600 ◦ C
(sample 600, Figure 2b). Additional thermal treatment of the sample 600 by means of thermal cycling
did not bring about any perceptible changes in the texture, which remained rather random and very
feeble. The other samples not shown in Figure 2 exhibited a similar or even weaker crystallographic
texture, in line with the observed random EBSD orientation maps in Figure 1. Note that the texture
was calculated from much larger areas than shown in Figure 1, comprising a sufficient number of
grains (see the Materials and Methods).

Figure 2. EBSD inverse pole figures of the selected samples: (a) AD sample, (b) sample 600, (c) sample
600-10c, and (d) sample 600-20c. The intensity scale bar (multiples of a random distribution) is valid
for all the images.

The compression tests were performed on each sample type, and the selected deformation curves
are presented in Figure 3; the insets show the details of the elasto-plastic transition. In Figure 3a,
the compression curves of the annealed samples, 600–1100, are depicted, demonstrating (i) an
increasing fracture strain (ε max ) of the samples (from ∼0.25 for AD to ∼0.38 for 900, 1000, and 1100),
(ii) a gradually declining yield strength σ0.2 of the samples (from ∼486 MPa for AD down to ∼269 MPa
for 1100), and (iii) a non-monotonous evolution of the compressive strength (σmax ) of the samples,
ranging from ∼960 to ∼1090 MPa, as a function of increasing annealing temperature (see also
Table 1, listing all the data). Figure 3b depicts several reference deformation curves from Figure 3a
(samples AD, 600, 800, and 1000) and the deformation curves of the respective samples subjected to the
thermal cycling of 10 cycles (samples 600-10c, 800-10c, and 1000-10c) and 20 cycles (samples 600-20c,
800-20c, and 1000-20c). It was observed that the thermal cycling of 10 cycles resulted in a decrease
in all the values of evaluated mechanical parameters—ε max , σmax , and σ0.2 —giving the values in the
Materials 2020, 13, 5775 6 of 13

relatively narrow ranges of ∼0.27–0.32, ∼885–926 MPa, and ∼257–287 MPa, respectively. No obvious
dependence of these values on the temperature of the preceding isothermal annealing was observed,
even though some trend was still noticed in the grain size values (Table 1). After the thermal cycling of
20 cycles, the values of all the mechanical parameters—ε max , σmax , and σ0.2 —exhibited a further slight
decrease to ∼0.25–0.28, ∼863–885 MPa, and ∼254–276 MPa, respectively, again with no clear trends in
relation to the thermal history (see Table 1).

1 2 0 0 1 2 0 0
(a ) (b )
1 0 0 0 1 0 0 0
A D
8 0 0 8 0 0 6 0 0
T ru e stre ss [M P a ]

T ru e stre ss [M P a ]
8 0 0
6 0 0 A D 6 0 0 1 0 0 0
6 0 0 6 0 0 -1 0 c
4 0 0 7 0 0 4 0 0 8 0 0 -1 0 c
8 0 0 1 0 0 0 - 1 0 c
2 0 0 9 0 0 2 0 0 6 0 0 -2 0 c
1 0 0 0 8 0 0 -2 0 c
1 1 0 0 1 0 0 0 - 2 0 c
0 0
0 .0 0 .1 0 .2 0 .3 0 .4 0 .0 0 .1 0 .2 0 .3 0 .4
T r u e p la s t ic s t r a in T r u e p la s t ic s t r a in
Figure 3. Deformation curves of the selected samples: (a) AD sample and samples 600–1000, (b) AD
sample and samples 600, 800, and 1000 (for the sake of comparison), samples 600-10c, 800-10c, 1000-10c,
and samples 600-20c, 800-20c, 1000-20c.

As presented above, besides the evolution of the grain size and mechanical properties, the fraction
of the residual fcc phase was another significantly affected quantity depending on different types of
thermal treatment. The reversible hcp↔fcc transition is a martensitic-type transformation [4,32] and,
thus, can be effectively investigated by conventional thermal analyses. For such observations, sample
1000 (i.e., the sample annealed for 1 h at 1000 ◦ C) was used, as its microstructure was already fairly
recrystallized, hence enabling the sole investigation of the fcc phase evolution without a significant
influence of the evolution of other microstructural features. Figure 4 shows the thermodilatometric (TD)
data recorded during thermal cycling. For the sake of clarity, only selected cycles—Cycles 1, 2, 5, 10, 15,
and 20—are presented. Most portions of the curves exhibit quasi-linear dependence on the temperature,
which is the typical behavior of pure metals. However, there are gradual jumps in this linear behavior
from around 450 to 470 ◦ C during the heating and from about 380 to 350 ◦ C during the cooling stages
of cycling related to the volume changes associated with the transition (the maximum theoretical
volume difference was ∼0.3–0.4% [7,33]). These temperatures thus correspond to the austenite start
(TAS ), austenite end (TAE ), martensite start (TMS ), and martensite end (TME ) temperatures, respectively.
The observed temperature intervals and thermal hysteresis agreed well with the previously reported
kinetic analyses of the hcp↔fcc transition in pure cobalt [21,34], as will be discussed in detail later.
It is also evident that the shape of the curves in the vicinity of the transition evolves with an increasing
number of cycles (for details, see Figure 4 insets). The TAS temperature successively shifts to slightly
higher values, whereas, on the other hand, TMS seems to be not affected. The TAE and TME temperatures
cannot be directly distinguished from the measured TD data, which are alone difficult to use for
quantitative analysis and, especially, for the precise determination of the transition temperatures.
To that end, differential scanning calorimetry (DSC) analysis was employed maintaining the same
conditions (thermal cycling program) and sample type (1000). Figure 5 presents the data of the selected
cycles as in the case of TD. There are well-defined peaks corresponding to the hcp↔fcc transition
during the heating and cooling stages of each cycle. The peak analysis allowed for quite accurate
calculations of the transition temperatures by means of finding the intersections of the baselines and
the tangent lines at the peak left and right inflection points. The evolution of the thus-determined
Materials 2020, 13, 5775 7 of 13

temperatures is shown in Figure 6. These data show that, indeed, the entire endothermic austenitic
peak (i.e., both TAS and TAE ) shifts to a somewhat higher temperature as the number of cycles increases
(to be precise, the TAE value is rather saturated already after five cycles). Moreover, the temperature
difference between TAS and TAE slightly decreases (Figure 6a) as the peak sharpens and its height
increases (Figure 5). The evolution of the transition temperatures of the exothermic martensitic
peak, TMS and TME , is less straightforward. The TMS temperature gradually decreases throughout
the entire cycling; however, TME decreases only up to five cycles and subsequently starts to rise
(Figure 6b). Therefore, this cooling peak also narrows during cycling (the latter stages) and its height
increases, in agreement with the observations for the heating peak. The cooling peak is, however,
less intense, and the martensitic fcc↔hcp transformation manifestly proceeds slower than the austenitic
(hcp↔fcc) one, this being well visible in Figure 5 and also demonstrated by the higher temperature
difference in Figure 6b compared to Figure 6a. Nevertheless, it can be clearly inferred that the thermal
hysteresis (∆T = TAS − TMS ) increases with a rising number of thermal cycles. Fairly similar shapes
and the evolution of the DSC peak in cobalt associated with the hcp↔fcc transition were observed
elsewhere [21,26,34].

8 0 6 0
lin g
5 5 c o o
7 0 5 0
L e n g th c h a n g e [μm ]

4 5
3 6 0 3 7 0 3 8 0 3 9 0 4 0 0 6 5
6 0
6 0

5 5
5 0
5 0
4 3 0 4 4 0 4 5 0 4 6 0 4 7 0

4 0 c y c le 1 c y c le 1 0
t in g
h e a
c y c le 2 c y c le 1 5
c y c le 5 c y c le 2 0
3 0
3 5 0 4 0 0 4 5 0 5 0 0 5 5 0
T e m p e ra tu re [° C ]
Figure 4. Selected thermodilatometric (TD) curves of the sample 1000 subjected to thermal cycling of
20 cycles.

0 .0 8
0 .1 0 e x o 0 .0 7

0 .0 8
0 .0 6
n g
0 .0 5
c o o li
/m g ]

0 .0 6 0 .0 4
3 7 0 3 7 5 3 8 0 3 8 5 3 9 0 3 9 5 4 0 0
H e a t flo w [ m W

0 .0 4
- 0 .0 4
- 0 .0 4

- 0 .0 6 - 0 .0 6 h e a t in g

- 0 .0 8 - 0 .0 8
c y c le 1 c y c le 1 0
- 0 .1 0 c y c le 2 c y c le 1 5
- 0 .1 0
4 3 0 4 4 0 4 5 0 4 6 0 4 7 0 c y c le 5 c y c le 2 0

3 5 0 4 0 0 4 5 0 5 0 0 5 5 0
T e m p e ra tu re [° C ]
Figure 5. Selected DSC curves of the sample 1000 subjected to thermal cycling of 20 cycles.
Materials 2020, 13, 5775 8 of 13

4 7 0
( a ) H e a t in g ( b ) C o o lin g
4 0 0
4 6 0
3 9 0
T e m p e ra tu re [°C ]

T e m p e ra tu re [°C ]
4 5 0
3 8 0
4 4 0
3 7 0
T T
4 3 0
A S M S
T A E T M E
3 6 0
0 5 1 0 1 5 2 0 0 5 1 0 1 5 2 0
N u m b e r o f c y c le s N u m b e r o f c y c le s
Figure 6. Evolution of the TAS , TAE , TMS , and TME temperatures determined from the DSC peaks as a
function of the increasing number of thermal cycles (a) heating, (b) cooling.

4. Discussion
The mechanical properties of crystalline materials are, apart from the chemical composition,
critically affected by their microstructural features. The flow stress and deformability of a material are
governed primarily by the grain size (through the Hall–Petch relation), the Taylor strengthening related
to the dislocation density (i.e., associated with the residual strain), the solid solution strengthening,
and the precipitation strengthening, out of which the latter two can be neglected in this study as the
impurity content was very low in the cobalt samples. In addition to these four basic strengthening
mechanisms, the proportion of different allotropic phases contained in pure metallic polycrystals,
which exhibit allotropic transformation involving a relatively low thermodynamic driving force for
the transition, can be of high significance.
Softening due to the increase in the grain size was observed in the investigated samples, as seen
from the comparison of Figures 1 and 3 and, especially, in Table 1. This dependence is also confirmed
by means of the conventional Hall–Petch plot shown in Figure 7 with points calculated for all the
sample types involved in this study. This graph, however, does not show a perfectly linear behavior of
1
the yield strength, σ0.2 , as a function of the reciprocal of the square root of grain size, d− 2 , as predicted
by the theory for such intermediate grain sizes [17,35,36]. The deviations might be brought about
by several factors: (i) there was a very limited number of samples with a small grain size (since the
applied heat treatment and cycling led to a rather expeditious grain growth) to ensure better statistics;
(ii) besides the grain growth, the initially deformed microstructure, as demonstrated by color variations
within individual grains in Figure 1a, recovered and recrystallized (Figure 1c–k); and (iii) the fraction of
the residual fcc phase significantly decreased due to the heat treatment. Nevertheless, there was a clear
dependence of σ0.2 for the samples AD and 600–1100 marked in red, which exhibited distinct grain
size variations (see also the inset in Figure 3a), in agreement with previously reported observations [7].
On the other hand, the thermally cycled sample sets (600-10c to 1100-10c and 600-20c to 1100-20c
marked in blue and green, respectively) featuring much narrower ranges of larger recrystallized grains
rather formed a “cluster”, with no clear grain size effect on the σ0.2 value.
The dependence of the compressive strength (σmax ) and fracture strain (ε max ) values was more
straightforward in some respects. The sample sets AD and 600–1100 exhibited a clear positive
dependence of ε max on the grain size (which itself depended positively on the annealing temperature),
as it gradually increased from ∼0.25 (AD) to ∼0.38 (1100), this being a typical behavior of cobalt and
metallic polycrystals in general [7]. Yet, very similar σmax values of ∼1000 MPa were found for all the
samples within this set, independent of the grain size. Another direct observation is the decrease in
both the ε max and σmax values due to the thermal cycling, whereas the values were rather constant
across the samples of the same set—ε max = ∼0.27, σmax = ∼900 MPa after 10 cycles and ε max = ∼0.30,
Materials 2020, 13, 5775 9 of 13

σmax = ∼880 MPa after 20 thermal cycles (cf. Figure 3 and Table 1). Decreasing σmax as a result of
thermal cycling can be readily related to the microstructure evolution, i.e., the recovery, recrystallization,
and grain growth (the grain growth in cobalt is typically promoted when temperatures above 500 ◦ C
during cycling are involved [7]). One would, however, expect that the recrystallized microstructure and
larger grain size (after 20 thermal cycles as opposed to 10) also would give rise to higher deformability
(ε max ), which was not the case here. In order to elucidate this observation, we must finally take into
account the content of the residual fcc phase, which was shown to substantially affect the deformability
of cobalt and the overall shape of its deformation curves [7]. During loading of polycrystalline cobalt
samples containing the fcc phase, deformation-induced martensitic fcc↔hcp transformation takes
place, contributing to the sample ductility. The driving force for such stress-induced transition is
actually orders of magnitude higher than that during the thermal cycling [7]. Hence, these results
strongly suggest that not only the grain size, but also the fcc phase content might be the determining
factor of the sample deformability. It should be, however, noted that even though there were observable
differences in the fcc fraction between the samples subjected to 10 and 20 thermal cycles (cf. Figure 1),
its content was already rather low (<1%) in both sets. Still, the most significant decrease in the
ε max value occurred after the thermal cycling (in comparison with the annealed samples 600–1100),
in accordance with a rather significant reduction in the residual fcc phase (cf. Table 1). The theoretical
volume change of ∼0.3–0.4% can accommodate only a very limited strain. However, it was asserted
that due to the abundance of slip systems in the fcc phase, the slip and transformation are closely
related and may bring about grain shape change, thus providing large amounts of shear [7]. It was
estimated that in this way, a strain of 0.25% can be theoretically attained by a full transformation of
1% of the retained fcc phase. The deformation-induced transformation in cobalt is, however, typically
completed by a strain of roughly 10% (depending on the materials and testing conditions [7,29]),
and in the latter stages of deformation, mechanical twinning takes place as the secondary deformation
mechanism [7,18]. Indeed, it was documented that only basal slip is typically active in cobalt with
no prominent secondary slip systems available, while a multiplicity of twinning systems can be
activated [7]. It is also believed that in hcp metals, twinning is more readily activated in materials with
larger grains [31,37]. The formation of twins gives rise to the abundance of high-angle grain boundaries,
which, in turn, strengthen the material and make it less ductile (since these new boundaries act as
barriers to dislocation slip, notwithstanding that the slip can be to some extent also promoted due to
crystal reorientation). In this manner, the ductility of the material featuring larger grains can be even
lower than in the case of materials with smaller grains [38,39]. We speculate that such an effect can
be present also in the investigated cobalt samples and, together with the influence of the residual fcc
phase, can account for the observed reduced fracture strain (ε max ) observed despite the microstructure
recovery, recrystallization, and grain growth. To shed more light on this matter, quantitative (semi) in
situ microstructural observations are needed, those being the subject of the ensuing separate study.
The dynamics of the hcp↔fcc, as evaluated by means of thermal analyses (Figures 4–6),
also confirms that the microstructure is not stable after 10 cycles and the changes in the TD and DSC
curves, even though they somewhat slow down, continue to take place up to 20 cycles. Indeed, it was
shown that a completely stable microstructure of polycrystalline cobalt can be obtained only after
more than ∼50 cycles around the transition temperature (although the exact evolution arguably
depends on the conditions, such as initial microstructure, temperature range, etc.) [21,26,28]. Therefore,
notwithstanding that the content of the fcc phase is very low after both 10 and 20 cycles, further
evolution of the microstructure occurs, as was also discernible in the (rather local) EBSD images in
Figure 1.
In order to comprehend the evolution of the microstructure during thermal cycling, especially in
relation to the phase transformation, the calorimetric curves must be examined in detail. Sharpening
and intensification of both the endothermic and exothermic peaks (i.e., during heating and cooling
stages, respectively) with an increasing number of thermal cycles signifies acceleration of the hcp↔fcc
transition. It was also documented in [21] that the TAS , TAE , TMS , and TME temperatures tend to
Materials 2020, 13, 5775 10 of 13

shift, and at the same time, the shape of the peaks changes during cycling. The thermal hysteresis,
i.e., the difference between the transformation temperatures upon heating and cooling, increases with
thermal cycling (apart from some slight deviations in TME ). This effect arises due to the strain energy
induced by volume changes and the reduction in the number of nucleation sites during cycling, as will
be discussed below [6,40].

5 0 0

4 5 0

4 0 0
[M P a ]

3 5 0
0 .2

3 0 0
A D −1 1 0 0
6 0 0 -1 0 c −1 1 0 0 -1 0 c
2 5 0 6 0 0 -2 0 c −1 1 0 0 -2 0 c
L in e a r fit
2 0 0
0 .1 0 0 .1 5 0 .2 0 0 .2 5 0 .3 0 0 .3 5
d [µ m ]
-1 /2 -1 /2

Figure 7. The Hall–Petch plot for all the investigated samples sets—initial sample, annealed samples,
and thermally-cycled samples.

It was shown as early as in the 1970s that the hcp↔fcc transition proceeds to completion only
rarely, e.g., when a single interface transformation takes place in single crystals [6]. Several theories
have been proposed to account for the incomplete fcc→hcp transition upon cooling through the TMS
and TME temperatures. They involved the passage of Shockley partial dislocations (SPs) over every
second closest-packed plane {111} in the fcc lattice to form the hcp structure, and these theories differ
only in the mechanisms of dislocation generation and movement [41–43]. Much later, these theories
were refined, and the following explanation summarized, e.g., by Bauer et al. [21] is currently rather
generally accepted. The fcc↔hcp transition (i.e., the transformation from the ABCABC . . . to ABABAB
. . . stacking sequence and back) is realized by an ordered glide of SPs with 16 h112i Burgers vectors on
every second {111} plane. The fcc→hcp transition takes place by dissociation of the perfect dislocations
into SPs and the hcp→fcc one by association of the same SPs (i.e., by a reverse dislocation reaction).
During the first thermal cycles, entire grains cannot fully transform back to the hcp structure due
to mutual blocking of the growing martensite along one of the four equivalent {111} fcc planes.
Continued thermal cycling, however, leads to the preferred growth along favorably oriented planes by
progressively seizing the nearby dislocations. As the hcp structure has only a single closest-packed
set of {0001} planes and the fcc structure features four equivalent sets of {111} planes, the dislocation
structure gradually rearranges so that only one active {111} set parallel to the {0001} plane remains
finally active (see also the comprehensive schematic views in [21]). In other words, the first cycles
involve the rearrangement of crystal defects, while several tens of cycles are needed to facilitate
a fully reversible hcp↔fcc transition. Along these lines, it is not surprising that there were more
significant changes in the TD (Figure 4) and DSC data (Figure 6) in the course of the initial thermal
cycles, whereas the process gradually stabilized with the increasing number of cycles. Yet, it might
then seem counterintuitive at first glance why the thermal hysteresis becomes larger during thermal
cycling. In fact, this delay in the transformation also interrelates with the microstructure evolution;
however, different processes are responsible for that effect. Specifically, transformation-induced defects
(mostly sessile dislocations), which are effectively generated during thermal cycling, interact with the
Materials 2020, 13, 5775 11 of 13

above-described interface dislocations. The expanded hysteresis corresponds to the amount of energy
needed to overcome these pinned transformation-induced defects during subsequent cycles [26].
The understanding of the physical mechanisms underlying the dynamics of the hcp↔fcc transition
can be thus effectively utilized for fine-tuning the material microstructure, e.g., via thermo-mechanical
treatment. In turn, the mechanical performance of polycrystalline cobalt, which was shown to be
significantly affected by the content of the retained high-temperature fcc phase, can be systematically
optimized to meet the application-specific requirements.

5. Conclusions
Pure polycrystalline cobalt was thermally treated in order to systematically modify its
microstructure and, in turn, to examine its effect on the mechanical properties in compression. The main
conclusions drawn from this study can be summarized as follows:

• The initial as-drawn samples feature a deformed and relatively fine microstructure with a grain
size of ∼9 µm. Isothermal annealing for one hour at different temperatures in the range of
600–1100 ◦ C leads to grain growth, yielding grain sizes ranging from ∼22 µm to ∼47 µm with
explicit dependence on the annealing temperature. Consequently, the yield strength (σ0.2 ) drops
from ∼486 MPa down to ∼269 MPa and the fracture strain (ε max ) increases from ∼0.25 to ∼0.38
as a function of rising grain size. There are, however, no straightforward trends observed in the
compressive strength (σmax ) with increasing grain size, having the values of ∼961–1091 MPa.
The content of the residual high-temperature fcc phase is ∼6–11% across this sample set.
• Subsequent thermal cycling in the temperature range of 300–550 ◦ C (10 or 20 cycles) is applied to
assess the evolution of the microstructure and mechanical performance. After 10 cycles, the grain
size further increases to ∼31–70 µm, still somewhat reflecting the microstructure differences after
the preceding annealing at various temperatures. On the other hand, no clear dependencies on
the grain size of ε max , σmax , and σ0.2 are observed, and their values are ∼0.27–0.32, ∼885–926 MPa,
and ∼257–287 MPa, respectively. The content of the fcc phase drops to below 1%. After 20 cycles,
the grain size of ∼60 µm is evidenced for all the samples. The ε max , σmax , and σ0.2 values decrease
slightly further to ∼0.25–0.28, ∼863–885 MPa, and ∼254–276 MPa, respectively. The fraction of
the fcc phase drops below 0.5%.
• An atypical decrease in ε max and ambiguous trends in the other mechanical quantities as a
function of grain size are explained in terms of decreasing fcc phase content upon thermal
cycling. Upon loading, the fcc phase exhibits a stress-induced fcc→hcp transformation,
thus accommodating a significant amount of plastic strain (by the associated volume change and
by providing additional slip systems).
• The processes involved in the hcp↔fcc transformation during thermal cycling are studied by
differential scanning colorimetry, and it is evidenced that a complete transformation does not
occur even after 20 cycles. Yet, the effectiveness of the transformation improves with an increasing
number of cycles owing to the evolving dislocation structure, which plays a crucial role in the
transformation dynamics.
• Apart from modifying the standard microstructural features, such as grain size and dislocation density,
tailoring the content of the residual fcc phase seems to be crucial for attaining the desired mechanical
performance of polycrystalline cobalt. Such an approach may, therefore, open new possibilities for
effective mechanical performance optimization with respect to particular applications.

Author Contributions: Conceptualization, M.K., P.M., and F.C.; methodology, M.K., P.D., and P.M.; formal
analysis, F.C. and E.H.; investigation, M.K., P.M., P.D., J.Š., and M.M.C.; data curation, M.K., P.M., and P.D.,
writing, original draft preparation, M.K.; writing, review and editing, P.M., P.D., J.Š., M.M.C., E.H., and F.C.;
supervision, F.C.; project administration, F.C. and E.H. All authors read and agreed to the published version of
the manuscript.
Materials 2020, 13, 5775 12 of 13

Funding: This research received financial support from the Czech Science Foundation, Grant No. 19-22604S and
MEYS Kontakt ME (Barrande), Grant No. 8J19FR023. M.K. and P.D. gratefully acknowledge the financial support
from OP RDE, MEYS, Grant No. CZ.02.1.01/0.0/0.0/16 013/0001794. P.M. received partial financial support from
ERDF, Grant No. CZ.02.1.01/0.0/0.0/15 003/0000485.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; nor in the decision to
publish the results.

References
1. Kiani, M.; Gandikota, I.; Rais-Rohani, M.; Motoyama, K. Design of lightweight magnesium car body
structure under crash and vibration constraints. J. Magnes. Alloy. 2014, 2, 99–108. [CrossRef]
2. Zheng, G. Grain-size effect on plastic flow in nanocrystalline cobalt by atomistic simulation. Acta Mater.
2007, 55, 149–159. [CrossRef]
3. Crone, W.C. A Brief Introduction to MEMS and NEMS. In Springer Handbook of Experimental Solid Mechanics;
Sharpe, W.N., Ed.; Springer: Boston, MA, USA, 2008; pp. 203–228. [CrossRef]
4. Betteridge, W. The properties of metallic cobalt. Prog. Mater. Sci. 1980, 24, 51–142. [CrossRef]
5. Mises, R.G.V. Mechanik der plastischen Formänderung von Kristallen. ZAMM-Z. FüR Angew. Math.
Und Mech. 1928, 8, 161–185. [CrossRef]
6. Holt, R.T. The High Temperature Deformation of Cobalt Single Crystals. Ph.D. Thesis, The University of
British Columbia, Vancouver, BC, Canada, 1968.
7. Sanderson, C.C. Deformation of Polycrystalline Cobalt. Ph.D. Thesis. University of British Columbia,
Vancouver, BC, Canada, 1972.
8. Marx, V.M.; Kirchlechner, C.; Breitbach, B.; Cordill, M.J.; Többens, D.M.; Waitz, T.; Dehm, G. Strain-induced
phase transformation of a thin Co film on flexible substrates. Acta Mater. 2016, 121, 227–233. [CrossRef]
9. Sort, J.; Nogués, J.; Suriñach, S.; Baró, M.D. Microstructural aspects of the hcp-fcc allotropic phase
transformation induced in cobalt by ball milling. Philos. Mag. 2003, 83, 439–455. [CrossRef]
10. Edalati, K.; Toh, S.; Arita, M.; Watanabe, M.; Horita, Z. High-pressure torsion of pure cobalt: Hcp-fcc
phase transformations and twinning during severe plastic deformation. Appl. Phys. Lett. 2013, 102, 181902.
[CrossRef]
11. Barry, A.H.; Dirras, G.; Schoenstein, F.; Tétard, F.; Jouini, N. Microstructure and mechanical properties
of bulk highly faulted fcc/hcp nanostructured cobalt microstructures. Mater. Charact. 2014, 91, 26–33.
[CrossRef]
12. Holt, R.T.; Teghtsoonian, E. The influence of the allotropic transformation on the deformation behavior of
pure cobalt. Metall. Trans. 1972, 3, 2443–2447. [CrossRef]
13. Korner, A.; Karnthaler, H.P. Weak-beam study of glide dislocations in h.c.p. cobalt. Philos. Mag. A 1983,
48, 469–477. [CrossRef]
14. Seeger, A.; Kronmüller, H.; Boser, O.; Rapp, M. Plastische Verformung von Kobalteinkristallen. Phys. Status
Solidi (b) 1963, 3, 1107–1125. [CrossRef]
15. Zhu, Y.T.; Zhang, X.Y.; Liu, Q. Observation of twins in polycrystalline cobalt containing face-center-cubic
and hexagonal-close-packed phases. Mater. Sci. Eng. A 2011, 528, 8145–8149. [CrossRef]
16. Zhang, X.Y.; Zhu, Y.T.; Liu, Q. Deformation twinning in polycrystalline Co during room temperature
dynamic plastic deformation. Scr. Mater. 2010, 63, 387–390. [CrossRef]
17. Fleurier, G.; Hug, E.; Martinez, M.; Dubos, P.A.; Keller, C. Size effects and Hall–Petch relation in
polycrystalline cobalt. Philos. Mag. Lett. 2015, 95, 122–130. [CrossRef]
18. Martinez, M.; Fleurier, G.; Chmelík, F.; Knapek, M.; Viguier, B.; Hug, E. TEM analysis of the deformation
microstructure of polycrystalline cobalt plastically strained in tension. Mater. Charact. 2017, 134, 76–83.
[CrossRef]
19. Edwards, O.S.; Lipson, H. An X-ray Study of the Transformation of Cobalt. J. Inst. Metals 1943, 96, 177–187.
20. Troiano, A.R.; Tokich, J.L. The transformation of cobalt. Trans. Am. Inst. Min. Metall. Eng. 1948, 175, 728–741.
21. Bauer, R.; Jägle, E.A.; Baumann, W.; Mittemeijer, E.J. Kinetics of the allotropic hcp–fcc phase transformation
in cobalt. Philos. Mag. 2011, 91, 437–457. [CrossRef]
22. Laughlin, D.E.; Hono, K. Physical Metallurgy, 5th ed.; Number sv. 1,sv. 3 in Physical Metallurgy; Elsevier
Science: Amsterdam, The Netherlands, 2014.
Materials 2020, 13, 5775 13 of 13

23. Schryvers, D. Electron Microscopy Studies of Martensite Microstructures. J. Phys. IV 1997, 7, 109–118.
[CrossRef]
24. Xie, G.; Song, M.; Mitsuishi, K.; Furuya, K. Transmission Electron Microscopy of Martensitic Transformation
in Xe-implanted Austenitic 304 Stainless Steel. J. Mater. Res. 2005, 20, 1751–1757. [CrossRef]
25. Chang, S.H.; Lin, P.T.; Tsai, C.W. High-temperature martensitic transformation of CuNiHfTiZr high- entropy
alloys. Sci. Rep. 2019, 9, 19598. [CrossRef] [PubMed]
26. Munier, A.; Bidaux, J.E.; Schaller, R.; Esnouf, C. Evolution of the microstructure of cobalt during diffusionless
transformation cycles. J. Mater. Res. 1990, 5, 769–775. [CrossRef]
27. Bidaux, J.E.; Schaller, R.; Benoit, W. Study of the hcp-fcc phase transition in cobalt by internal friction
and elastic modulus measurements in the kHz frequency range. J. Phys. Colloq. 1987, 48, C8–477–C8–482.
[CrossRef]
28. Kuang, Z.Q.; Zhang, J.X.; Zhang, X.H.; Liang, K.F.; Fung, P.C.W. Latent heat in the thermoelastic martensitic
transformation of Co. Scr. Mater. 2000, 42, 795–799. [CrossRef]
29. Dubos, P.A.; Fajoui, J.; Iskounen, N.; Coret, M.; Kabra, S.; Kelleher, J.; Girault, B.; Gloaguen, D. Temperature
effect on strain-induced phase transformation of cobalt. Mater. Lett. 2020, 281, 128812. [CrossRef]
30. Lejček, P. Grain Boundary Segregation in Metals, 1st ed.; Springer Series in Materials Science; Springer: Berlin,
Germany, 2010; Volume 136.
31. Dobroň, P.; Chmelík, F.; Yi, S.; Parfenenko, K.; Letzig, D.; Bohlen, J. Grain size effects on deformation
twinning in an extruded magnesium alloy tested in compression. Scr. Mater. 2011, 65, 424–427. [CrossRef]
32. Sebastian, M.T.; Krishna, P. Random, Non-Random, and Periodic Faulting in Crystals; Gordon and Breach
Science Publishers: Yverdon, Switzerland; Langhorne, PA, USA, 1994.
33. Sargent, P.M.; Malakondaiah, G.; Ashby, M.F. A deformation map for cobalt. Scr. Metall. 1983, 17, 625–629.
[CrossRef]
34. Ray, A.E.; Smith, S.R.; Scofield, J.D. Study of the phase transformation of cobalt. J. Phase Equilibria 1991,
12, 644–647. [CrossRef]
35. Louchet, F.; Weiss, J.; Richeton, T. Hall-Petch Law Revisited in Terms of Collective Dislocation Dynamics.
Phys. Rev. Lett. 2006, 97, 075504. [CrossRef]
36. Hug, E.; Keller, C. Size effects and magnetoelastic couplings: A link between Hall–Petch behaviour and
coercive field in soft ferromagnetic metals. Philos. Mag. 2019, 99, 1297–1326. [CrossRef]
37. Bohlen, J.; Dobroň, P.; Meza Garcia, E.; Chmelík, F.; Lukáč, P.; Letzig, D.; Kainer, K.U. The Effect of Grain
Size on the Deformation Behaviour of Magnesium Alloys Investigated by the Acoustic Emission Technique.
Adv. Eng. Mater. 2006, 8, 422–427. [CrossRef]
38. Wang, X.; Jiang, L.; Luo, A.; Song, J.; Liu, Z.; Yin, F.; Han, Q.; Yue, S.; Jonas, J.J. Deformation of twins in a
magnesium alloy under tension at room temperature. J. Alloy. Compd. 2014, 594, 44–47. [CrossRef]
39. Krajňák, T.; Minárik, P.; Stráská, J.; Gubicza, J.; Máthis, K.; Janeček, M. Influence of the initial state on the
microstructure and mechanical properties of AX41 alloy processed by ECAP. J. Mater. Sci. 2019, 54, 3469–3484.
[CrossRef]
40. Votava, E. Electron microscopic investigation of the phase transformation of thin cobalt samples. Acta Metall.
1960, 8, 901–904. [CrossRef]
41. Cottrell, A.H.; Bilby, B.A. A mechanism for the growth of deformation twins in crystals. Lond. Edinb. Dublin
Philos. Mag. J. Sci. 1951, 42, 573–581. [CrossRef]
42. Sumino, K. Surface dislocations and the growth of deformation twins. Acta Metall. 1966, 14, 1607–1615.
[CrossRef]
43. Christian, I.W. Some Comments on the Burgers and Bogers-Burgers Transformation Mechanisms and Their
Relation to Mathematical Theories of Martensite Crystallography. J. Less Common Met. 1972, 28, 8. [CrossRef]

Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

Common questions

Powered by AI

After 20 thermal cycles, polycrystalline cobalt exhibits a grain size growth to approximately 60 µm, with little variation from preceding annealing temperatures . The mechanical properties such as εmax, σmax, and σ0.2 show slight reductions to ranges of 0.25–0.28, 863–885 MPa, and 254–276 MPa, respectively . This decline is primarily driven by the significant reduction in the fcc phase content, which falls below 0.5%, limiting the stress-induced transformation capacity essential for maintaining ductility and strength . The evolution of transformation-induced defects and their interaction with the microstructure further contributes to the mechanical characteristics observed , indicating the interplay of thermal history, microstructural evolution, and phase transformations in determining the resultant material properties.

The grain size alone does not straightforwardly affect εmax and σmax values during thermal cycling, as opposed to its predictable effect in other materials. Instead, these properties are significantly influenced by the residual fcc phase content . While larger grains are typically expected to increase deformability (εmax), the observed εmax decreases with additional cycling due to the reduced fcc phase, which facilitates less stress-induced transformation and less plastic accommodation . Therefore, in polycrystalline cobalt, both grain size and fcc phase content critically determine mechanical performance during thermal cycling, overshadowing the isolated impact of grain size .

Grain size reduction can significantly impact the high-temperature deformation behavior of cobalt by affecting the interaction between slip and transformation processes. During isothermal annealing, grain growth occurs due to elevated temperatures, leading to reduced yield strength but increased deformability . Conversely, thermal cycling reduces grain size while also altering the fcc phase content and creating more pinned defects, affecting material stability during deformation . The combination of grain size reduction and altered microstructural features from different treatments influences how cobalt responds under high-temperature conditions, dictating the deformation pathways, strength, and ductility achievable .

Thermo-mechanical treatment can be strategically designed to modify the microstructure of cobalt to meet application-specific performance requirements by controlling grain size, dislocation density, and the content of the residual fcc phase. By fine-tuning the annealing temperature and the number of thermal cycles, one can influence grain growth and phase content to optimize strength and ductility according to specific needs . The integration of thermal cycling can further refine the microstructure by inducing defect formation and phase transformations that serve to stabilize or enhance desired properties, such as toughness or work hardenability, crucial for different industrial applications .

Stress-induced martensitic transformation plays a crucial role in the mechanical performance of cobalt by allowing plastic accommodation through the hcp↔fcc transformation. This transformation provides additional slip systems for deformation, thus enhancing ductility and facilitating substantial shear deformation . The extent of this transformation and its efficacy in enhancing mechanical performance are heavily dependent on the residual content of the fcc phase, which diminishes with increasing thermal cycling . As the fcc phase content decreases, the potential for transformation diminishes, leading to reduced material deformability and mechanical properties .

The residual fcc phase content critically determines cobalt's deformability under mechanical loading since it underpins the stress-induced fcc↔hcp transformation, which accommodates significant plastic strain by providing additional slip systems . Materials with higher fcc phase content can facilitate more extensive transformation under stress, leading to increased ductility and shear capacity. However, as thermal cycling reduces the fcc phase content, the ability of the material to undergo transformation and accommodate strain diminishes, negatively affecting deformability and overall mechanical performance . Thus, maintaining an optimal fcc content is vital for maximizing cobalt's mechanical capability in applications requiring high deformability .

Thermal cycling results in a decrease in all the evaluated mechanical parameters of polycrystalline cobalt, such as fracture strain (εmax), compressive strength (σmax), and yield strength (σ0.2). After 10 thermal cycles, these values lie within the ranges of approximately 0.27–0.32, 885–926 MPa, and 257–287 MPa, respectively. However, after 20 cycles, they further decrease slightly to around 0.25–0.28, 863–885 MPa, and 254–276 MPa, respectively . This decline can be attributed to the reduction in the residual face-centered cubic (fcc) phase content, which reduces the material's ability to undergo deformation-induced martensitic transformation, thus limiting its plastic strain capability .

Transformation-induced defects formed during thermal cycling, such as sessile dislocations, contribute to increased thermal hysteresis. These defects interact with interface dislocations and require additional energy to overcome during subsequent cycles, thereby expanding the hysteresis loop . As these defects pin transformation interfaces, more energy is needed for the reverse transformation, leading to larger hysteresis and affecting the material's thermal-structural stability .

Differential scanning calorimetry (DSC) studies provide crucial insights into the hcp↔fcc transformation by revealing the energy involved in the phase transitions during thermal cycling. DSC data demonstrate significant changes during initial cycles, highlighting the transformation dynamics and the evolving defect structures within the material . As the transformations stabilize over increasing cycles, DSC assists in identifying the thermal hysteresis and the residual energy barriers due to pinned defects, helping to understand the complex interplay between microstructure evolution and phase transformation behavior .

Isothermal annealing at temperatures between 600 °C and 1100 °C results in grain growth proportional to the annealing temperature, with grain sizes increasing from roughly 22 µm to 47 µm . This grain growth leads to a reduction in yield strength (σ0.2) from approximately 486 MPa to 269 MPa and an increase in fracture strain (εmax) from about 0.25 to 0.38 . However, no consistent trend is observed for compressive strength (σmax), which remains between 961 MPa and 1091 MPa across various grain sizes .

You might also like