You are on page 1of 11

Materials Science and Technology

ISSN: 0267-0836 (Print) 1743-2847 (Online) Journal homepage: https://www.tandfonline.com/loi/ymst20

Understanding the yield behaviour of L12-ordered


alloys

Y. M. Wang-Koh

To cite this article: Y. M. Wang-Koh (2017) Understanding the yield behaviour of L12-ordered
alloys, Materials Science and Technology, 33:8, 934-943, DOI: 10.1080/02670836.2016.1215961

To link to this article: https://doi.org/10.1080/02670836.2016.1215961

Published online: 31 Aug 2016.

Submit your article to this journal

Article views: 1864

View related articles

View Crossmark data

Citing articles: 10 View citing articles

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


https://www.tandfonline.com/action/journalInformation?journalCode=ymst20
Understanding the yield behaviour of L12-
ordered alloys
Y. M. Wang-Koh∗
Nickel superalloys exhibit a remarkable characteristic. Their yield stress that required to cause the
onset of plastic deformation increases with temperature. This typically occurs up to a temperature
of around 800°C. This effect is thought to originate from the precipitates of the microstructure, which
have an L12-ordered crystal structure. A number of other L12-based alloys exhibit similar yield
properties. It is generally accepted that this is caused by the exhaustion of dislocations by
cross-slip from {111} glide planes to {010} planes on which they are sessile. However, the
underlying mechanisms that control this cross-slipping process are yet to be fully understood,
with little consistency between empirical results and theory. A critical review of the various
theories surrounding nickel superalloys is offered.
Keywords: Anomalous yield, Dislocations, Superalloys, Intermetallics, Cross-slip, Anti-phase boundary

This review was submitted as part of the 2016 Materials Literature Review Prize of the Institute of Materials, Minerals and Mining run by the Editorial
Board of MST. Sponsorship of the prize by TWI Ltd is gratefully acknowledged.

Introduction decade since the inception of the first engine in 1940.


The highest TET achieved by a gas turbine is around
Ni-based superalloys are the industry standard 1600°C.1 This has placed great demand on the turbine,
materials for applications in both marine and aerospace and it is here that yield properties are most critical for
turbine engines, where resistance to creep, fatigue defor- high stress applications. To further improve strength, the
mation, corrosion and oxidation, is required. These volume fraction of precipitates in the superalloys has
materials possess a unique ability to maintain material increased in both the turbine discs and blades.
properties within the harsh environment and tempera- As the volume fraction of γ′ of these alloys is increased,
tures experienced during operation. This is thought to deformation mechanisms that were thought to be only a
be due to precipitates present in their microstructure. feature of turbine discs have begun to appear in the blades.
The peak temperature within the turbine of an aircraft Although not a primary limiting factor in blade alloy
engine exceeds that in which titanium alloys are useful design, yield is becoming increasingly important,
due to their poor oxidation resistance. Conversely, especially where stress concentrations result in low cycle
excellent oxidation and creep resistance of ceramics is fatigue challenges, such as at the turbine blade root.
offset by poor toughness and ductility. Ni-based super- There are many observations of dislocation pairs in two-
alloys have emerged as the choice material for such phase alloys. Through transmission electron microscopy
applications. (TEM), the anti-phase boundary (APB) energies on the
A gas turbine is composed of four sections: the com- active slip plane and cross-slip plane have been measured
pressor, combustion chamber, turbine and exhaust noz- in an attempt to deduce the deformation mechanism.
zle. Hot gases drawn in by the compressor are ignited However, the mechanism governing the locking and
in the combustion chamber and expand within the tur- unlocking processes identified in single-phase intermetal-
bine, which extracts mechanical work in order to drive lics is yet to be confirmed nor translated across to two-
the compressor. The gas turbine is an example of a phase superalloys.
heat engine, whose efficiency is given by: η = 1 − (T1/
T2), where T1 and T2 are the initial and final tempera-
tures, respectively. The most practical way to increase The L12 phase
engine efficiency is to try and raise T1, the temperature Nickel superalloys consist primarily of two phases: a Ni3Al
at which gases enter the turbine, known as the turbine γ′ -phase is present in the microstructure as precipitates,
entry temperature (TET). often cuboidal (Fig. 1), surrounded by a pure Nickel matrix
Great emphasis has been placed on increasing the TET, γ phase. This γ phase has a face-centred cubic (f.c.c.) struc-
which has led to improvements of around 100°C every ture, compared to the Ni3Al, γ′ phase displays the primitive
cubic, ordered, L12 crystal structure, with Al atoms at the
cube corners and Ni atoms at the centre of each face (Fig.
Department of Materials Science and Metallurgy, University of Cambridge, 2). The properties of the superalloys are thought to depend
27 Charles Babbage Road, CB3 0BN, UK on the coherency of the γ/γ′ interface, as will be shown in

Corresponding author, email ymw24@cam.ac.uk ‘Effect of lattice misfit’. A distinct cube–cube orientation

© 2016 Institute of Materials, Minerals and Mining


Published by Taylor & Francis on behalf of the Institute
Received 29 January 2016; accepted 15 July 2016
934 DOI 10.1080/02670836.2016.1215961 Materials Science and Technology 2017 VOL 33 NO 8
Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

has consequences on the deformation mechanisms of


Ni3Al.
As the dislocation penetrates further into the γ′ precipi-
tate, it experiences resistance as the width of the APB
grows. The passage of a second dislocation with the
same Burgers vector would remove the APB, restoring
the structure to an undefected state. Therefore, dislo-
cations move in pairs through γ′ , as supported by images
from TEM.

The cross-slip process


The yield stress anomaly
1 The microstructure of a single-crystal superalloy after
Unlike in most alloy systems, the Ni superalloys and a
standard heat treatment showing cuboidal γ′ precipitates
select group of L12-based compounds exhibit a remark-
encompassed within a pure Ni γ-matrix2
able characteristic: their yield stress, defined here as the
stress required to initiate plastic deformation by dislo-
relationship exists between γ and γ′ , which can be described cation glide, increases or shows minimal variation with
as follows: increasing temperature, from room temperature up to
around 800°C, before dropping off rapidly. Pure nickel,
{100}g //{100}g′
like most materials, exhibits a decrease in yield strength
with increasing temperature, suggesting the remarkable
, 010 .g // , 010 .g′ mechanical properties stem from the L12-ordered phase.
However, a comparison between alloyed-γ′ and a two-
phase Ni superalloy, Mar-M200, shows the increase in
If the lattice misfit between the γ and γ′ phases is not too
strength is much more substantial in the former (Fig. 3).
large, the γ/γ′ interface remains coherent, minimising the
Through the experimentation of various L12 alloys, the
interfacial energy. The γ′ phase may deform while remain-
yield stress anomaly is principally characterised by the
ing geometrically compatible with the γ-phase.
following features:
In an f.c.c. material, the slip systems are of the type
(i) Below the peak temperature, where the yield stress is
a/2<110>{111}. The Burgers vector of a dislocation
increasing with increasing temperature, the active slip
must be a lattice vector and is normally the shortest
system is predominantly <110>{111}. Beyond the
vector. However, a/2<110> is not a lattice vector in
peak temperature, the active slip system predominantly
the ordered L12 structure of Ni3Al. If a dislocation
becomes <110>{010}.
with said Burgers vector were to glide through, it
(ii) The yield stress usually begins to increase beyond
would move Ni atoms onto Al sites, and Al atoms
room temperature, though exceptions do exist; the
onto Ni sites, resulting in the formation of ‘forbidden’
yield stresses of Ni3Ge5 and Ni3(Si, Ti)6 are known to
Ni–Ni and Al-Al bonds. The slip plane becomes a
increase from liquid helium temperatures (77 K).
boundary between two γ′ structures. This boundary is
(iii) The peak position is dependent upon load
known as an APB. With higher energy Ni–Ni and
orientation.
Al–Al bonds replacing favourable Ni–Al bonds, an
(iv) The yield stress is reversible; prior deformation at
APB requires energy to form. In Ni3Al, APBs on the
T1 and then at T2 < T1, exerts only a modest effect
{111} plane have an energy of around 0.1 J m−2.3 This
higher than a virgin sample deformed at T2.
(v) The yield stress is strain rate insensitive.

2 The crystal structures of (left) fcc γ-phase and (right) L12


γ′ -phase. Substitution of Ni for Al in the L12 crystal struc-
ture reduces the shortest Burger’s vector from [110] to 1/2 3 A comparison of the CRSS of the Ni-based superalloy Mar-
[110]. Two shears are required to put the atoms back to M200 and the individual constituent phases of a superal-
their correct atomic positions loy. Adapted from Pollock and Tin4

Materials Science and Technology 2017 VOL 33 NO 8 935


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

(vi) The work hardening rate is abnormally high, peaks temperature, causing a transition in the mechanism con-
within the temperature range where anomalous yield is trolling flow stress: from exhaustion hardening at low
observed, and is orientation-dependent. temperature, where mobile dislocations are immobilised
(vii) The yield stress anomaly is not a feature of all L12 by interactions with any forest dislocations, debris and
alloys. solute impurities, to debris hardening at high tempera-
ture; dislocations are mobilised by multiplication from
It is thought that all of these characteristics must be old sources and the activation of fresh ones. No transition
included into any mechanism describing the family of of mechanisms was found to account for the increase in
L12 alloys. This is reflected in the significant effort spent yield stress and an alternate theory was proposed by
iterating various models to account for each principal Takeuchi et al., which has now become known as the
characteristic. However, recent experiments have shown TK model, in which segments along the length of a
that for stoichiometric single-crystal Ni3Al the yield stress screw dislocation cross-slip, becoming pinning points
is independent of alloy stoichiometry and orientation- (Fig. 4).13 The temperature dependence is accounted for
independent.7–9 This demonstrates complexities involved by defining the probability of cross-slip as a function of
in producing a coherent, adaptable mechanism to explain temperature:
such atypical phenomenon.  
Ht
f = f0 exp −
Kear–Wilsdorf locking kT
One common agreement between all theories is the cause of
the positive temperature dependence is due to the cross-slip where f is the probability of cross-slip, f0 is the probability
of dislocations from octahedral {111} planes onto cubic constant, Ht is the activation energy of cross-slip, k is
{100} planes. Since {100} is not a glide plane, dislocations Boltzmann’s constant and T is the temperature. This
in this cross-slipped configuration become immobile, form- suggests a greater number of pinning points along a dislo-
ing Kear–Wilsdorf locks.10 Dislocations find it favourable cation length are generated at decreasing pinning point
to cross-slip onto {100} planes because the APB energy is spacing as the temperature increases, and a higher level
lower on these planes compared to the active {111} slip of effective stress is required to keep dislocations in a
plane. With increasing temperature, a greater proportion steady-state motion.
of dislocations become immobile by cross-slip, resulting These pinning points unlock when forces either side of
in an increase in strength. Slip activation on the cube planes the kink cause the segments of dislocations adjacent of the
is thought to lead to the decline in stress past the tempera- pinning point to collapse. At a critical angle, unpinning
ture corresponding to peak stress in L12 alloys. would occur athermally, with the dislocation free to
However, this explanation for the deformation behav- glide again until further cross-slip occurs. Paidar et al. 15
iour is incomplete and unable to explain some character- developed this model further, accounting for why the dis-
istics thought to be principle to the yield stress anomaly of location segments become immobile, proposing the APB
L12 alloys. Namely, the anomalous increase in stress anisotropy originated from the spread of the dislocation
begins as low as −196°C for some alloys,5,6,11 far too core into a non-planar configuration. They were then
low for a thermally activated mechanism. able to explain the tension/compression asymmetry and
orientation dependence, though adapting the unpinning
mechanism to a thermally activated process.
Modelling The tension–compression asymmetry of the critically
resolved shear stress (CRSS) was justified by the Escaig
Pinning models
Various models have attempted to explain the above
characteristics, all of which are thought to be principal
features of anomalous yielding, but have been met with
difficulty in relation to experimental observations. The
earliest model proposed a mechanism where screw dislo-
cations cross-slip from {111} planes, on which they are
mobile, to {100} planes, becoming immobile; the driving
force being the differential in APB energy between {111}
and {100} planes.12
The positive temperature dependence is attributed to an
increased propensity for {010} slip with increasing

5 Schematic of compact cross-slip mechanism, also known


4 Schematic of the successive positions for a dislocations as Escaig mechanism. The dislocation immediately re-dis-
moving on the (111) plane by bowing between pinning sociates onto the cross-slip plane, represented by the
points14 darkly shaded region bound by points A and B17

936 Materials Science and Technology 2017 VOL 33 NO 8


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

mechanism for cross-slip, shown schematically in Fig. 5.16 The steady-state configuration for the unlocking–
A pair of partial dislocations is compressed against each locking sequence is shown in Fig. 6. The locks are
other on the primary slip plane. The Escaig mechanism considered to be stronger, and harder to unlock, than the
analyses the dislocation core width of this pair before pinning points of the PPV model. Unlocking occurs by a
and after cross-slip. If it is energetically favourable for process that is thermally activated but also involves a
the faulted ribbon between the two partials to expand large athermal component, to account for the small strain
more widely on the cross-slip plane than the slip plane, rate sensitivity.14 A sudden increase in strain rate decreases
the total superpartial pair constricts, before spon- the critical length, producing a large number of mobile
taneously cross-slipping and dissociates on the cross- superkinks, rapidly increasing the number of mobile super-
slipped plane. The difference between the normalised dislocations. These superdislocations would then move
stresses on the slip and cross-slip planes gives rise to the rapidly to accommodate the new strain rate level.
tension/compression asymmetry. This asymmetry, a key Hirsch’s model is able to explain the orientation
characteristic of the anomalous yield phenomenon, dependence and tension/compression asymmetry and
would have been missed by Takeuchi and Kuramoto non-Schmid effects by adopting the activation enthalpy
since their experiments involved only compression tests. and Escaig mechanisms of the PPV model, despite the dif-
A further contribution to the energy balance of cross- fering origins of the locked screws configuration between
slip was made by Yoo and co-workers.18 The elastic aniso- the two models. Furthermore, the Escaig model fails to
tropy exerts an additional torque on a dislocation pair. In quantitatively explain the orientation effects because
the case of two parallel screw dislocations, this torque Escaig himself admits that the empirical data used to sup-
facilitates the cross-slip from {111} onto {010}, where port the model is unreliable.16 Nevertheless, it must be
the torque disappears due to symmetry. emphasised that this is not the focus of Hirsch’s model,
Both the TK and PPV models have been adopted to rather it does a good job of explaining the small strain
explain anomalous yield because of their strong agree- rate dependence, another key feature of yield behaviour
ment with the macroscopic, mechanical observations in L12-based compounds.
associated with anomalous yield behaviour. However, Hirsch’s model is deemed one of the most successful
currently no TEM investigations – neither in situ nor superkink unlocking models, due to the depth of detail
post-mortem samples have succeeded in producing describing the locking and unlocking mechanisms. Unlike
further evidence in support of the pinning process. the PPV model where unlocking is a purely athermal
Instead, the microstructures of L12-based alloys exhibit process, unlocking occurs via a hybrid thermal/athermal
long, screw dislocations adopting either incomplete or process. Furthermore, PPV assumed that jogs move
complete Kear–Wilsdorf (KW) locking configurations.19 slowly and screws are pinned only locally compared to
Furthermore, although the model of Pope et al. does a the fast movement of jogs on {010} planes, which requires
good job rationalising the sensitivity of flow stress to crys- a finite, albeit unknown, amount of activation energy and
tallographic orientation, it struggles to account for the thus can only occur above a certain temperature.
independence of strain rate to the yield stress. A balance However, both the PPV pinning model and Hirsch’s
in activation energy between the formation and unlocking superkink model possess two underlying assumptions
of dislocation-locks needs to be established because it is that hinder their ability to account for all empirical fea-
the activation energy of the unlocking mechanism that tures of yield behaviour. First, both models assume a
is responsible for the strain rate sensitivity. Microstruc- characteristic length from which no basis is explained;
tural observations of dislocation features that are left for the PPV model, this is the characteristic distance
behind after flow show small, elongated dipole loops.20 between pinning points, whereas for superkinks, it is
Explaining their formation led to the development of an the length of the superkinks that is assumed to be con-
alternative unlocking mechanism and different set of stant. Second, the models are based on a steady-state
models. fluctuation. This means the density of obstacles along a
superdislocation remains constant, and only upon reach-
Kink expansion models ing a critical stress, can these superdislocations become
unlocked. Beyond this stress, all obstacles become
The presence of long, straight screw superdislocations is mobile and are eliminated simultaneously. This would
one of the most significant microscopic observations in translate to a sharp discontinuity in the flow stress at
L12 compounds. Edge dislocations mainly act as links the yield point. However, yielding in L12-based com-
between long screw dislocations and their cross-slipped pounds occurs gradually. Furthermore, since all the
parts, and are mobile.21 These step segments have been cross-slipped segments dissolve at a critical stress, flow
termed ‘superkinks’.19 These superkinks can then shuffle continues indefinitely at a uniform stress. Empirically,
along the dislocation, forming long segments of screw single crystals exhibit significant work hardening beyond
orientation cross-slipped on (010) planes, which are yield.
locked by KW locks. As a result, the superkink model was revised.14,22 The
adapted model assumes the superkink height is statisti-
cally distributed and must lie between upper and lower
bound values to become mobile. Above the upper
bound critical height, the superkinks not only become
mobile but can also generate new superkinks. Below the
6 The steady-state configuration for the unlocking–locking lower value, the superdislocation is locked. An increase
sequence. Unlocking occurs at A by superkink AB and in temperature shifts these critical values so that a greater
progresses through the sequence 1 to 6. The screw is portion of dislocations are locked than mobile upon
eventually locked again along CD21 increasing temperature. The mechanism and causation

Materials Science and Technology 2017 VOL 33 NO 8 937


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

for the superkink height variation with temperature how- and pure Ni3Al. As a result, the flow stress behaviour of
ever is not clear. many Ni3(Al, X) alloys has been analysed.25–29
The mobility of a screw dislocation is proportional to Various hypotheses relating the mechanical properties
the average height of the superkinks on it. If the average to the yield stress anomaly have been proposed, though
superkink height falls below a critical value, a mobile the wide variety of factors thought to affect the anomaly
dislocation will experience ‘exhaustion of mobility’ makes forming a coherent argument difficult. Neverthe-
and cease. The mobile dislocation density and kink less, it is commonly agreed that the yield stress anomaly
height evolution are dependent on one another as well is a result of competition between the exhaustion (lock-
as temperature and applied stress. Including these revi- ing) and multiplication of mobile dislocations. The most
sions, the prediction of L12 compounds’ yield behaviour commonly adopted theories are outlined below, with
improves. Some discrepancies between experimental and accompanying experimental justifications. It should be
theoretical results still exist. Since the density of mobile noted here that for alloys in which the yield stress increase
superdislocations can change rapidly, the dynamic in accompanied by a change on long range order par-
changes in superdislocation mobility needs to be fac- ameter, such as Cu3Au30 and β-brass,31 the following
tored into the model. However, at the time of design, mechanisms and discussion do not apply and have been
this could not occur due to the limited experimental omitted from the current discussion.
understanding of the dynamical process of exhaustion The Kear–Wilsdorf locking mechanism is the most
in L12 compounds. commonly quoted explanation for observed, anomalous
This section highlights the inconsistencies present yield behaviour in L12-based alloys. Post-mortem TEM
between proposed models and experimental observations. studies of dislocations lead to quick judgement calls on
The origin of the yield anomaly by exhaustion of dislo- what controls anomalous yield. Ni3Ga is the intermetallic
cation motion remains to be elucidated. In particular, from which the TK model is founded. Similar to Ni3Al,
the specific mechanisms and causality of dislocation lock- the yield stress of Ni3Ga under compression reaches a
ing and unlocking. Significant work is required, both maximum around 425°C, though the increase in yield
experimentally and theoretically before an appropriate stress is steeper than Ni3Al.32 This behaviour was
model should be proposed. accounted for by a higher {111} to {010} APB energy
ratio in Ni3Ga compared to Ni3Al, though no measure-
ments or theoretical calculations were conducted and
Anomalous yield in intermetallic L12 the justification seems dubious.
APB energies are notoriously difficult to calculate, both
alloys theoretically and experimentally. The most common
Stoichiometric, single crystals of Ni3Al are difficult to syn- method experimentally is to measure the dissociation
thesise due to the peritectic reaction during solidification, widths between partial dislocations by TEM under
described by the reaction: liquid + NiAl → Ni3Al, because weak-beam condition; this technique is thought to give
grains that nucleate during the peritectic reaction are diffi- the best compromise between adequate resolution and
cult to suppress.23 Elemental additions to Ni3Al or devi- comfortable working conditions. However, sources of
ations from stoichiometry towards the Ni-rich side avoid error make it difficult to ascertain exact values for APB
this reaction,24 and alloy designers have considered repla- energies. One problem arises from the fact that any
cing some high-temperature, high-strength alloys with observed dislocation configurations may not be under
intermetallics due to the similarities between the γ′ -phase complete equilibrium. Furthermore, since images are
taken under diffracting conditions, a dislocation is shifted
with respect to its line and the shift is unique to each par-
tial, the resultant separation does not correlate directly
with the actual separation in the crystal.33 Baluc et al.
have shown the APB energy ratio is often overestimated
prior to considering contributions from elastic
anisotropy.34

Dislocation cores
The immobilisation of screw dislocations, observed by
TEM, was thought to be analogous to the plastic defor-
mation of body-centred cubic (b.c.c.) metals.35,36 It is
now commonly accepted that low temperature plastic
deformation of b.c.c. metals is controlled by screw dislo-
cations which can cross-slip very easily but are simul-
7 Schematic of the possible core configurations. The oval taneously sessile due to the non-planar structure of the
shape represents the core of the ½[−101] superpartial. dislocation cores.37,38 It was thought that screw dislo-
When the APB lies on the (111) plane, the core also lies cations behave in a similar manner in the L12 alloys, lead-
on that plane (top position); here the superpartial is glis- ing to work analysing the nature of dislocation cores in
sile. When the APB lies on the (010) plane, the core model L12-ordered alloys by computer simulation
spreads onto the (1−11) or (111) plane (middle and bottom techniques.
ovals, respectively), depending upon the position of the In the L12 structure, two {111} planes and one {010}
core. In both of these configurations, the superpartial is plane share a common line along any <110> direction,
sessile15 providing an opportunity for screw dislocations to spread

938 Materials Science and Technology 2017 VOL 33 NO 8


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

into several non-parallel crystallographic planes. Two dis- reducing the average electron atom ratio or increasing
sociation configurations are possible: a glissile one on the atomic radius ratio.5,39–42 The effects of compo-
{111} planes and a sessile one on {010} planes. When sitional deviations from stoichiometry32,43 and ternary
the dislocation core is planar and spread in the plane of additions44–46 also fit the predicted behaviour of this
the APB, as shown in Fig. 7, the dislocation is glissile. concept.
The core can also be non-planar, extending in both the Pt3Al and Pt3Ga have an L12 crystal structure yet pos-
(111) and (1−11) planes; the dislocation possessing this sess drastically differing flow behaviour.47–50 These alloys
core configuration is sessile. Two further symmetry- exhibit an increase in flow stress with decreasing tempera-
related configurations are also possible, where the APB ture (see Fig. 9); a phenomenon termed the low tempera-
is on the (010) plane with the core spread onto (111) ture anomaly (LTA). In the L12 structure, Pt atoms form
and (1−11) planes, respectively. Both of these configur- an octahedron within the unit cell as shown in Fig. 10.
ations are sessile. The D0c and D0c′ structures are obtained by distorting
Before cross-slip, the APB lies entirely on the (111) or rotating this Pt octahedra, respectively.49 Both trans-
plane and the core is coplanar with the APB. After formations are accompanied by a tetragonal distortion
cross-slip, the APB lies on the (010) plane, the core of the lattice. The D0c′ structure was found to be more
spreads on the (1−11) or (111) plane. In both cases, energetically favourable than the L12 structure in the
after cross-slip onto (010), the dislocation core is not con- temperature region where the LTA is observed, while
fined to the plane determined by the APB resulting in a the D0c structure was unstable across the whole tempera-
sessile dislocation with respect to glide on the (010).15 It ture range.
was proposed that [−101] superdislocations dissociate This instability has electronic origins. A transition from
into two ½[−101] superpartials separated by an APB L12 to D0c′ alters the charge distribution in the D0c′ phase
(the case for Ni3Al, Ni3Ga and Ni3Ge), or into ⅓[−211] to become more covalent. Additional electron density
and ⅓[−1−12] superpartials separated by super intrinsic bridges form between Pt and Al atoms, enhancing chemi-
stacking fault, the case for Pt3Al and Pt3Ga, which cal bonding, stabilising the D0c′ structure. The stability
show a dramatic flow stress increase at low temperatures ratio between the D0c′ and L12 structures is also sensitive
(see below). The anomalous increase in flow stress is to stoichiometry. Iterative additions of Al atoms to Pt3
then explained by an increasing quantity of core trans- −xAl1+x cause the Pt octahedra to rotate back to the origin
formations from the glissile to sessile forms. position of the L12 structure. The L12 structure was found
to become stable at compositions in excess of 6 at.-% Al.48
Despite the consistency of argument to justify anoma-
Phase stability lous yield through phase stability, further calculations
The phase stability of the L12 crystal structure with based on phase stability for Ni3(Al, X) single crystals
respect to isomorphous crystal structures is also thought and other L12-based alloys are yet to be fulfilled.
to affect anomalous yield behaviour. This in particular
has gained support following analysis of the yield stress
upon systematic stoichiometric deviations where it is
hypothesised that increasing levels of non-stoichiometry
affect the dislocation cores.
The phase stability theory was first proposed to account
for why the yield stress anomaly was not a feature of all
L12-ordered alloys.39,40 It was shown that a lower L12
phase stability with respect to the isomorphous ordered
f.c.c. D022 phase (Fig. 8) reduced the APB energy on
cube planes to almost zero, facilitating cross-slip from
{111}. In general, substitution of ternary elements away
from the stoichiometric composition reduces the compo-
sition region in which the L12 structure is stable by

9 Temperature dependence of CRSS for three different


orientations under compression for single crystals of
8 The crystal structure of L12 (left), and D022 (right) Pt3Al. All three orientations show evidence of the LTA47

Materials Science and Technology 2017 VOL 33 NO 8 939


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

10 From top left, clockwise: 3D crystal structure of L12 Pt3Al, {001} projection of L12, D0c, D0c′ crystal structures49

Extending towards two-phase comparison between the constituent phases of a superal-


loy, γ and γ′ , and a commercial superalloy Mar-M200.
superalloys Furthermore, upon reaching the peak temperature, the
Despite the abundant source of information regarding sharp drop-off in yield stress is thought to occur via differ-
how specific elemental additions to Ni3Al affect ent mechanisms; in Ni3Al-base alloys, there is a transfer
flow stress behaviour, the behaviour of Ni3(Al, X) and from octahedral slip to cube slip as more cubic slip sys-
Ni3−xAlx compounds is considerably different from that tems are activated, while in a two-phase superalloy the
of a two-phase Nickel superalloy. First, the yield stress decrease is attributed to a decrease in γ′ volume fraction.
increase with increasing temperature is not as steep, and The dislocation structures observed through TEM are
occurs over a much narrower temperature range in a also quite different. Instead of the long lengths of screw
two-phase superalloy. This is shown in Fig. 3, a dislocations, divided by superkinks, in a two-phase super-
alloy, the dislocations grow out from built-in dendritic
cores in the γ-phase and are constrained within the narrow
γ-matrix channels. Upon application of sufficient stress,
the dislocations enter in the γ′ precipitates, where they
split into partial dislocations. Dislocations may therefore
possess a different Burgers vector since they originate as
non-APB-bound dislocations. The configurations
adopted by partial dislocations within the γ′ precipitates
vary in shape, and are sometimes accompanied by stack-
ing faults, as shown in Fig. 11.
To further develop the performance and usefulness of
superalloys, our understanding of anomalous yielding
needs to extend to Nickel-based superalloys. Upon shift-
ing to a two-phase γ/γ′ system, the additional effects on
the strength must be taken into account: the volume frac-
tion and size of γ′ , the misfit between γ and γ′ and the APB
11 The deformation structure of CMSX-4 subject to creep energy of γ′ . Clear relations describing those effects, theor-
conditions shows stacking faults accompanying dislo- etical or empirical, are still yet to be fully understood and
cations in γ and dislocation pairs in γ′ precipitates51 significant data supporting them are still lacking.

940 Materials Science and Technology 2017 VOL 33 NO 8


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

Shah and Duhl52 attempted to relate the precipitate size creep. For alloys with a negative lattice misfit, the precipi-
to yield strength. The yield strength has been hypothesised tates inhibit deformation in adjacent vertical γ channels,
to be controlled by the distance between cube cross-slip while horizontal γ channels are deformed quite easily.
events and γ′ precipitate size in the superalloy PWA- Zhang et al. 56 attempted to clarify the effect of lattice mis-
1480 by the following relation: fit on dislocation motion during creep in TMS-75 and
TMS-138A. The misfit values were determined by X-ray
1 1
s= + analysis as TMS-75 is found to have a smaller misfit
R l than TMS-138: −0.16 to −0.33%, respectively. TEM
where sigma is the yield strength, R is the precipitate size observations during primary creep showed dislocation
and λ is the distance between cube cross-slip events within loops gliding between narrow matrix channels during pri-
the precipitates upon shearing. In the temperature region mary creep.
room temperature to peak yield temperature, the γ′ size
was thought to control strength since the distance Effects of elemental partitioning
between cross-slip events is the smaller of the two par-
The effect of Re and Ru on the creep strength of Ni-based
ameters. Around the peak temperature, the distance
superalloys has been extensively studied, due to the ben-
between cross-slip events becomes very small, in turn
eficial additions each element has provided to superal-
becoming the dominant factor controlling strength. It is
loys.58 Re partitions to the γ-matrix, hardening the γ-
yet to be determined what controls a specific segment of
phase by solid solution strengthening.55,59 The effect of
a dislocation line cross-slipping. This theory breaks
Ru however is still contentious.60 It has been reported
down when stacking faults become the dominant shearing
that additions of Ru homogenise the distribution of Re
mechanism.
between the γ and γ′ , known as reverse partitioning, in
turn slightly reducing the strength of the γ-phase.61 How-
Effect of lattice misfit ever, there has been significant disagreement on this
In addition to the deformation of both the γ and γ′ phases, effect.62,63
the γ/γ′ interface is thought to play a crucial role in the Additions of Re have been shown to make the lattice
deformation mechanism of two-phase superalloys. Poly- misfit more negative, while additions of Ru were found
crystalline Ni-based superalloys usually have low lattice to have little effect. Re-containing alloys have been
misfit values and a γ′ volume fraction of around 50%.53 thought to improve creep strength. High lattice misfit,
This is in stark contrast to the grain-boundary-free micro- associated with the presence of Rhenium, leads to the for-
structures of single-crystal superalloys with up to 70% γ′ mation of fine, elongated rafted microstructures and
volume fraction and, relatively, much higher lattice misfit. dense interfacial dislocation networks during creep.
In both cases, the misfit between matrix and precipitates These aid creep strength at elevated temperatures.
creates significant internal stresses which are thought to Whether these produce a similar effect on the yield stress
be a contributing factor to the mechanical high-tempera- remains to be investigated.
ture behaviour, though the resultant, specific effect on dis-
location behaviour is contentious.
A coherent interface exists between the γ and γ′ phases Conclusion
within a Ni-based, single-crystal superalloy. Nevertheless, It is known that the Ni3Al, L12-phase imparts strength to
the small difference in their respective lattice parameters a Nickel superalloy and allows anomalous yield
results in a lattice misfit, denoted δ, and defined as behaviour to be exhibited. Therefore, the study of yield
behaviour has up to now focused predominantly on
2(ag′ − ag )
d= L12-ordered alloys and ternary Ni3(Al, X) compounds.
a g ′ + ag However, the mechanism controlling anomalous yielding
where aγ and aγ′ are the lattice parameters of the γ and γ′ is still to be elucidated. The most commonly adopted
phases, respectively. For all commercial Ni-based superal- mechanism involves the cross-slip of dislocations from
loys, the misfit value is negative at engine-operating {111} planes to {100} planes, where they become locked.
temperature.54,55 The locking and unlocking processes and detailed mech-
The effect of lattice misfit on the dislocation motion in anisms that govern such processes are also still unclear.
single-crystal superalloys is currently unclear because it is Various models have been proposed, with increasing com-
difficult to isolate and test the lattice parameters against patibility with the various macroscopic mechanical
the strength of superalloys in situ; lattice misfit can be characteristics observed in L12-ordered alloys. Both still
changed by altering the alloy composition that concomi- lack the versatility to explain all characteristics thought
tantly modifies a number of other properties. Neverthe- to be principal features of the anomalous yield behaviour.
less, two lines of argument exist: the first suggests Though present, the anomalous yield effect is not as
keeping misfit as small as possible to minimise the coher- pronounced in two-phase γ/γ′ systems. Factors such as
ency stresses, producing a more stable microstructure.56 volume fraction and size of γ′ , the misfit between γ and
Alternatively, Harada and co-workers57 propose a larger γ′ and the APB energy of γ′ additionally need to be
misfit favours the formation of denser dislocation net- accounted for.
works at the γ/γ′ interfaces, protecting the shearing of γ′
precipitates, in turn stabilising the microstructure.
Experiments attempting to elucidate the effect lattice
Acknowledgements
misfit has on creep strength have shown that a combi- The author gratefully appreciates the support of his super-
nation of the sign of the lattice misfit and the direction visor Professor Cathie Rae of the University of Cam-
of applied stress are both factors in high-temperature bridge, and Dr Neil Jones of Rolls Royce plc. along

Materials Science and Technology 2017 VOL 33 NO 8 941


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

with the EPSRC and Rolls Royce plc. for financial 25. S. Takeuchi and E. Kuramoto: ‘Anomalous temperature dependence
support. of the yield stress in Ni3Ga’, J. Phys. Soc. Jpn., 1971, 31, (4), 1282–
1282.
26. T. Saburi, T. Hamana, S. Nenno and H.-R. Pak: ‘Temperature and
orientation dependence of the yield strength of Ni3(Al, W)’,
ORCiD Jpn. J. Appl. Phys., 1977, 16, (2), 267–272.
27. J. Fang, E. M. Schulson and I. Baker: ‘The dislocation structure
Y. M. Wang-Koh http://orcid.org/0000-0003-4094-6434 in L12 ordered alloy Ni3Ge’, Philos. Mag. A, 2006, 70, (6),
1013–1025.
28. C. Lall, S. Chin and D. P. Pope: ‘The orientation and temperature
dependence of the yield stress of Ni3(Al, Nb) single crystals’,
References Metall. Trans. A, 1979, 10, (9), 1323–1332.
1. Mitsubishi heavy industries, ltd. ‘Archiveorg’, 2016, available at 29. R. D. Rawlings and A. E. Staton-Bevan: ‘The alloying behaviour
https://web.archive.org/web/20131113162749/http://www.mhi.co.jp/ and mechanical properties of polycrystalline Ni3Al (γ′ phase) with
en/news/story/1105261435.html, (accessed 19 June 2016). ternary additions’, J. Mater. Sci., 1975, 10, (3), 505–514.
2. Z. X. Shi, S. Z. Liu and J. R. Li: ‘Rejuvenation heat treatment of the 30. D. P. Pope: ‘The flow stress of Cu3Au’, Philos. Mag. 1972, 25, 917–
second-generation single-crystal superalloy DD6’, Acta Metall. Sin. 925.
(Engl. Lett.), 2015, 28, (10), 1278–1285. 31. G. W. Ardley and A. H. Cottrell: ‘Yield points in brass crystals’,
3. J. Douin, P. Veyssière and P. Beauchamp: ‘Dislocation line stability in Proc. R. Soc. London Ser. A – Math. Phys. Sci., 1953, 219, (1138),
Ni3AI’, Philos. Mag. A, 1986, 54, (3), 375–393. 328–341.
4. T. M. Pollock and S. Tin: ‘Nickel-based superalloys for advanced 32. O. Noguchi, Y. Oya and T. Suzuki: ‘The effect of nonstoichiometry
turbine engines: chemistry, microstructure and properties’, J. Propul. on the positive temperature dependence of strength of Ni3AI and
Power, 2006, 22, (2), 361–374. Ni3Ga’, Metall. Trans. A, 1981, 12, (9), 1647–1653.
5. T. Suzuki, Y. Oya and D. M. Wee: ‘Transition from positive to nega- 33. P. Veyssière and D. G. Morris: ‘Comment on ‘Transmission
tive temperature dependence of the strength in Ni3Ge-Fe3Ge solid electron microscopy of dislocation structures in iron-doped
solution’, Acta Metall., 1980, 28, (3), 301–310. Al3Ti with the L12 structure’’, Philos. Mag. A 1993, 67, (2), 491–495.
6. T. Y. Takasugi and M. Yoshida: ‘Strength anomaly and dislocation 34. N. Baluc, R. Schäublin and K. J. Hemker: ‘Methods for determining
structure at 4.2 K in Ni3(Si, Ti) single crystals’, Philos. Mag. A, precise values of antiphase boundary energies in Ni3Al’, Philos.
1992, 65, (3), 613–624. Mag. A, 1991, 64, (5), 327–334.
7. D. Golberg, M. Demura and T. Hirano: ‘Effect of Al-rich off- 35. M. Yamaguchi, V. Paidar, D. P. Pope and V. Vitek: ‘Dissociation and
stoichiometry on the yield stress of binary Ni3Al single crystals’, core structure of <110> screw dislocations in L12 ordered alloys I.
Acta Mater., 1998, 46, (8), 2695–2703. Core structure in an unstressed crystal’, Philos. Mag. A, 1982, 45,
8. D. Golberg, M. Demura and T. Hirano: ‘Single crystal growth and (5), 867–882.
characterization of binary stoichiometric and Al-rich Ni3Al’, 36. V. Paidar, M. Yamaguchi, D. P. Pope and V. Vitek: ‘Dissociation and
J. Cryst. Growth, 1998, 186, (4), 624–628. core structure of <110> screw dislocations in L12 ordered alloys II.
9. T. Hirano, M. Demura and D. Golberg: ‘Compliance to Schmid’s Effects of an applied shear stress’, Philos. Mag. A, 1982, 45, (5),
law in the stress anomaly regime of binary stoichiometric Ni3Al’, 883–894.
Acta Mater., 1999, 47, (12), 3441–3446. 37. J. W. Christian: ‘The interaction between dislocations and point
10. B. H. W. Kear and H. G. Wilsdorf: ‘Dislocation configurations in defects’, Proc. 2nd Int. Conf. on ‘Strength of metals and alloys’,
plastically deformed polycrystalline Cu3Au alloys’, Trans. TMS- Metals Park, OH, 1970, American Society of Metals, 31.
AIME, 1962, 224, 382–386. 38. V. Vitek: ‘Theory of core structures of dislocations in body-centred-
11. K. Suzuki, M. Ichihara and S. Takeuchi: ‘Dissociated structure of cubic metals’, Cryst. Lattice Defects, 1974, 5, (1), 1–34.
superlattice dislocations in Ni3Ga with the L12 structure’, Acta 39. D. M. Wee, O. Noguchi, Y. Oya and T. Suzuki: ‘New L12 ordered
Metall., 1979, 27, (2), 193–200. alloys having the positive temperature dependence of strength’,
12. P. H. Thornton and R. G. Davies: ‘The temperature dependence of Trans. Jpn. Inst. Met., 1980, 21, (4), 237–247
the flow stress of gamma prime phases having the Ll2 structure’, 40. D. M. Wee and T. Suzuki: ‘The temperature dependence of hardness
Metall. Trans., 1970, 1, (2), 549–550. of Ll2 ordered alloys’, Trans. Jpn. Inst. Met., 1979, 20, (11),
13. S. Takeuchi and E. Kuramoto: ‘Temperature and orientation depen- 634–646.
dence of the yield stress in Ni3Ga single crystals’, Acta Metall., 41. T. Suzuki and Y. Oya: ‘The temperature dependence of the strength
1973, 21, (4), 415–425. of pseudo-binary platinum-based L12 alloys with B-subgroup
14. M. J. Mills and D. C. Chrzan: ‘Dynamical simulation of dislocation elements’, J. Mater. Sci., 1981, 16, (10), 2737–2744.
motion in Ll2 alloys’, Acta Metall. Mater., 1992, 40, (11), 3051– 42. A. Fujita, Y. Mishima and T. Suzuki: ‘Temperature dependence of
3064. strength in a Cu3Au-5% Ni alloy and its relevance to the APB mor-
15. V. Paidar, D. P. Pope and V. Vitek: ‘A theory of the anomalous yield phology’, J. Mater. Sci., 1983, 18, (6), 1881–1886.
behaviour in L12 ordered alloys’, Acta Metall., 1983, 32, (3), 43. T. Suzuki, Y. Oya and S. Ochiai: ‘The mechanical behavior of non-
435–448. stoichiometric compounds Ni3Si, Ni3Ge, and Fe3Ga’, Metall.
16. B. Escaig: ‘Sur le glissement dévié des dislocations dans la structure Trans. A, 1984, 15, (1), 173–181.
cubique à faces’, J. Phys., 1968, 29, (2–3), 225–239. 44. Y. Kuriki, S. Ochiai, M. Yodogawa and T. Suzuki: ‘The effect of zinc
17. J. Bonneville and B. Escaig: ‘Cross-slipping process and the stress addition on the mechanical properties of Ni3Al’, Trans. Jpn. Inst.
orientation dependence in pure copper’, Acta Metall., 1979, 21, Met., 1985, 26, (3), 213–214.
(9), 1477–1486. 45. Y. Mishima, S. Ochiai, M. Yodogawa and T. Suzuki: ‘Mechanical
18. C. L. Fu, Y. Y. Ye and M. H. Yoo: ‘Theoretical investigation of the properties of Ni3Al with ternary addition of transition metal
elastic constants and shear fault energies of Ni3Si’, Philos. Mag. elements’, Trans. Jpn. Inst. Met., 1986, 27, (1), 41–50.
Lett., 2006, 67, (3), 179–185. 46. S. Ochiai, Y. Mishima, M. Yodogawa and T. Suzuki: ‘Mechanical
19. Y. Q. Sun and P. M. Hazzledine: ‘A TEM weak-beam study of dis- properties of Ni3AI with ternary addition of B-subgroup elements’,
locations in γ′ in a deformed Ni-based superalloy’, Philos. Mag. Trans. Jpn. Inst. Met., 1986, 27, (1), 32–40.
A, 1988, 58, (4), 603–617. 47. D. M. Wee, D. P. Pope and V. Vitek: ‘Plastic flow of Pt3Al single crys-
20. C. T. Chou, P. B. Hirsch, M. Mclean and E. Hondros: ‘Anti-phase tals’, Acta Metall., 1984, 32, (6), 829–836.
domain boundary tubes in Ni3Al’, Nature, 1982, 300, 621–623. 48. N. L. Okamoto, Y. Hasegawa and H. Inui: ‘Plastic deformation of
21. P. B. Hirsch: ‘A new theory of the anomalous yield stress in L12 single crystals of Pt3Al with the L12 structure having a far Al-rich
alloys’, Philos. Mag. A, 1992, 25, (7), 1725–1730. off-stoichiometric composition of Pt-29at.%Al’, Philos. Mag.,
22. F. Louchet: ‘Dislocation exhaustion and stress anomaly in 2014, 94, (12), 1327–1344.
L12 alloys: the ‘E.L.U.’ model’, J. Phys. III, 1995, 5, (11), 1803– 49. Y. N. Gornostyrev, O. Y. Kontsevoi, A. F. Maksyutov, A. J.
1807. Freeman, M. I. Katsnelson, A. V. Trefilov and A. I. Lichtenshtein:
23. R. W. W. Guard and J. H. Westbrook: ‘Alloying behaviour of Ni3Al ‘Negative yield stress temperature anomaly and structural instability
(γ′ phase)’, Trans. AIME, 1959, 215, 807–814. of Pt3Al’, Phys. Rev. B, 2004, 70, 014102.
24. B. R. McDonnell, R. T. Pascoe, G. F. Hancock and C. W. A. Newey: 50. Y. Oya-Seimita, T. Shinoda and T. Suzuki: ‘Low temperature
‘The growth of single crystals of the intermediate phases NiAl and strength anomaly of L12 type intermetallic compounds Co3Ti and
Ni3Al’, J. Mater. Sci., 1967, 2, (4), 365–370. Pt3Al’, Mater. Trans., JIM, 1996, 37, (9), 1464–1470.

942 Materials Science and Technology 2017 VOL 33 NO 8


Wang-Koh Understanding the yield behaviour of L1 2 -ordered alloys

51. V. Sass, U. Glatzel and M. Feller-Kniepmeier: ‘Anisotropic creep temperature low-stress creep’, Acta Mater., 2005, 53, (17), 4623–
properties of the nickel-base superalloy CMSX-4’, Acta Mater., 4633.
1996, 44, (5), 1967–1977. 58. A. Heckl, R. Rettig, R. F. Singer: ‘Modeling of precipitation kinetics
52. D. M. Shah and D. N. Duhl: ‘The effect of orientation, temperature of TCP-phases in single crystal nickel-base superalloys’, Adv. Mater.
and gamma prime size on the yield strength of a single crystal nickel Res., 2010, 278, 339–344.
base superalloy’, Superalloys, 1984. doi:10.7449/1984/Superalloys_ 59. X.-X. Yu and C. Y. Wang: ‘The effect of alloying elements on the
1984_105_114. dislocation climbing velocity in Ni: a first-principles study’, Acta
53. R. C. Reed: ‘The superalloys: fundamentals and applications’, 2006, Mater., 2009, 57, (19), 5914–5920.
New York, Cambridge University Press. 60. S. Neumeier, F. Pyczak and M. Göken: ‘The influence of ruthenium
54. H. Mughrabi: ‘The importance of sign and magnitude of γ/γ′ lattice and rhenium on the local properties of the γ - and γ′ -phase in nickel-
misfit in superalloys—with special reference to the new γ′ -hardened base superalloys and their consequences for alloy behaviour’, 2008.
cobalt-base superalloys’, Acta Mater., 2014, 81, 21–29. doi:10.7449/2008/Superalloys_2008_109_119v.
55. T. M. F. Pollock and R. D. Field: ‘Chapter 63: dislocations and high- 61. K. W. O’Hara, W. Ross and R. Dorolia: ‘Nickel base superalloy’,
temperature plastic deformation of superalloy single crystals’, 2002, US Patent, 1995.
Amsterdam, Elsevier. 62. J. Chen, Q. Feng and Z. Sun: ‘Topologically close-packed phase pro-
56. J. X. Zhang, T. Murakumo, Y. Koizumi and H. Harada: ‘The influ- motion in a Ru-containing single crystal superalloy’, Scr. Mater.,
ence of interfacial dislocation arrangements in a fourth generation 2010, 63, (8), 795–798.
single crystal TMS-138 superalloy on creep properties’, J. Mater. 63. R. A. Hobbs, L. Zhang, C. M. F. Rae and S. Tin:
Sci., 2003, 38, (24), 4883–4888. ‘Topologically close-packed phase promotion in a Ru-contain-
57. J. Zhang, J. Wang, H. Harada and Y. Koizumi: ‘The effect of ing single crystal superalloy’, Mater. Sci. Eng.: A, 2008, 489,
lattice misfit on the dislocation motion in superalloys during high- (1–2), 65–76.

Materials Science and Technology 2017 VOL 33 NO 8 943

You might also like