Strain Hardening in Aerospace Alloys
Strain Hardening in Aerospace Alloys
Strain Hardening in Aerospace Alloys
net/publication/275273103
CITATIONS READS
20 3,226
3 authors, including:
Some of the authors of this publication are also working on these related projects:
To find out the effectiveness of flow forming technique for Aluminium alloy AA2219 tubes View project
All content following this page was uploaded by R.K. Gupta on 14 January 2016.
Strain hardening is one of the important strengthening mechanisms, which plays significant role in processing and application
of metals and alloys. For non‐heat treatable alloys, it becomes more important. Its effect is different in different metals and
alloys and accordingly specific process and application regime are selected. A large variety of metals and alloys from the family
of light alloys (Al, Ti based), high strength steels and high temperature alloys (Co, Ni and Nb based) are used in aerospace
systems. This paper analyses importance of strain hardening phenomenon in these alloys. Attempts are made to explain the
differential behaviour of various alloys in governing the tensile to yield strength ratio along with % elongation. Role of
temperature in this behaviour is also included.
Keywords
Strain Hardening; Aerospace Alloys; Ti6Al4V; Maraging Steel; Al Alloys
Introduction
Strain hardening or work hardening is one of the most commonly used means of improving strength of an alloy. In
a simple way it is the use of permanent deformation to increase the strength of the metal. It is either denoted by
tempers (as half hard, full hard, spring temper, etc.) in case of steels or by % cold work in case of light metals/
alloys (aluminium/ titanium). Strain hardening in metals, i.e. the capacity of the material where flow stress
increases with increasing plastic strain, is being studied since the discovery of dislocations, and is still a matter of
current interest. Understanding the strain‐hardening capability of structural alloys is of practical importance also
since it controls their fracture properties and deformability.
Strain hardening is an important industrial process that is used to harden metals or alloys that do not respond to
heat treatment. For alloys strengthened by solid‐solution addition, rate of strain hardening may be either increased
or decreased as compared to pure metals. Dieter explained that the final strength of a cold worked solid solution
alloy is mostly greater than that of pure metal cold worked to the same extent at the same temperature. Increasing
temperatures reduces strain hardening and accordingly strength. The rate of strain hardening can be assessed from
the slope of the true stress‐true strain graph (flow curve). Generally, the rate of strain hardening is lower for hcp
metals than for cubic metals.
Increase in strength can be seen from schematic Fig. 1 explaining strain hardening phenomenon. It also explains
that material shows increase in strength in unloading and reloading. The degree of cold working determines the
strength of a metal. As the amount of cold work increases, so does the strength. However, the total elongation also
changes with amount of cold work. Harder tempers like spring and super spring have high strength and low
ductility (bad formability), whereas softer tempers like annealed and 1/4 hard have low strength and high ductility
(very good formability). When choosing a material for specific application, it is best to use the highest strength
material that still meets the formability requirements of the design.
As the percentage cold work increases, there is a diminishing increase (increase with lower rate) on strength.
Generally it is seen that increase in strength with increase in cold work results in lowering of % elongation. This is
due to the fact that at higher amount of reductions, there are fewer free dislocations to become entangled. Since the
material is less able to plastically deform, fracture becomes much more likely. At high levels of cold work, the
material becomes very difficult to further process or form. If it must be formed, or reduced further in thickness,
then annealing becomes necessary. Extent of reduction/ changes in ductility due to cold work is different for
Frontiers in Aerospace Engineering, Vol. 4 No. 1‐May 2015 1
2325‐6796/15/01 001‐13, © 2015 DEStech Publications, Inc.
doi: 10.12783/fae.2015.0401.01
2 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
different alloys. It depends on several other factors like type of alloy, stacking fault energy, number of phases,
strain hardening behaviour and strengthening mechanisms in the alloy.
Strain hardening results from the interaction of dislocations with each other and with the various constituents of
the microstructure, such as, grain boundaries, precipitates and solutes, etc. When a material is permanently
deformed, the dislocations move until they are stopped in the crystalline lattice. The most effective dislocation
barriers are another dislocation. Where dislocations run on different planes and intersect, they cannot pass through
each other. Finally, dislocations pile up against each other, and become inter‐twined. This dislocation entanglement
prevents further permanent deformation of that particular grain, without use of significantly greater energy. This
greatly increases the strength of the material in subsequent loading.
FIG. 1: SCHEMATIC REPRESENTATION OF STRAIN HARDENING
Varieties of alloys in various forms are used in aerospace systems, which are realised through different processing
techniques. Strain hardening has important role in processing as well as in application of these materials. This
phenomenon has been extensively studied and reported. However, its role and importance in aerospace alloys has
not been discussed adequately. The present paper attempts to highlight, analyse and review this important
phenomenon in aerospace alloys.
Strain Hardening and Plastic Instability
Plastic instability is another condition, which is closely associated with strain hardening phenomenon. It is an
established fact that necking is the result of a trade‐off (during deformation) between the strain hardening process
within the material and the level of the applied stress. Metals generally show a reduced strain hardening with
increasing strain while the level of the stress continues to increase. At some point, the two values cross and the
material is plastically unstable and the smallest of defects is sufficient to promote localized deformation at that
position. Essentially all deformation is localized thereafter in the necked region until fracture occurs. The greater
the strain hardening coefficient, the plastic instability is delayed. Delaying plastic instability to higher strain levels
is desirable for forming operations and also for increased toughness. Plastic instabilities continue to be a real
concern in the design of structures. Strain localization is mostly promoted by large inclusions, by triaxial stresses
(e.g. sharp notches and cracks) and are pronounced in materials with low strain hardening and low strain rate
hardening.
Strain Hardening in Aerospace Alloys 3
Kujawski and Ellyin showed that for a strain hardening material, the plastic zone size in plane stress is affected as a
result of strain hardening. The larger the strain hardening, the greater the strain to instability. It would thus appear
that the ductility is increased by decreasing the volume fraction and size of particles (fewer voids nucleated) and
that it is especially effective in materials of high strain hardening.
Strain Hardening and Fracture
Griffith explained the fracture problem arising from small cracks and computed the fracture stress, while Orowan
applied this result to materials which had some degree of plasticity by adding an additional term to the true
surface energy with the plastic deformation. Thus more realistic picture that develops for most structural materials
is that some plasticity occurs at the crack tip and with continued loading, the size of the plastic zone at the crack tip
increases, eventually leading to fracture. Of course while the size of the zone increases, the strain at the crack tip
increases as well. Eventually a critical strain is reached and fracture takes place.
The plastic strain has a significant effect on the energy that may be absorbed at the tip of a sharp crack. It means
plastic strain (localization) plays a key role in the fracture resistance of metals. Strain localization on a macroscopic
scale is delayed in smooth bars by increasing the strain hardening coefficient to promote more uniform
deformation. This implies rather directly that all things being equal, the fracture toughness of a material increases
with increasing strain hardening. Petch and Armstrong described the strain hardening phenomenon for steel and
other reasonably cleavage‐prone materials that, on loading a pre‐cracked material, the plastic zone at the tip of the
crack strain hardens until the cleavage stress is reached and the material fractures.
Most of the strengthening methods of material raise the yield stress without changing the cleavage fracture stress
and so the fracture toughness decreases. But strengthening through strain hardening phenomenon increases
fracture stress also and correspondingly fracture toughness. For example, grain refinement raises the material yield
stress but the fracture stress is raised even more and so the fracture toughness. In case of low strain hardening,
very little hardening occurs ahead of crack and deformation is localized to a small strip. However, an alloy having
significant strain hardening will have a large plastic zone because once the material directly in front of the crack
deforms, it becomes harder than the non‐deformed material and deformation spreads in both the lateral and
forward directions. All factors being equal, the material with high strain hardening exhibits high fracture
toughness.
The work of Hahn and Rosenfield provides a simple example of the factors affecting the fracture toughness:
K IC n cE ys f
where KIC is the plane strain fracture toughness, n the strain hardening coefficient, σys the yield strength, f fracture
strain, E the Young’s modulus, and c is a constant depending on the state of stress.
Strain Hardening in Nano Grained Material
The small grain size of nano materials makes them quite different from the bulk material in a number of ways. For
example, at high and moderate strain rates, the nanomaterials exhibit very low ductility at room temperature.
However, at very low strain rates (less than 10‐4 s‐1), very large strains are usually observed in fcc nanomaterials.
This is due to an extremely high strain rate sensitivity coefficient ‘m’ which is analogous to that observed in high‐
temperature superplasticity of micro‐grained materials. Consequently, the deformation mechanism at room
temperature and low strain rate resembles that for high‐temperature superplasticity. This results in large amount
of deformation. It means high strain hardening assists in large deformation at very low strain rates. Indeed, Coble
creep (enabling both the grain boundary sliding and grain rotation) can operate according to the Ashby–Verall.
Strain Hardening and Other Properties
Material having good strain hardening behaviour also produces changes in other physical properties through cold
working. There is usually a small decrease in density of the order of few tenths of a per cent, decrease in electrical
conductivity due to an increased number of scattering centres, and a small increase in the of thermal expansion.
4 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
Because of the increased internal energy of the cold worked state, chemical reactivity is increased. This leads to a
general decrease in corrosion resistance and in certain alloys introduces the possibility of ‘stress corrosion
cracking’.
Strain Hardening Coefficient and Strength
Dieter et al. and Ebrahimi et al. explained that the strain hardening coefficient has major role in the forming
operation, which controls the amount of uniform plastic strain in the material before strain localization or necking.
A simple power curve relation express the flow curve of many metals in the region of uniform plastic deformation,
that is, from yielding to maximum load σ= Kεn, where n is the strain‐hardening coefficient, and K is the strength
coefficient. Zhang et al. also described that the strain hardening coefficient and strength are the basic mechanical
parameters of metallic material performance. Ebrahimi et al. described that maximum amount of uniform plastic
deformation in tensile straining is given by the strain‐hardening coefficient (n). A standard method to evaluate it is
based on stress–strain data obtained from uniaxial tensile test. Stress–strain curves are usually represented by the
Holloman equation. Therefore, by plotting stress–strain data on logarithmic coordinates, it can be shown that the
slope of the line in the fully plastic region defines the strain‐hardening coefficient (n).
Similarly, the intersection between the stress axis of the stress strain curve and the extrapolated line of the true
stress strain curve gives the value of the strength. The relationship between K value and temperature and strain
rate is similar to the strain hardening coefficient value. It decreases with the temperature and increases as the strain
rate increases. Further, mechanical behaviour of superplastic materials is normally described by the power‐law
relationship between the equivalent stress σ, equivalent strain and equivalent strain‐rate ε as K n m where n
is the hardening index, m the strain‐rate sensitivity index and K a material constant.
Strain, Strain Hardening and Strain Hardening Rate
Luo et al. explained that the strain hardening coefficient n is the result of balance between strain hardening
mechanism depending mainly on strain, and softening mechanism depending on time. The strain hardening
coefficient increases with the increase in strain. The yield strength and the strain hardening increases with the
increasing strain rate at a fixed temperature but decreases with increasing temperature for a given strain rate. Role
of strain rate and temperature is highly significant in determining strain hardening behaviour and formability of
material. Lee et al. have shown that at constant temperature, the flow stress, strain hardening rate and strain rate
sensitivity increases with increase in strain rate.
While increases in the strain hardening rate will increase the strain at instability εu (true tensile strain at plastic
instability), it is not true that increases in σi (Friction stress for yielding in Hall–Petch equation) have the same
effect. In fact, increasing σi without a corresponding increase in the strain hardening rate will lower εu, as is known
from the Considere criteria.
Grain size of material also has important effect on strain hardening. As the grain size decreases, the yield and flow
stresses increase without a corresponding increase in the strain hardening and some decrease in ductility is
observed. The effects even for a given material are not always obvious. For example, Morrison and Miller showed
that for titanium and Fe–Ni alloy, there was a decrease in uniform strain with decreasing grain diameter as
compared for other steel and titanium materials in which there was an increase in the reduction in area with
decreasing grain diameter. These two opposing behaviours can be understood with different degree of H–P
dependence.
Dynamic strain ageing (DSA) is another important phenomenon, which initially produces a rapid increase in the
strain‐hardening rate with straining. However, when the supply of interstitial atoms (for dislocation locking) is no
longer sufficient, the strain hardening decreases. Increased interstitial content delays the onset of instability. In
high nitrogen steels (0.03%), neither decrease in strain hardening rate nor any reduction in ductility was noted.
Strain Hardening in Aerospace Alloys 5
In fcc material, when σi is lower, favouring higher values of the instability strain, εu. There is no major effect of
reducing the temperature or increasing the strain rate on σi. However, both of these factors lead to increased
dislocation storage and increased strain hardening, thus leading to increasing the strain at instability. Composition
changes can also be very important. As an example, the case of 70–30 brass is instructive. In this material, alloying
tends to increase the strength. However, the SFE is also lowered delaying the onset of cross slip which maintains a
relatively high strain hardening rate. Thus compared to Cu, where the strain hardening decreases with increasing
strain, εu is greater.
In hcp materials like Mg, Zn, Cd and Be, the primary slip system, basal, is similar to fcc. However, there are only
two independent slip systems on this plane. Thus other non‐close packed systems such as prism and pyramidal as
well as possible twinning must be activated for generalized polycrystalline deformation. The stress for basal slip is
not strongly temperature dependent. However, the stresses for the secondary systems are strongly temperature
dependent. The effect of these behaviours on the polycrystalline yield stress is that σi, shows little change with
temperature. However, the microstructural stress intensity parameter is related to slip accommodation between
neighboring grains and thus it becomes temperature dependence and yield increases with decreasing temperature
favouring a reduction in εu at lower temperatures. Opposing this trend is the fact that the strain hardening becomes
higher at lower temperatures. The net effect of these two factors is that εu increases at lower temperatures. For slip
in Zr and α‐Ti, the primary slip system is prismatic. In these metals, σi is very much higher than for those metals
whose primary deformation mode is basal slip. The PN stress again dominates and there is a large increase in σi at
low temperatures. The net effect is that εu is low and decreases even more at low temperatures. εu can be
understood in terms of the two competing terms in the Hall–Petch equation and details of the deformation process.
Ti and Ti Alloys
1) Commercially Pure Ti
At room temperature CP‐Ti has moderate formability and limited ductility. Although it has hcp structure only a
few slip systems are active. Therefore twin systems are necessary in order to obtain good amount of
deformation. However, twinning is also limited at {1012} and {1011}. Accordingly deformation is carried out at
elevated temperatures where basal slip systems can be activated. Chun and Wen‐feng conducted a hot
compression experiment on commercially pure Titanium under high temperatures ranging from 700 to 9000C
with 500C interval and at a strain rate 3.6 to 40 per minute. It was noted that dynamic recrystallization occurs at
900°C, dynamic recovery takes place at temperature range from 750 to 850°C, and the strain hardening
phenomenon occurs at lower temperature of 700 °C.
Furhter, Tsao et al. conducted a study on its flow stress behaviour during warm tensile deformation of sheet
specimen. It shows (Table 1) that the strain hardening coefficient of the CP‐Ti sheet increases with the increase
in strain rate and decrease in deformation temperature.
TABLE 1: STRAIN HARDENING COEFFICIENT OF CP‐TI AT VARIOUS TEMPERATURES AND STRAIN RATE
Temperature (0C) n value at different Strain rates (per sec)
5×10‐2 1.6 ×10‐2 2.5 ×10‐3 8.3× 10‐4
350 0.27114 0.23338 0.21536 0.1897
400 0.25602 0.21521 0.19058 0.16218
450 0.22366 0.19813 0.16278 0.13668
500 0.21171 0.15973 0.13912 0.11974
Ayman et al. conducted a simple compression test, plane strain compression test and simple shear test. In
simple shear, the strain hardening rate is much lower than in other two simple tests. In simple compression and
in plane strain compression, the strain hardening behavior is found to be in three stages. First stage consists of a
dynamic recovery regime and the second stage a region with increasing strain hardening rate. The third stage is
6 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
characterized by a falling strain hardening rate. The increase in the strain hardening rate in first stage is due to
deformation twinning and not due to dynamic strain aging. The falling strain hardening rate in third stage can
be expected to be a result of increasing the difficulty of producing deformation twin.
Also, Purceka et al. explained that CP Ti could be uniformly processed through equal channel angular extrusion
(ECAE) upto 8 passes. Significant improvement in the strength levels with adequate ductility is because of the
grain refinement and the strain hardening without alloying.
2) Ti Alloy Ti6Al4V
Luo et al. explained the flow stress–strain curves of Ti–6Al–4V alloy in α β two‐phase region and β single‐
phase region at different strain rates and deformation temperatures. In both cases, isothermal compression
behaviour exhibits a peak flow stress at a very low strain followed by extensive flow softening where softening
rate is higher at low strains and considerably less at higher strains. This phenomenon is related to the fact that
the dominant deformation mechanism is due to slip of dislocations at low strains. Thus, the dislocation density
increases quickly with increasing plastic strain. Then, the softening effect plays an important role in dislocation
density evolution and with further increase of strain, flow stress sharply decreases to a steady value when the
dynamic softening effect is sufficient to counteract the strain hardening effect. The strain hardening coefficient
‘n’ results from a balance between hardening mechanisms depending mainly on strain and softening
mechanisms depending mainly on time. The strain affects significantly the strain hardening coefficient ‘n’ due
to the change in grain size of primary phase, and the competition between thermal softening and strain
hardening.
It is reported that the strain hardening coefficient is not constant but it changes as the strain rate changes. It
decreases as the strain rate decreases due to decrease in the dislocation density.
Lee et al. conducted a study on the impact behaviour of Ti‐6Al‐4V alloy at ‐150 0C, 0 0C and 25 0C and at
different strain rates. The strain hardening coefficient value for the low temperature testing was found to be
larger as given in the Table 2. In this study, it has been also confirmed that the flow stress depends on both the
strain and the strain rate. Specifically, for a constant strain, the flow stress increases rapidly with increasing
strain rate, while for a constant strain rate, the flow stress increases gradually with increasing strain. Also it
explains that the strain hardening coefficient (n) increases with increasing strain rate at a fixed temperature, but
decreases with increasing temperature for a given strain rate. This result suggests that a lower deformation
temperature increases the density and multiplication rate of the dislocations within the Ti–6Al–4V
microstructure, and prompts a corresponding improvement in the resistance of the alloy to plastic flow.
TABLE 2: STRAIN HARDENING COEFFICIENT VALUES AT VARIOUS TEMPERATURES AND STRAIN RATE
fairly sensitive to temperature. It is noted that under the present test conditions, the yield strength and the
strain hardening rate decrease with increasing temperature. This can be explained by the fact that increasing
temperature decreases the density and multiplication rate of dislocations, and consequently, resulting in a loss
of resistance to plastic flow, leading to, the material becomes softer and more ductile. When a material deforms
under high temperature and high strain‐rate conditions, its flow behavior is predominated by a competition
process between the rate of strain hardening and the rate of thermal softening.
Generally, high‐rate deformation can cause the enhancement of a material’s strength due to the high rate of
strain hardening. By contrast, an increase of temperature will lead to a rapid reduction in the rate of strain
hardening. When thermal softening becomes dominant, the plastic flow is controlled completely by the
temperature effect, which results in a rapid drop of flow stress. Therefore, although the plastic deformation
proceeds at a high strain rate, not only is the effect of the strain rate, but also the effect of temperature, found on
strength: however, the latter has more pronounced effect than the former.
Morita showed STQ‐treated (ST+ quenched) material indicating strong strain hardening. When the material at
the STQ condition was further aged, the degree of strain hardening decreased as the aging temperature
increased. This tendency was reflected in the change in the ratio of YS/ TS, namely, this ratio was lowest at the
STQ condition and recovered with increasing aging temperature.
Steels
1) Carbon Steel and HSLA Steel
Strain hardening behaviour of Ti, micro‐alloyed steels and ferritic‐pearlitic steels depends significantly on the
microstructure. There is a very significant relationship between the microstructure of steels and their strain
hardening behaviour. Small precipitates or increasing volume fraction of hard second phase increase the strain
hardening rate. Strain hardening rate increases with decreasing particle size of hard Ti(C, N) precipitates or
increasing volume fraction of pearlite in ferrite.
Influence of pearlite morphology on strain hardening behaviour is rather small. b.c.c, steels have a smaller
strain hardening rate than f.c.c, steels. The microstructure dependence of the strain hardening rate depends on
the stability of the steels. The temperature influence on the strain hardening rate is larger in f.c.c, steels than in
ferritic b.c.c. steels for both small and large deformations.
Steel characterized by high strain hardening at room temperature and having low strain‐hardening capability
behave differently under cold forming. It shows smoother surface finish for the former and barreling/ bulging
for latter indicating clear phenomenon of uniform elongation by strain hardening.
2) Austenitic Stainless Steel (300 Series)
Austenitic stainless steel is thermodynamically stable over a wide range of temperatures and it can be hardened
only by cold working. T.S. Byun et al. conducted a study on the plastic behaviour and the strain hardening
dependence on temperature. The tensile test was done on annealed 304, 316, 316LN and 20 % cold worked
316LN. The steels considered for the study showed strong temperature dependence on the strength and
ductility. The strength increases as the test temperature decreases. The cold worked specimen tested at room
temperature showed high uniform elongation of 40% to 100% and relatively small necking elongations. In the
annealed and cold worked conditions the strength of stainless steel is found to be decreasing with increasing
temperature while the ductility shows peak at room temperature.
Above room temperature the true strain hardening rate decreases monotonically with strain in the uniform
deformation region. At room temperature or below steel shows a two stage hardening behaviour. It is reported
to decrease with strain below 3‐20% and then it experiences an additional increase‐decrease cycle before plastic
instability. This characteristics of two stage hardening is due to stress induced transformation of austenite to
martensite.
Park et al. conducted a study on the strain rate effects on the mechanical behaviour of AISI 300 series of
8 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
austenitic steels. In this study it has been explained that at low temperatures (below room temperature) the
stainless steel shows two stage sigmoidal deformation which is non‐ linear hardening behaviour that is
dependent on temperature. It is found to have strong dependency between the two stage strain hardening and
the temperature.
3) Maraging Steel
Maraging steels in general have a low strain‐hardening ability. The engineering stress‐strain curve of the
maraging steel has a peak stress at relatively small plastic strains followed by continued decrease of stress,
leading to a low tensile to yield‐strength ratio. However, the observed total elongation is found to be moderate
(~10%). A low tensile to yield strength ratio indicates that the tolerance to plastic overload is small as the strain
hardening after yield is minimal. Such a characteristic may become a problem in a safe design against sudden
overload. However, there is a possibility by which strain hardening of maraging steel (in solutionised
condition) can be increased i.e. through transformation induced plasticity (TRIP), which results in higher tensile
to yield strength ratio. This phenomenon occurs when it has presence of retained austenite. But, in aged
condition resultant improvement in strength due to strain hardening is minimised since primary mechanism of
strengthening is through precipitation of intermetallic in martensitic matrix while such precipitation is
restricted in presence of austenite. Although, retained austenite transforms to martensite during straining, it is
not as strong as maraged martensite. In such condition no significant strain hardening is observed in maraging
steel.
Ultrahigh strength steels, maraging M250 and D6AC have been tested at room temperature to determine both
monotonic and cyclic stress‐strain curves. For the tempered or aged condition, steels had strain hardening
coefficients less than 0.10 and showed cyclic softening.
Aluminium Alloys
Strain hardening behaviour of aerospace Al alloys, like AA6061, AA2024, AA7010, AA2014, AA7075 and
commercial purity aluminium has been reported. Tensile test samples are tested at a strain rate of 10–3 s–1 at room
temperature. The strain hardening coefficient values are given in Table 3. It is clear from Table 3 that Al alloys have
very low strain hardening among aerospace alloys. However, formability of alloy is found to be good where SFE
plays important role in deformation of alloy along with fcc crystal structure.
TABLE 3: STRAIN HARDENING COEFFICIENT OF VARIOUS AL ALLOYS
High Temperature Materials
Among the materials used for high temperature applications, Ni based superalloy Inconel 718, Co based superalloy
Hayens‐25 and refractory alloy C‐103 are the important materials for space application. Zhang et al. studied and
explained that strain hardening and flow softening occurs when Inconel 718 is superplastically deformed. It is well
known that grain growth plays an important role in the hardening of superplastic materials where, dynamic
recovery, cavitation formation, grain refinement and dynamic precipitation contribute to flow softening. This is a
tradeoff between hardening and softening. Strain hardening and superplastic phenomenon are closely related and
strain hardening helps to get high elongation under slow strain rate tension.
Lee fitted stress–strain relations using the simple model proposed by Ludwik. The results show that the yield
strength and strain hardening coefficient increase with increasing strain rate at a constant temperature (n varies
Strain Hardening in Aerospace Alloys 9
between 0.28‐0.44), but decrease with increasing temperature at a constant strain rate. This suggests that a lower
deformation temperature increases the density and multiplication rate of the dislocations within the Inconel 718
microstructure, and prompts a corresponding increase in the plastic flow resistance.
Due to low stacking fault energy of Co, resulting in interaction between stacking faults, high strain hardening rates
are reported (strain hardening coefficient of L‐605: 0.83). It results in high degree of resistant to necking. Cross slip
is difficult in these alloys. However, C‐103 alloy has low rate of strain hardening. It has 1% Ti, which improves the
ductility of alloy. Although rate of strain hardening is low, Nb being bcc structure, having large number of slip
system, some of which get active in the alloy and results in good ductility.
Discussion
The aerospace sector has traditionally been a promoter for the development and application of advanced
engineering materials, where component performance is primarily determined by mechanical properties such as
strength, stiffness, and damage tolerance as well as by physical and chemical properties such as density, corrosion
resistance at ambient and high temperatures. However, behaviour of alloys is primarily governed by their crystal
structure and deformation mechanisms. Considering these facts, structural materials for aerospace systems are
selected on the basis of general requirements of high specific strength alongwith functional requirements of specific
systems and processing feasibility. A list of important structural alloys with their properties (Table 4) and
characteristic stress strain curve (Fig. 2) are presented here.
Though strengthening is primarily contributed by precipitation hardening in several aerospace alloys (especially
Al alloys), role of strain hardening has significant importance in most of the alloy systems. Even some of these
alloys are not designed with the benefit of strain hardening, it has contribution especially in deciding the amount
of elongation through uniform plastic deformation prior to necking. It delays the necking by shifting local
deformation towards un‐necked region (since starting of necked region becomes strain hardened due to dislocation
pile‐up). In this way deformation proceeds uniformly throughout the length until the stress crosses the threshold
level required to break the dislocation pile‐up and then necking (local deformation) starts and develops till break.
It means that higher the strain hardening, higher shall be the uniform elongation and correspondingly resulting in
higher elongation (total). This aspect has important application for all the materials listed in Table 4. Extent of
strain hardening is also indicated (indirectly) through the ratio/ difference in ultimate tensile strength and yield
strength. Higher the ratio between these, higher shall be the elongation. In such materials role of strain hardening
is higher during deformation. This can be seen from Table 4 that elongation of stainless steels and Haynes‐25 is
higher. It is interesting to note that, these alloys are single phase (non‐heat treatable) alloys and strengthened by
solid solution hardening and has important scope of strengthening through strain hardening.
Strain hardening behaviour is very useful in fabrication as well as in application. For example, fabrication methods
are designed and decided on the basis of material properties and the extent to allowable deformation with or
without intermittent heat treatment (annealing). Stainless steels, Nb alloy and Haynes‐25 properties (yield strength
and % El) clearly indicate that these alloys have relatively higher formability as compared to other alloys.
Examining further it emerges that it has relation to their close packed crystal structure and available deformation
mode. Whereas stainless steel being fcc crystal structure (and single phase austenitic structure) can take large
amount of deformation, Nb alloy and Haynes‐25 also can take large deformation due to single phase structure even
though their crystal structure (bcc and hcp respectively) is not as beneficial like stainless steel (with respect to close
packing). In C‐103 alloy some more slip planes become active and result in good formability. In Haynes‐ 25 high
strain hardening helps in getting large deformations. Metals and alloys with wider stacking faults (low SFE) like
stainless steels strain harden more rapidly and twin easily on annealing. But, in case of Al alloys although it has fcc
crystal structure (close packed), it shows relatively lower strain hardening behaviour. Primarily it is due to
multiple phase alloys where strengthening is mainly by precipitation hardening and role of strain hardening is
limited to the extent of allowable forming during tube stretching/ cold rolling/ roll bending. However, here another
phenomenon of deformation i.e. cross slip helps in increasing the formability of the alloy in cold deformation and
10 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
cryogenic ductility. This is due to very high stacking fault energy. Its benefits are used in plate/ sheet forming/ roll
bending to realise large capacity propellant tanks. It means large deformation is possible for a material to either
have high strain hardening (n) or have large number of active slip system/ close packed planes or promoted by
cross slip.
Observing the properties of high strength steels (maraging steel and Afnor 15CDV6 steel), it can be seen that ratio/
difference in ultimate tensile strength and yield strength is very low and correspondingly it results in lower
elongation. Here also strengthening mechanism is through precipitation and fine distribution of second phase.
Shifting of deformation due to dislocation pile up/ strain hardening (as in the case of stainless steel) does not take
place to large extent. Also crystal structure of these alloys are basically martensitic (bcc/ bct) based where limited
slip systems (unlike austenitic stainless steels) are available and cross slip/ twin assisted slip is difficult (unlike Al
alloys) due to relatively low stacking fault energy. These alloys are altogether designed for special purpose high
strength materials and have other specific benefits, which are used in fabrication and heat treatment. Also design
criterion preferred for such alloys are mostly fracture based. Since the strain hardening is low and the UTS to yield
ratio/ difference is very low, it is not good for structural design to consider yield stress as limiting criterion.
Accordingly structural designer considers fracture toughness as limiting criterion for such materials.
Role of strain hardening in Ti alloys is significant. Though the important alloy Ti6Al4V is a two phase alloy and its
cold formability is poor due to spring back property, it has important role during application. The alloy is selected
for its specific advantages and primarily used in pressure vessel area where localised deformation is not permitted
and cyclic tests are conducted to qualify the component for application at various service conditions. Since it has
relatively moderate extent of strain hardening, it is carefully pressurised at a specific safe rate to avoid localised
yielding. For example, in a specific instance, pressure vessel was pressurised to a particular level and then
depressurised, and in the subsequent pressurization it had shown lower strain as compared to previous
pressurization. This is a clear effect of strain hardening. In this alloy, although twinning plays important role in
deformation but assistance to slip is limited and hence large deformation in cold condition is limited without
intermediate annealing.
TABLE 4: IMPORTANT AEROSPACE ALLOYS AND THEIR TYPICAL MECHANICAL PROPERTIES
Mechanical properties
Alloys Condition
E (GPa) UTS (MPa) YS (MPa) %El Impact (J)
High strain hardening behaviour alloys
Stainless steel 304/ 316 Annealed 200 500 210 50 100
Stainless steel 321 Annealed 200 550 210 50 100
Annealed + subzero
Stainless steel 301 200 1300 1050 23 60
tested
Stainless steel 202 Annealed 200 700 400 50 200
Co based alloy Haynes‐25 Annealed 244 1030 485 50 ‐
Moderate strain hardening behaviour alloys
α‐Ti alloy ELI
Annealed 110 724 665 10 30
Ti5Al2.5Sn
Ti6Al4V Annealed 110 895 825 12 25
Ti6Al4V Solution treated +aged 114 1100 1030 10 23
Nb alloy, C‐103 Annealed 87 405 310 26 156
Low strain hardening behaviour alloys
Al alloy AFNOR 7020 Precipitation hardened 71 370 310 12 66
Al alloy AA2014 Precipitation hardened 72.4 470 414 12 60
Al alloy AA2219 Precipitation hardened 73.8 450 350 10 60
M250 Maraging steel Solution annealing +aging 190 1800 1750 12 95*
AFNOR 15CDV6 steel Hardened + tempered 210 980 835 13 60
*Fracture Toughness in (MPa√m)
It is clear from the above discussion that strain hardening phenomenon has very important role in various stages of
Strain Hardening in Aerospace Alloys 11
processing/ application of alloy especially for the systems used in aerospace where performance is demanded
through alloy selection/ design and reliability is ensured through characterization/ testing.
Considering the above, aerospace structural alloys can be classified under three groups on the basis of strain
hardening behaviour. Austenitic stainless steel and Hayens‐25 superalloy show highest response to strain
hardening, which may be due to single phase fcc structure whereas materials like single phase bcc C‐103 alloy and
two phase Ti alloys show moderate strain hardening behaviour. Al alloys and high strength steels have lower
strain hardening behaviour. Material processing methods are selected accordingly to obtain optimum properties
and useful products.
FIG. 2: REPRESENTATIVE STRESS STRAIN CURVE OF VARIOUS AEROSPACE ALLOYS
Summary
Role of strain hardening in aerospace alloys has been reviewed. The strain hardening behaviour of Ti‐6Al‐4V, CP‐
Ti, Al alloys, austenitic steels, maraging steels, and high temperature alloys are considered and following points
are noted.
1. CP‐Ti and Ti‐6Al‐4V alloy are found to have moderate strain hardening behaviour. The strain hardening
coefficient (n) increase with increasing strain rate at a fixed temperature, but decrease with increasing
temperature for a given strain rate.
2. Austenitic steels have been identified as materials with high strain hardening. Above room temperature, the
true strain hardening rate decreases monotonically with strain in the uniform deformation region. In
maraging steel, strain hardening phenomenon is not significant, leading to smaller ratio/ difference in YS to
UTS.
3. For Al alloys, strain hardening phenomenon is not significant and coefficient is found to be low.
4. Cobalt based superalloy Haynes‐25 has very good response to strain hardening and it provides significant
improvement in ductility of the alloy. Nb alloy C‐103 has low strain hardening rate but its formability is
good due to other mechanisms. Inconel 718 also shows moderate strain hardening behaviour with
decreasing trend with increase in temperature.
ACKNOWLEDGEMENT
Authors are thankful to DD, MME, VSSC for necessary guidance and Director VSSC for permission to publish this
work.
REFERENCES
[1] Antolovich S. D., Armstrong R. W., Plastic strain localization in metals: origins and consequences, Progress in Materials
Science 59 (2014) 1–160
12 R. K. GUPTA, CHRISTY MATHEW, P. RAMKUMAR
[2] Ashby MF, Verrall RA, Acta Metall 21 (1973) 149–63.
[3] ATI Wah Chang Technical data sheet
[4] Ayman A, Salem 1, Surya R. Kalidindi, Roger D. Doherty, Acta Mater 51 (2003) 4225–4237.
[5] Baird JD, Jamieson A, J Iron Steel Inst 204 (1966) 793–803.
[6] Byun T. S, Hashimoto N., Farrell K., Acta Mater 52 (2004) 3889‐3899.
[7] Champion Y, Nowak S, Guerin S, Duhamel C., Mater Sci Forum, 567–568 (2007) 29–32.
[8] Chen FK, Chiu KH, J Mater Process Technol, 170 (2005)181–186.
[9] Cheng L, Poole W, Embury J, Lloyd D. Metall Mater Trans A, 34A (2003) 2473–81.
[10] Chun XU, Wen‐feng ZHU, Trans. nonferrous Met. Soc. China 20 (2010) 2162 ‐2167.
[11] Considere A. Memoire sur l’emploi du fer et de l’acier dans les constructions (On the use of iron and steel in construction)
Ann Ponts Chauss 9 (1885) 574–775.
[12] Cottrell AH, Bilby BA, Proc Phys Soc Lond A62 (1949) 49–62.
[13] Dieter G.E., Mechanical Metallurgy, McGraw‐Hill company (1988) 231‐232, ISBN 0‐07‐100406‐8
[14] Dieter George E., Kuhn Howard A. S. Semiatin Lee, ASM International. Handbook of Workability and Process Design.
(2003) pp 25.
[15] Ebrahimi R., Pardis N., Determination of strain‐hardening coefficient using double compression test, Materials Science and
Engineering 518 A (2009) 56–60.
[16] Eom J.G., Son Y.H., Jeong S.W., Ahn S.T., Jang S.M., Yoon D.J., Joun M.S., Materials and Design,
http://dx.doi.org/10.1016/j.matdes.2013.08.101
[17] Fang X.F., Dahl W., Materials Science and Engineering A203 (1995) 14‐25
[18] Fang X.F., Dahl W., Materials Science and Engineering A203 (1995) 36‐45
[19] Fang X.F., Gusek C.O., Dahl W., Materials Science and Engineering A203 (1995) 26‐35.
[20] Fribourg G., Bre´chet Y., Deschamps A., Simar A., Acta Mater 59 (2011) 3621–3635
[21] Gil Sevillano J. Flow stress and work hardening. In: Cahn RW, Haasen P, Kramer EJ, editors. Material science and
technology: a comprehensive treatment, vol. 6 (1993) p. 19–88.
[22] Giuliano, Materials and Design 29 (2008) 1330–1333.
[23] Griffith AA, Philos Trans Roy Soc Lond A 221 (1921) 163–98.
[24] Hahn GT, Rosenfield AR. Sources of fracture toughness: the relation between KIc and the ordinary tensile properties of
metals. In: Applications related phenomena in titanium alloys. ASTM STP, vol. 432. Philadelphia (PA, USA): ASTM
International (1968) p. 5–32.
[25] Hanlon T, Suresh S., Fatigue of nanostructured metals and alloys. In: Carpinteri A, editor. International conference on
fracture (ICF 11), Politechnico di Torino, Italy (2005).
[26] Herchenroeder R.B. and Ebihara W.T., Metals Engineering Quarterly, May (1969) 313‐323
[27] Jiao Luo, Li Miaoquan, Yu Weixin, Li Hong, Materials and Design 31 (2010) 741–748.
[28] Jonas J.J., Sellars C.M., McG Tegart W.J., Metall. Rev. 130 (1969) 1–24.
[29] Kashyap K. T., Ramachandra C, Dutta C and Chatterji B, Bull. Mater. Sci., 23, 1 (2000) 47–49.
[30] Kocks U, Mecking H. Prog Mater Sci, 48 (2003) 171–273.
[31] Kujawski D, Ellyin F. On the size of plastic zone ahead of crack tip. Eng Fract Mech, 25 (1986) 229–36.
[32] Lee Woei‐Shyan, Chen Tao‐Hsing, Huang Sian‐Cing, Journal of Nuclear Materials 402 (2010) 1–7.
[33] Lee Woei‐Shyan, Chen Tao‐Hsing, Huang Sian‐Cing, Journal of Nuclear Materials 402 (2010) 1–7.
[34] Lee Woei‐Shyan, Chi‐Feng Lin, Journal of Materials Processing Technology 75 (1998) 127–136
[35] Lee Woei‐Shyan, Lin Chi‐Feng, Chen Tao‐Hsing, Hong‐Wei Chen, Materials Science and Engineering A 528 (2011) 6279–
6286
Strain Hardening in Aerospace Alloys 13
[36] Mara NA, Bhattacharyya D, Hirth JP, Dickenson T, Misra A. Mechanism of shear banding in nanolayered composites.
Appl Phys Letts (2010) 97. http://dx.doi.org/10.1063/1.3458000.
[37] Meyers MA, Mishra A, Benson DJ., Prog Mater Sci; 51 (2006) 427–556.
[38] Mithun Kuruvila, Srivatsan T. S., Petraroli M., Park L., Sadhana, 33, 3, (2008) 235–250.
[39] Morrison W.B., Miller R.L. The ductility of ultrafine‐grain alloys. In: Burke JJ, Weiss V, editors. Ultrafine grain metals. NY:
Syracuse University Press (1970) p. 183–211.
[40] Morrison W.B., Trans Am Soc Metals 59 (1966) 824–829.
[41] Mukhopadhyay C., Rajkumar K., Jayakumar T., Raj B., J. Mater. Sci. 45 (2010) 1371–1384.
[42] Nieh T.G., Wadsworth J., Mater. Sci. Eng. A 239–240 (1997) 88–96.
[43] Nieh T.G., Wadsworth J., Sherby O.D., Superplasticity in Metals and Ceramics, Cambridge University Press (1997) p. 137.
[44] Orowan E., in ʺFatigue and Fracture of Metals”, Symposium at Massachusetts Institute of Technology, John Wiley & Sons,
Inc New York (1952).
[45] Park D, Niewczas M. Mater Sci Eng A, 491 (2008) 88–102.
[46] Park W. S, Seong Won Yoo, Myung Hyun Kim, Jae Myung Lee, Materials and design 31 (2010) 3630‐3640.
[47] Petch NJ, Armstrong RW. Work hardening in cleavage fracture toughness. Acta Metall 37 (1989) 2279–85.
[48] Purceka G., Yapicib G.G., Karaman I., Maier H.J., Materials Science and Engineering A 528 (2011) 2303–2308.
[49] Seetharaman V., Krishnan R., J. Mater. Sci. 16 (1981) 523–530.
[50] Simar A, Bre´chet Y, de Meester B, Denquin A, Pardoen T., Acta Mater 55 (2007) 6133–43.
[51] Tatsuro Morita, Kei Hatsuoka, Takashi Iizuka and Kazuhiro Kawasaki, Materials Transactions, 46, 7 (2005) 1681 ‐ 1686
[52] Titanium a technical guide, Edited by Matthew J. Donachie, Jr. ASM internationals (1988) 32‐35. ISBN: 0‐87170‐309‐2.
[53] Titanium and Titanium Alloys. Fundamentals and Applications, Edited by Christoph Leyens, Manfred Peters, WILEY‐
VCH Verlag GmbH & Co. KGaA, Weinheim (2003) pp 333‐349
[54] Tsao L.C., Wub H.Y., Leong J.C., Fang C.J., Materials and Design 34 (2012) 179–184.
[55] Wang Songtao, Yanga, Ke, Shan Yiyin, Laifeng Li, Materials Science and Engineering A 490 (2008) 95–104
[56] Zener C, Holloman JH, J Appl Phys 15 (1944) 22–32.
[57] Zhang B., Mynorsa D.J., Mugarra A., Ostolaza K., Journal of Materials Processing Technology 153–154 (2004) 694–698
[58] Zhang Zhongping, Qiang Sun, Chunwang Li, and Wenzhen Zhao, Theoretical Calculation of the Strain‐Hardening
Coefficient and the Strength of Metallic Materials, ASM International JMEPEG (2006) 15:19‐22.