Strengthening
Mechanisms
THE STRENGTHENING OF IRON AND ITS ALLOYS
• Although pure iron can be weak, steels cover a wide range of
the strength spectrum from low yield stress levels (around 200
MN m-2) to very high levels (approaching 5500 MN m-2),
without compromising toughness.
• There are many ways of strengthening steels, which is why
they are able to offer such a wide range of properties.
• It is also possible to combine several strengthening
mechanisms, and in such circumstances it is often difficult to
quantify the variety of contributions to the overall strength.
• On the other hand, there has been considerable progress in
methods for mathematically modelling of properties, and
hence of deconvoluting the overall strength into its
components.
• The basic ways in which iron can be strengthened are
discussed first, by reference to simple systems. These results
should then be helpful in examining the behaviour of more
• Like other metals, iron can be strengthened by several
mechanisms, the most important of which are:
1. Work hardening.
2. Solid solution strengthening by interstitial atoms.
3. Solid solution strengthening by substitutional atoms.
4. Refinement of grain size.
5. Dispersion strengthening, including lamellar and random dispersed
structures.
The most distinctive aspect of strengthening of iron is the role of the
interstitial solutes carbon and nitrogen. These elements also play a
vital part in interacting with dislocations, and in combining
preferentially with some of the metallic alloying elements used in
steels.
WORK HARDENING
• Work hardening is an important strengthening process in steel, particularly in
obtaining high strength levels in rod and wire, both in plain carbon and alloy
steels. For example, the tensile strength of a 0.05 wt% C steel subjected to 95%
reduction in area by wire drawing, is raised by no less than 550 MN m-2, while
higher carbon steels are strengthened by up to twice this amount. Indeed, without
the addition of special alloying elements, plain carbon steels can be raised to
strength levels above 1500 MN m-2 simply by the phenomenon of work
hardening.
• The slip plane in α-iron is not unique. Slip occurs on several planes, {110}, {112}
and {123}, but always in the close packed 〈 111 〉 direction which is common to
each of these planes. The diversity of slip planes leads to rather irregular wavy
slip bands in deformed crystals, as the dislocations can readily move from one
type of plane to another by cross slip, provided they share a common slip
direction.
• Since plastic flow occurs by a dislocation mechanism the fact that work-hardening
occurs means that it becomes difficult for dislocations to move as the strain
increases. All theories of work-hardening depend on this assumption, and the basic
idea of hardening, put forward by Taylor in 1934, is that some dislocations become
‘stuck’ inside the crystal and act as sources of internal stress which oppose the
motion of other gliding dislocations.
• The dislocation density in a metal increases with deformation or cold work, due to
dislocation multiplication or the formation of new dislocations, as noted previously.
Consequently, the average distance of separation between dislocations decreases—
the dislocations are positioned closer together. On the average, dislocation–
dislocation strain interactions are repulsive. The net result is that the motion of a
dislocation is hindered by the presence of other dislocations. As the dislocation
density increases, this resistance to dislocation motion by other dislocations
becomes more pronounced. Thus, the imposed stress necessary to deform a metal
increases with increasing cold work.
• All the properties of a metal that are dependent on the lattice, structure are
affected by plastic deformation or cold working. Tensile, strength, yield
strength, and hardness are increased, while ductility, represented by percent
elongation, is decreased.
• Although both; strength and hardness increase, the rate of change is not the
same. Hardness generally increases most rapidly in the first 10 percent
reduction, whereas the tensile strength increases more or less linearly. The yield
strength increases more rapidly than the tensile strength, so that, as the amount
of plastic deformation is increased, the gap between the yield and tensile
strengths decreases (Fig. 3.18). This is important in certain forming operations
where appreciable deformation is required. In drawing, for example, the load
must be above the yield point to obtain appreciable deformation but below the
tensile strength to avoid failure. If the gap is narrow, very close control of the
load is required.
Ductility follows a path opposite to that of
hardness, a large decrease in the first 10
percent reduction and then a decrease at a
slower rate.
The influence of cold work
on the stress–strain
behavior for a low-carbon
Such changes in mechanical properties are, of course, of interest theoretically, but they
are also of great importance in industrial practice. This is because the rate at which the
material hardens during deformation influences both the power required and the
method of working in the various shaping operations, while the magnitude of the
hardness introduced governs the frequency with which the component must be
annealed (always an expensive operation) to enable further working to be continued.
SOLID SOLUTION STRENGTHENING BY
INTERSTITIALS
Carbon and nitrogen have a
disproportionate influence on the
strength of ferritic iron and a relatively
minor effect on that of austenitic iron.
Solid solution strengthening occurs
when the strain fields around misfitting
solutes interfere with the motion of
dislocations.
Atoms which substitute for iron cause
local expansions or contractions; these
strains are isotropic and therefore can
only interact with the hydrostatic
components of the strain fields of
dislocations.
In contrast, an interstitial atom located in the irregular octahedron interstice in
ferrite causes a tetragonal distortion (Fig. 2.3) which has a powerful interaction
with the shear which is the dominant component of a dislocation strain field. This
is why interstitial solid solution strengthening is so potent in ferrite.
The corresponding interstitial site
in austenite is the regular
octahedron. An interstitial atom
in austenite therefore behaves
like a substitutional solute, with
only hydrostatic strains
(volumetric strain or dilatation)
surrounding it. This is why
carbon is much less effective in
strengthening austenite.
• Ferritic iron has such a small solubility (0.02% at 723 °C and
<0.00005 at 20 °C) for carbon that there is an overwhelming
tendency for the carbon to segregate to defects. This leads to
another significant effect in α-iron, that carbon and nitrogen can
promote heterogeneous deformation by making it difficult to
initiate plastic flow. This is the yield point effect, described next.
The yield point
Carbon and nitrogen, even in
concentrations as low as 0.005 wt%, in
iron lead to a sharp transition between
elastic and plastic deformation in a
tensile test performed on ferritic iron
(Fig. 2.4a). Decarburization of the iron
results in the elimination of this sharp
transition or yield point, which implies
that the solute atoms are in some way
responsible for this striking behaviour.
Frequently the load drops dramatically
at the upper yield point (A) to another
value referred to as the lower yield
point (B). Under some experimental Fig. 2.4 (a) Schematic diagram of yield
conditions, yield drops of about 30% of phenomena as shown in a tensile test,
the upper yield stress can be obtained. (b) Luders bands in deformed steel
specimens.
• Following the lower yield point, there is frequently a
horizontal section of the stress–strain curve (BC) during which
the plastic deformation propagates at a front which can move
uniformly along the specimen. This front is referred to as a
Luders band (Fig. 2.4b), and the horizontal portion, BC, of the
stress–strain curve as the Luders extension.
• The development of Luders bands can be much less uniform
and, e.g., in pressings where the stress is far from uniaxial,
complex arrays of bands can be observed. These are often
referred to as stretcher strains, but they are still basically
Luders bands.
When the whole specimen has yielded, general work hardening commences and
the stress–strain curve begins to rise in the normal way. If, however, this
deformation is interrupted, and the specimen allowed to rest either at room
temperature or for a shorter time at 100–150◦C, on reloading a new yield point is
observed (D). This return of the yield point is referred to as strain ageing.
The role of interstitial elements in yield
phenomena
• The sharp upper and lower yield point in iron is eliminated by annealing in wet
hydrogen, which reduces the carbon and nitrogen to very low levels. However,
substantial strain ageing can occur at carbon levels around 0.002 wt%, and as little
as 0.001–0.002 wt% N can result in severe strain ageing. Nitrogen is more
effective in this respect than carbon, because its residual solubility near room
temperature is substantially greater than that of carbon (<0.0001 at 20 C).
• Cottrell and Bilby first showed that interstitial atoms such as carbon and nitrogen
would interact strongly with the strain fields of dislocations. The interstitial atoms
have strain fields around them, but when such atoms move within the dislocation
strain fields, there should be an overall reduction in the total strain energy.
• This leads to the formation of interstitial concentrations or atmospheres in the
vicinity of dislocations, which in an extreme case can amount to lines of
interstitial atoms along the cores of the dislocations (condensed atmospheres), e.g.
in edge dislocations at the region of the strain field where there is maximum
dilation (Figs 2.5a, b).
Fig. 2.5
Interstitial atoms
in the vicinity of
an edge
dislocation (a)
random
atmosphere
(b) condensed.
• The binding energy between a dislocation in iron and a carbon atom is about 0.5
eV. Consequently, dislocations can be locked in position by strings of carbon
atoms along the dislocations, thus substantially raising the stress which would be
necessary to cause dislocation movement. A particular attraction of this theory is
that only a very small concentration of interstitial atoms is needed to produce
locking along the whole length of all dislocation lines in annealed iron. For a
typical dislocation density of 108 lines cm-2 in annealed iron, a carbon
concentration of 10-6 wt% would be sufficient to provide one interstitial carbon
atom per atomic plane along all the dislocation lines present, i.e. to saturate the
dislocations. Consequently, this theory can explain the observation of yield
phenomena at very low carbon and nitrogen concentrations.
• The formation of interstitial atmospheres at dislocations requires diffusion of the
solute. As both carbon and nitrogen diffuse very much more rapidly in iron than
substitutional solutes, it is not surprising that strain ageing can take place readily
in the range 20–150◦C.
• The zone of yielding is well defined at the lower testing temperatures, becoming
less regular as the temperature is raised, until it is replaced by fine serrations along
the whole stress–strain curve. This phenomenon is referred to as dynamic strain
ageing, in which the serrations represent the replacement of the primary yield
point by numerous localized yield points within the specimen. These arise because
the temperature is high enough to allow interstitial atoms to diffuse during
deformation, and to form atmospheres around dislocations generated throughout
the stress–strain curve. Steels tested under these conditions also show low
ductilities, due partly to the high dislocation density and partly to the nucleation of
carbide particles on the dislocations where the carbon concentration is high. The
phenomenon is often referred to as blue brittleness, blue being the interference
colour of the steel surface when oxidized in this temperature range.
• An alternative theory was also developed which assumed that, once condensed carbon atmospheres
are formed in iron, the dislocations remain locked, and the yield phenomena arise from the
generation and movement of newly formed dislocations (Gilman and Johnston). Consequently, if at
the upper yield stress the density of mobile dislocations is low, e.g. as a result of solute atom
locking, a large drop in yield stress will occur if a large number of new dislocations is generated.
Observations indicate that the dislocation density just after the lower yield stress is much higher
than that observed at the upper yield point.
• To summarize, the occurrence of a sharp yield point depends on the occurrence of a sudden
increase in the number of mobile dislocations. However, the precise mechanism by which this
takes place will depend on the effectiveness of the locking of the pre-existing dislocations. If the
pinning is weak, then the yield point can arise as a result of unpinning. However, if the dislocations
are strongly locked, either by interstitial atmospheres or precipitates, the yield point will result
from the rapid generation of new dislocations. Under conditions of dynamic strain ageing, where
atmospheres of carbon atoms form continuously on newly generated dislocations, it would be
expected that a higher density of dislocations would be needed to complete the deformation, if it is
assumed that most dislocations which attract carbon atmospheres are permanently locked in
position. Electron microscopic observations have shown that in steels deformed at 200◦C, the
dislocation densities are an order of magnitude greater than those in specimens similarly deformed
at room temperature. This also accounts for the fact that increased work hardening rates are
obtained in the blue-brittle condition as compared to tests at room temperature.
Strengthening at high interstitial concentrations
• Austenite can take into solid solution up to 10 at% carbon which can be retained in
solid solution by rapid quenching. However, in these circumstances the phase
transformation takes place, not to ferrite but to a body-centred tetragonal structure
referred to as martensite. This phase forms as a result of a diffusionless shear
transformation leading to characteristic laths or plates, which normally appear
acicular in polished and etched sections. If the quench is sufficiently rapid, the
martensite is essentially a supersaturated solid solution of carbon in a tetragonal
iron matrix, and as the carbon concentration can be greatly in excess of the
equilibrium concentration in ferrite, the strength is raised very substantially. High
carbon martensites are normally very hard but brittle, the yield strength reaching
as much as 1500 MN m-2; much of this increase can be directly attributed to
increased interstitial solid solution hardening, but there is also a contribution from
the high dislocation density which is characteristic of martensitic transformations
in iron–carbon alloys. By subjecting martensite to a further heat treatment at
intermediate temperatures (tempering), a proportion of the strength is retained,
with a substantial gain in the toughness and ductility of the steel.
SUBSTITUTIONAL SOLID SOLUTION STRENGTHENING OF IRON
Many metallic elements form solid solutions
in γ- and α-iron. These are invariably
substitutional solid solutions, but for a
constant atomic concentration of alloying
elements there are large variations in strength.
Using single crystal data for several metals,
Fig. 2.7 shows that an element such as
vanadium has a weak strengthening effect on
α-iron at low concentrations (<2 at%), while
silicon and molybdenum are much more
effective strengtheners. Other data indicate
that phosphorus, manganese, nickel and
copper are also effective strengtheners.
However, it should be noted that the relative
strengthening may alter with the temperature Fig. 2.7 Solid solution strengthening
of testing, and with the concentrations of of α-iron crystals by substitutional
interstitial solutes present in the steels. solutes. Ratio of the critical resolved
shear stress τ0 to shear modulus µ as
a function of atomic concentration
• The strengthening achieved by substitutional solute atoms is, in general, greater
the larger the difference in atomic size of the solute from that of iron, applying the
Hume-Rothery size effect. However, from the work of Fleischer and Takeuchi it is
apparent that differences in the elastic behaviour of solute and solvent atoms are
also important in determining the overall strengthening achieved. In practical
terms, the contribution to strength from solid solution effects is superimposed on
hardening from other sources, e.g. grain size and dispersions. Also it is a
strengthening increment, like that due to grain size, which need not adversely
affect ductility. In industrial steels, solid solution strengthening is a far from
negligible factor in the overall strength, where it is achieved by a number of
familiar alloying elements, e.g. manganese, silicon, nickel, molybdenum, several
of which are frequently present in a particular steel and are additive in their effect.
These alloying elements are usually added for other reasons, e.g. Si to achieve
deoxidation, Mn to combine with sulphur or Mo to promote hardenability.
Therefore, the solid solution hardening contribution can be viewed as a useful
bonus.
GRAIN SIZE Hall–Petch effect
• The refinement of the grain size of ferrite
provides one of the most important
strengthening routes in the heat treatment of
steels. The first scientific analysis of the
relationship between grain size and strength,
carried out on ARMCO iron by Hall and Petch,
led to the Hall–Petch relationship between the
yield stress σy and the grain diameter d,
• where σ0 and ky are constants. This type of
relationship holds for a wide variety of irons and
steels as well as for many non-ferrous metals
and alloys. A typical set of results for mild steel
is given in Fig. 2.7, where the linear relationship Fig. 2.8 Dependence of the
lower yield stress of mild steel
between σy and d-½ is clearly shown for the three on grain size
test temperatures.
• The constant σ0 is called the friction stress. It is the intercept on the stress axis,
representing the stress required to move free dislocations along the slip planes in
the bcc crystals, and can be regarded as the yield stress of a single crystal (d-½ =
0). This stress is particularly sensitive to temperature (Fig. 2.7) and composition.
The ky term represents the slope of the σy - d-½ plot which has been found not to be
sensitive to temperature (Fig. 2.8), composition and strain rate.
• In line with the Cottrell–Bilby theory of the yield point involving the break away
of dislocations from interstitial carbon atmospheres, ky has been referred to as the
unpinning parameter. However, the insensitivity of ky to temperature suggests that
unpinning rarely occurs, and emphasizes the theory that new dislocations are
generated at the yield point. This is consistent with the theories explaining the
yield point in terms of the movement of new dislocations, the velocities of which
are stress dependent (Section 2.3.2).
• The grain size effect on the yield stress can therefore be explained by assuming
that a dislocation source operates within a crystal causing dislocations to move and
eventually to pile-up at the grain boundary. The pile-up causes a stress to be
generated in the adjacent grain, which, when it reaches a critical value, operates a
new source in that grain. In this way, the yielding process is propagated from grain
• The grain size determines the distance dislocations have to move to form grain
boundary pile-ups, and thus the number of dislocations involved. With large grain
sizes, the pile-ups will contain larger numbers of dislocations which will in turn
cause higher stress concentrations in neighbouring grains. The shear stress τi at the
head of a dislocation pile-up is equal to nτ, where n is the number of dislocations
involved and τ is the shear stress on the slip plane. This means that the coarser the
grain size, the easier it will be to propagate the yielding process.
• In practical terms, the finer the grain size, the higher the resulting yield stress and,
as a result, in modern steel working much attention is paid to the final ferrite grain
size. While a coarse grain size of d-½ = 2, i.e. d = 0.25 mm, gives a yield stress in
mild steels of around 100 MN m-2, grain refinement to d-½ = 20, i.e. d = 0.0025
mm, raises the yield stress to over 500 MN m-2, so that achieving grain sizes in
the range 2–10 µm is extremely worthwhile. Over the last 40 years, developments
in rolling practice and the addition of small concentrations of particular alloying
elements to mild steels, have resulted in dramatic improvements in the mechanical
properties of this widely used engineering material (Chapter 9).
Nanostructured steels
• Modern technologies allow steels to be made routinely and in large quantities with grain
sizes of about 1 µm. Limited processes, generally involving severe thermomechanical
processing, have been developed to achieve nanostructured ferrite grains in steel, with
a size in the range 20–100 nm. Experiments have revealed that the Hall–Petch equation
holds down to some 20 nm, confirming that enormous strengths can be achieved by
refining the grain size. The equation begins to fail at grain sizes less than about 20 nm,
possibly because other mechanisms of deformation, such as grain boundary sliding,
begin to play a prominent role.
• Although the nanostructured steels are strengthened as expected from the Hall–Petch
equation, they tend to exhibit unstable plasticity after yielding. The plastic instability
occurs in both tension and in compression testing, with shear bands causing failure in
the latter case. It is as if the capacity of the material to work harden following yielding
diminishes. The consequence is an unacceptable reduction in ductility as the grain size
is reduced in the nanometre range. At very fine grain sizes, the conventional
mechanisms of dislocation multiplication fail because of the proximity of the closely
space boundaries. It then becomes impossible to accumulate dislocations during
deformation. Grain boundaries are also good sinks for defects. This would explain the
observed inability of nanostructured materials to work harden. One way of overcoming
this difficulty is described in Chapter 6.
• The difficulty that nanocrystalline grains have in deforming by a dislocation mechanism
is highlighted in recent experiments3 where nanocrystals of ferrite were forced to
deform in shear. The crystals underwent a shear transformation into austenite.
Fig. 2.9 (a)The volume fraction of grain boundary within a given volume,
as a function of the grain size. (b) Loss of ductility as the strength is
increased by dramatically reducing the grain size in aluminium and iron
alloys
DISPERSION STRENGTHENING
• In all steels there is normally more than one phase present, and indeed it is often
the case that several phases can be recognized in the microstructure. The matrix,
which is usually ferrite (bcc structure) or austenite (fcc structure) strengthened by
grain size refinement and by solid solution additions, is further strengthened, often
to a considerable degree, by controlling the dispersions of the other phases in the
microstructure. The commonest other phases are carbides formed as a result of the
low solubility of carbon in α-iron. In plain carbon steels this carbide is normally
Fe3C (cementite) which can occur in a wide range of structures from coarse
lamellar form (pearlite), to fine rod or spheroidal precipitates (tempered steels). In
alloy steels, the same range of structures is encountered, except that in many cases
iron carbide is replaced by other carbides which are thermodynamically more
stable. Other dispersed phases which are encountered include nitrides,
intermetallic compounds and, in cast irons, graphite.
• Most dispersions lead to strengthening, but often they can have adverse effects on
ductility and toughness. In fine dispersions, ideally small spheres randomly
dispersed in a matrix, there are well-defined relationships between the yield stress,
or initial flow stress, and the parameters of the dispersion. The simplest is that due
to Orowan relating the yield stress of the dispersed alloy τ0 to the interparticle
spacing :
• where τs is the yield strength of the matrix, T is the line tension of a dislocation
and b is the Burgers vector. This result emerges from an analysis of the movement
of dislocations around spherical particles, showing that the yield stress varies
inversely as the spacing between the particles.
• If the dispersion is coarsened by further heat treatment, the strength of the alloy
falls. A more precise form of the Orowan equation, due to Ashby, takes into account
the radius r of the particles:
• where φ is a constant and G is the shear modulus.
• These relationships can be applied to simple dispersions sometimes found in steels,
particularly after tempering, when, in plain carbon steels, the structure consists of
spheroidal cementite particles in a ferritic matrix. However, they can provide
approximations in less ideal cases, which are the rule in steels, where the
dispersions vary over the range from fine rods and plates to irregular polyhedra.
Perhaps the most familiar structure in steels
is that of the eutectoid pearlite, usually
approximated as a lamellar mixture of
ferrite and cementite. This can be
considered as an extreme form of
dispersion of one phase in another, and
undoubtedly provides a useful contribution
to strengthening. The lamellar spacing can
be varied over wide limits, and again the
strength is sensitive to such changes (see
Chapter 3).When the coarseness of the
pearlite is represented by a mean
uninterrupted free ferrite path (MFFP) in
the pearlitic ferrite, it has been shown that
the flow stress is related to MFFP-½, i.e.
there is a relationship of the Hall–Petch Fig. 2.10 Dependence of the flow
type (Fig. 2.10). stress at several strains on the
MFFP in a pearlitic steel
AN OVERALLVIEW
• Strength in steels arises from several phenomena, which usually contribute
collectively to the observed mechanical properties. The heat treatment of steels is
aimed at adjusting these contributions so that the required balance of mechanical
properties is achieved. Fortunately the γ/α change allows a great variation in
microstructure to be produced, so that a wide range of mechanical properties can
be obtained even in plain carbon steels. The additional use of metallic alloying
elements, primarily as a result of their influence on the transformation, provides an
even greater control over microstructure, with consequent benefits in the
mechanical properties.
• We have not discussed in this chapter the strengthening and deformation behaviour
of mixed microstructures, such as the dual-phase steels which consist of ferrite and
harder martensite. This can radically alter the stress versus strain behaviour with
the deformation being heterogeneous on a microscopic scale with complex
constraint and compatibility issues governing plasticity. Some of these aspects of
mixed microstructures are described in Chapter 14 as one of the two case studies.
SOME PRACTICAL ASPECTS
• The presence of a sharp yield point in a steel can be detrimental to its behaviour,
e.g., when used for pressings, where complex patterns of Luders bands can produce
rough surfaces and lead to poor workability. The severity of the yield point is
directly related to the amount of carbon and nitrogen in solution in the ferrite, so that
steps taken to reduce these concentrations are helpful. Unfortunately, yield points
can be obtained with very low concentrations of carbon and nitrogen, and it is
impracticable in industrial conditions to obtain steels below these limits. However,
any heat treatment which reduces interstitial solid solution is beneficial, e.g. slow
cooling from annealing treatments. The yield point can be more reliably eliminated
prior to working by a small amount of cold rolling (0.5–2%), referred to as temper
rolling. As both nitrogen and carbon diffuse appreciably in ferrite at ambient
temperatures, it is desirable to fabricate steels soon after rolling and annealing.
• While carbon and nitrogen can both cause strain ageing and consequently a yield
point, the higher solubility of nitrogen in ferrite means that it provides the greater
problem in steels used for deep drawing and pressing. Steps are taken during
steelmaking to keep the nitrogen level down, but to minimize its effects, the easiest
solution is to add small concentrations of strong nitride formers such as aluminium,
titanium or vanadium, which reduce the nitrogen in solution to very low levels.
• The occurrence of strain ageing can, by increasing both the yield
stress and ultimate tensile stress, benefit mild steels which are
used for constructional purposes. Furthermore, the fatigue
properties are improved, both at room temperature and in the
range up to 350◦C. The existence of a well-defined fatigue limit in
steels, i.e. a fatigue stress limit below which failure does not
occur, has been linked to the occurrence of strain ageing during
the test, but even very pure iron shows the same behaviour. It
should be emphasized that even in a relatively simple low carbon
steel, the strength arises not only from these effects of carbon
and nitrogen, but also from the solid solution hardening of
elements such as silicon and manganese, and potentially from
the refinement of the ferrite grain size by various means.
Useful web links
• Fushun Special / Professional Manufacturer of Special St
eel
• Online Materials Information Resource – MatWeb