Precipitation Hardening in Metals - T Gladman

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Precipitation hardening in metals

T. Gladman

Precipitation hardening has long been used to increase the strength of commercial alloys, such as quenched and tempered steels
and the duralumin type aluminium alloys. T he theoretical treatments of precipitation hardening are briefly considered. T he
equations for strengthening by ‘hard’ indeformable particles and by ‘soft’ deformable particles are presented, and the implications
are discussed. T hese lead to the concept of an optimum particle size for a given system, but the optimum can vary from system to
system depending upon the particle characteristics. A broad comparison is made between the increments in strength that occur
due to precipitation in commercial alloys and the predictions of the theories; an important contribution to these increments in
strength is shown to derive from variations in the volume fraction of precipitated particles that can be employed in the various
systems. MST /4157

T he author has recently retired from the Department of Materials, University of L eeds, L eeds L S2 9JT , UK. T his paper
formed part of the A
‘ dvances in physical metallurgy’ sessions of Materials Congress ’98 organised by T he Institute of Materials
and held at Cirencester on 6–8 April 1998.
` 1999 IoM Communications L td.

As F increases, so the bowing of the dislocation increases,


Introduction i.e. h increases. The magnitude of the resistance force is
important in controlling the sequence of events. The
Precipitation strengthening has been used extensively for a dislocation line tension force is a maximum when sin h=
considerable period of time, and in a wide range of metallic 1, or h=p/2 radians. If the particle is hard, such that the
systems. Essentially, precipitation strengthening is achieved resistance force F can be greater than 2T , then dislocations
by producing a particulate dispersion of obstacles to will bypass the particle either by Orowan looping or cross-
dislocation movement, the particulate dispersion being slip and the particle will remain unchanged, i.e. non-
achieved by a second phase precipitation process. The deformed (Fig. 2). The actual strength of the particle under
degree of strengthening obtained is highly dependent upon these circumstances becomes irrelevant, as the bypassing
the metallic system involved, the volume fraction and size operation becomes dependent only upon the interparticle
of the particles, and the nature of the interaction of the spacing. If, however, the strength of the particle is such
particles with dislocations. Precipitation strengthening is that the maximum resistance force is attained before sin h=
used extensively in ferrous systems as, for example, in the 1, then particles will be sheared and the dislocation will
quench aging of low carbon steels, in microalloyed steels, pass through the particle (Fig. 3). It therefore follows that,
and in the tempering of martensite. This strengthening for a given interparticle spacing (given volume fraction and
mechanism is arguably more important in non-ferrous particle size), hard particles will give the maximum
systems, the most widely known being the Al–Cu and precipitation strengthening, and this condition defines the
Al–Si based alloys, and more recently the Al–Li based maximum degree of strengthening attainable. Soft particles
alloys, where the opportunities for structural refinement will give a lesser degree of strengthening.
are restricted by the absence of allotropic changes. The
detailed interactions of the precipitated particles with dis-
locations is important to the magnitude of strengthening HARD PARTICLES AND THE
observed. Various mechanisms have been proposed, involv- OROWAN RELATIONSHIP
ing particle bypassing by Orowan looping, or bypass slip, Consideration of the relationship between the applied stress
or particle shearing. Theoretical treatments of these mechan- and the dislocation bowing, following Orowan,1 leads to
isms are available and can be used, albeit with varying the Orowan equation
degrees of success, to predict the degree of strengthening
to be expected from a given volume fraction of particles Dt =Gb/L . . . . . . . . . . . . . . (1)
y
and precipitated particle size. In practice, complications
arise from coherent particles with their attendant coher-
ency strains. line
The purpose of the present paper is to examine tension particle
the nature of the various precipitation processes and T T
their attendant effects on the strength of the metal, and
to compare experimental observations with theoretical pre- dislocation
dictions.

Precipitation hardening mechanisms

DISLOCATION–PARTICLE INTERACTION
The forces acting on a mobile dislocation in a stressed
metal containing a dispersion of second phase particles are
shown schematically in Fig. 1. Consideration of the balance
of forces between the line tension of the dislocation T and resisting
the resistance force of the second phase particle F shows force F
that 1 Balance of forces acting during particle resistance to
F=2T sin h dislocation movement

30 Materials Science and Technology January 1999 Vol. 15 ISSN 0267–0836


Gladman Precipitation hardening in metals 31

4 Ashby–Orowan relationship

particle size, are virtually independent of the system


provided that the particle shape remains essentially spheri-
cal. Geometric considerations may introduce changes in
the proportionality factors for needle or plate shaped
precipitates.
The increases in yield strength to be expected from small
volume fractions ( less than 0·5%) of particles of various
sizes in an iron matrix are illustrated in Fig. 4. The root
dependency of yield strength on the volume fraction is
2 Dislocation meets hard undeformable second phase
particles: dislocation release at higher stresses may
reflected by the slope of 0·5 in logarithmic coordinates. The
occur by Orowan looping or by cross-slip dominant effect of the reciprocal of particle diameter in
equation (3) over the logarithmic term is clearly shown in
the diagram. Also shown in Fig. 4 is a comparison of the
where Dt is the increase in yield stress due to the particles, predictions of the Orowan and Ashby–Orowan equations.
y
G the shear modulus of the matrix, b the Burgers vector of The effect of using the effective spacing in the Ashby–
the dislocation, and L the particle spacing. The value of L Orowan equation rather than the minimum spacing in the
in the Orowan equation is usually considered to be the Orowan equation is to reduce the predicted precipita-
distance between particles arranged on a square grid in the tion strengthening contribution. Experimentally observed
slip plane. precipitation strengthening levels in microalloyed steels
Ashby2 further developed this equation to take into have shown excellent agreement with the predictions of
account the interparticle spacing, and the effects of the Ashby–Orowan equation.6
statistically distributed particles (after Foreman and Makin3 The implications of equation (3) for the precipitation
and Kocks4). The Ashby–Orowan relationship is given as strengthening of other metals by hard particles is that the
Dt =0·84(1·2Gb/2pL ) ln(x/2b) . . . . . . . (2) magnitude of the strengthening contribution will be almost
y proportional to the product Gb in equation (3) for any
Application of the Taylor factor for polycrystalline mater- given metal matrix (Table 1).
ials, expressing the microstructural parameters in terms of It can be seen that the major source of strengthening
the volume fraction and real diameter and converting shear variation derives from variations in the elastic rigidity
stress to tensile stress, yields5 modulus, the Burgers vector being very similar for all
Ds =(0·538Gbf 1/2/X) ln(X/2b) . . . . . . . (3) of the metals given in Table 1. Broadly, the degree of
y strengthening varies by a factor of 4, from the lowest
where Ds is the increase in yield strength (MPa), G is the modulus metal shown (magnesium) to the highest shown
y
shear modulus (MPa), b is the Burgers vector (mm), f is (iron). Some variation in the degree of strengthening can
the volume fraction of particles, and X is the real (spatial ) be expected from variations in the Burgers vector in the
diameter of the particles (mm). logarithmic term in equation (3). Given a very small
In the case of hard particles, therefore, the system particle diameter of, say, 10 nm, variations between the
parameters involved are the shear modulus of the matrix
and the Burgers vector of the slip dislocations. The
microstructural parameters, i.e. the volume fraction and Table 1 Shear moduli G and Burgers vectors b in
metal matrixes

G, b, Gb,
Matrix MPa nm Slip system MPa mm

Al 26200 0·286 (111) 110 7·5


Cu 48300 0·255 (111) 110 12·3
Fe 81600 0·248 (110) 111 20·2
Mg 17300 0·321 (0001) 112: 0 5·6
Ni 76500 0·249 (111) 110 18·9
3 Dislocation motion may continue through second Ti 45600 0·295 (0001) 112: 0 13·5
phase particles (particle cutting)

Materials Science and Technology January 1999 Vol. 15


32 Gladman Precipitation hardening in metals

5 Low energy position of flexible dislocation

Burgers vectors given in Table 1 would affect the degree of


strengthening only by about 9%, and with larger particles
this effect is reduced. Bearing in mind the opposing effects
of the Burgers vector in the linear and the logarithmic
terms, the elastic modulus of the matrix becomes even
more dominant than the data in Table 1 would suggest.

SOFT PARTICLES AND PRECIPITATION


STRENGTHENING
If the particle is ‘soft’, i.e. it is sheared by the dislocation,
as shown in Fig. 3, there are several effects which may be
involved in raising the stress level required for plastic
yielding. The deformation of a particle results in an increase 6 Coherency strain hardening
in the particle/matrix interfacial energy. The passage of a
dislocation through a particle may produce an antiphase In the case of chemical hardening the microstructure/
boundary with an associated disordering energy, or may property relationships derived are typified by that for
produce a stacking fault within the particle with its order hardening10
associated stacking fault energy. The latter effect would
naturally require the stacking fault energy in the particles Dt =(c1/2/b)(4r f/pT )1/2 . . . . . . . . . . (6)
y s
to be significantly different from that of the matrix, where c is the antiphase boundary energy, the boundary
and must take into account the changes in the width resulting from the disordering effect of the penetrating
of the stacking fault from the matrix to the particle. dislocation, and r is the mean radius of the particle in the
These strengthening effects are generally termed chemical s
slip plane.
hardening. The variable T is the free line tension (De Wit–Koehler)
A further important effect observed in precipitation of the dislocation in an isotropic matrix,9 and is pro-
hardened alloys is associated with the strain fields around portional to the product Gb2. Equation (6) thus contains
coherent particles, particularly in the early stages of parameters relevant to the matrix, arising from the effects
precipitation. The atomic displacements associated with of the dislocation and the attendant force exerted by the
the dislocation interact with the coherency strain fields dislocation on the particle, and parameters relevant to the
thus providing a greater strengthening effect than would nature of the particle (in this case the antiphase boundary
be expected from the precipitates/zones/clusters alone. Mott energy), as well as the geometric terms relevant to
and Nabarro7 have described the features involved in this the dispersion (particle size r ) and volume fraction. The
type of hardening. Briefly, a dislocation will bend to lie in s
importance of this equation lies in the increase in strength
low energy positions with respect to the strain fields that accompanies an increased particle size at a given
(Fig. 5), and consequently any force applied must overcome volume fraction. A similar form of equation also applies to
an enhanced number of obstacles without the assistance of stacking fault strengthening, but not to the simple increase
segments of dislocation in high energy positions before in interfacial area between the particles and matrix due to
stressing. The introduction of coherency strain hardening particle shearing11 where
introduces complications into the interpretation of age
hardening curves, as the effective obstacle volume fraction Dt =2(c /b)3/2/(bL T )1/2 . . . . . . . . . . (7)
y s
is greater than the volume fraction of second phase particles. The equation indicates an increase in strengthening as the
These mechanisms have been reviewed extensively8,9 and particle size decreases (i.e. as L decreases). In general,
attention has been drawn to the effect of the maximum however, the interfacial energy is often regarded as being
restraining force (that can be exerted by a particle) on the small, when compared with antiphase boundary energies,
angle through which the dislocation can be turned (Fig. 1). and is therefore neglected.
The significance of the critical angle to the increase in yield In the case of coherency strain hardening, the dislocation
stress is quite obvious, but a second and less obvious effect interacts with the coherency strain field in the matrix
is in the definition of the particle spacing. With a very around the (coherent) particle. When the particles are
weak restraining force from the particle, the angle h (see relatively small, one solution, relevant to very limited
Fig. 1) approaches zero, and the number of particles per dislocation bending, yields
unit length, n is given by
L Dt =4·1Ge3/2(rf /b)1/2 . . . . . . . . . . . (8)
n =3f/4r . . . . . . . . . . . . . . . (4) y
L In the case of larger particles, where considerable flexing
or the mean spacing between particles L is given by of the dislocations occurs as a result of the spacing of
L =n−1 =4r/3f . . . . . . . . . . . . . (5) individual particles (Fig. 5), any movement of the dislo-
L cation will have to overcome a larger number of obstacles
As the restraining force increases, so the convolutions of per unit length
the dislocation increase up to the limit where Orowan
Dt =0·7Gf 1/2e1/4(b/r)3/4 . . . . . . . . . . (9)
looping occurs. The models developed for strengthening y
must therefore relate the effective particle spacings to the In both Equations (8) and (9), the misfit strain e is the
general parameters of the dispersoid, e.g. volume fraction fractional dilation of the coherent lattice irrespective of
and particle size. sign. For small particles, coherency strain hardening

Materials Science and Technology January 1999 Vol. 15


Gladman Precipitation hardening in metals 33

Work hardening rate MPa


Shear proof stress MPa
contraction

Contraction %
shear
proof stress

°C
work
hardening

Temperature
rate

as quenched Aging time h


7 Variation of shear proof stress (tensile 0·1%PS/3) and
initial work hardening ((tensile 0·1%PS–0·05%PS)/3)
with time for Ni–Fe–Cr alloy containing L1 type
2
particles

increases with increasing particle size, whereas for larger


particles, the coherency strain hardening decreases with
increasing particle size, giving rise to a maximum in the Cu wt-%
strengthening at a critical particle size which is of the 8 Aluminium rich Al–Cu binary diagram
order of
r/b=Be−1 . . . . . . . . . . . . . . (10) the dislocation interactions may be increased only slightly
giving rise to little difference in the work hardening rate.
irrespective of the volume fraction. As pointed out pre- However, with Orowan looping and cross-slip, a consider-
viously,9 the upper limit of this form of strengthening is able enhancement of work hardening can be found due to
imposed by the transition to the classical Orowan looping the enhanced dislocation interactions (Fig. 7).
process (Fig. 2). Again, this process is not only dependent
upon the matrix properties and geometric feature of the
dispersion, but also on the nature of the precipitated Precipitation hardening processes
particle – in this case, the fractional misfit strain.
The net effect of increasing the particle size in these
systems is illustrated in Fig. 6, using the ‘small particle’ and The precipitation hardening process requires that the
‘large particle’ coherency strengthening equations. Initially second phase is sufficiently soluble to allow extensive
the shear strength in the slip plane increases with increasing dissolution at an elevated temperature – the solution
size, and this would be true not only for fine particle treatment temperature – and that the solubility is consider-
coherency strain hardening, but also for stacking fault ably reduced at lower temperatures – the aging temperature.
hardening, and for order hardening. At some transitional The extent of precipitation is thus dependent on the
stage, e.g. as defined by equations (8)–(10), the mechanism equilibrium phase diagram for the alloy system which
changes and the precipitation strengthening is reduced, controls the extent of dissolution and precipitation at
whether this is by the ‘large particle’ model shown in Fig. 6, appropriate solution and aging temperatures, and hence
or by the classical Orowan looping process described controls the volume fraction of fine precipitated particles
above. As illustrated in Fig. 6, a maximum strengthening that contribute to precipitation hardening by the various
will be attained at the critical particle size, and the mechanisms described above.
magnitude of this maximum will be dependent upon The variations in permissible volume fractions of particles
the magnitude of the fractional coherency strain and the in different systems can be about 2 orders of magnitude.
volume fraction of particles. As each of the strengthening For example, in the Al–Cu system, up to 4 wt-%Cu can be
mechanisms show a square root dependency of the strength dissolved in the aluminium at near eutectic temperatures,
increment on the volume fraction of particles, the peak giving rise to about 5 vol.-% of h (CuAl ) phase after
2
strength will occur at a given particle size (all other things precipitation. In the case of microalloyed steels, however,
being equal) irrespective of the volume fraction. Enhanced the solubility of most of the microalloy carbides and
hardening by increasing the coherency strain will reduce nitrides is limited, generally to less than 0·1 wt-%, giving
the particle size at which the maximum strength occurs rise to 0·1 vol.-% of carbide or nitride after precipitation.
and will therefore increase the maximum strength attainable Thus the extent of precipitation hardening in the peak aged
at a given volume fraction. Similar effects are predicted for condition can not be judged from the shear modulus of the
the other dispersion strengthening mechanisms involving matrix alone (see Table 1), but depends on the solubility of
‘soft’ particles, together with the limiting strength imposed the second phase.
by the ‘hard’ particle Orowan looping mechanism, which
is independent of the particle type. CLASSICAL PRECIPITATION HARDENING
The strength of the dislocation–particle repulsion force SYSTEMS
in connection with ‘soft’ particle strengthening is important The development of the microstructure is not often confined
in dictating the critical particle size at which the transition to simple precipitation of the equilibrium phase. The classic
to ‘hard’ particle strengthening occurs and, therefore, the precipitation sequences in the case of Al–Cu alloys have
magnitude of the strengthening contribution at this tran- been established for at least 60 years, and the story of the
sitional particle size. Nicholson12 has suggested that, development of Guinier–Preston zones, followed by h◊, h∞,
irrespective of the ‘hardness’ of hard particles, there will be and the equilibrium h, is well established. The solvus
a transition to particle cutting mechanisms at very small temperatures of these metastable phases (Fig. 8) provides a
particle sizes. firm basis for successful heat treatment of the duralumin
One other consequence of the change from particle type alloys. These complex precipitation sequences are seen
cutting to Orowan looping is found in the work hardening in virtually all precipitation hardened alloys. In Al–Cu–Mg
characteristics of the material. When particle cutting occurs, alloys, GP zones are formed, followed by the ordered S∞,

Materials Science and Technology January 1999 Vol. 15


34 Gladman Precipitation hardening in metals

before the equilibrium S (Al CuMg) phase. In Al–Mg–Si when tempering is carried out at temperatures above 300°C.
2
alloys, GP zone formation is followed by the ordered b∞, The volume fraction of e-carbide would be somewhat less,
before the formation of the equilibrium Mg Si precipitate. about 2%, when tempering is carried out at low temper-
2
A similar sequence is also observed in Cu–Be alloys, where atures (100–200°C) due to the relatively high solubility of
c (CuBe) is the stable intermetallic compound. A direct e-carbide, but of course the particle size is much finer and
consequence of the formation of these intermediate meta- capable of very large precipitation strengthening contri-
stable stages is that the age hardening curve may reflect butions. One of the main problems in assessing the degree
these changes, giving rise either to multistage hardening of precipitation hardening in these steels is the coincident
or to a single hardening stage, depending on the aging variation in effective grain size, whether this is the
temperature relative to the solvus temperature for a martensite plate dimension or the subgrain structure, and
particular intermediate stage. in dislocation density.
The role of vacancies has long been acknowledged in the From the standpoint of precipitation, microalloyed steels
precipitation process. The enhanced vacancy concentration provide an illustration of the development of a dispersoid
formed by quenching from the solution temperature leads which is exclusively confined to metals showing allotropic
to an enhanced rate of solute diffusion, and therefore an transformation, i.e. interphase precipitation. The microalloy
increased rate of clustering. As the vacancy concentration carbides/nitrides are wholly or partially dissolved at
decreases, so the rate of clustering decreases. The vacancies elevated temperatures in the austenite temperature range,
may eventually condense to form dislocation loops, which usually 1150–1300°C. The solubilities of the various micro-
can act as nucleation sites for precipitation or may escape alloy carbides and nitrides are defined by the solubility
to various sinks, such as grain boundaries. The loss of product
vacancies at these sinks can give rise to precipitate free
k =a a /a . . . . . . . . . . . . . (10)
zones. The effect of solute additions to these systems, e.g. s M X MX
the addition of lithium13 or of beryllium,14 is to retard the where a is the activity of the microalloy in solution and
M
escape of vacancies and also to reduce the rate of formation a is the activity of the interstitial in solution, in equilibrium
X
of dislocation loops by forming solute–vacancy pairs, with the compound MX. The activity coefficients ( f and
M
thereby changing the rates of clustering and of precipitate f ), and the activity of the reaction product (MX), are
X
formation. In the case of lithium, additions of 0·7 wt-%Li commonly assumed to be unity, and the solubility product
introduce an additional precipitated phase, AlLiSi, which can be written as
can coexist with Si–Mg solute clusters.
k =[M][X] . . . . . . . . . . . . . (11)
Double aging treatments can be used to produce a fine s
distribution of GP1 zones at a temperature below the zone where [M] and [X] are solute concentrations (in wt-%).
solvus, followed by zone coarsening at a higher temperature. The solubility product increases with increasing temper-
Such treatments are important in reducing precipitate free ature according to the Arrhenius equation
zones, and the decision to use double aging treatments or
log k =B−(A/T )
to use alloying in order to optimise the structures and s
properties may well be influenced by economic factors. where B and A are constants and T is the absolute
The complexities of the many classic precipitation temperature. The values of A and B for various microalloy
hardening sequences have been well documented, but the carbides and nitrides are shown in Table 2, together with
coexistence of clusters, and of one or more coherent or the solubility products in austenite at 1200°C. It can be
semicoherent intermediate phases, can lead to difficulties in seen that most of the simple compounds have a solubility
relating the change in properties to the microstructure in a product of 10−3–10−2, but exceptions are vanadium carbide
quantitative manner. with a solubility product in excess of 1, and titanium nitride
with a value less than 10−6.
In commercial steels, due account has to be taken of the
PRECIPITATION HARDENING IN simultaneous presence of carbon and nitrogen, and also
ALLOTROPIC SYSTEMS the fact that many steels may contain more than one
Although there are examples of classic precipitation microalloy together with aluminium. Aluminium nitride,
hardening reactions in ferrous systems, such as the with its hcp structure, forms a separate precipitate, while
precipitation of iron carbides in a-Fe during the quench the microalloy carbides and nitrides enjoy mutually com-
aging of ferrite, and the use of c∞ (Ni Al Ti ) in austenitic plete solid solubility, all the structures being fcc with similar
3 x 1−x
stainless steels, the allotropic changes in transformable but not identical lattice parameters. Treatments of the
steels introduce variants in the development of precipitation initial solid solubility have been presented by Hudd et al.16
hardened microstructures. One of the most extensively used and by Adrian,17 and have been considered elsewhere.5
classes of steels involving precipitation hardening has the On cooling through the austenite range, precipitation
tempered martensitic structure. These transformable engin- of microalloy carbonitrides (and aluminium nitride) is
eering steels are heated into the austenitic temperature extremely sluggish, and can be neglected at most practical
range, where the solubility of iron carbide is increased cooling rates. However, during the austenite to ferrite, or
dramatically ( by a factor of at least 40) when compared austenite to pearlite, transformation there is a step change
with the maximum solubility of 0·02 wt-%C in the bcc in solubility, all microalloy carbides and nitrides being less
a-Fe. Quenching to avoid ferrite formation allows the soluble in the ferrite than in the austenite at these low
development of the hard but brittle bct martensitic
structure, which is effectively a heavily supersaturated
solution of carbon in a-Fe. Tempering such structures Table 2 Solubility products k of microalloy carbides and
s
nitrides in austenite at 1200?C
allows the precipitation of the metastable hexagonal
e-carbide (Fe C), or the stable cementite (Fe C), or, in Component A B k (1200°C) Structure
2·4 3 s
the case of secondary hardening steels, alloy carbides such
as Mo C or V C . In many ways, these precipitation VC 9500 6·72 1·9 fcc
2 4 3 VN 8700 3·63 5·2×10−3 fcc
systems mirror those described for the aluminium alloys, NbC 7900 3·42 1·1×10−2 fcc
i.e. from supersaturated solution, through solute clusters NbN 8500 2·80 1·1×10−3 fcc
and metastable phases, to the equilibrium phase. TiC 10475 5·33 1·7×10−2 fcc
Ultrahigh strength steels usually contain about 0·4 TiN 15020 3·82 4·2×10−7 fcc
AlN 6770 1·03 2·7×10−4 hcp
wt-%C which will allow precipitation of some 6%Fe C
3

Materials Science and Technology January 1999 Vol. 15


Gladman Precipitation hardening in metals 35

9 Interphase precipitation as result of ferrite growth by


‘ledge’ mechanism

temperatures. Also, the mobility of the various solutes in


ferrite is much higher than that in austenite at the same 10 Dispersion strengthening in low carbon steels
temperature. The result of these changes is the precipitation
of microalloy carbonitrides at the austenite/ferrite inter-
face, giving the characteristic interphase distribution with conditions (Table 3). Obviously these estimates are very
rows of fine and relatively uniform precipitated particles. crude, neglecting the changes in solute concentration, and
Honeycombe and his co-workers18 have suggested that the the possible changes in coherency strain as particles grow
interphase morphology arises as a result of the ledgewise (due to the incorporation of interface dislocations), but
growth of the ferrite grains (Fig. 9), the size and spacing of show some reasonable agreement with the prediction. More
the particles being dependent upon the transformation precise comparisons would require specific knowledge of
temperature and the rate of advance of the ferrite ledges the strengthening mechanism and the parameters relevant
(or growth rate of the ferrite). The particles developed in to that mechanism, e.g. coherency strains, surface energies,
this way are fairly uniform in size, and the sizes commonly stacking fault energies, disorder energies, or modulus
range from 5 to 20 nm. The particle size can be controlled differences.
by controlling the temperature, and rate of transformation. Again, with martensitic steels, the prediction of yield
It is interesting to note that this same interphase precipi- strength is difficult. In the early stages of carbide precipi-
tation can be seen in higher carbon steels where the tation, the strength changes are clouded by the dramatic
austenite transformation product is the eutectoid, pearlite, loss of strength to be expected from carbon depletion in
rather than the ferrite of low carbon steels. the matrix; because carbon is an interstitial solute giving a
large tetragonal distortion, precipitation hardening causes
only a marginal increase in strength over the solid solution
Strengthening increments (martensite). One may also question the additivity of
strengthening mechanisms in such structures with their
In general, the observed strengthening increments observed submicrometre plate sizes, forest dislocations, and varying
in ‘hard’ particle systems are described by the Ashby– solute contents. Further complications may arise in very
Orowan relationship (equation (3)). In ‘soft’ particle sys- lightly tempered martensite due to the transition to carbide
tems, although particle shear is clearly observed, there is cracking and/or brittle fracture. In more heavily tempered
often some considerable doubt as to the detailed mechanism martensites, the particle spacing is increased dramatically
of strengthening. The experimental evidence for the smaller and yielding occurs by the Orowan mechanism, the yield
particles often shows the root dependency on the volume stress decreasing with increasing particle size. Some indi-
fraction, and an increase in strength as the particle size cations of the possible contribution in martensitic steels
increases (after the completion of precipitation), often with may be obtained from the absolute strength level of
the root radius dependency. In the case of aluminium alloy, ultrahigh strength lightly tempered martensitic steels, where
for example, with a volume fraction of particles of about 0·2% proof stress levels of 1·4–1·7 GPa are obtained. In a
0·05 and a coherency strain of 0·01 (a 1% misfit), the 0·4%C steel, the formation of 6 vol.-%Fe C ( f =0·06),
3
maximum yield stress (in shear) that would be expected would produce a precipitation strengthening contribution
from coherency strengthening is of the order of 130 MPa. of some 0·8 GPa, with a similar contribution arising from
Conversion to a tensile strength in polycrystalline material the fine submicrometre grain size.
would produce an expected proof stress increment of some Precipitation hardening is used to great advantage in
200–300 MPa. An examination of data from the specifica- microalloyed steels. Although very fine particles can be
tions for aluminium alloys allows an assessment of the sheared, as pointed out by Nicholson, the strengths of these
strengthening increment to be obtained from the difference fine alloy carbonitrides and their small volume fraction are
in proof stress between the annealed and the age hardened such that the transition from particle deformation to
Orowan looping must occur at extremely small particle
sizes. Thus 5 nm particles are capable of resisting deforma-
Table 3 Strengthening increments in aluminium alloys tion, or require near looping stresses to cause shear. The
yield stresses of these steels therefore contain a contribution
0·2% proof stress, MPa
from the precipitation strengthening which is described
Alloy Annealed Age hardened Difference quantitatively by the Ashby–Orowan equation. The small
volume fractions and particle sizes are indicated in Fig. 10,
2024 75 395 320 which illustrates the magnitude of the precipitation contri-
2219 75 290 215
bution. The increase in yield strength for the low carbon

Materials Science and Technology January 1999 Vol. 15


36 Gladman Precipitation hardening in metals

11 Precipitation strengthening in medium carbon steels: after Sawada et al.19

miroalloyed steels shows a remarkable agreement with the precipitation temperatures. Consideration is given to broad
Ashby–Orowan equation to the extent that the equation is classes of commercial alloys, showing the very significant
used directly in alloy design to produce steels of given yield strength increments that can arise from precipitation
properties.6 Interestingly, the use of vanadium microalloying hardening.
in medium and high carbon steels produces very similar
levels of precipitation hardening,19 as shown in Fig. 11,
when due account is given to secondary effects such as
fraction of pearlite, pearlite spacing, and ferrite grain size. References

1. . : ‘Internal stress in metals and alloys’, 451; 1948,


Summary London, The Institute of Metals.
2. . . : in ‘Oxide dispersion strengthening’, (ed. G. S.
Ansell et al.), 143; 1958, New York, Gordon and Breach.
Precipitation strengthening in metals has long been a 3. . . .  and . . : Philos. Mag., 1966, 14, 911.
common strengthening method. This brief overview has 4. . . : Philos. Mag., 1966, 12, 541.
discussed the theoretical approaches which make the 5. . : ‘The physical metallurgy of microalloyed steels’;
distinction between so called ‘hard’ particles, which are 1997, London, The Institute of Materials.
bypassed by dislocations, and the ‘soft’ particles, which 6. . , . , and . . c: in ‘Microalloying
75’, (ed. M. Korchynsky), 32; 1977, New York, Union Carbide.
become deformed when plastic yielding occurs. The 7. . .  and . . . : in ‘Bristol conference on the
Ashby–Orowan equation describes the precipitation strength of solids’; 1948, London, The Physical Society.
strengthening due to ‘hard’ particles, the strength incre- 8. . .  and . : Prog. Mater. Sci., 1963, 10, (3).
ment being controlled by the matrix characteristics and 9. . .  and . . : in ‘Strengthening methods in
the particle distribution. Other relationships exist for the crystals’, (ed. A. Kelly and R. B. Nicholson), 12; 1971, London,
strengthening contributions of ‘soft’ particles, and these are Applied Science Publishers.
dependent on the characteristics of the precipitates as well 10. . . : Proc. Int. Conf. on ‘Strength of metals and alloys’,
as the above factors. Soft particle strengthening increases Tokyo, Japan, 1968.
11. .  and . . : Acta Metall., 1957, 5, 365.
with increasing particle size, whereas hard particle strength-
12. . . : ‘Effect of second phase particles on the
ening decreases with increased particle spacing (increasing mechanical properties of steel’, 1; 1971, London, The Iron and
particle size). This gives rise to the noted transition from Steel Institute.
particle shearing to dislocation looping at some critical 13. . . , . . , . . , and . :
size. The critical size of the precipitated particles is Mater. Sci. T echnol., 1994, 10, 869.
decreased as the stiffness of particles increases, so that 14. . .  and . : Mater. Sci. T echnol., 1994, 10,
the transition to Orowan looping occurs with very small 1031.
particles in the case of alloy carbides in steel, but at a 15. . : T rans. Iron Steel Inst. Jpn, 1975, 15, 145.
significantly larger critical particle size when the particles 16. . . , . , and . . : J. Iron Steel Inst., 1971,
are soft, such as is the case for many classic precipitation 209, 121.
17. . : in ‘Microalloying 1995’, 285; 1995, Warrendale, PA,
hardening aluminium alloys. Iron and Steel Society.
The extent of precipitation hardening generally increases 18. . . . : Metall. T rans. A, 1976, 7A, 915.
as the square root of the volume fraction of particles (other 19. . , . . , . . , and . :
things being equal), and phase equilibrium becomes Proc. 35th Mechnical Working and Steel Processing Conf.,
important from the standpoint of solubility at solution and Pittsburgh, PA, October 1993; Proc. ISS–AIME, 1994, 31, 263.

Materials Science and Technology January 1999 Vol. 15

You might also like