You are on page 1of 19

HEAT TREATMENT OF CAST ALUMINIUM ALLOYS

A. K. Bhargava
Department of Metallurgical and Materials Engineering,
Malaviya National Institute of Technology Jaipur

Introduction

The heat treatment of articles made of aluminium alloys is considered one of the most important
operations in the whole manufacturing process. Its role is extraordinarily great in imparting high
mechanical strength properties to these alloys. In this respect aluminium alloys are second only to
steels and have become the most important materials (after steels) used for structural purposes in the
modern machine-building industry solely through heat treatment. As structural materials, the
aluminium alloys even have some advantages over steels. In particular, their specific strength is
higher than that of steels, with the result that in some instances the weight of aluminium structures
may be less than that of steel structures with the same strength and rigidity.

In its broadest sense, heat treatment refers to as any operation involving heating and cooling
performed for the purpose of changing the metallurgical structure and hence the properties or the
residual stress of a metallic product. When the term is applied to aluminium alloys, it is frequently
confined to the specific operations employed to increase strength and hardness of precipitation-
hardenable wrought and cast alloys. These alloys are generally referred to as “heat treatable” alloys to
distinguish them from those aluminium alloys in which no significant strengthening can be achieved
by heating and cooling. Although this condition is met by most binary aluminium alloy systems,
many exhibit very little precipitation hardening, and these alloys ordinarily are not considered heat
treatable. The alloys in which a significant strengthening cannot be achieved by heating and cooling
are termed as “non-heat-treatable” alloys. Such alloys, however, can be strengthened by cold working.
For example, binary Al-Si and Al-Mn alloy systems which show increasing solid solubility of Si and
Mn in aluminium, exhibit relatively insignificant changes in mechanical properties as a result of heat
treatment that produce considerable precipitation.

Alloying Elements in Aluminium

The alloying elements commonly used in commercial aluminium alloys include copper, silicon,
magnesium, manganese, and occasionally zinc, nickel, and chromium. The overall effect of alloy
additions is to enhance the yield strength, tensile strength and hardness with corresponding reduction
in percentage elongation. Alloying elements are added extensively to aluminium castings to improve
casting qualities as well as mechanical properties. In addition, machinability of the castings are also
improved by alloying additions.

Copper has been the principal alloying element in aluminium for many years. It is used in amounts
up to 4.5% in wrought alloys and up to 8% in castings. Its effect is to decrease shrinkage and hot
shortness and to provide the basis for age hardening in many aluminium alloys.

Silicon is probably second to copper in its importance as an alloying element, particularly in casting
alloys. It is used in amounts from about 1 to 14% as a primary or secondary alloying element. Silicon
improves casting qualities, such as fluidity and freedom from hot shortness, in addition to providing
corrosion resistance, low thermal expansion, and high thermal conductivity (which assist heat transfer
in water cooling and lubrication systems). Casting alloys containing silicon are noted for good impact
toughness, and pressure tightness. Silicon usually appears as an impurity in all commercial aluminium
alloys. It is used as secondary alloying element in Al-Mg alloys which depend on the formation of
Mg2Si as a hardening phase.

Magnesium is alloyed with aluminium in amounts ranging from about 1 to 10%. Such alloys are
lighter than aluminium and possess good mechanical properties. Magnesium improves machinability
and when present in sufficient amount impart resistance against salt water and alkaline solutions to
aluminium base alloys.

Zinc in amounts up to about 10%, usually associated with other elements, may be added to improve
mechanical properties through the formation of hard intermetallic phase, such as Mg 2Zn.

Manganese and chromium are added in small amounts to increase both the strength and the corrosion
resistance of aluminium alloys. Nickel is also added sometimes to improve the strength at elevated
temperatures with some decrease in corrosion resistance. In the absence of silicon, iron decreases the
hardening capacity of the Al-Cu alloys by removing copper from the solid solution. Lead and bismuth
are sometimes added to improve machinability of aluminium alloys.

Aluminium alloys

Aluminium alloys can be divided up into four categories as shown below:

Non-heat-treatable alloys: As the name implies, these alloys do not respond significantly to heat
treatment processes beyond annealing and stress relief after casting or cold working.

Heat-treatable alloys: As the name implies, these alloys do respond the heat treatment and in
particular to the processes known as solution treatment and precipitation hardening.

Casting alloys are alloys which can be used successfully for casting by a variety of processes
including sand-casting, permanent mould casting and die casting. They may be heat-treatable or non-
heat treatable.

Wrought alloys are those whose mechanical properties allow them to be formed by a variety of
processes including forging, rolling, extrusion and drawing. They may be heat treatable or non-heat
treatable.

Heat-treatable aluminium casting alloys are usually complex alloys containing copper, silicon or even
nickel in significant amounts plus other alloying elements in lesser amounts. A number of these alloys
contain up to 4% Cu, whilst others contain up to 2% Ni. Silicon in silicon containing alloys usually
present up to 5% with exception where it can be as high as 11%. Table 1 lists the examples of these
heat-treatable casting alloys with their composition and some applications. The heat treatment process
principally consists of solution treatment followed by precipitation or ageing treatment.

Aluminium Casting alloys

Aluminium alloys are widely used for producing castings by processes such as sand casting,
permanent mould casting and pressure die-casting. Shrinkage allowances between 3.5 and 8.5% have
to be given in the mould design. Generally, cast aluminium alloys require much longer soaking time
at solutionising temperature as compared to wrought alloys. The pressure die casting alloys such as
LM2 (Al-10Si-1.5Cu, wt%) and LM24 (al-8.5Si-3.5Cu) are not heat treated because in these alloys air
or gas is entrapped during casting process and this air or gas may expand and distort the casting and
give rise to surface blistering at solution sing temperature. Castings, in general, except for creep
resistance, have inferior mechanical properties as compared to wrought products.

To designate the cast alloys of different compositions, American Aluminium Association used four
digit numerical systems in which the first digit indicates the alloy group with 0 and 1 denoting
castings and ingot respectively. The temper designations for castings only too are same as for
wrought products. British standard uses prefix LM. Most cast aluminium alloy components are made
from only four alloys called LM2, LM4, LM6 and LM21. Those compositions that are hardeanble by
heat treatment through precipitation are given the letter “T” in accordance with the following
classification:

-T Heat treatable to stable temper

-T2 Annealed (cast only)

-T4 Solution treated followed by natural ageing at room temperature

-T5 (or TE) Artificially aged

-T6 (or TF)Solution treated and artificially aged

-T7 Solution treated and stabilized to control growth and distortion

-TE Precipitation treated

-TF Fully heat treated: Solutionised, quenched and aged.

Al-Si-Cu Alloys

Binary Al-Si alloys are not susceptible to precipitation hardening but addition of copper made them
susceptible to age hardening since copper has precipitation hardening effect. Addition of copper to Al-
Si alloys results in improvement of strength as well as machinability of alloys but the corrosion
resistance, ductility and castability are adversely affected. LM4 alloy (Al-5Si-3Cu, in wt%) is a
typical alloy in this category. The heat treatment cycle of this alloy is:

Solutionising: Heating to 505 – 520 °C, soaking for 6 - 16 hours, followed by water quenching at 70-
80 °C.

Ageing: Artificially aged at 150 – 170 °C for 6 – 8 hours


Since cupper addition to Al-Si alloy reduces castability, an improved alloy containing higher content
of silicon has been developed. This alloy consists of 11% Si and 3% Cu. This alloy is able to maintain
its useful strength and hardness at high temperatures and therefore can be used safely for piston
application which experience elevated temperature. In the as cast condition this alloy has a T.S. of
190 MPa with 1.5% elongation. After ageing at 200 – 210 °C for 7 – 9 hours, the respective values are
increased to 250 – 320 MPa with 2 – 8% elongation.

Table 1 Heat treatment cycle and applications of some representative commercial cast aluminium alloys

Type Composition, wt Heat treatment Properties Applications


% 0.2% T.S., % El.
P.S. MPa
MPa
LM4 Al-3Cu-5Si-0.5Mn- Solution treated at 252 294 1 Cylinder heads,
0.8Fe-0.2Ti 520º C for 6 hrs. crank cases,
Aged at170ºC for junction boxes, gear
12 hrs. boxes, switch gear
covers, etc.
LM8 Al-4.5Si-0.5Mg- Solution treated at ----- 280 2 Applications
0.5Mn-0.15Ti 465ºC for 8 hrs. requiring good
Aged at 165ºC for casting properties
10 hrs. and corrosion
resistance couples
with good
mechanical
properties.
LM13 Al-12Si-1Cu-1Mg- Solution treated at 280 310 1 Pistons of all types
1.5Ni-1Fe 515ºC-525ºC, for 8 of diesel and petrol
hrs.Quench in hot engines and for
water. Aged other engine parts
hardened at 160- operating at
180ºC, for 4-16 hrs. elevated
Stabilised at 200- temperatures, for
250ºC, 4-16 hrs. pulleys and others
requiring good wear
resistance.
LM14 Al-4.0Cu-0.3Si- Solution treated at 215 280 ---- Pistons and cylinder
1.5Mg-2.0Ni-0.2Ti 510ºC. Age heads for liquid and
hardened in boiling air cooled engines.
water for 2 hrs. A good general
purpose alloy
LM16 Al-1.2Cu-5.0Si- Solution treated at 182 231 1 Water cooled
0.5Mn-0.25Ni 520ºC for 12 cylinder heads,
hrs.water valve bodies, water
quenched. Aged at jackets, cylinder
150ºC for 10 hrs. blocks, air
compressor pistons,
fuel pump bodies,
etc.
4L35, Al-4Cu-0.3Si- Solution treated at ---- 270 Pistons and
“Y” 1.5Mg-2Ni 510ºC, quenched cylinder-heads for
alloy and aged in boiling liquid– and air-
water for 2 hrs or cooled engines.
aged at room Heavy duty pistons
temperature for 5 for diesel engines.
days
Al-Si-Mg Alloys

Like copper, small amount of magnesium addition to binary Al-Si alloys also make them precipitation
hardenable. LM8 (Al-5.5Si-0.6Mg) and LM25 (Al-7Si-0.3Mg) are important alloys fall in this
category. Strengthening is produced by Mg 2Si precipitates. Besides Mg and Si, the other elements
which are invariably present in these alloys as impurities include Fe, Mn, Cu and Zn. The castability
and other properties of these alloys are superior to Al-Si-Cu alloys. They find applications in
carburetor parts, pump castings, and so on. The heat treatment cycle for LM25 alloy is:

Solution treatment: Heating to 535 °C, soaking for 2 – 6 hours and subsequent quenching

Ageing: Artificial ageing at 150 – 180 °C for 3 – 5 hours.

Al-Si-Cu-Mg Alloys

The most common alloy of this group contains Al-4.5 – 5.5% Si, 1.0 - 1.5% Cu, and 0.35 0.6% Mg.
Fe and Zn are present as impurities. These alloys possess relatively good casting properties. Gas
absorption and cracking tendency are low. Precipitation hardening occurs due to CuAl2 and Mg2Si
precipitates. Normally the alloys having higher copper content have CuAl 2 precipitates and those with
higher magnesium (lower copper) content have Mg 2Si precipitates. When iron impurity content is
higher, AlCuSiFe phase may form. This phase neither dissolves properly during solution treatment
nor does it contribute to the process of precipitation hardening. On the contrary, CuAl 2 and Mg2Si
phases precipitate fully and contribute to hardening. The heat treatment involves the following steps:

Solution treatment: Heating to 525 °C, soaking for 6 hours, finally quenched in boiling water.

Ageing: Room temperature ageing for 5 days to obtain T.S. of 227 MPa and % elongation of 0.8. If
aged for 5 hours at 225 °C, its ductility improves to 1.4 % elongation and strength enhanced to 270
MPa.

Precipitation from Solid Solution

One of the important characteristics of precipitation hardenable alloys is the increase of equilibrium
solid solubility of solute element with increasing temperature up to certain limit as illustrated in Fig. 1
for aluminium-copper system which is a typical example of a precipitation hardening system.
Conversely, as the temperature decreases the solid solubility of the solute element decreases with
decrease in temperature, i.e. the phase diagram should show a solvus. The alloys susceptible to
precipitation hardening are those which can form supersaturated solid solutions and then reject a
finely dispersed precipitate when aged at room temperature or intermediate temperatures.

Thus the major precipitation hardening aluminium alloys are:

1. Aluminium-copper alloys with strengthening from CuAl 2 precipitates


2. Aluminium-copper-magnesium systems (magnesium intensifies precipitation)
3. Aluminium-magnesium-silicon systems with strengthening from Mg 2Si precipitates
4. Aluminium-zinc-magnesium systems with strengthening from MgZn 2 precipitates
5. Aluminium-zinc-magnesium-copper systems
Thus, the general requirement for precipitation strengthening of supersaturated solid solutions
involves the formation of finely dispersed precipitates during ageing heat treatments (which may
include either natural ageing or artificial ageing). The ageing must be accomplished not only below
the equilibrium solvus temperature, but below a metastable miscibility gap called the Guinier-Preston
zone solvus line (Fig. 1).

Fig. 1 Solvus lines for metastable transition precipitates as well as stable phase in aluminium-rich end of Al-Cu
phase diagram.

Figure 1 shows that the equilibrium solid solubility of copper in aluminium increases as the
temperature increases and reaches to a maximum of 5.7% at the eutectic melting temperature of 548
ºC. For aluminium-copper alloys containing up to 5.6% Cu, two distinct equilibrium solid states are
possible. At temperatures above the solvus curve, the copper is completely soluble in aluminium, and
when the alloy is held at such temperatures for sufficient time to permit diffusion, all the copper will
be taken completely into solid solution. At temperatures below the solvus, the equilibrium state
consists of two solid phases: solid solution  and an intermetallic compound  (CuAl2). When an
alloy of about 4% Cu in aluminium is heated to a temperature at B, the structure consists of
homogeneous grains of unsaturated solid solution  as shown in Fig. 2. If this alloy is slowly cooled
down from this temperature to room temperature, the microstructure will consists of grains of  solid
solution with precipitation of coarse theta phase preferentially at grain boundaries. A near network of
theta phase at grain boundaries results. This is because, sufficient time for precipitation of copper is
available due to slow cooling. Some precipitation within the grains may also take place if the rate of
cooling is moderate. This microstructure gives rise to poor mechanical properties. On the other hand,
if the alloy is rapidly cooled down, as by water quenching, from temperature at B, there will not be
enough time available for precipitation of copper atoms to take place and form theta phase and the
high temperature solid solution  is retained at room temperature (Fig. 2). Since copper is still
retained in the solid solution of  phase beyond its solubility limit, the solid solution is termed

Fig. 2 Al-rich end of Al-Cu system showing microstructural changes during different heat treating cycles of
precipitation hardening alloys

as supersaturated. This is metastable phase and has tendency to decompose to two phase equilibrium
condition over a period of time: the second phase tends to form by solid-state precipitation. The
supersaturation is the driving force for rejection of the excess solute in the form of precipitates. At
low temperatures, diffusivity of atoms is limited and is promoted by the presence of nonequilibrium
quench in vacancies. The supersaturation of vacancies allows diffusion, thus zone formation, to occur
much faster than expected from equilibrium diffusion coefficients. In the precipitation process, the
saturated solid solution first develops solute clusters, which then become involved in the formation of
transitional (nonequilibrium) intermediate precipitates. This is called ageing. The precipitation occurs
by the nucleation and growth process. The fluctuations in the solute concentration provide small
clusters of atoms in the crystal lattice, which act like nuclei for precipitation. The growth rate of these
nuclei is controlled by the rate of atomic migration, so that precipitation increases. Precipitation
increases either by increasing temperature of ageing or by increasing time a given temperature of
ageing. Lower is the temperature of ageing finer will be the precipitates. The precipitates must be
coherent in nature.

Steps in Age-hardening Treatment

Controlled precipitation from a supersaturated solid solution hardens the alloys, while its critical
dispersion causes extensive hardening. The following are the steps in precipitation or age hardening of
an alloy of proper composition. Figure 3 shows the basic steps involved in precipitation hardening of
alloys.

1. Solutionising: It is the process of heating the alloy just above the solvus temperature, such as
-phase region (Fig. 1) and soaking for some time to obtain a single phase homogeneous
solid solution, i.e., to obtain the complete solution of all alloying elements at a temperature
within the single phase equilibrium solid solution range for the given alloy. Overheating or
under-heating should be avoided. The alloy should not be heated above the solidus
temperature (or initial eutectic melting temperature), as melting or oxidation at the grain
boundaries shall occur which will cause adverse effect on ductility and other mechanical

Fig. 3 Steps involved in precipitation hardening of alloys

properties. Heating the alloy much above the solvus temperature is also not advisable as it
causes grain growth in the alloy. Once grain growth occurs, its refinement at a later stage is
difficult as no phase change occurs during heating these alloys.

Fig. 4 Aluminium rich portion of Al-Cu binary phase diagram showing temperature ranges for
annealing, precipitation heat treatment and solution heat treating. The range for solution treating is
below the eutectic melting point of 548 ºC at 5.65 wt% Cu.

The lower limiting temperature should be above the solvus temperature (Fig. 4). Many
problems are faced while effectively solutionising the Al-Cu-Mg alloys within a few degrees
of the solvus temperature. If coring is present in the alloy, then the solutionising temperature
close to the solidus causes “burning”, i.e. melting and oxidation at grain boundaries seriously
decreases the ductility of the alloy. As the soaking temperature is raised, the rate of
dissolution of alloying elements into solution increases. The soaking time should be just
enough to dissolve the solute completely. The soaking time should be just minimum in thin
clad sheets to avoid the diffusion of solutes from the alloy to the aluminium outside sheet.
Salt bath takes less time than air furnaces. The atmosphere of air furnace generally contains
water vapour which reacts with aluminium alloy resulting in the formation of hydrogen. This
hydrogen diffuses in the surface of the castings and cause surface blistering. The blistering is
normally due to overheating of the alloy. Overheating also causes grain growth resulting in
inferior properties. The surface reaction with water vapour can be decreased by adding
fluoride salt in the furnace. The soaking time, apart from the thickness and shape of the
component, depends on conditions under which the alloy is cast. It is basically dependent on
the coarseness, or fineness of the microstructure. Sand cast parts have coarser structure than
permanent mould cast parts, and thus, former need longer soaking time (12 hours) to dissolve
phases than latter parts (8 hours). The earlier thermal history of alloy also affects the
dissolution of the alloying elements during solution-heat treatment. For example, the rate of
dissolution increases if the component has been given repeated solution heat treatment. Thus,
the soaking time is decreased. Full-annealed components, having coarse precipitate particles,
show decreased rate of dissolution of the alloying elements (precipitates), thus, demand
higher soaking time. A microstructure having fine and uniform dispersion of precipitates is
easy and quick to be solutionised in solution heat treatment.

2. Quenching: The solutionised alloy is cooled rapidly to retain the high temperature single
phase solid solution at room temperature as metastable supersaturated solid solution (SSS).
Care must be exercised to avoid any precipitation during cooling which otherwise will be
detrimental to mechanical properties and corrosion resistance. As this supersaturation is
normally required even in the centre (i.e. the core) of the section of a part, cold, hot or boiling
water or even air cooling may be used. As cold water gives maximum supersaturation, it may
be used for thick sections. The fastest precipitation occurs in the temperature range 200º to
400ºC. Thus the part should get completely immersed in quenchant before it attains a
temperature of about 410 ºC during cooling from solutionising temperature to avoid
precipitation. There should be enough coolant so that the temperature of the coolant does not
raise to 200º to 400 ºC. The quench delay time, i.e., the maximum allowable transfer time
from furnace to quench bath should be determined to fix cooling rate, the temperature of
coolant, etc. A cooling rate of about 315 ºC/sec in the critical range, or more is needed to
avoid precipitation, though it depends on the alloy, size of the part, coolant, etc. A cooling
rate faster than that required to avoid precipitation, enhances unnecessary chances of
distortion of parts. Milder coolants like hot water (60 – 80 ºC), boiling water, water spray, air
blasts can be used to reduce distortion. Al-Zn-Mg alloys are quenched in salt bath at
temperature, 180 ºC, soaked there for some time, and then cooled to room temperature. Now-
a-days organic quenchants are also being used. If an alloy is quenched slowly, the quench in
vacancies will be less in number, and thus precipitation kinetics may be different. The
precipitation may require more time.
Straightening, if required, of the aluminium alloy part may be done in the as-quenched state.
The precipitation characteristics at room temperature vary from alloy to alloy. Thus in such
cases, the parts may be refrigerated after solution treatment and before the ageing in done.

3. Ageing: It is the process of controlled decomposition of SSS to form finely dispersed


precipitates at one or sometimes two intermediate temperatures for a suitable time period. The
selection of time-temperature cycle for ageing is therefore very important. Depending on the
type of alloy, the ageing is classified as:
(a) Natural ageing: If the quenched alloy is aged at room temperature then ageing is called
natural ageing.
(b) Artificial ageing: Some alloys are not hardened fully by natural ageing and therefore
aged artificially by heating at a temperature higher than room temperature. Artificial
ageing is normally carried out between 100 to 200 ºC for 5 to 48 hours. The precipitation
process chosen has an appreciable effect on the strength and hardness of the alloy as can
be seen from Fig. 5.

Fig. 5 Effects of time and temperature on the precipitation hardening of aluminium alloys

Most alloys require heating for a time interval (i. e. ageing) at one or more elevated temperatures.
When single ageing is to be done, a temperature is selected for which ageing time to develop high
strength properties is of convenient duration. Precipitation hardening of castings should be based on
preliminary tests. The foundry practice affects lots of properties. Soaking times for solution treatment
of castings are much longer as castings have very coarse structures. Castings have to be quenched
more slowly, and thus, boiling water or milder quench may be used.

Types of Precipitates

Depending on the structure of the boundary between the precipitate and the surrounding matrix,
precipitate can be classified as:

(i) Coherent precipitate


(ii) Semi- or partially coherent precipitate, and
(iii) Incoherent precipitate.

A precipitate is said to be coherent if there is one-to-one matching of lattice planes across the
interface, that is, the atomic planes of the matrix lattice more or less continue in the precipitate phase.
This generally produces elastic lattice strains called coherency strains around the boundary where the
lattice planes must be ‘bent’ to give this one-to-one matching as is illustrated in Fig.6.
Fig. 6 Coherent precipitate in Al-Cu precipitation hardened alloy

In an incoherent boundary, there is no regularity of lattice plane matching across the boundary, i.e.,
there is no coherent boundary. An incoherent boundary is a large angle boundary between the
precipitate and the matrix, i.e., normal interface boundary. A semi-coherent precipitate is one having
partial coherency at the interface with the matrix. Some coherency strains are relaxed in creating edge
dislocations at the interface resulting in only partial atomic plane matching at the interface. The edge
dislocations are created in the precipitate phase at the point where the lattice planes of precipitate
phase are located symmetrically between two matrix () – planes (Fig. 7). Figure 7 clearly illustrates
that on both sides of this edge dislocation, there are coherency strains. On such an interface, the edge
dislocation is obtained almost periodically with spacing between the two neighbouring dislocations.
Thus, a semi coherent boundary consists alternately of regions of coherency and regions of disregistry
(region around dislocation).

Precipitation Sequence during Ageing of Alloy

The decomposition of SSS during ageing is usually a complex process. The equilibrium precipitate, ,
normally does not form directly from the SSS at commonly used ageing temperatures. This is because
the nucleation barrier for its formation is too high (i.e. large energy is needed because of the surface
energy required to create the surface of the critical sized precipitate). The precipitation occurs in steps
involving several transitions (metastable) precipitates before the equilibrium precipitate forms. To
illustrate the precipitation process, consider the most popular precipitation hardening aluminium
alloy, namely, Al- 4.5% Cu alloy.
Fig. 7 Semi-coherent or ’ precipitate (atoms are not being shown as spheres but only planes of atoms are
shown). Lightly dark region is the precipitate surrounded by matrix (light region).

The ageing of SSS begins with the formation of clusters of atoms on certain specific lattice planes
(generally of low index) in the matrix. In Al-Cu alloy system, these are the planes {1 0 0}. Size of the
clusters is too small to be detected by common structural methods. These clusters are fully coherent
with the matrix (Fig. 8). These clusters have no definite shape, but once they attain a definite shape
they are called G.P. Zones (A. Guiner and G.D. Preston Zones). Difference between clusters and G.P.
zones is only nomenclature. G.P. zones also called G.P.-I, is cluster of solute atoms having the shape
of a disk with about 10 nm diameter and about 1 nm (i.e. 2-3 atomic planes) thick in Al-Cu alloys
(Fig. 8). Figure 9 shows transmission electron micrograph of GP zones.
 1 nm

 10 nm (100) plane

 1 0 nm

(b )

Fig. 8 Guinier-Preston (GP) - zones showing clustering of atoms in one of {1 0 0} planes


Fig. 9 Transmission electron micrograph of Al-4.5% Cu alloy aged to GP zones.

Strengthening occurs due to G.P. zones and gives rise to first hardening peak in Al-4.5 Cu alloy when
aged at 130ºC (Fig. 10). The elastic strain field resulting due to coherency extends into the matrix
such that effective size of the zone, in impeding dislocation motion, is much larger than its actual
physical size. A G.P. zone does not have a well-defined crystal structure and chemical composition. It
is often considered as a portion of initial solid solution that is enriched with respect to solute atoms.
G.P. zones have a distinct solvus line (Fig. 1) and dissolve above this line on raising the temperature
or do not form if the alloy is not undercooled below this line. In this sense, these are called second
phase coherent particles. Crystal lattice of G.P. zones considered to be the same as that of the matrix
but deformed owing to the difference in atomic diameter.

There is no distinct interface between G.P. zones and the surrounding matrix i.e. G.P. zones are fully
coherent. Spacing between G.P. zones ranges from 7.5 to 10 nm. Density of GP zones is of the order
of 1018 cm-3 (number of zones per unit volume). G.P. zones have been observed in Al-Cu, Al-Mg, and
Al-Mg-Si systems. On further ageing intermediate precipitate phase having structure between matrix
and stable phase (CuAl2 in case of Al-Cu alloys) is formed. In Al-Cu alloys, the intermediate phases
θ’’ and θ’ are formed. All intermediate phases must have at least one coherent boundary with the
matrix.

Fig. 10 Ageing curves for Al-Cu alloys aged at 130 ᵒC


The metastable θ’’ phase is almost fully coherent but has tetragonal structure and is commonly called
as G.P. II having diameter in the range 100-150 nm and thickness of about 10 nm to 15 nm (Fig. 11).
Theta double prime nucleate and begin to replace GP zones. Figure 12 is an electron micrographs of
” phase. These particles are more in number and produce greater distortion in the matrix than any
other type of transition phase and therefore give maximum strengthening (Fig. 10). On continued
ageing of Al-Cu alloy, θ’’ → θ’ → θ. Theta prime (θ’) is semicoherent CuAl 2 phase. θ’ is associated
with smaller distortional field than θ’’. Overageing results in the formation of θ phase with no
coherency with the matrix. Means, these precipitates are completely incoherent and lower the strength
of the alloy. During overageing, larger precipitate particles grow at the expense of smaller ones in the
same way as grain growth occurs during annealing of a cold worked material. This is called Ostwald
ripening. This results in increase of interparticle spacing l and hence the lower yield strength of the
alloy in accordance with the following relationship:

Gb
τ=
l

Fig. 11 Metastable Theta double prime (”) phase corresponding to CuAl2 composition.

The overall sequence of precipitation in Al-Cu alloys can be represented as:

SSS  GP zones  ”  ’   (CuAl2)


Fig. 12 Transmission electron micrograph of Al-4.5% Cu alloy aged to ”.

Mechanism of Precipitation Hardening

In age hardened alloys, hardening is governed by the interaction of moving dislocations with precipitates.
In precipitation hardening alloys, the obstacles which hinder the motion of dislocations may be either or
both of the following:
(i) the elastic coherency strains (and hence the stress field) around the coherent precipitates (such
as GP-zones in the matrix);
(ii) the GP-zones themselves or precipitates,

In case of GP-zones themselves or intermediate precipitates, the moving dislocations can either cut through
them or to bent around them and by pass. Thus there appears to be three causes of hardening: internal
strain hardening, chemical hardening that occurs when the moving dislocations shear through the
precipitate particles, and/or dispersion hardening through dislocations as they bent around (circumvent)
precipitate particles and by pass.

Internal strain hardening:  An elastic stress field always exists in the matrix upon formation of a
coherent or a semicoherent precipitate because of an elastic distortion (or strain) at the matrix-
precipitate interface for the reasons mentioned above and is represented in Fig. 6 and 7. The
magnitude of elastic stress fields is greater at a larger size misfit between the structures of the matrix
and precipitate. In order a gliding dislocation to move through elastically deformed matrix, a stress
greater than the stress associated with the elastic strain field around the precipitates must be applied
i.e., the applied stress must be at least equal to the average internal stress. Thus, the greater the elastic
stress field around the precipitates in the matrix more will be the stress required for dislocation motion
and harder will be the material.

Chemical hardening:  Once the elastic stress field around the coherent precipitates has overcome by
the applied stress, the dislocations pass (or cut) through the precipitates by shearing them (Figs. 13 and
14) because of the atomic plane matching between the matrix and coherent precipitates. As a
consequence the precipitates get deformed together with the matrix. Another reason for dislocations to
cut through
Fig. 13 Schematic representation showing shear of coherent precipitates by a gliding dislocation .

Fig. 14 Electron micrograph showing the shearing of a precipitate particle by dislocation

coherent precipitates such as GP-zones instead of avoiding them is that the density of GP-zones is very
high and the spacing between them is so small (of the order of 10nm) that the stress required for
dislocation to by-pass them is appreciably higher than that needed for shear. For the given volume
content of precipitate phase, an increase in particle size will result in a decrease of particle density and
an increase of the interparticle spacing. There is a critical particle diameter dc above which the particles
are not sheared, but bypassed. When particle cutting occurs, the extent of strengthening depends on the
characteristics of the precipitates, which influence the nature of the dislocation-precipitate interaction.
When a particle is sheared by a dislocation, a step which is one Burgers vector (i.e. equal to b) larger is
produced at the particle-matrix interface (Figs. 13 and 14). As a consequence the surface area of the
particle is increased. If the surface-to-volume ratio of the particles is relatively high, then this increase in
surface area can be significant and can represent a substantial increase in the overall energy. This
increased surface area is associated with an increase in surface energy which must be supplied by the
external stress. This reflects an increase in strength of the alloy. This effect is referred to as chemical
hardening and is of principal importance in those systems which form thin, plate-like zones or
precipitates as in Al-Cu and Cu-Be alloys.
Hardening through circumvention of precipitated particles by dislocations:  When the precipitate
particles become strong enough (as on overageing) and lose coherency so that they cannot be sheared by
dislocations (due to non-coherency, e.g.), hardening occurs as the precipitate particles are by-passed by
dislocations and leaving behind dislocation loops around them. For this to occur, the stress applied must be
increased so as to bend the dislocation between the particles. However, this bending of dislocation through
the pair of particles depends on the spacing l between the precipitate particles. The critical stress required
to squeeze the dislocation between the particles is inversely proportional to l in according to the
relationship:

(6.15)
where, G is the shear modulus of the matrix and b is the Burgers vector of the dislocation. Thus, for an
alloy in the overaged condition, smaller is the interparticle spacing stronger is the alloy. As the dislocation
bows between the pair of particles, its portion at both sides of the particle bent, join, and eventually form
loop (also called Orowan’s loop) around the particle (Fig. 15 & 16). Upon detaching from the loop the

Fig. 15 Interaction of an edge dislocation with pair of precipitate particles.

dislocation continue to glide in the matrix. When a second dislocation approaches a particle encircled by a
dislocation it will need a greater magnitude of stress to squeeze between such particles. This is because the
effective distance between two adjacent precipitates l is decreased due to first dislocation loop around the
particles. As more number of dislocations glide on the slip plane they all will form loops around the
particles with increasing magnitude of applied stress. Progressive development of more and more
dislocation loops around precipitate particles ultimately blocks the motion of succeeding dislocations
through the lattice. This is because the dislocation loops exert a back stress on dislocation sources which
must be overcome for additional slip to take place. This requires an increase of applied shear stress, and
hence, to result in strain hardening. The overaged precipitates are larger in size. During overageing of the
alloy, the larger precipitate particles grow at the expense of the smaller ones. As a consequence, for a given
volume fraction of the precipitate particles, larger the size of particles less will be their number and more
will be the mean distance between them such that the radius of curvature, R, of the dislocation becomes
much smaller than the mean spacing l between the particles. It is thus easier for dislocations to bend round
the particles at lower flow stress. That is, the alloy in the overaged condition displays lower yield strength
but rate of strain hardening increases with strain. The hardness of a completely overaged alloy can be
lower than the hardness of the supersaturated solid solution from which precipitate phase has been derived.
Fig. 16 Electron micrograph showing a moving dislocation formed a loop around precipitate particle and
bypassed

However, even when the dispersion of the precipitate is coarse a greater applied stress is necessary to force
a dislocation past the obstacles as compared to the case when no obstruction exists for dislocation motion.
Thus, in overaged alloy, dislocations avoid the precipitate particles by bending round them. Dislocations
can also avoid precipitates by cross-slipping. When a dislocation is blocked by a pair of dislocation looped
particles, the screw part of the moving dislocation can cross-slip to a plane where there is no obstacle. The
edge part of the dislocation can climb if sufficient thermal energy is available. The stress needed to
overcome an obstacle in this way, thus, decreases with increasing temperature. This way the dislocations
avoid precipitate particles.

The magnitude of hardening in precipitation hardenable alloys also depends on composition which is
apparent from Figure 10. At concentration of Cu in aluminium lower than 4 wt% hardening is lower.
This can be correlated to low density of precipitate particles to hinder the dislocation motion. In
addition, with low density of particles the interparticle spacing will also be larger to make dislocation
bending between them easier. Part of the strengthening is also associated with coherency strains
around the particles.

Tensile strength, yield strength and hardness increase considerably due to the strengthening by
precipitation hardening. Tensile strength values as high as 6 to 20 times of the base metal can be attained in
precipitation hardenable alloys. The increase in strength and hardness depends on solution heat treatment
temperature, aging temperature and aging time. Higher the solution treatment temperature, shorter is the
ageing time required to reach maximum strength or hardness for a given aging temperature. A similar
trend is observed with aging temperature. However, the extent of strengthening (maximum strength or
hardness) also decreases with increasing ageing temperature which is not desirable. By lowering ageing
temperature, maximum strength or hardness value can be raised reasonably. Ageing time required to reach
maximum strength or hardness will be certainly more with lower ageing temperatures.

An optimum combination of aging temperature and aging time is selected in actual practice. Ductility
of the precipitation hardenable alloys decreases with increase in strength and hardness.

Conclusion

Heat treatment of aluminium alloy casting is an art and challenge. Age hardening has its origin in
complex precipitation processes provides a good example of the transition of metallurgy from an art
to a science. In order to achieve a good combination of mechanical and physical properties in as cast
commercial aluminium alloys, the control of heat treating cycle at all the stages is of great concern in
industries. The presence of impurities or intentionally added elements in traces in aluminium alloys
also have great effect on solutionising and quenching cycles of heat treatments. After artificial ageing
to achieve peak hardening by primary precipitate phase, one should also look for any secondary
precipitation that may result over a period of time at room temperature and ultimately affect the
mechanical properties of the alloy, in particular, ductility and toughness.

Bibliography

1. T.V. Rajan, C. P. Sharma and Ashok Sharma, Heat Treatment : Principles and Techniques,
PHI Learning Pvt. Ltd., New Delhi, II Ed., 2011.
2. Vijendra Singh, Heat Treatment of Metals, Standard Publishers Distributers, Delhi, 1998.
3. A. K. Bhargava and C. P> Sharma, Mechanical Behaviour and Testing of Metals, PHI
Lerning Pvt. Ltd., New Delhi, Reprint, 2014.
4. R. L. Timings, Engineering Materials, Vol. I, Longman Scientific & Technical, UK, 1989.
5. I. Novikov, Theory of Heat Treatment of Metals, Mir Publisher, Moskow, 1978.
6. H.R. Shercliff, Aluminium Alloys, Material, Handouts 2, Part IB, Paper 3, October, 2001.
7. A.M.A. Mohamed and F.H. Samuel, A review on the Heat Treatment of Al-Si-Cu/Mg Casting
Alloys, INTECH, http://dx.doi.org/10.5772/79832.
8. Raymond A. Higgins, Engineering Metallurgy, Pt.I: Applied Physical Metallurgy, 5 th Ed.,
English Language Book Society (ELBS)/Edward Arnold, 1983.
9. Donald S. Clark and Wilbur R. Varney, Metallurgy for Engineers, D. Van Nostrand Co.,
INC., Reprint, 1963.
10. I.J. Polmear, Aluminium Alloys-A Centuray of Age Hardening, Materials Forum, Vol. 28,
2004, pp 114.
11. James P. Schaffer, Ashok Saxena, Stephen D. Antolovich, Thomas H. Sanders, Jr., Steven B.
Warner, The Science and Design of Engineering Materials, Richard D. Irwin, 1995.

You might also like