Sie sind auf Seite 1von 18

Processing of Superplastic Aluminium Alloys

John A. Wert
Center for Fundamental Research: Metal Structures in Four Dimensions
Materials Research Department
Ris National Laboratory
4000 Roskilde

53

Abstract
Metallic alloys often achieve extensions exceeding 1000% during superplastic deformation.
Attaining such large extensions requires avoiding plastic instability. In metallic alloys, this is
achieved by processing the material to a grain size smaller than 10 m and deforming at a
moderate strain rate at high temperatures such that grain boundary sliding makes a large
contribution to the overall tensile strain. Material models reveal the strong influence of grain
size on superplastic flow. The required grain sizes can be achieved in aluminium alloys via
several processes. Many conventional heat treatable alloys can be thermomechanically
processed such that the recrystallised grain size is between 5 and 10 m. Alternatively,
special alloy compositions can be processed to finer grain sizes, approaching 1 m. Such
alloys exhibit superplastic deformation at much higher strain rates, but the service properties
may be compromised to some extent by the special microstructures that allow finer grain size
to be attained. Cavitation during superplastic straining can be suppressed by hydrostatic
pressure, avoiding the deleterious effects of voids in formed components. Finally, potential
applications for superplastically formed aluminium components are considered.

54

54

1. Introduction
Superplastic deformation describes a condition in which a material can be extended to very
large strains. A uniaxial tensile strain of 300%, corresponding to a logarithmic strain of 1.4,
is normally considered to be a lower limit for superplasticity, but this lower limit is often
exceeded. For example, many superplastic metallic alloys exhibit tensile elongations
exceeding 1000% and superplastic ceramics typically exhibit elongations of several hundred
percent. The main requirement for attaining superplasticity is avoiding processes that limit
elongation in tensile deformation or biaxial stretch forming: plastic instability (necking) and
fracture. In most metallic alloys, plastic instability is a more severe limitation. The objective
of Section 2 is to analyse plastic instability and to assess the methods available to avoid it.
The conclusion of Section 2 is that superplasticity requires elevated temperature deformation
of materials with grain sizes less than about 10 m [1-4]. Such fine grain sizes are readily
produced in many engineering alloys, but are difficult to attain in aluminium alloys. Thus,
novel processes and special alloy types have been invented to manufacture aluminium alloys
tailored for superplastic forming. In Section 3 of the present article, the processes for
manufacturing superplastic aluminium alloys are summarised. In general, processes
applicable to conventional heat treatable alloys produce grain sizes larger than 5 m, limiting
superplastic deformation to strain rates below 10-3 s-1. Modified alloy compositions combined
with novel processes decrease the grain size to the neighbourhood of 1 m, allowing faster
forming rates but compromising service properties.
When aluminium alloys are subjected to large extensions, internal cavities develop [5,6].
These cavities can limit the elongation achieved during high temperature deformation and,
more importantly, they have a deleterious effect on service properties of superplastically
formed components. Fortunately a method has been devised for preventing formation of
cavities during superplastic flow; Section 4 describes this method briefly. Finally, some
potential applications for superplastic forming of aluminium alloys are mentioned in Section
5.
2. Mechanism of Superplastic Flow
The large extension characteristic of superplastic deformation requires averting plastic
instability and fracture during uniaxial extension. Most alloys of engineering importance
neck prior to fracture, indicating that plastic instability is the factor limiting elongation. Thus,
design of alloy microstructures that promote superplasticity must begin with an understanding
of the mechanics of plastic instability in tensile deformation.
In the gauge section of a tensile test specimen, small zones of reduced cross section occur
randomly during straining. These are zones of slightly greater-than-average extension, and by
the principle of conservation of volume during plastic flow, they are also zones of slightly
reduced cross-section. Since every cross-section in a tensile sample must bear the same load,
the zones of reduced cross-section correspond to zones of higher stress. One can imagine that
flow would be concentrated the zone of greatest stress, further reducing the cross-sectional
area in this zone, leading to a higher stress concentration, creating an avalanche effect that
causes failure as a consequence of the slightest flow perturbation. This effect may be
prevalent in nanocrystalline metallic alloys, for example, with the result that uniform
elongation is virtually zero even though the materials flow plastically in compression. To
55

attain large uniform elongation, flow must be prevented from further concentrating in the
zones of higher-than-average stress. Work hardening and strain rate hardening are the
mechanisms available for preventing further concentration of flow.
For the case of work hardening in the absence of strain rate effects, the well-known result for
predicting the onset of plastic instability was first derived by Considre [7] who showed that
for a material characterised by an exponential hardening law

= K n

[1]

( = true stress, = logarithmic strain (true strain), n = strain hardening exponent, K =


constant), the maximum uniform strain is
u = n .

[2]

For metallic alloys, work hardening exponents approach 0.5 only in rare cases, irrespective of
strain rate and deformation temperature. Thus, attaining the large extensions characteristic of
superplasticity by a strain hardening mechanism is not feasible.
For the case of strain rate hardening in the absence of work hardening, the constitutive law for
material behaviour is

= C& m

[3]

( = true stress, & = logarithmic strain rate = d/dt, m = strain rate exponent, C = constant).
Mathematical analysis of this situation shows that the maximum uniform strain is
approximately proportional to m [8].
The analysis results are consistent with the general observation that dilatant (m > 1) materials
as well as materials that exhibit Newtonian flow (m = 1) are readily extended to very large
elongations. Examples include glass heated above the glass transition temperature (Tg) and
the mixture of silicone and boric acid known as Silly Putty, both of which are easily
stretched into long thin fibres.
The strain rate exponent, m, is on the order of 0.1 or less for metals deforming by dislocation
glide, far below the level of m = 0.3, which corresponds roughly to an elongation of 300%.
However, higher values of m are characteristic of metallic alloys deformed at elevated
temperatures; values approaching 1 are found for some metallic alloys deformed at elevated
temperatures and slow strain rates. Such alloys are resistant to plastic instability under such
deformation conditions and therefore exhibit large tensile elongation; i.e., they are
superplastic.
To understand the material and deformation factors that affect elongation, mechanism based
models of deformation are more useful than phenomenological expressions such as Eqs. 1 and
3. Several mechanisms contribute to creep strain in crystalline materials deformed at elevated
temperatures [9-11]. The three primary mechanisms are:
climb-controlled dislocation motion
diffusional creep
grain boundary sliding accommodated by diffusional or glide.
56

56

To a first approximation, these mechanisms operate independently in parallel. Therefore, the


total creep strain rate is approximately the sum of the rates contributed by the individual
mechanisms. Constitutive models based on atomic scale processes have been developed for
each of these deformation mechanisms; the individual constitutive equations are combined to
give an overall constitutive equation for creep [11,12]

& =

100D
0.72 3.3D b AD v Gb

1 +
+
2
d
dD v
kT G
kTd

[4]

The parameters appearing in Eq 4 are defined in Table 1 where values appropriate for
aluminium are also shown.
The relationship between flow stress and strain rate represented by Eq. 4 is illustrated in Fig.
1 for typical high temperature deformation conditions for Al. The strain rate sensitivity, m, is
the tangent to the curve at any point. All three deformation mechanisms contribute to the
total strain rate for any specified stress, but usually one mechanism makes a dominant
contribution. Thus, the curve is divided into regions of differing slope (m):
Region I: climb-controlled creep dominates, m is often near 0.2 for metallic alloys.
Region II: grain-boundary sliding accommodated by diffusional or glide dominates, m is
usually above 0.3.
Region III: there is much debate about the dominant mechanism and characteristics m in
this region. Definitive experiments are challenging because the creep rate is extremely
slow and tests are therefore of long duration. Microstructural changes such as grain
growth occur during the course of such tests and these changes complicate interpretation
of mechanical test results. Region III is of purely academic interest for application of
superplastic forming.

Table 1: Definition of parameters appearing in Eq. [4] and values for Al. From Ref. [12]
Parameter
Symbol
Value for Al
Units
-6
Molar volume
7.4x10
m3/mol

Grain boundary energy


0.3
N/m

10
Shear modulus
G
2.3x10
Pa
Burger's vector
b
2.9x10-10
m
Grain boundary width
8.6x10-10
m

Volume diffusion pre-exponential


D0v
3.5x10-6
m2/s
5
Volume diffusion activation energy
Qv
1.2x10
J
-5
Grain boundary diffusion pre-exponential
D0b
1.0x10
m2/s
4
Grain boundary diffusion activation energy
Qb
6.0x10
J
29
Power law creep constant
A
2.1x10
-Power law creep exponent
n
4.8
--

57

Clearly, superplastic extensions are achieved by deforming a material in Region II where the
primary mechanism of deformation is grain boundary sliding accommodated by diffusional or
glide. The model predicts a value of m approaching 1 for such deformation conditions. This
is sometimes observed in practice, but values near 0.5 are more frequently observed. This
implies that the models giving rise to Eq. 4 are oversimplified. However, they correctly
identify the strain rate regime where superplastic elongation can be expected and the
microscopic deformation mechanism.

II

Log( , MPa)

III

Al
T = 700K
d = 3 m
Log( , s )

-1

Log( , s )
-1

Fig. 1 The upper plot shows the flow stress strain rate relationship prescribed by Eq. 4
for the deformation conditions indicated on the plot. Deformation in Region II gives
rise to large elongations as a result of the high m value, shown in the lower plot.

58

58

Log( , MPa)

30

10

0.3

d, m

Al
T = 700K

Log( , MPa)

Log( , s -1 )

30

10

0.3

d, m

Al
T = 750K
Log( , s )
-1

Fig. 2 The flow stress strain rate relationship prescribed by Eq. 4 for the deformation
conditions and grain sizes indicated on the plots. The strong effect of grain size on the
strain rate for Region II is clearly revealed. Note that grain sizes smaller than about 3 m
are predicted to give rise to high strain rate superplasticity at 750 K.

59

Note that only deformation with the range of strain rates where m > 0.3 (approximately) gives
rise to large elongations. Thus, superplastic behaviour is only predicted for strain rates in the
range 2x10-4 s-1 to 5x10-1 s-1 the case shown in Fig. 1. The largest elongation is predicted at
strain rates near the maximum m, about 10-2 s-1. For commercial application, it is obvious that
increasing the strain rate corresponding to maximum m is highly desirable, since this reduces
the time required to form a given component geometry.
Examining the list of variables shown in Table 1 reveals that all parameters except grain size
are fixed by choice of the main alloy constituent; i.e., all alloys based on Al have the values
given in Table 1. Thus, the model indicates that grain size is the only microstructural variable
that can be manipulated to affect superplastic properties. While this conclusion is also an
oversimplification, the broad conclusion that grain size changes have a stronger effect on
superplastic properties than any other microstructural parameter is generally true. (An
exception is known to this rule. Some solutes in Ti alloys have exceptionally high diffusivity,
small additions of these solutes have been shown to enhance the superplastic response [12].)
The predicted effect of grain size on the flow stress strain rate curve is illustrated in Fig. 2
for a range of grain sizes in Al alloys. The figure shows that a factor of 3 decrease in grain
size raises by about an order of magnitude the strain rate regime where superplastic
deformation occurs. For grain sizes greater than 10 m, the maximum strain rate at which
superplastic deformation can be expected is less than 10-4 s-1, which is too slow for
manufacturing operations. Thus, even a simplified model reveals that practical superplastic
alloys must have a grain size smaller than about 10 m.
3. Processing Aluminium alloys for Superplastic Properties

The analysis described previously leads to the conclusion that grain size control is the
essential objective of processes used to develop alloys with superplastic properties. Although
the analysis is oversimplified, the conclusion is generally supported by decades of scientific
research spanning a wide variety of alloy compositions and applications.
Fig.3 shows the results of an experiment in which the flow properties of alloy AA7475 were
measured as a function of grain size and temperature [13]. The similarity between the model
results in Fig. 2 and the experimental flow curves is obvious and provides confidence that the
model results approximate the behaviour of real materials in superplastic conditions. Fig. 4
emphasises the strong effect of grain size on the superplastic elongation attained in tensile
tests [14,15]. The grain size listed in Fig. 4 is the value measured parallel to the rolling
direction; the grains in commercial aluminium alloys have a disk-shaped morphology and the
grain size in the direction normal to the plane of the rolled sheet is on the order of 10 30 m
in all cases.
Both modelling and experimental results reveal the strong effect of grain size on superplastic
extension. Thus, reducing the grain size to the 10 m scale and preventing grain growth
during the superplastic forming interval, are critical requirements. In this context, it is
essential to note that deformation microstructures, composed of dislocation cells or subgrains
with primarily low angle dislocation boundaries, are not suitable for superplastic deformation.
This is because the boundary sliding process does not occur for boundaries with
60

60

misorientations less than about 10 [16,17]. Thus, the requirement is a stable grain structure,
with a grain size less than about 10 m.
For alloys that can be superplastically formed in a temperature regime where two deformable
phases are present in somewhat equal volume fractions, such as Fe and Ti alloys, grain sizes
suitable for superplastic forming can be readily attained by phase transformation or large
plastic deformation in the two phase field [14,15]. Examples of such alloys include Ti alloys
processed and superplastically formed in the two phase + phase field, duplex stainless
steels deformed in the + phase field, and hypereutectic Fe-C alloys deformed in the +
Fe3C phase field. Both deformation processing and an allotropic phase transformation can be
exploited in the grain refinement process, and exceptional grain size stability attributable to
chemical segregation between the phases since grain growth in this case requires long range
chemical diffusion.

Fig. 3 The effect of grain size on the flow stress strain rate relationship in
AA7475, a commercial Al-Zn-Mg-Cu alloy. Lower plot shows the m value as a
function of grain size, a commercial Al-Zn-Mg-Cu alloy. (From [13])

61

Fig. 4 Effect of grain size on superplastic elongation of AA7475, a commercial AlZn-Mg-Cu alloy. The grain size designated on the figure, dL, is the mean linear
intercept size measured in the longitudinal direction.
In contrast, grain size control in alloys that do not contain 2 deformable phases is difficult.
Several broad approaches have been pursued in efforts to engender superplastic properties in
aluminium alloys. These broad approaches are outlined in the following subsections.
3.1 Conventional Recrystallisation of Quasi-Single Phase Heat Treatable Alloys

The density of nucleation sites for recrystallised grains during rolling and annealing of
aluminium alloys is far too small to enable the desired grain size to be achieved. Thus, a
mechanism is needed to provide additional nucleation sites. Work at Rockwell International
in the early 1980s showed that such processes could be developed for conventional heat
treatable aluminium alloys [18-20]. While many variants of the process have been developed,
the underlying metallurgical principles are similar.
Heat treatable alloys belonging to the 2000 series (Al-Cu-Mg), 6000 series (Al-Mg-Si), 7000
series (Al-Zn-Mg), and 8000 series (Al-Cu-Li) are normally heat treated to produce fine
dispersions of particles to provide high yield strength. The usual heat treatment procedure
involves solutionising to take the solute elements into solid solution, quenching to retain the
62
62

(supersaturated) solid solution at room temperature, and ageing at a low temperature (T <
200C) to precipitate the solute in a controlled manner, creating dispersions of roughly 10 nm
diameter particles with a spacing on the order of 20 nm.
The thermomechanical process for grain size control employs the same solute elements in a
different manner, as described in the sequence of steps listed below.
1. Solutionising.
condition.

This dissolves any particles present thus providing a known starting

2. Annealing only slightly below the solvus temperature (T 400C). This precipitates the
solute as coarse particles, on the order of 1 or 2 m diameter with a spacing of several
micrometers. At this size, the particles are much too coarse and far apart to provide
significant strengthening.
3. Cold or warm rolling. The dislocation substructure that provides the driving force for
recrystallisation is introduced during this step. In addition, a special zone of deformation is
created around each nondeformable particle. These zones contain small dislocation cells
having crystal orientations not present in the general deformation structure.
The
incompatibility of deformation between the nondeforming particles and the plastically
deforming matrix creates the deformation zones.
4. Recrystallisation. The rolled alloy is heated rapidly to the solutionising temperature. Two
things happen simultaneously. One is dissolution of the coarse particles, returning the solute
to solid solution. The other is nucleation of recrystallised grains in the deformation zones that
were created around the coarse particles. While not every particle creates a recrystallised
grain, many do so, with the result that the recrystallised grain size is about twice as large as
the particle spacing.
5. Grain growth limitation. During superplastic forming, grain growth must not occur rapidly.
This is prevented by small particles present in some aluminium alloys, due to additions of Zr,
Mn or Cr. These particles do not dissolve at the solutionising (superplastic forming)
temperature and they are able to impede migration of boundaries after recrystallisation such
that grain growth is very effectively prevented. Alloys that do not contain small additions of
Zr, Cr or Mn can be processed by the method described here, but they are less useful for
superplastic deformation since grain growth is rapid.
The solute used to form the large particles is returned to solid solution during superplastic
deformation, and can therefore be re-used in the form of small particles which strengthen the
alloy after the required shape is attained.
One might imagine that the process could be improved by precipitating the solute as finer
particles with a smaller spacing, thus providing more nucleation sites for recrystallising
grains. However, the special deformation zone characteristics required for nucleation of a
recrystallising grain are only attained for particle sizes larger than about 1 m [22]. Thus for
a given alloy composition, the minimum particle diameter is set by this requirement and the
volume fraction of second phase is set by alloy composition and annealing temperature
particle diameter. The particle spacing is then determined by the relationship between particle
size, volume fraction and spacing. This particle spacing fixes the minimum attainable grain
63

size for the alloy being considered. Thus, using phase diagrams it is possible to estimate the
recrystallised grain size and to identify alloys particularly suited to this process.
Alloys processed by this method attain grain sizes down to about 5 m and thus can be used
for superplastic forming at rates about 2 x 10-4 s-1. While suitable for moderate production
rates, such as that required for aircraft components, this strain rate corresponds to a forming
time on the order of 15 minutes for usual component geometries. An advantage of this
method is that standard alloys are used, and after processing and superplastic forming the
alloys recover the standard set of properties.
3.2 Continuous Recrystallisation of Quasi-Single Phase Alloys

In the late 1970s and early 1980s, a different avenue for recrystallising aluminium alloys
containing particles was invented in Great Britain [23-25]. This process requires nonstandard alloy compositions. In particular, the alloys contain larger Zr additions than are
present in standard alloys. The Zr precipitates in the form of Al3Zr particles that effectively
resist migration of boundaries.
These alloys processed in the following way.
1. Cold or warm rolling. The dislocation substructure that provides the driving force for
recrystallisation is introduced during this step.
2. Superplastic deformation. Such alloys contain mainly low angle boundaries as a result of
rolling, they are therefore not superplastic at the start of deformation. However, by a
mechanism that is still not completely understood, recovery occurs simultaneously with
straining, with the result that the microstructure contains mainly high angle boundaries after
about 50% extension. Subsequent deformation is by the superplastic mechanism. The
mechanism of conversion of the deformation microstructure (mainly low angle boundaries)
into a microstructure containing mainly high angle boundaries, is not a conventional
recrystallisation process involving nucleation and growth of recrystallising grains. It is
thought that the mechanism involves gradual rotation of subgrains driven by the simultaneous
recovery and slow strain rate straining. This mechanism has become known as continuous
recrystallisation.
The grain sizes attainable by continuous recrystallisation are somewhat smaller than attained
by conventional recrystallisation, typically in the range 2 4 m in the alloys developed in
the late 1970s. Thus, the superplastic strain rate is thus near 10-3 s-1, about 5 times higher than
that attainable in alloys processed by conventional recrystallisation. Note that this increase in
strain rate is in accord with the prediction of the model described earlier, grain size reduction
by a factor of 2 increases the superplastic strain rate by about 5 times.
A disadvantage of this process is that the alloy compositions are nonstandard, and in many
applications they require qualification prior to use. In addition, the starting microstructure is
not stable. If deformation does not start soon after the alloy is heated to the superplastic
forming temperature, it will begin to recrystallise to a coarse grain size by the conventional
mechanism, and the subsequent superplastic deformation step will fail.
64

64

After the original development of this alloy/process combination, a wide variety of alloy
compositions and processes were developed that rely on the same basic metallurgical
principle. However, the need for special alloy compositions is a severe limitation and
adoption of these processes in general manufacturing operations has not succeeded to the
degree originally envisioned.
3.3 Duplex Alloys with Nondeformable Particles

The duplex alloys mentioned earlier contain 2 deformable phases. The only alloy of that type
with an appreciable aluminium content is the Zn 22% Al monotectic alloy which undergoes
spinodal decomposition to produce two deformable phases of roughly equal volume fraction
[26]. The superplastic properties of Zn - Al monotectic alloys have been explored from an
academic point of view. The grain size is on the order of 1 m and thus the alloys are readily
deformable. But the deformation temperature is so low that creep at room temperature after
forming is a serious limitation and the alloys have not been adopted for manufacturing
purposes.
Alloy with greater promise contain two phases, but one of the phases is not deformable.
Microstructures suitable for superplastic forming are generated in such alloys in the following
way.
1. Forming the particle dispersion. A process is used to introduce volume fractions of about
20% nondeformable particles with a spacing of about the desired grain size. Alloys of two
types are known. Metal matrix composites with an aluminium matrix and ceramic
reinforcement particles introduced by powder metallurgy [27,28], mechanical alloying [29,30]
or special casting processes. The ceramic particles must be small and uniformly distributed in
such composite materials if superplasticity is to be attained. Alternatively, alloys of eutectic
composition are cast, and the lamellar- or rod-like eutectic phase is broken up and
redistributed during subsequent processing [31-33]. Al-Cu, Al-Ni and Al-Ca-Zn alloys were
specially developed for this processing method.
Note that alloy requirements are rather severe. The volume fraction of the nondeformable
phase cannot approach 50%, as in true duplex alloys, but it must be high enough so that a
particle occupies most grain corners. This effectively prevents grain growth. If the volume
fraction of particles is lower, grain growth occurs up to the grain size at which most grain
corners are occupied by a particle. Thus, a balance must be achieved between volume
fraction, size and spacing of the nondeforming particles.
2. Hot deformation and recrystallisation. This creates a microstructure containing high angle
grain boundaries in which grains have a nondeforming insoluble particle at most grain
corners. Grain sizes vary, but alloys having grain sizes in the neighbourhood of 1 m are
attainable.
As Fig. 3 indicates, the superplastic strain rate can be higher than 10-1 s-1 for such alloys. In
early observations of deformation of such alloys, the strain rate was so much higher than
expected for other superplastic alloys, it was thought that a special deformation mechanism
operated.
Superplastic deformation of these alloys is thus called high-strain-rate
superplasticity [34,35]. Fig. 3 shows that the standard deformation mechanism can readily
65

account for the observations. However, the appeal of a special mechanism has proved so
great that bringing the high-strain-rate superplasticity observations back into perspective in
the standard model has proven difficult. Despite the high strain rates and short component
forming times attainable, duplex aluminium alloys with nondeforming particles have not been
adopted for manufacturing because the alloys
are so highly developed for the specific
purpose of high superplastic forming rate that
they lack balanced service properties. Thus,
Ph + P
these alloys have remained largely in the
research domain.
Ph

Ph + P

Ph

Ph

Fig. 5 A schematic of the gas pressure


forming process being conducted with
superposed hydrostatic pressure. Ph is
the hydrostatic pressure preventing
cavitation, P is the pressure differential
that forms the sheet.

66

3.4 Cavitation

Superplastic
deformation
of
many
engineering alloys, including aluminium
alloys, causes formation of microscopic
cavities (voids) [5,6]. The volume fraction of
cavities can approach 10 volume % after large
strains. In the early 1980s, it was been
discovered that imposing a sufficient
hydrostatic pressure during superplastic flow
effectively prevents cavitation [36-38]. In
tensile testing experiments, the hydrostatic
pressure is created by gas pressure in a
chamber surrounding the sample during
straining. The hydrostatic pressure required
to avoid appearance of cavities during
straining depends on the strain state (i.e.,
uniaxial tension, biaxial tension, plane strain
tension, etc.), but flow under conditions of
negative maximum principal stress is
sufficient to avoid cavitation in all situations
[36]. The hydrostatic pressure that prevents
cavitation is readily implemented in a bulge
forming experiment by imposing a back
pressure during the forming operation. Fig. 5
shows that the hydrostatic pressure (Ph)
suppresses cavitation while the pressure
differential (P)allows forming to proceed.
Practical considerations are the main
limitation to use of hydrostatic pressure for
suppressing cavitation during superplastic
flow. In particular, for gas pressure forming,
the clamping force used to hold the split die
together must be sufficient to resist the gas
pressure. For small components this is not a
significant limitation, but for forming of large
66

sheets it can be an important consideration.


5. Applications

Gas pressure forming of aluminium sheet components with complex shapes could be
attractive for a wide variety of applications in the aerospace, automotive and medical device
industries. The advantage for aerospace and transport sectors is the reduction of material loss
and time required for machining components from thick plates. In advanced aerospace
applications, 70 to 90% of the material purchased is machined away to create the complex
geometry that satisfies the mechanical requirement with the minimum possible component
weight. Gas pressure forming enables redesigned components to be fabricated with little or
no loss of material to machining. Various demonstration components have been fabricated
for aerospace and automotive applications.
In medical applications, the mechanical requirements are less severe but the need for
sterilisation is important. In components assembled from multiple pieces, overlapping or butt
joints are locations where infectious agents may persist after cleaning. The ability to fabricate
components with complex geometries as single pieces with no joints is an advantage in such
situations. Less obvious applications with similar requirements are public telephone boxes
and slot machine covers where a joint represents an opportunity to pry open the cover and
steal the contents.
Despite the apparent advantages of superplastically formed aluminium alloy components,
large scale production of such components has yet to be realised. The question of why
superplastic forming of aluminium alloys is not widespread as a manufacturing technology is
perhaps as intriguing as any of the scientific or technological questions addressed in this
article. The technology base is in place, the trade-offs between forming speed and postforming properties are approximately known, and a workable solution to the problem of
cavitation has been developed. One possible explanation involves the near simultaneous
development of large-scale manufacturing of carbon fibre composite materials and
superplastic forming technology for aluminium alloys. The attractive service properties of
carbon fibre components for aerospace applications captured interest in the aerospace
community, possibly displacing superplastic forming of aluminium alloys for aerospace
manufacturing. In other domains, the economic benefit was apparently insufficient to drive
large scale production of superplastically formable aluminium alloys. Lacking a ready supply
of material and an installed base of forming presses, the economic driving pressures favour
continuing use of machining as the main manufacturing method for aluminium alloy
components having complex geometry. It is possible that the situation could be reversed by
coupling superplastic forming with another emerging technology such as friction stir welding,
or if energy costs raise the price of aluminium and favour components manufactured by net
shape processes.

67

6. Conclusion

Superplastic forming of aluminium alloy components is a viable manufacturing technology.


The science and technology information base for material processing and forming is
developed, the trade-offs between forming speed and post-forming properties are
approximately known, and a workable solution to the problem of cavitation has been
developed. Large scale manufacturing of aluminium components by superplastic forming is
clearly a technology for the future. The question is: will it be a technology for the future
forever.
7. Acknowledgements

Support for preparing this article was provided by the Danish National Research Foundation
through the Center for Fundamental Research: Metal Structures in Four Dimensions.
8. References

J.W. Edington, K.N. Melton and C.P. Cutler, Progress in Materials Science, 1976, 21, 61.
J.W. Edington, Metallurgical Trans., 1982, 13A, 703.
A.K. Ghosh and R. Raj, Acta Met., 1981, 29, 607.
N. Ridley, Materials Science and Technology, 1990, 6, 1145.
M.J. Stowell, in Superplastic Forming of Structural Alloys, N.E. Paton and C.H.
Hamilton, The Metallurgical Society of AIME, Warrendale, PA, 1982, p. 321.
6. N. Ridley and J. Pilling, in Superplasticit, B. Baudelet and M. Sury (eds.) CNRS, Paris,
1985, p. 8.1.
7. A. Considre, Ann. ponts et chausses, 1885, 9, 574.
8. W.E. Hosford and R.M. Caddell, Metal Forming: Mechanics and Metallurgy, PrenticeHall, Englewood Cliffs, 1983, p. 84.
9. T. Langdon, Metallurgical Trans., 1982, 13A, 689.
10. R.C. Gifkins, in Superplastic Forming of Structural Alloys, N.E. Paton and C.H.
Hamilton, The Metallurgical Society of AIME, Warrendale, PA, 1982, p. 3.
11. M.F. Ashby and R.A. Verrall, Acta Met., 1973, 21, 149.
12. J.A. Wert, in Mechanical Properties and Phase Transformations in Engineering
Materials, S.D. Antolovich, R.O. Ritchie and W.W. Gerberich (eds), The Metallurgical
Society of AIME, Warrendale, 1986, p. 307.
13. C.H. Hamilton, A.K. Ghosh, and J.A. Wert, Metals Forum, 1985, 8, 172.
14. J.A. Wert, in Superplastic Forming of Structural Alloys, N.E. Paton and C.H. Hamilton
(eds), The Metallurgical Society of AIME, Warrendale, PA, 1982, pp. 69.
15. J.A. Wert, Journal of Metals, 1982, 34 (9), 35.
16. M. Biscondi and C. Goux, Mem. Sci. Revue Metallurgie, 1968, 65, 167.
17. P. Lagarde and M. Biscondi, Canad. Metall. Q., 1974, 13, 167.
18. J.A. Wert, N.E. Paton, C.H. Hamilton and M.W. Mahoney, Metallurgical Transactions,
1981, 12A, 1267.
19. C.C. Bampton, J.A. Wert and M.W. Mahoney, Metallurgical Transactions, 1982, 13A,
193.
20. J.A. Wert, in Microstructural Control in Aluminum Alloys, E.H. Chia and H.J. McQueen
(eds), The Metallurgical Society of AIME, Warrendale, PA, 1986, p. 67.
1.
2.
3.
4.
5.

68

68

21. F.J. Humphreys and M. Hatherly, Recrystallization and Related Annealing Phenomena,
Pergamon, Oxford, 1995, p. 269.
22. F.J. Humphreys, Acta Met., 1977, 25, 1323.
23. B.M. Watts, M.J. Stowell, B.L. Bakie and D.G.E. Owen, Metal Science, 1976, 10, 189.
24. R.H. Bricknell and J.W. Edington, Met. Trans., 1979, 10A, 1257.
25. H. Gudmundsson, D.D. Brooks and J.A. Wert, Acta Met. Mater., 1991, 39, 19.
26. H. Ishikawa, F.A. Mohamed and T.G. Langdon, Phil. Mag., 1975, 32, 1443.
27. D.J. Lloyd, Int. Mater. Rev., 1994, 39, 1.
28. R.S. Mishra, T.R. Bieler and A.K. Mukherjee, Acta Met. Mater., 1995, 43, 877
29. T.R. Bieler and A.K. Mukherjee, Materials Transactions JIM, 1991, 32, 1149.
30. G. Piatti, G. Pellegrini and R. Trippodo, J. Mat. Sci., 1976, 11, 186.
31. R.L. Morris, in Proc. Fourth Int. Conf. Strength of Metals and Alloys, Nancy, France, 30
Aug-3 Sep. 1976, p. 649.
32. N.J. Grant and G. Rai, Met. Trans., 1975, 6A, 385.
33. T. Hasegawa, T. Yasuno, T. Nagai and T. Takahashi T, Acta Mater., 1998, 46, 6001.
34. H. Hosokawa and K. Higashi, Mat. Sci. Res. International, 2000, 6, 153.
35. R. Grimes, R.J. Dashwood and H.M. Flower, Mat. Sci. Forum, 2001, 357, 357.
36. C.C. Bampton and R. Raj, Acta Met. Mater., 1982, 30, 2043.
37. C.C. Bampton, M.W. Mahoney, C.H. Hamilton, A.K. Ghosh and R. Raj, Met. Trans.,
1983, 14A, 1583.
38. J. Pilling and N. Ridley, Res Mech., 1988, 23, 31.

69

70

70

Das könnte Ihnen auch gefallen