review: achieving superplasticity in metals processed by high-pressure torsion

10
ULTRAFINEGRAINED MATERIALS Review: achieving superplasticity in metals processed by high-pressure torsion Megumi Kawasaki Terence G. Langdon Received: 24 January 2014 / Accepted: 26 March 2014 Ó Springer Science+Business Media New York 2014 Abstract It is now well established that processing by equal-channel angular pressing (ECAP) leads to grain refinement and produces materials having the potential for exhibiting extensive superplastic flow at elevated tempera- tures. High-pressure torsion (HPT) is also an effective pro- cedure for refining the grain sizes of polycrystalline metals to the submicrometer or even the nanometer level, and recent results show that this processing method also gives materials that exhibit excellent superplastic characteristics. This report examines the various publications describing superplasticity in metallic alloys processed by HPT. A comprehensive tabulation is presented listing all of the results to date showing true superplastic elongations of at least 400 % after processing by HPT. Examples of superplastic elongations are described for tensile tests conducted using specimens cut from either disk or ring samples. An analysis shows that the flow behavior of various Al and Mg alloys is in good agreement with the predicted flow behavior for conventional superplastic materials. Introduction When a polycrystalline metal is pulled in a laboratory in tensile testing, it generally breaks at a relatively low strain. However, under some testing conditions at elevated tem- peratures, it is possible to achieve exceptionally high elongations prior to failure. Superplasticity is defined for- mally as the ability of a material to exhibit a tensile elongation of at least 400 % before breaking, where this is accompanied by a measured strain rate sensitivity, m (= q ln r/q ln _ e), of *0.5, where r is the flow stress and _ e is the strain rate [1]. The ability of some metallic alloys to pull out to very high elongations forms the basis for the superplastic forming industry where sheet metals are formed into complex shapes and curved parts for use in applications ranging from aerospace to automotive and architectural [2]. It is now well established that there are two basic requirements for achieving superplastic elongations in polycrystalline materials [3]. First, the grain size must be very small and, typically, this means a value \ 10 lm. Second, since superplastic flow is a diffusion-controlled process, the testing temperature must be relatively high and generally at or above *0.5 T m , where T m is the absolute melting temperature of the material. Since grain growth occurs at elevated temperatures in pure metals and simple solid solution alloys, superplastic metals are usually two- phase alloys or they contain a fine dispersion of a second phase to act as a grain refiner. Detailed analysis has established that the physical flow process characterizing superplasticity is grain boundary sliding in which the individual grains of a polycrystalline matrix flow over each other in response to the applied stress [4]. In practice, however, sliding cannot occur in isolation without opening up cavities in the material, and M. Kawasaki (&) Division of Materials Science and Engineering, Hanyang University, Seoul 133-791, South Korea e-mail: [email protected] M. Kawasaki T. G. Langdon Departments of Aerospace & Mechanical Engineering and Materials Science, University of Southern California, Los Angeles, CA 90089-1453, USA T. G. Langdon e-mail: [email protected] T. G. Langdon Materials Research Group, Faculty of Engineering and the Environment, University of Southampton, Southampton SO17 1BJ, UK 123 J Mater Sci DOI 10.1007/s10853-014-8204-5

Upload: terence-g

Post on 24-Jan-2017

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Review: achieving superplasticity in metals processed by high-pressure torsion

ULTRAFINEGRAINED MATERIALS

Review: achieving superplasticity in metals processedby high-pressure torsion

Megumi Kawasaki • Terence G. Langdon

Received: 24 January 2014 / Accepted: 26 March 2014

� Springer Science+Business Media New York 2014

Abstract It is now well established that processing by

equal-channel angular pressing (ECAP) leads to grain

refinement and produces materials having the potential for

exhibiting extensive superplastic flow at elevated tempera-

tures. High-pressure torsion (HPT) is also an effective pro-

cedure for refining the grain sizes of polycrystalline metals to

the submicrometer or even the nanometer level, and recent

results show that this processing method also gives materials

that exhibit excellent superplastic characteristics. This report

examines the various publications describing superplasticity

in metallic alloys processed by HPT. A comprehensive

tabulation is presented listing all of the results to date

showing true superplastic elongations of at least 400 % after

processing by HPT. Examples of superplastic elongations

are described for tensile tests conducted using specimens cut

from either disk or ring samples. An analysis shows that the

flow behavior of various Al and Mg alloys is in good

agreement with the predicted flow behavior for conventional

superplastic materials.

Introduction

When a polycrystalline metal is pulled in a laboratory in

tensile testing, it generally breaks at a relatively low strain.

However, under some testing conditions at elevated tem-

peratures, it is possible to achieve exceptionally high

elongations prior to failure. Superplasticity is defined for-

mally as the ability of a material to exhibit a tensile

elongation of at least 400 % before breaking, where this is

accompanied by a measured strain rate sensitivity, m (= qln r/q ln _e), of *0.5, where r is the flow stress and _e is the

strain rate [1]. The ability of some metallic alloys to pull

out to very high elongations forms the basis for the

superplastic forming industry where sheet metals are

formed into complex shapes and curved parts for use in

applications ranging from aerospace to automotive and

architectural [2].

It is now well established that there are two basic

requirements for achieving superplastic elongations in

polycrystalline materials [3]. First, the grain size must be

very small and, typically, this means a value \10 lm.

Second, since superplastic flow is a diffusion-controlled

process, the testing temperature must be relatively high and

generally at or above *0.5 Tm, where Tm is the absolute

melting temperature of the material. Since grain growth

occurs at elevated temperatures in pure metals and simple

solid solution alloys, superplastic metals are usually two-

phase alloys or they contain a fine dispersion of a second

phase to act as a grain refiner.

Detailed analysis has established that the physical flow

process characterizing superplasticity is grain boundary

sliding in which the individual grains of a polycrystalline

matrix flow over each other in response to the applied

stress [4]. In practice, however, sliding cannot occur in

isolation without opening up cavities in the material, and

M. Kawasaki (&)

Division of Materials Science and Engineering, Hanyang

University, Seoul 133-791, South Korea

e-mail: [email protected]

M. Kawasaki � T. G. Langdon

Departments of Aerospace & Mechanical Engineering and

Materials Science, University of Southern California,

Los Angeles, CA 90089-1453, USA

T. G. Langdon

e-mail: [email protected]

T. G. Langdon

Materials Research Group, Faculty of Engineering and the

Environment, University of Southampton,

Southampton SO17 1BJ, UK

123

J Mater Sci

DOI 10.1007/s10853-014-8204-5

Page 2: Review: achieving superplasticity in metals processed by high-pressure torsion

this means that the sliding is accommodated by slip within

the adjacent grains so that ultimately the rate-controlling

step is the climb of these intragranular dislocations into the

adjacent grain boundaries. Direct experimental evidence is

available to confirm the movement of intragranular dislo-

cations during superplastic flow through measurements of

the densities of dislocations trapped in coherent twin

boundaries in a Cu-based alloy [5], from measurements of

the intragranular strains in a Pb–62 % Sn eutectic alloy [6]

and through the use of a Zn–22 % Al eutectoid alloy

containing a nanometer-scale dispersion of particles [7].

A theoretical model was developed for superplastic flow

based on grain boundary sliding accommodated by intra-

granular slip [8]. This model predicts a steady-state strain

rate, _e, which is given by a relationship of the form

_e ¼ ADgbGb

kT

b

d

� �2 rG

� �2

; ð1Þ

where A is a dimensionless constant having a value of

*10, Dgb is the coefficient for grain boundary diffusion,

G is the shear modulus, b is the Burgers vector, k is

Boltzmann’s constant, T is the absolute temperature and

d is the grain size.

Conventional superplastic metals are usually produced

by thermo-mechanical processing and have grain sizes

within the range of *3–5 lm. However, it is now known

that the processing of metals through the application of

severe plastic deformation (SPD) leads to exceptional

grain refinement so that the grain sizes typically lie within

the submicrometer or even the nanometer range. It is

reasonable to anticipate that these SPD materials will

exhibit excellent superplastic properties. In practice, sev-

eral SPD processing methods are now available [9], but

the two most important techniques, and the procedures

most widely used, are equal-channel angular pressing

(ECAP) [10] and high-pressure torsion (HPT) [11]. A

comprehensive review published in 2007 tabulated the

large numbers of results available at that time describing

superplasticity in metals processed by ECAP [12]. More

recently, numerous publications have also become avail-

able describing the occurrence of superplastic flow in

metals processed by HPT. Thus, the objective of this

report is to examine the fundamental characteristics of

superplastic flow in materials processed by HPT and to

provide a comprehensive tabulation of the results avail-

able to date. Accordingly, the following section summa-

rizes the historical development of superplasticity in SPD

processing, the next sections describe some of the prin-

ciple results obtained using disk and ring samples, and the

final section tabulates the data and examines the flow

process in terms of the fundamental mechanism for

superplasticity in conventional materials.

Historical developments for superplasticity using SPD

techniques

An early report, published over 25 years ago, sets the scene

for establishing SPD processing as a major tool in producing

materials that exhibit excellent superplastic properties. In a

paper entitled ‘‘Low-temperature superplasticity of metallic

materials,’’ Valiev et al. [13] used an Al–4 % Cu–0.5 % Zr

alloy, processed the alloy by HPT and then obtained an

elongation of 250 % at the very low temperature of 493 K.

Although the latter elongation is not within the current def-

inition of superplasticity, nevertheless the result raised

considerable interest because of the high elongation that was

achieved at an unusually low temperature [14]. Thus, this

early result prompted additional experiments on samples

processed by ECAP [15] and, when two review articles were

published describing extensive results in SPD processing

[16, 17], interest in the field developed further so that pro-

cessing by SPD methods became a regular feature of many

materials laboratories around the world.

As demonstrated by Eq. (1), the strain rate varies

inversely with the grain size raised to a power of two in

conventional superplasticity and it was first noted in 1996

that provided the ultrafine grain sizes were reasonably

stable at elevated temperatures, materials produced by SPD

techniques may exhibit excellent superplasticity with the

high elongations occurring at very rapid strain rates [18]. In

the following year, in 1997, superplasticity was achieved in

two commercial aluminum alloys with elongations up to

[1000 % at a strain rate of 10-2 s-1 [19]. This strain rate

is generally defined as the point of transition to high strain

rate superplasticity [20].

For many years, the development of superplasticity in

SPD materials was devoted almost exclusively to metals

processed by ECAP. The reason is due to the relatively

large samples that are easily produced using the ECAP

method. Typically, ECAP samples are in the form of bars

or rods with dimensions of the order of 10 9 10 mm2 and

lengths of at least 65 mm. These large sizes mean that it is

a simple procedure to machine tensile specimens from the

as-processed samples. Processing by HPT is advantageous

because it produces materials generally having smaller

grain sizes and higher fractions of high-angle boundaries

than in ECAP [21–23]. Nevertheless, the processing gen-

erally uses thin disks and these samples are not directly

amenable to conventional tensile testing.

An additional problem is that the equivalent von Mises

strain imposed on an HPT disk, eeq, is given by the rela-

tionship [24–26]

eeq ¼2pNr

hffiffiffi3p ; ð2Þ

J Mater Sci

123

Page 3: Review: achieving superplasticity in metals processed by high-pressure torsion

where r and h are the radius and height (or thickness) of the

disk, respectively, and N is the number of HPT turns. It

follows from Eq. (2) that the imposed strain varies from

zero at the disk center to a maximum at the periphery. In

practice, however, extensive experiments show that there is

a gradual evolution toward a reasonable homogeneity after

a large number of turns [27, 28]. Because of the problems

associated with inhomogeneities in the centers of the disks,

it has become a standard practice to machine two tensile

specimens from each HPT disk using off-center positions,

as illustrated schematically in Fig. 1 where the central

point of each specimen is located at 2 mm from the center

of the disk [29]. It is then possible to pull these specimens

in tension at an elevated temperature in order to evaluate

the superplastic properties after HPT processing. Numer-

ous results are now available confirming the feasibility of

using miniature specimens cut from HPT disks to measure

the mechanical properties of materials after processing by

HPT [30, 31].

Characteristics of superplasticity after HPT processing

of disks

Equation (1) delineates the strain which is imposed on the

disk when processing by HPT. In practice, the strain

increases with additional torsional revolutions, but ulti-

mately a stable microstructure is attained, usually after a

relatively small number of turns, and thereafter there is

little or no additional grain refinement with increasing

straining. Figure 2 shows the elongations achieved in a Zn–

22 % Al eutectoid alloy after processing by HPT at room

temperature under an applied pressure of 6.0 GPa with a

rotational speed of 1 rpm [32]. The tensile tests were

conducted at 473 K using an initial strain rate of

1.0 9 10-1 s-1, the upper specimen is untested, and the

other three specimens were processed through 1, 3 and 5

turns, respectively. Inspection shows that all specimens in

Fig. 2 are superplastic but the maximum elongation of

1800 % was attained after a total of N = 5 turns.

There is also a significant dependence on the testing

strain rate as shown in Fig. 3 where specimens of the Zn–

22 % Al alloy were processed through 5 turns in HPT and

then pulled to failure at initial strain rates from 1.0 9 10-3

to 1.0 s-1 [32]. These results show that, as in conventional

superplastic materials not processed by SPD techniques,

superplasticity occurs over a limited range of strain rates,

and there are decreases in the elongations to failure at both

faster and slower strain rates [33]. The behavior visible in

Fig. 3 corresponds to the three regions of flow in super-

plasticity where region II is true superplasticity with high

Fig. 1 Procedure for cutting miniature tensile specimens from off-

center positions in disks processed by HPT [29]

Fig. 2 The variation of the elongations to failure with the numbers of

revolutions applied in torsional straining for a Zn–22 % Al alloy

tested at a strain rate of 1.0 9 10-1 s-1 at 473 K [32]

Fig. 3 The variation of the elongations to failure with the imposed

strain rates for a Zn–22 % Al alloy tested at a temperature of 473 K

[32]

J Mater Sci

123

Page 4: Review: achieving superplasticity in metals processed by high-pressure torsion

elongations at intermediate strain rates, region III corre-

sponds to a transition to conventional dislocation creep at

faster strain rates, and region I represents a regime of flow

which is impurity dependent at the slowest strain rates [34].

It is important to note that the maximum elongation of

1800 % appears to represent the highest elongation

achieved to date in any material processed by HPT. Fur-

thermore, this specimen exhibits the fundamental charac-

teristics of true superplastic flow including uniform

deformation without any evidence for necking within the

gauge length and with failure occurring by pulling down to

a point [35].

An important question concerns the relative values of

the elongations that may be achieved in the same alloy

when processing either by HPT or ECAP. In principle,

HPT produces a smaller grain size, and this should displace

the superplastic region II to faster strain rates and at the

same time produce larger elongations to failure. This

comparison is shown in Fig. 4 using HPT data for the Zn–

22 % Al alloy after processing through 5 turns [32] and

ECAP data for the same alloy processed through 8 passes

at 473 K [36]. For the ECAP processing, the billets were

pressed through a die with a channel angle of 90� using

processing route BC, in which the ECAP billet is rotated in

the same sense by 90� between each pass [37]. The mea-

sured grain sizes in the Zn–Al alloy were *900 nm after

ECAP and *350 nm after HPT so that the HPT specimens

are anticipated to show a peak ductility at a faster strain

rate and with elongations which are larger than after

ECAP. Inspection of Fig. 4 shows, however, that the HPT

datum points are correctly displaced to faster rates but the

maximum elongation in HPT is lower than in ECAP.

The reason for this apparent discrepancy lies in the

dimensions of the two sets of tensile specimens. For ECAP

the initial cross-sectional areas of the specimens were

3 9 2 = 6 mm2, whereas in HPT the specimens had a

width of 1 mm from Fig. 1 and a thickness of *0.8 mm so

that the cross-sectional areas were *0.8 mm2. Various

experiments have shown that the elongations achieved in

tensile testing are dependent upon the cross-sectional areas,

and specifically the measured elongations increase when

using samples having larger cross-sectional areas [38–40].

For example, tensile tests at room temperature on ultrafine-

grained Cu samples produced by ECAP processing showed

that the overall strains increased significantly as the spec-

imen thickness was increased from 250 lm to 1.0 mm

[41]. In addition, a later study used miniature tensile

specimens of various sizes and geometries and compared

the experimental data with predictions derived from finite

element modeling [42]. The general conclusions from an

examination of these earlier experiments are that the lower

elongations recorded for the superplastic HPT specimens in

Fig. 4 are a direct consequence of the very small cross-

sectional areas available in these specimens prior to testing.

Figure 5 shows another example of superplastic elon-

gations after HPT where specimens of an Al–33 % Cu

eutectic alloy were tested in tension at 723 K: a detailed

description of these results was given earlier [43]. These

results are similar to the Zn–Al alloy in Fig. 3 except that

the maximum elongation in this alloy occurs at a slower

initial strain rate and there is no evidence for high strain

rate superplasticity.

The testing temperature also affects the measured

elongations to failure and this can be seen in Fig. 6 where

all specimens of the AZ61 magnesium alloy were pulled at

Fig. 4 A comparison of superplasticity in a Zn–22 % Al alloy

processed by either HPT [32] or ECAP [36] and then tested at a

temperature of 473 K

Fig. 5 The variation of the elongations to failure with the imposed

strain rates for an Al–33 % Cu alloy tested at a temperature of 723 K

[43]

J Mater Sci

123

Page 5: Review: achieving superplasticity in metals processed by high-pressure torsion

the same initial strain rate of 3.3 9 10-3 s-1 over a range

of temperatures from 373 to 523 K [44]. The results show

an optimum elongation of 620 % at 473 K and smaller

elongations at both lower and higher testing temperatures.

A careful microstructural analysis showed that the decrease

in elongation at 523 K was due to grain growth at this

higher temperature.

Characteristics of superplasticity after HPT processing

of rings

The preceding results were obtained for HPT specimens

cut from conventional thin disks using the specimen con-

figurations shown in Fig. 1. However, it is also possible to

conduct HPT processing using ring samples. This approach

was adopted many years ago in experiments using very low

pressures [45, 46] but more recently ring samples were

used to evaluate the potential for achieving superplastic

elongations in an Al–3 % Mg–0.2 % Sc alloy [47].

The tests were conducted using rings with widths of

3 mm and outer diameters of 40 mm. After processing

through 1 turn under a pressure of 1.25 GPa, the rings were

used to provide both disks having diameters of 3 mm for

inspection using transmission electron microscopy (TEM)

and tensile specimens having gauge widths and lengths of

1 mm where these dimensions are identical to the disk

specimens as shown in Fig. 1: the procedure for cutting

TEM and tensile samples is shown schematically in Fig. 7

[47]. The tensile specimens were pulled to failure at a

temperature of 573 K at two different strain rates and the

result is shown in Fig. 8 where it is apparent that these

miniature specimens exhibit excellent superplastic prop-

erties including an elongation of[1000 % in the regime of

high strain rate superplasticity [47]. It follows, therefore,

that superplastic specimens may be achieved using both

disk and ring configurations.

The flow process in superplastic HPT specimens

The preceding sections on disk and ring samples show that

excellent superplastic properties can be achieved in

metallic alloys processed by HPT. Several superplastic

results are now available for HPT specimens, and a com-

prehensive summary is given in Table 1. In this Table, the

HPT processing conditions are listed in columns 2–4 as the

maximum numbers of turns, the applied pressure and the

processing temperature, column 5 gives the grain sizes

achieved by HPT, details of the superplastic tests are given

in columns 6–8 including the testing temperature, imposed

strain rate and the maximum reported elongation to failure,

Fig. 6 The variation of the elongations to failure with the testing

temperature for an AZ61 magnesium alloy tested at a strain rate of

3.3 9 10-3 s-1 [44]

Fig. 7 Procedure for cutting TEM samples and miniature tensile

specimens from rings processed by HPT [47]

Fig. 8 The variation of the elongations to failure with the imposed

strain rates for an Al–3 % Mg–0.2 % Sc alloy processed as rings by

HPT and tested at a temperature of 573 K [47]

J Mater Sci

123

Page 6: Review: achieving superplasticity in metals processed by high-pressure torsion

Table 1 Reports of superplasticity in ultrafine-grained materials

Alloy or

composition

(wt%)

HPT Grain

size (nm)

Superplasticity Reference

Number

of turns

Pressure

(GPa)

Temperature

(K)

Testing

temperature

(K)

Strain rate

(s-1)

Maximum

elongation (%)

Al1420a 3n 1.2 RT 100 523 1.0 9 10-2 440 Mishra et al. [48]

573 1.0 9 10-1 775

Al1421b ? 6 293 1080 623 1.0 9 10-2 537 Islamgaliev et al. [49]

393 700–1000 508

573 1160 489

613 930 664

643 570 1030

Al1570c 5 6 RT 120 573 1.0 9 10-1 580 Perevezentsev et al.

[50]623 1.8 9 10-1 800

673 1.0 9 10-2 1460

673 1.0 600

Al2024d,j 5 6 293 140 673 1.7 9 10-3 570 Dobatkin et al. [51]

673 290 450

Al2024d,k 293 160 410

673 300 440

Al7034e 5 4 RT (293) *600 673 1.0 9 10-2 550 Xu et al. [52]

703 1.0 9 10-2 750

473 *800 673 1.0 9 10-3 550

703 1.0 9 10-2 650

Al7075f 5 6 RT *250 623 1.0 9 10-4 *700 Sabbaghianrad et al.

[53]10 1.0 9 10-3 *570

1.0 9 10-4 *680

Al–33Cu 5 6 RT *3000 723 3.3 9 10-5 *1250 Kawasaki et al. [43]

10 1.0 9 10-4

Al–3 Mg–0.2Sc 5 6 RT *150 673 3.3 9 10-2 500 Sakai et al. [54]

Al–3 Mg–0.2Sc 2 1 RT *130 573 3.3 9 10-3 *1600q Horita and Langdon

[55]

Al–3 Mg–0.2Sc 1 1.25 623 *220 573 3.3 9 10-2 1030r Harai et al. [47]

3.3 9 10-3 1510r

AZ61g 5 3 423 230 473 3.3 9 10-3 620 Harai et al. [44]

ZK60h 5 2 RT (296) 900–1000 473 1.0 9 10-3 415 Torbati-Sarraf and

Langdon [56]523 477

573 508

Mg–9All 5 3 RT (298) 210 473 1.0 9 10-3 600 Kai et al. [57]

423 330 5.0 9 10-3 550

1.0 9 10-3 620

5.0 9 10-4 810

Mg–9Alm 423 370 1.0 9 10-3 590

Mg–10Gd 5 6 RT *100 673 1.0 9 10-2 470 Kulyasova et al. [58]

1.0 9 10-3 580

Ni3Ali 7o 7 RT *50 998 1.0 9 10-3 560 Mishra et al. [59],

Valiev et al. [60]

Ti–6Al–4 V 5p 2.2 RT (293) 100-200 923 1.0 9 10-4 530 Sergueeva et al. [61]

5 1.0 9 10-3 575

J Mater Sci

123

Page 7: Review: achieving superplasticity in metals processed by high-pressure torsion

and finally the reference for each report is listed in the last

column. The results summarized in Table 1 cover only

those investigations where the measured elongations to

failure fulfill the requirement of a tensile elongation of at

least 400 % for superplastic flow: reports of tensile elon-

gations less than 400 % are not included. Additional

information is provided after the Table 1 on the composi-

tions of the various alloys and there are also notes on the

HPT processing.

The results documented in this report show that excel-

lent superplastic properties may be achieved in HPT sam-

ples and it is appropriate to consider whether these

specimens exhibit flow properties which are consistent with

the theoretical model for conventional superplasticity [8].

An earlier report examined data for superplastic Al and Mg

alloys after processing by ECAP and showed that for both

sets of materials there was excellent agreement, to within

an order of magnitude of strain rate, with the theoretical

model as given by Eq. (1) [65].

The same approach was adopted in the present analysis,

and Fig. 9 shows the results for aluminum-based alloys

where the temperature and grain size compensated strain

rate, ( _ekT/DgbGb)(d/b)2, is plotted against the normalized

stress, r/G, using data taken from reports on various Al

alloys [43, 47, 48, 50–54]. The experimental points were

plotted using the following values for all aluminum alloys:

Table 1 continued

Alloy or

composition

(wt%)

HPT Grain

size (nm)

Superplasticity Reference

Number

of turns

Pressure

(GPa)

Temperature

(K)

Testing

temperature

(K)

Strain rate

(s-1)

Maximum

elongation (%)

Ti–6Al–4 V 5p 5 RT (293) 100–200 923 1.0 9 10-3 568 Sergueeva et al. [62]

1.0 9 10-4 513

998 1.0 9 10-2 504

1.0 9 10-3 676

Zn–22Al 5 6 RT *600 473 1.0 1170 Kawasaki and

Langdon [32, 63]1.0 9 10-1 1800

1.0 9 10-2 1370

1.0 9 10-3 900

Zn–22Al 1 6 RT *770 473 1.0 9 10-2 1800 Cho et al. [64]

3 *650 1.0 9 10-1 1680

5 *600 1.0 9 10-1 1800

a Al1420: Al–5 Mg–2Li–0.1Zrb Al1421: Al–5.5 Mg–2.2Li–0.12Zr–0.2Scc Al1570: Al–5.6 Mg–0.4Mn-0.32Scd Al2024: Al–4.4Cu–1.5 Mg–0.6Mne Al7034: Al–11.5Zn–2.5 Mg–0.9Cu–0.2Zrf Al7075: Al–5.6Zn–2.5 Mg–1.6Cug AZ61: Mg–6.4Al–0.74Zn–0.35Mn-0.0012Ni-0.001Cu–0.001Fe-0.015Sih ZK60: Mg–5.5Zn–0.5Zri Ni3Al: Ni-8.5A1-7.8Cr-0.6Zr–0.02Bj The Al2024 was processed by HPT and annealed for 10 min at 673 K followed by air coolingk The Al2024 was processed by HPT and solution treated for 10 min at 768 K followed by quenching into waterl The Mg–9Al was cast prior to HPTm The Mg–9Al was cast and subsequently extruded prior to HPTn The total number of turns was calculated from the true logarithmic strain of 6 at a distance of 3 mm from the center of a disk with a diameter of

20 mm and a thickness of *0.5 mmo The total number of turns was calculated from the true logarithmic strain of 7 at a distance of 3 mm from the center of a disk with a diameter of

12 mm and a thickness of *0.3 mmp The total number of turns was calculated from the true logarithmic strain of 7 at a distance of 3 mm from the center of a disk with a diameter of

15 mm and a thickness of *0.5 mmq Samples were taken from the mid-section of a bulk HPT sampler Specimens were taken from ring HPT samples

J Mater Sci

123

Page 8: Review: achieving superplasticity in metals processed by high-pressure torsion

b = 2.86 9 10-10 m [66], G (MPa) = (3.022 9 104) –

16T where the temperature is in degrees Kelvin [66],

Dgb = Do(gb) exp (-Qgb/RT) where R is the gas constant,

Do(gb) = 1.86 9 10-4 m2 s-1, Qgb = 86 kJ mol-1 [66], a

value for the dimensionless constant in Eq. (1) of A = 10

[8] and converting from the reported mean linear intercept

grain size, �L, to the spatial grain size, d, using the

expression d = 1.74 �L [67]. The prediction for the theo-

retical model is shown in Fig. 9 by the line labeled _esp

having a slope of n = 2 where n is the stress exponent.

Inspection of Fig. 9 shows that all of the experimental

points are mutually consistent and they are generally in

reasonable agreement with the theoretical model except

that, for any imposed strain rate, the measured stress levels

tend to be too high by about one-half order of magnitude on

the stress axis. This discrepancy is due to the use of

exceptionally small tensile samples.

A similar plot is shown in Fig. 10 for magnesium-based

alloys processed by HPT [44, 56–58] where

b = 3.2 9 10-10 m [68], G (MPa) = 1.92 9 104 -

8.6T [68], Do(gb) = 7.8 9 10-3 m2 s-1 [69], and

Qgb = 92 kJ mol-1 [69]. Again, all experimental points

are consistent with the theoretical model but they are dis-

placed to slightly higher stresses which is attributed to the

use of very small tensile specimens.

All of these results confirm that HPT processing, using

either disk or ring samples, is as effective as processing by

ECAP in producing specimens that exhibit excellent

superplasticity when pulled in tension at elevated temper-

atures. As anticipated from the very small grain sizes

attained in HPT, the peak superplastic flow tends to occur

at faster strain rates than when testing ECAP samples but

the total elongations are slightly lower due to the use of

specimens having very small cross-sectional areas.

Summary and conclusions

1. Samples processed by HPT exhibit excellent super-

plastic properties when tested in tension at elevated

temperatures. Several results are now available docu-

menting this behavior for both disks and rings and

these results are tabulated and described.

2. The strain rates inherent in superplastic flow provide a

reasonable match to the anticipated behavior from the

theoretical mechanism for conventional superplasticity.

However, the measured stresses tend to be slightly high

because of the use of very small tensile specimens.

Acknowledgements This work was supported in part by the

National Science Foundation of the United States under Grant No.

DMR-1160966 and in part by the European Research Council under

ERC Grant Agreement No. 267464-SPDMETALS.

References

1. Langdon TG (2009) Seventy-five years of superplasticity: historic

developments and new opportunities. J Mater Sci 44:5998–6010

2. Barnes AJ (2007) Superplastic forming 40 years and still grow-

ing. J Mater Eng Perform 16:440–454

Fig. 9 Temperature and grain size compensated strain rate versus the

normalized stress for various Al alloys processed by HPT and

exhibiting superplastic behavior [43, 47, 48, 50–54]: the solid line

shows the theoretical prediction for superplastic flow in conventional

metals not processed using SPD techniques [8]

Fig. 10 Temperature and grain size compensated strain rate versus

the normalized stress for various Mg alloys processed by HPT and

exhibiting superplastic behavior [44, 56–58]: the solid line shows the

theoretical prediction for superplastic flow in conventional metals not

processed using SPD techniques [8]

J Mater Sci

123

Page 9: Review: achieving superplasticity in metals processed by high-pressure torsion

3. Langdon TG (1982) The mechanical properties of superplastic

materials. Metall Trans A 13A:689–701

4. Langdon TG (1994) An evaluation of the strain contributed by

grain boundary sliding in superplasticity. Mater Sci Eng A

174:225–230

5. Falk LKL, Howell PR, Dunlop GL, Langdon TG (1986) The role

of matrix dislocations in the superplastic deformation of a copper

alloy. Acta Metall 34:1203–1214

6. Valiev RZ, Langdon TG (1993) An investigation of the role of

intragranular dislocation strain in the superplastic Pb–62 % Sn

eutectic alloy. Acta Metall Mater 41:949–954

7. Xun Y, Mohamed FA (2003) Slip-accommodated superplastic

flow in Zn–22 wt% Al. Phil Mag 83:2247–2266

8. Langdon TG (1994) A unified approach to grain boundary sliding

in creep and superplasticity. Acta Metall Mater 42:2437–2443

9. Valiev RZ, Estrin Y, Horita Z, Langdon TG, Zehetbauer MJ, Zhu

YT (2006) Producing bulk ultrafine-grained materials by severe

plastic deformation. JOM 58(4):33–39

10. Valiev RZ, Langdon TG (2006) Principles of equal-channel

angular pressing as a processing tool for grain refinement. Prog

Mater Sci 51:881–981

11. Zhilyaev AP, Langdon TG (2008) Using high-pressure torsion for

metal processing: fundamentals and applications. Prog Mater Sci

53:893–979

12. Kawasaki M, Langdon TG (2007) Principles of superplasticity in

ultrafine-grained materials. J Mater Sci 42:1782–1796

13. Valiev RZ, Kaibyshev OA, Kuznetsov RI, Musalimov RSh,

Tsenev NK (1988) Low-temperature superplasticity of metallic

materials. Dokl Akad Nauk SSSR (Proc USSR Acad Sci)

301:864–866

14. Langdon TG (2007) Ultrafine-grained materials: a personal per-

spective. Int J Mater Res 98:251–254

15. Wang J, Horita Z, Furukawa M, Nemoto M, Tsenev NK, Valiev

RZ, Ma Y, Langdon TG (1993) An investigation of ductility and

microstructural evolution in an Al–3 % Mg alloy with submicron

grain size. J Mater Res 8:2810–2818

16. Valiev RZ, Krasilnikov NA, Tsenev NK (1991) Plastic defor-

mation of alloys with submicron-grained structure. Mater Sci Eng

A 137:35–40

17. Valiev RZ, Korznikov AV, Mulyukov RR (1993) Structure and

properties of ultrafine-grained materials produced by severe

plastic deformation. Mater Sci Eng A 168:141–148

18. Ma Y, Furukawa M, Horita Z, Nemoto M, Valiev RZ, Langdon

TG (1996) Significance of microstructural control for superplastic

deformation and forming. Mater Trans JIM 37:336–339

19. Valiev RZ, Salimonenko DA, Tsenev NK, Berbon PB, Langdon

TG (1997) Observations of high strain rate superplasticity in

commercial aluminum alloys with ultrafine grain sizes. Scripta

Mater 37:1945–1950

20. Higashi K, Mabuchi M, Langdon TG (1996) High-strain-rate

superplasticity in metallic materials and the potential for ceramic

materials. ISIJ Intl 36:1423–1438

21. Zhilyaev AP, Lee S, Nurislamova GV, Valiev RZ, Langdon TG

(2001) Microhardness and microstructural evolution in pure

nickel during high-pressure torsion. Scripta Mater 44:2753–2758

22. Zhilyaev AP, Nurislamova GV, Kim BK, Baro MD, Szpunar JA,

Langdon TG (2003) Experimental parameters influencing grain

refinement and microstructural evolution during high-pressure

torsion. Acta Mater 51:753–765

23. Wongsa-Ngam J, Kawasaki M, Langdon TG (2013) A compari-

son of microstructures and mechanical properties in a Cu–Zr

alloy processed using different SPD techniques. J Mater Sci

48:4653–4660

24. Valiev RZ, Ivanisenko YuV, Rauch EF, Baudelet B (1996)

Structure and deformation behaviour of Armco iron subjected to

severe plastic deformation. Acta Mater 44:4705–4712

25. Wetscher F, Vorhauer A, Stock R, Pippan R (2004) Structural

refinement of low alloyed steels during severe plastic deforma-

tion. Mater Sci Eng A 387–389:809–816

26. Wetscher F, Pippan R, Sturm S, Kauffmann F, Scheu C, Dehm G

(2006) TEM investigations of the structural evolution in a

pearlitic steel deformed by high-pressure torsion. Metall Mater

Trans A 37A:1963–1968

27. Langdon TG (2013) Twenty-five years of ultrafine-grained

materials: achieving exceptional properties through grain refine-

ment. Acta Mater 61:7035–7059

28. Kawasaki M (2014) Different models of hardness evolution in

ultrafine-grained materials processed by high-pressure torsion.

J Mater Sci 49:18–34

29. Loucif A, Figueiredo RB, Kawasaki M, Baudin T, Brisset F,

Chemam R, Langdon TG (2012) Effect of aging on microstruc-

tural development in an Al–Mg–Si alloy processed by high-

pressure torsion. J Mater Sci 47:7815–7820

30. An XH, Lin QY, Wu SD, Zhang ZF, Figueiredo RB, Gao N,

Langdon TG (2011) The influence of stacking fault energy on the

mechanical properties of nanostructured Cu and Cu–Al alloys

processed by high-pressure torsion. Scripta Mater 64:954–957

31. An XH, Wu SD, Zhang ZF, Figueiredo RB, Gao N, Langdon TG

(2012) Enhanced strength-ductility synergy in nanostructured Cu

and Cu–Al alloys processed by high-pressure torsion and sub-

sequent annealing. Scripta Mater 66:227–230

32. Kawasaki M, Langdon TG (2011) Developing superplasticity and

a deformation mechanism map for the Zn–Al eutectoid alloy

processed by high-pressure torsion. Mater Sci Eng A

528:6140–6145

33. Ishikawa H, Mohamed FA, Langdon TG (1975) The influence of

strain rate on ductility in the superplastic Zn–22 % Al eutectoid.

Phil Mag 32:1269–1271

34. Langdon TG (1991) The physics of superplastic deformation.

Mater Sci Eng A 137:1–11

35. Langdon TG (1982) Fracture processes in superplastic flow.

Metal Sci 16:175–183

36. Kawasaki M, Langdon TG (2008) Grain boundary sliding in a

superplastic zinc-aluminum alloy processed using severe plastic

deformation. Mater Trans 49:84–89

37. Furukawa M, Iwahashi Y, Horita Z, Nemoto M, Langdon TG

(1998) The shearing characteristics associated with equal-channel

angular pressing. Mater Sci Eng A 257:328–332

38. Morrison WB (1968) Elongation of superplastic alloys. Trans

TMS-AIME 242:2221–2227

39. Ahmed MMI, Langdon TG (1977) Exceptional ductility in the

superplastic Pb–62 pct Sn eutectic. Metall Trans A

8A:1832–1833

40. Ma Y, Langdon TG (1994) Factors influencing the exceptional

ductility of a superplastic Pb–62 Sn alloy. Metall Mater Trans A

25A:2309–2311

41. Zhao YH, Guo YZ, Wei Q, Dangelewicz AM, Xu C, Zhu YT,

Langdon TG, Zhou YZ, Lavernia EJ (2008) Influence of speci-

mens dimensions on the tensile behavior of ultrafine-grained Cu.

Scripta Mater 59:627–630

42. Zhao YH, Guo YZ, Wei Q, Topping TD, Dangelewicz AM, Zhu

YT, Langdon TG, Lavernia EJ (2009) Influence of specimen

dimensions and strain measurement methods on tensile stress-

strain curves. Mater Sci Eng A 525:68–77

43. Kawasaki M, Foissey J, Langdon TG (2013) Development of

hardness homogeneity and superplastic behavior in an aluminum-

copper eutectic alloy processed by high-pressure torsion. Mater

Sci Eng A 561:118–125

44. Harai Y, Kai M, Kaneko K, Horita Z, Langdon TG (2008)

Microstructural and mechanical characteristics of AZ61 magne-

sium alloy processed by high-pressure torsion. Mater Trans

49:76–83

J Mater Sci

123

Page 10: Review: achieving superplasticity in metals processed by high-pressure torsion

45. Erbel S (1979) Mechanical properties and structure of extremely

strain-hardened copper. Metals Technol 6:482–486

46. Saunders I, Nutting J (1984) Deformation of metals to high

strains using combination of torsion and compression. Metal Sci

18:571–575

47. Harai Y, Edalati K, Horita Z, Langdon TG (2009) Using ring

samples to evaluate the processing characteristics in high-pres-

sure torsion. Acta Mater 57:1147–1153

48. Mishra RS, Valiev RZ, McFadden SX, Islamgaliev RK, Muk-

herjee AK (2001) High-strain-rate superplasticity from nano-

crystalline Al alloy 1420 at low temperatures. Phil Mag A

81:37–48

49. Islamgaliev RK, Yunusova NF, Valiev RV (2006) The influence

of the SPD temperature on superplasticity of aluminium alloys.

Mater Sci Forum 503–504:585–590

50. Perevezentsev VN, Shcherban MYu, Murashkin MYu, Valiev RZ

(2007) High-strain-rate superplasticity of nanocrystalline alumi-

num alloy 1570. Tech Phys Lett 33:648–650

51. Dobatkin SV, Bastarache EN, Sakai G, Fujita T, Horita Z,

Langdon TG (2005) Grain refinement and superplastic flow in an

aluminum alloy processed by high-pressure torsion. Mater Sci

Eng A 408:141–146

52. Xu C, Dobatkin SV, Horita Z, Langdon TG (2009) Superplastic

flow in a nanostructured aluminum alloy produced using high-

pressure torsion. Mater Sci Eng A 500:170–175

53. Sabbaghianrad S, Kawasaki M, Langdon TG (2012) Micro-

structural evolution and the mechanical properties of an alumi-

num alloy processed by high-pressure torsion. J Mater Sci

47:7789–7795

54. Sakai G, Horita Z, Langdon TG (2005) Grain refinement and

superplasticity in an aluminum alloy processed by high-pressure

torsion. Mater Sci Eng A 393:344–351

55. Horita Z, Langdon TG (2008) Achieving exceptional superplas-

ticity in a bulk aluminum alloy processed by high-pressure tor-

sion. Scripta Mater 58:1029–1032

56. Torbati-Sarraf SA, Langdon TG (2014) Mechanical properties of

ZK60 magnesium alloy processed by high-pressure torsion. Adv

Mater Research (in press)

57. Kai M, Horita Z, Langdon TG (2008) Developing grain refine-

ment and superplasticity in a magnesium alloy processed by high-

pressure torsion. Mater Sci Eng A 488:117–124

58. Kulyasova OB, Islamgaliev RK, Kil’mametov AR, Valiev RZ

(2006) Superplastic behavior of magnesium-based Mg–10 wt%

Gd alloy after severe plastic deformation by torsion. Phys Met

Metall 101:585–590

59. Mishra RS, Valiev RZ, McFadden SX, Mukherjee AK (1998)

Tensile superplasticity in a nanocrystalline nickel aluminide.

Mater Sci Eng A 252:174–178

60. Valiev RZ, Song C, McFadden SX, Mukherjee AK, Mishra RS

(2001) TEM/HREM observations of nanostructured superplastic

Ni3Al. Phil Mag A 81:25–36

61. Sergueeva AV, Stolyarov VV, Valiev RZ, Mukherjee AK (2000)

Enhanced superplasticity in a Ti–6Al–4V alloy processed by

severe plastic deformation. Scripta Mater 43:819–824

62. Sergueeva AV, Stolyarov VV, Valiev RZ, Mukherjee AK (2002)

Superplastic behaviour of ultrafine-grained Ti–6A1–4V alloys.

Mater Sci Eng A 323:318–325

63. Kawasaki M, Langdon TG (2014) Microstructure development

and superplasticity in a Zn–22 % Al eutectoid alloy processed by

severe plastic deformation. Mater Sci Forum (in press)

64. Cho T-S, Lee H-J, Ahn B, Kawasaki M, Langdon TG (2014)

Microstructural evolution and mechanical properties in a Zn–Al

eutectoid alloy processed by high-pressure torsion. Acta Mater.

doi:10.1016/j.actamat.2014.03.026

65. Kawasaki M, Balasubramanian N, Langdon TG (2011) Flow

mechanisms in ultrafine-grained metals with an emphasis on

superplasticity. Mater Sci Eng A 528:6624–6629

66. Mohamed FA, Langdon TG (1974) Deformation mechanism

maps based on grain size. Metall Trans 5:2339–2345

67. Mohamed FA, Langdon TG (1975) Creep at low stress levels in

the superplastic Zn 22 % Al eutectoid. Acta Metall 23:117–124

68. Vagarali SS, Langdon TG (1982) Deformation mechanisms in

h.c.p. metals at elevated temperatures—II. Creep behavior of a

Mg–0.8 % Al solid solution alloy. Acta Metall 30:1157–1170

69. Frost HJ, Ashby MF (1982) Deformation-mechanism maps: the

plasticity and creep of metals and ceramics. Pergamon Press, Oxford

J Mater Sci

123