You are on page 1of 50

SEVERE PLASTIC

DEFORMATION
SEVERE PLASTIC
DEFORMATION
Methods, Processing
and Properties

GHADER FARAJI
HYOUNG SEOP KIM
HESSAM TORABZADEH KASHI
Elsevier
Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom
50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

Copyright r 2018 Elsevier Inc. All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopying, recording, or any information storage and retrieval
system, without permission in writing from the publisher. Details on how to seek permission, further
information about the Publisher’s permissions policies and our arrangements with organizations such
as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our
website: www.elsevier.com/permissions.

This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).

Notices
Knowledge and best practice in this field are constantly changing. As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical treatment
may become necessary.

Practitioners and researchers must always rely on their own experience and knowledge in evaluating
and using any information, methods, compounds, or experiments described herein. In using such
information or methods they should be mindful of their own safety and the safety of others, including
parties for whom they have a professional responsibility.

To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume
any liability for any injury and/or damage to persons or property as a matter of products liability,
negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas
contained in the material herein.

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library

Library of Congress Cataloging-in-Publication Data


A catalog record for this book is available from the Library of Congress

ISBN: 978-0-12-813518-1

For Information on all Elsevier publications


visit our website at https://www.elsevier.com/books-and-journals

Publisher: Matthew Deans


Acquisition Editor: Christina Gifford
Editorial Project Manager: Peter Jardim
Production Project Manager: Sruthi Satheesh
Cover Designer: Christian Bilbow

Typeset by MPS Limited, Chennai, India


Introduction
I.1 THE ULTRAFINE-GRAINED AND NANOSTRUCTURED
MATERIALS
All materials are composed of atoms. Depending on the order in which
the atoms are arranged with one another, solid material can be classified
into crystalline and noncrystalline or amorphous materials. In crystalline
materials, the atoms are arranged in a repetitive or periodic array over
large atomic distances, while in amorphous materials this long-range
atomic order is absent. The crystalline metals with different crystal
structures, such as face-centered cubic, hexagonal closed pack, or body-
centered cubic, are divided into two categories: single crystal and
polycrystalline. In single crystal metals, all unit cells are interlocked in the
same way and have the same orientation, while the polycrystalline metals
are formed from a large number of single crystals with different crystal
orientations. Conversely, most polycrystalline metals are composed of a
collection of many small single crystals named grains and are similar to
pomegranate fruit, which is made up of many small sweet seeds (Fig. I.1).
The grains are separated from each other by grain boundaries while pre-
serving the integrity of the metal. Grain size is one of the most important
factors that determines most of the physical, chemical, and mechanical
properties of metals. It is well known that the room temperature mechan-
ical properties, high-temperature superplastic behavior, chemical activity,
bioactivity, corrosion behavior, hydrogen storage capacity, and most other
properties of polycrystalline metals depend on the microstructural charac-
teristics, especially the grain size [1]. In general, the decrease in grain size
enhances the strength of the metal at room temperature, according to the
well-known Hall Petch relationship (Eq. (I.1)) [2] as follows:

σy 5 σ0 1 Kd22
1
(I.1)
Thus, metals with smaller grain size having high strength and moder-
ate ductility at room temperature will be very promising for load-bearing
structural applications. At elevated temperatures, the fine grain size pro-
vides an excellent superplastic forming capability, a key process for
manufacturing complicated parts that are difficult to deform.

Severe Plastic Deformation © 2018 Elsevier Inc.


DOI: https://doi.org/10.1016/B978-0-12-813518-1.00020-5 All rights reserved. 1
2 Severe Plastic Deformation

Figure I.1 (A) Pomegranate fruit (Saveh city, Iran) composed of a large number of
sweet seeds, (B) optical microscopy microstructure of cast commercially pure Al con-
sisting of grains [3].

Coarse-grained Nanograined
Grain size > 10 µm 100 nm > Grain size

Fine-grained Ultrafine-grained
1 µm < Grain size < 10 µm 1 µm > Grain size > 100 nm

Figure I.2 Different category of metals based on the grain size.

The micrograph of a different category of metals based on the grain


size is illustrated in Fig. I.2. The most common grain size of industrial
metals is .10 µm and is often referred to as coarse-grained (CG) metals.
The CG metals are mainly produced by the casting processes. Fine-grained
metals with a grain size of B1 10 µm are produced via industrial ther-
momechanical processing in which the metal-forming process is carried
out at elevated temperatures [4]. The third is ultrafine-grained (UFG) metals
with grain size below 1 µm and larger than 100 nm. The metals with
grain size below 100 nm are called nanostructured (NS) or nanograined
metals. The UFG or nanograined metals are hard to produce using con-
ventional thermomechanical and metal-forming processes [1,2,5]. This
can be attributed to the limited amount of plastic strain that results from
limitations in the reduced cross-section, limited values of hydrostatic
compressive stresses, and the lack of high-angle grain boundaries [6,7].
Introduction 3

Over the last decade, the development of UFG and NS materials has
become one of the major advances in modern materials science [6,8].
The enhanced physical and mechanical properties of UFG and NS mate-
rials are of great interest to further research in order to obtain unique
properties for a variety of structural and functional applications [2,9].
The investigation of mechanical properties of nanostructured materials
is one of the principal directions that has been developed in recent years.
As mentioned earlier, the formation of UFG and NS in metals and alloys
should lead to high strength of these materials according to the well-
known Hall Petch relationship [10,11]. Apart from a few exceptions,
both high strength and high ductility have been reported in UFG and NS
metals [12 17]. In some cases, the ductility of nanostructured and UFG
metals is reduced due to the low strain-hardening rate. Thus, considerable
effort has been made to enhance the microstructure and deformation
mechanisms to provide the desired strain-hardening rate, some of which
has led to an increase in both strength and ductility [2,18]. Superplasticity
is an important feature in UFG and NS materials and is associated with
very high ductility, with elongation of greater than 400% in tension [19].
The development of a polycrystalline material with a superplastic
capability is significant because of the increasing importance of the
superplastic-forming industry in the manufacture of complex parts with
curved surfaces [20,21]. The formation of nanostructures greatly improves
fatigue strength and durability, but knowledge of fatigue behavior of
nanostructured materials is rather poor and requires further research
[2,22,23]. The investigation of corrosion behavior has already shown con-
flicting results in UFG and NS materials. It is reported that the corrosion
rate of the AZ31 alloy decreases after the equal channel angular pressing
(ECAP) process [24], while the decrease in grain size due to the ECAP
process increases the corrosion rate in AE21 due to increased chemical
activity at the grain boundaries [25]. On the other hand, significant deg-
radation of corrosion resistance in nanocrystalline Ni-P has been reported
for grain sizes of 8.4 and 44.6 nm [26]. Moreover, the same researchers
have shown that the good corrosion properties of pure nickel remain
unchanged in the nanostructured state [27]. Because of this ambiguity,
the corrosion properties of NS materials are presumed to depend strongly
on processing technology, thermal history, and material purity, and further
investigation is required to elucidate this ambiguity [2]. Essential require-
ments for all biomedical applications are corrosion resistance and excellent
biocompatibility, and both these properties are observed in titanium and
4 Severe Plastic Deformation

most titanium alloys. Preliminary investigations of the corrosion behavior


of nanotitanium have shown improved corrosion resistance by
introducing a UFG microstructure [28,29].
These superior mechanical and physical properties make UFG and NS
materials attractive for numerous advanced applications in medical, aero-
space, energy, sporting goods, transportation, and high-tech industries.
Various methods for producing these materials have been developed,
which, based on their approaches, can be classified into two main catego-
ries: bottom-up and top-down methods [30].
As shown in Fig. I.3, the first approach involves a method of stacking
atoms together to produce UFG and NS materials. The second approach
involves a method of slicing or successive cutting of a bulk material to get
nanosized particles or a bulk metal with nanograins [31,32]. The severe

Top-down

Bulk

Powder

Nanoparticles

Cluster

Atoms

Bottom-up

Figure I.3 Schematic representation of the building up of nanostructures.


Introduction 5

plastic deformation (SPD) methods, which are based on the second


approach, are the main subject of this book. The different methods of the
first and second approaches are presented in the next section. In the sub-
sequent section, the main advantages of SPD methods will be explained.

I.2 FIRST APPROACH: BOTTOM-UP


In the bottom-up synthesis method, atoms are placed together, which
results in the formation of crystal planes, and these crystal planes are fur-
ther stacked on each other to form a nanostructure. The bottom-up tech-
niques include [33]:
1. Chemical synthesis;
2. Self-assembly;
3. Positional assembly.
In the following, some methods of the bottom-up approach are
described.

I.2.1 Inert Gas Condensation


The inert gas evaporation method has been extensively used by the
Japanese school since the 1960s, and, in 1991, Uyeda published an excel-
lent and comprehensive summary of the Japanese literature [34]. It has
been shown that a wide variety of metals with very fine particles can be
synthesized in a low-pressure inert gas atmosphere and that their sizes can
be controlled by varying the gas pressure (in the range of 1 30 Torr).
Inert gas condensation (IGC) is a bottom-up approach method for syn-
thesizing nanostructured materials with two basic steps [35].
The first step involves evaporation of the material, and the second step
involves a rapidly controlled condensation to produce the required parti-
cle size. In this unit, the chamber is evacuated to a pressure of about
2 3 1026 Torr by an oil diffusion pump. The crucible, which contains the
metal to be evaporated, is slowly heated via radiation from the graphite
heater element. The temperature is set to a predetermined value. After
evacuation, an inert gas (He, Xe, or Ar) is leaked into the chamber at a
low pressure, typically about 0.5 4 Torr, and the crucible is heated
rapidly (at constant temperature and inert gas pressure). The ultrafine
metal particles that nucleate and grow in the gas phase are collected on a
water-cooled surface. The powder particles are collected on the carbon-
coated electron-microscope grid attached to the center of the water-
cooled surface and can be observed directly in the transmission electron
6 Severe Plastic Deformation

Liquid nitrogen

Scraper

Main vacuum
chamber

A B
Evaporation Gas inlet
sources Funnel

Vacuum pumps
Bellows
Fixed piston

Anvil
Low-pressure
Slide compaction unit

Sleeve
High-pressure
Piston Piston compaction unit

Figure I.4 Schematic of the inert-gas condensation chamber for the synthesis of
nanostructured materials [35].

microscope. Samples with a thickness less than a monolayer of particles


can be collected on the grid using a shutter. Fig. I.4 shows a schematic of
a typical apparatus which is used nowadays [35]. The basic components of
the apparatus are similar to those employed by Granqvist and Buhrman
[36], except that the new unit: (1) contains a scraper to collect the pow-
der particles into a container and (2) allows in situ compaction of the
powders into bulk [35].

I.2.2 Spray Conversion Processing


Spray conversion processing consists of three sequential steps. In the first
step, an aqueous solution of the precursor compound is prepared and
mixed to fix the composition of the starting solution. In the second step,
the starting solution is spray-dried to form a chemically homogeneous
precursor powder. In the final step, the precursor powder is thermoche-
mically converted to the desired nanostructured end-product powder
[37]. A schematic illustration of this process is shown in Fig. I.5.
Introduction 7

Cyclone
chemicals
Precursor

Afterburner
Water

Feed
bin

Filter
Fuel
Air

Fluid bed
reactor
Solution
mix tank Spray
dryer
Reactant gases Product
Figure I.5 Schematic of the spray conversion process, developed by Nanodyne Inc.
[38]. Lassner E., Schubert WD. (1999) Tungsten in Hardmetals. In: Tungsten. Springer,
Boston, MA.

The latter step can be carried out in a fixed bed reactor when the
amount of powder being processed is small, but for the thermochemical
processing of large quantities of powder, it is advisable to use a fluid bed
reactor to ensure a uniform conversion rate for all the particles in the
bed. All three steps in the process are readily scalable [37].
Spray drying is an essential step in the process when dealing with
starting solutions that contain two or more precursor compounds. Rapid
drying of the aerosol droplets, which is accompanied by rapid precipita-
tion of the solute, produces chemically homogeneous precursor powder
even from a complex starting solution. In other words, spray drying tends
to suppress the phase separation that normally occurs during conventional
crystallization of the solution mixture. Typically, the spray-dried precursor
particles are spherical shells about 10 50 µm in diameter and have amor-
phous or microcrystalline structures [37].
Thermochemical conversion of the precursor powders in fluid bed
reactors is also an important step in the integrated process. This is
because the local environment of temperature and gas concentration in
the fluid bed reactor is the same for all parts of the bed, which ensures
uniform conversion of the precursor powder to the end-product
powder. This is not the case for fixed bed reactors where the uniformity
8 Severe Plastic Deformation

of gas percolation and temperature is difficult to maintain throughout


the powder aggregate [37].

I.2.3 Chemical Vapor Condensation


A modification of the conventional IGC system, called chemical vapor
condensation, was used to synthesized silicon base ceramics in 1994 [39].
A schematic of the apparatus is shown in Fig. I.6. The gas stream is
introduced at a controlled rate into the dynamically pumped vacuum
chamber through the needle valve. The pressure in the chamber is main-
tained at a constant low pressure (1 50 mbar) by high-speed pumping.
The heated tubular reactor of high-purity A1203 provides a heat source
for the controlled decomposition of the precursor. During the short
residence time of the precursor in the heated tube, the individual
molecules of the precursor start to decompose and combine to form small
clusters or particles. At the outlet of the reactor, the rapid expansion of
the cluster or particle beam serves to mitigate particle growth and
agglomeration. Finally, the particle beam condenses out on a liquid
nitrogen-cooled rotating substrate from which the powders can be
scraped off and collected [39,40].

Chamber pressure
(1–50 mbar) Control
valve
To
pump
Carrier
gas Cold
Substrate
Needle
valve Heated tubular reactor
Scraper
Precursor Gas Particles
source
Mass flow
controller

Funnel

Collector

Figure I.6 A schematic of the chemical vapor condensation system.


Introduction 9

I.3 TOP-DOWN METHODS


As mentioned earlier, the top-down approach starts from a bulk material
that incorporates critical nanoscale details. In this method, a material is
engineered by scaling down a complex entity into its parts, such as creat-
ing small crystals from a bulk mineralized hard tissue [41]. This type of
fabrication is based on a number of tools and methodologies, which con-
sist of three major steps:
1. The deposition of thin films/coatings on a substrate;
2. Obtaining the desired shapes via photolithography;
3. Pattern transfer using either a lift-off process or selective etching of
the films.
In the following, some methods of the top-down approach are
described.

I.3.1 High-Energy Ball Milling


The synthesis of materials by high-energy ball milling of powders was first
developed by John Benjamin (1970) and his coworkers at the
International Nickel Company in the late 1960s [42,43]. It was found
that this method, called mechanical alloying, could successfully produce
fine and uniform dispersions of oxide particles (Al2O3, Y2O3, ThO2) in
nickel-base superalloys which could not be made by conventional powder
metallurgy methods.
It is a ball milling process where a powder mixture placed in the ball
mill is subjected to high-energy collision from the balls. Fig. I.7 shows
the motions of the balls and the powder. Since the rotation directions of
the bowl and balls are opposite, the centrifugal forces are alternately syn-
chronized. Thus, friction resulted from the hardened milling balls and the
powder mixture being ground alternately rolling on the inner wall of the
bowl and striking the opposite wall. The impact energy of the milling
balls in the normal direction attains a value of up to 40 times higher than
that due to gravitational acceleration. Hence, the planetary ball mill can
be used for high-speed milling [44].
During the high-energy ball milling process, the powder particles are
subjected to high energetic impact. Microstructurally, the mechanical
alloying process can be divided into four stages: (1) initial stage, (2) inter-
mediate stage, (3) final stage, and (4) completion stage [44].
1. At the initial stage of ball milling, the powder particles are flattened
by the compressive forces caused by the impact of the balls.
10 Severe Plastic Deformation

Horizontal section

e
forc
Fracturing &

gal
flattening
trifu
Cen

Rotation of the milling bowl


Figure I.7 Schematic view of the motion of the ball and powder mixture.

Microforging leads to changes in the shapes of individual particles, or


clusters of particles being repeatedly impacted by the balls with high
kinetic energy. However, such deformation of the powders shows no
net change in mass.
2. At the intermediate stage of the mechanical alloying process, a signifi-
cant change occurs as compared to the initial stage. Cold welding
becomes significant. The intimate mixture of the powder constituents
decreases the diffusion distance to the micrometer range. Fracturing
and cold welding are the dominant milling processes at this stage.
Although some dissolution may take place, the chemical composition
of the alloyed powder is still not homogeneous.
3. At the final stage of the mechanical alloying process, more refinement
and reduction in particle size becomes evident. The microstructure of
the particle also appears to be more homogeneous in microscopic scale
than those at the initial and intermediate stages. True alloys may have
already been formed.
4. At the completion stage of the mechanical alloying process, the pow-
der particles possess an extremely deformed metastable structure. At
this stage, the lamellae are no longer resolvable by optical microscopy.
Further mechanical alloying beyond this stage cannot physically
improve the dispersoid distribution. Real alloy with a composition
similar to the starting constituents is thus formed [44].
Introduction 11

I.3.2 Physical Vapor Deposition


Physical vapor deposition (PVD) is a vaporization coating technique
that involves the transfer of material at the atomic level. The process
can be described according to the following sequence of steps. (1) The
material to be deposited is converted into a vapor by physical means
(high-temperature vacuum or gaseous plasma), (2) the vapor is trans-
ported to a region of low pressure from its source to the substrate, and
(3) the vapor undergoes condensation on the substrate to form a thin
film. Typically, PVD processes are used to deposit films with thicknesses
in the range of a few nanometers to thousands of nanometers.
However, they can also be used to form multilayer coatings, graded
composition deposits, very thick deposits, and freestanding structures
[45]. A typical PVD process is shown in Fig. I.8.
PVD thin-film technology covers a rather broad range of deposition
techniques, including electron-beam or hot-boat evaporation, reactive
evaporation, and ion plating. PVD techniques also include processes based
on sputtering, whether by plasma or by an ion beam. PVD is also used to
describe the deposition from arc sources that may or may not be filtered.
In general, this process can be divided into two groups: evaporation and
sputtering. Evaporation refers to thin films being deposited by thermal
means, whereas in the sputtering mode the atoms or molecules are
dislodged from the solid target through the impact of gaseous ions
(plasma). Both methods have been further developed into several specific
techniques [45].

U–

Sputtering target
Sputtered
target
Ar+ atoms

Substrate

Sputtering gas

Thin film

Figure I.8 Schematic illustration of the physical vapor deposition process.


12 Severe Plastic Deformation

I.3.3 Sputtering
Sputtering is one of the most important PVD techniques in which the
physical vaporization of atoms from a surface occurs by momentum trans-
fer from bombarding, energetic, and atomic-sized particles. Sputter depo-
sition permits better control of the composition of multielement films
and greater flexibility in the types of materials that may be deposited [45].
Although first reported by Wright in 1877, sputter deposition of films
became feasible only because of the relatively poor vacuum required for
its operation. Despite the fact that Edison patented a sputter deposition
process for depositing silver onto wax photograph cylinders in 1904, the
process was not widely used in the industry until the advent of magnetron
sputtering in 1974. The application of sputter deposition led to an accel-
eration in the development of reproducible, stable, long-lived vaporiza-
tion sources for production purposes. Following the use of a magnetic
field which confines the motion of the secondary electrons close to the
target surface, planar magnetron sputtering has become the most widely
used sputtering configuration. It was derived originally from the develop-
ment of the microwave klystron tube in World War II, from the investiga-
tions of Kesaev and Pashkova (in 1959) on confining arcs, and of Chapin
(in 1974) on developing the planar magnetron sputtering source [46,47].
The operating principles of both direct current and radio frequency
sputtering systems are illustrated schematically in Fig. I.9 [48]. Effective
sputter deposition can be achieved in [45]:

Matching
network
13.56 MHz
–V(DC)
Insulation

Target

Glow discharge Substrates Glow discharge

Anode

Sputtering Vacuum Sputtering Vacuum


gas gas
(A) (B)
Figure I.9 Schematic diagram of the principles of (A) direct current (DC) and (B)
radio-frequency sputtering systems [48].
Introduction 13

• A good vacuum (,1025 Torr) using ion beams;


• A low-pressure gas environment, where sputtered particles are trans-
ported from the target to the substrate without gas-phase collisions
(i.e., a pressure less than about 5 3 1023 Torr), using a plasma as the
ion source of ions; and
• A higher-pressure gas, where gas phase collisions and “thermalization”
of the ejected particles occurs but the pressure is low enough that gas-
phase nucleation is not important (i.e., a pressure greater than about
5 mTorr but less than about 50 mTorr).
Currently, plasma-based sputtering is the most common form of sput-
tering in which a plasma is present and positive ions are accelerated to the
target, which is at a negative potential with respect to the plasma. At higher
pressures, the ions suffer physical collisions and charge-exchange collisions,
so that there is a spectrum of energies of the ions and neutrals bombarding
the target surface. At low pressures, the ions reach the target surface with
an energy given by the potential drop between the surface and the point in
the electric field where the ions are formed. In vacuum-based sputtering,
however, an ion or plasma beam is formed in a separate ionization source,
and is then accelerated and extracted into a processing chamber which is
maintained under good vacuum conditions. In this process, the mean
bombarding energy is higher than plasma-based bombardment, and the
reflected high-energy neutrals are more energetic [45].
Sputter deposition can be used to deposit films of elemental materials,
and also to deposit alloy films and maintain the composition of the target
material. This is possible because the material is removed from the target
in a layer-by-layer fashion, which is one of the main advantages of the
process. This allows the deposition of some rather complex alloys such as
Al Si Cu for semiconductor metallization [49] and metal Cr Al Y
alloys for aircraft turbine blade coatings [45].
Deposition of the composite material film by sputtering can be
achieved by sputtering from a compound target or by sputtering from the
elemental target at the partial pressure of a reactive gas (i.e., “reactive
sputter deposition”). In most cases, the sputter deposition of a compound
material from a compound target results in a loss of some of the more
volatile material (e.g., oxygen from SiO2). However, this loss is often
made up by deposition in an ambient containing a partial pressure of the
reactive gas—a process known as “quasi-reactive sputter deposition.” In
the latter case, the required partial pressure of the reactive gas is less than
the partial pressure used for reactive sputter deposition [45].
14 Severe Plastic Deformation

However, the above techniques are often limited to the manufacturing


of small-scale specimens which are commonly used in electronic devices
but are generally not suitable for large-scale structural applications.
Furthermore, the finished products of these techniques invariably contain
residual porosity and contamination. According to recent studies, large
bulk metals in a fully dense state may be manufactured by combined
cryo-milling, hot isostatic pressing, and subsequent extrusion. However,
this method is expensive, and it is hard to be adapted for the structural
alloys used in large-scale industrial applications [5]. The “top-down”
approach is different because it is dependent upon taking a bulk solid
with a relatively coarse grain size and processing the bulk solid to produce
UFG and NG microstructures through heavy straining. This approach
avoids the small product sizes and the contamination which are inherent
features of materials produced using the “bottom-up” approach, and it
has the additional advantage that it can be readily applied to a wide range
of preselected alloys [5].

I.3.4 Severe Plastic Deformation Methods


SPD is a generic term for a group of metal-machining techniques involv-
ing very large strains which are imposed without significant changes in
the overall dimensions of the specimen or workpiece. Processing by SPD
provides the opportunity to achieve remarkable grain refinement in crys-
talline solids [30]. The SPD methods are the main subject of this book,
and the fundamentals are described completely in the next chapters. In
general, SPD processes should have the following two characteristics:
• The section area of the sample must be kept constant during the SPD
process so that the high strain can be applied to the material cyclically.
• The high hydrostatic pressure is applied to prevent crack growth and
other defects in the sample.
To respond as to why SPD methods should be used, the advantages of
these methods are introduced. Some benefits of SPD methods are as
follows:
• They have the capability to produce different shapes of UFG and NS
samples such as bulk, tube, and sheet.
• The different materials and various crystal structures can be refined by
SPD methods.
• The capability to apply extreme strain and enhance saturated grain
refinement.
Introduction 15

• The homogeneous microstructure can be improved by using different


processing routes and die geometry.
• The methods are relatively simple and low-cost to refine the material
microstructure.

REFERENCES
[1] Langdon TG. The processing of ultrafine-grained materials through the application
of severe plastic deformation. J Mater Sci 2007;42:3388 97.
[2] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nanostructured materials from
severe plastic deformation. Prog Mater Sci 2000;45:103 89.
[3] Ahmadkhanbeigi M, Shapourgan O, Faraji G. Microstructure and mechanical prop-
erties of Al tube processed by Friction Stir Tube Back Extrusion (FSTBE). Trans.
Ind. Inst. Metals 2017;70(7):1849 56.
[4] Figueiredo RB, Langdon TG. Grain refinement and mechanical behavior of a
magnesium alloy processed by ECAP. J Mater Sci 2010;45:4827 36.
[5] Valiev RZ, Langdon TG. Principles of equal-channel angular pressing as a proces-
sing tool for grain refinement. Prog Mater Sci 2006;51:881 981.
[6] Zehetbauer MJ, Stüwe HP, Vorhauer A, Schafler E, Kohout J. The role of hydro-
static pressure in severe plastic deformation. Adv Eng Mater 2003;5:330 7.
[7] Xu C, Xia K, Langdon TG. The role of back pressure in the processing of pure
aluminum by equal-channel angular pressing. Acta Mater 2007;55:2351 60.
[8] Gleiter H. Nanostructured materials: basic concepts and microstructure. Acta Mater
2000;48:1 29.
[9] Valiev R. Nanostructuring of metals by severe plastic deformation for advanced
properties. Nat Mater 2004;3:511 16.
[10] Weertman JR. Hall-Petch strengthening in nanocrystalline metals. Mater Sci Eng: A
1993;166:161 7.
[11] Valiev RZ. Structure and mechanical properties of ultrafine-grained metals. Mater
Sci Eng: A 1997;234:59 66.
[12] Valiev RZ, Alexandrov IV, Zhu YT, Lowe TC. Paradox of strength and ductility in
metals processed by severe plastic deformation. J Mater Res 2002;17:5 8.
[13] Höppel HW, Zhou ZM, Mughrabi H, Valiev RZ. Microstructural study of the
parameters governing coarsening and cyclic softening in fatigued ultrafine-grained
copper. Philos Mag A 2002;82:1781 94.
[14] Wang Y, Chen M, Zhou F, Ma E. High tensile ductility in a nanostructured metal.
Nature 2002;419:912 15.
[15] Youssef KM, Scattergood RO, Murty KL, Horton JA, Koch CC. Ultrahigh strength
and high ductility of bulk nanocrystalline copper. Appl Phys Lett 2005;87:091904.
[16] Mungole T, Kumar P, Kawasaki M, Langdon TG. A critical examination of the par-
adox of strength and ductility in ultrafine-grained metals. J Mater Res
2014;29:2534 46.
[17] Mungole T, Kumar P, Kawasaki M, Langdon TG. The contribution of grain bound-
ary sliding in tensile deformation of an ultrafine-grained aluminum alloy having
high strength and high ductility. J Mater Sci 2015;50:3549 61.
[18] Zhao Y, Zhu Y, Lavernia EJ. Strategies for improving tensile ductility of bulk nanos-
tructured materials. Adv Eng Mater 2010;12:769 78.
[19] Langdon TG. Seventy-five years of superplasticity: historic developments and new
opportunities. J Mater Sci 2009;44:5998.
16 Severe Plastic Deformation

[20] Barnes AJ. Superplastic forming 40 years and still growing. J Mater Eng Perform
2007;16:440 54.
[21] Valiev RZ, Estrin Y, Horita Z, Langdon TG, Zehetbauer MJ, Zhu YT.
Fundamentals of superior properties in bulk NanoSPD materials. Mater Res Lett
2016;4:1 21.
[22] Agnew SR, Weertman JR. Cyclic softening of ultrafine grain copper. Mater Sci
Eng: A 1998;244:145 53.
[23] Vinogradov A, Kaneko Y, Kitagawa K, Hashimoto S, Stolyarov V, Valiev R. Cyclic
response of ultrafine-grained copper at constant plastic strain amplitude. Scr Mater
1997;36:1345 51.
[24] Wang H, Estrin Y, Fu H, Song G, Zúberová Z. The effect of pre-processing and
grain structure on the bio-corrosion and fatigue resistance of magnesium alloy
AZ31. Adv Eng Mater 2007;9:967 72.
[25] Minárik P, Král R, Janeček M. Effect of ECAP processing on corrosion resistance of
AE21 and AE42 magnesium alloys. Appl Surf Sci 2013;281:44 8.
[26] Rofagha R, Erb U, Ostrander D, Palumbo G, Aust KT. The effects of grain size and
phosphorus on the corrosion of nanocrystalline Ni-P alloys. Nanostruct Mater
1993;2:1 10.
[27] Rofagha R, Langer R, El-Sherik AM, Erb U, Palumbo G, Aust KT. The corrosion
behaviour of nanocrystalline nickel. Scr Metall Mater 1991;25:2867 72.
[28] Balyanov A, Kutnyakova J, Amirkhanova NA, Stolyarov VV, Valiev RZ, Liao XZ,
et al. Corrosion resistance of ultra fine-grained Ti. Scr Mater 2004;51:225 9.
[29] Valiev RZ, Zehetbauer MJ, Estrin Y, Höppel H W, Ivanisenko Y, Hahn H, et al.
Cover picture: the innovation potential of bulk nanostructured materials. Adv Eng
Mater 2007;9:527.
[30] Raj KH, Sharma RS, Singh P, Dayal A. Study of friction stir processing (FSP) and
high pressure torsion (HPT) and their effect on mechanical properties. Procedia Eng
2011;10:2904 10.
[31] Iqbal P, Preece JA, Mendes PM. Nanotechnology: the “top-down” and “bottom-
up” approaches. Supramolecular chemistry. Chichester: John Wiley & Sons, Ltd;
2012.
[32] Pourhashemi A. Engineering materials: applied research and evaluation methods.
England, Taylor and Francis: Apple Academic Press; 2014.
[33] Karkare M. Nanotechnology: fundamentals and applications. New Delhi, India: I.K.
International Publishing House Pvt. Limited; 2008.
[34] Uyeda R. Studies of ultrafine particles in Japan: crystallography. Methods of prepara-
tion and technological applications. Prog Mater Sci 1991;35:1 96.
[35] Suryanarayana C, Prabhu B. 2 Synthesis of nanostructured materials by inert-gas
condensation methods A2. In: Koch CC, editor. Nanostructured Materials. 2nd ed.
Norwich, NY: William Andrew Publishing; 2007. p. 47 90.
[36] Granqvist CG, Buhrman RA. Ultrafine metal particles. J Appl Phys
1976;47:2200 19.
[37] Kear BH, McCandlish LE. Chemical processing and properties of nanostructured
WC-Co materials. Nanostruct Mater 1993;3:19 30.
[38] Lyshevski SE. Nanomaterials: recent advances in technology and industry. Dekker
encyclopedia of nanoscience and nanotechnology. 3rd ed. England, Taylor and Francis;
2004. p. 6. Seven volume set.
[39] Chang W, Skandan G, Hahn H, Danforth SC, Kear BH. Chemical vapor condensa-
tion of nanostructured ceramic powders. Nanostruct Mater 1994;4:345 51.
[40] Chang W, Skandan G, Danforth SC, Rose M, Balogh AG, Hahn H, et al.
Nanostructured ceramics synthesized by chemical vapor condensation. Nanostruct
Mater 1995;6:321 4.
Introduction 17

[41] Liu Y, Mai S, Li N, Yiu CKY, Mao J, Pashley DH, et al. Differences between top-
down and bottom-up approaches in mineralizing thick, partially-demineralized col-
lagen scaffolds. Acta Biomater 2011;7:1742 51.
[42] Benjamin JS. Dispersion strengthened superalloys by mechanical alloying. Metall
Trans 1970;1:2943 51.
[43] Yadav TP, Yadav RM, Singh DP. Mechanical milling: a top down approach for the
synthesis of nanomaterials and nanocomposites. Nanosci Nanotechnol
2012;2:22 48.
[44] Rajeshkanna S, Nirmalkumar O. Synthesis and characterization of Cu nanoparticle
using high energy ball milling route and compare with Scherrer Equation. Int J Sci
Eng Res (IJSER) 2014;2:30 5.
[45] Chi L. Nanotechnology: Volume 8: Nanostructured Surfaces. Chichester: John
Wiley & Sons; 2010.
[46] Wehner GK. Sputtering by ion bombardment. Adv Electron Electron Phys
1955;7:239 98.
[47] Kay E. Impact evaporation and thin film growth in a glow discharge. Adv Electron
Electron Phys 1963;17:245 322.
[48] Ohring M, Baker SP. Materials science of thin films: deposition and structure. San
Diego, CA: Elsevier Science & Technology Books; 2016.
[49] Gadepally KV, Hawk RM. Integrated circuits interconnect metallization for the sub-
micron age. J Arkansas Acad Sci 1989;43(Article 9).
CHAPTER 1

Fundamentals of Severe Plastic


Deformation
1.1 INTRODUCTION
This chapter reviews the historical aspects and fundamentals of severe
plastic deformation (SPD) and grain refinement processes for processing
ultrafine grain (UFG) and nanograin (NG) metals.

1.2 HISTORY
It follows from inspection of the Hall Petch relationship (Eq. (I.1)) at
room temperature that a small grain size is preferred in load-bearing
components and structural applications because it leads to significantly
higher strength [1]. Additionally, in high-temperature regimes, the creep
rate under steady-state conditions ε_ is expressed by a relationship in the
form [2,3]:
 
ADGb b p  s n
ε_ 5 (1.1)
kT d G

where D is the appropriate diffusion constant ( 5 D0 exp 2 Q=RT ,
where D0 is the frequency factor, Q is the activation energy for the flow
process, R is the gas constant, and T is the absolute temperature), G is
the shear modulus, b is the Burger vector, k is Boltzmann’s constant, σ is
the flow stress, n is the stress exponent, p is the inverse grain size expo-
nent, and A is a dimensionless constant. It is evident from Eq. (1.1) that
the smaller grain size leads to faster strain rates, and this provides the pos-
sibility of achieving a superplastic forming capability at high strain rates
that may be readily employed in industrial forming operations. Besides
excellent superplastic behavior, the other exceptional properties of nano-
materials are higher strength, toughness, fatigue life, and wear resistance
[4]. Thus, grain refinement is an important processing tool for achieving
excellent properties in metallic materials. The advantages of grain refine-
ment were recognized many years ago and led to the development of
Severe Plastic Deformation © 2018 Elsevier Inc.
DOI: https://doi.org/10.1016/B978-0-12-813518-1.00001-1 All rights reserved. 19
20 Severe Plastic Deformation

thermomechanical processing operations wherein materials are subjected


to mechanical straining and annealing treatments in order to reduce the
grain size in the range of B3 10 µm.
In 1988, a landmark report was published that demonstrated the
potential of achieving smaller grain sizes within the submicrometer range
through the application of intense plastic deformation to a bulk coarse-
grained metal [5]. This approach is now generally named SPD processing.
Publications on SPD processing have attracted much attention and led to
the development of plenty of research activities around the world to pro-
cess and measure the characteristics of exceptionally small grain size mate-
rials. It is interesting to note that in this field research activities have
continued to evolve to the present day and have greatly expanded.
It is often not recognized that the general concept of SPD processing
has a long history that dates back more than 2000 years. Though the idea
of SPD processing was introduced 30 years ago, lay not only with the
processing methods but rather with the existence of advanced micro-
scopic and analytical tools. This provided direct evidence of the enhanced
mechanical properties of these processed metals due to the exceptional
grain refinement [5].
The historical background of SPD processing consists of three separate
periods. These periods are briefly outlined in the following section to
provide a detailed description of the latest developments in this field of
research.

1.2.1 The Ancient Age


Although SPD processing in its modern form is relatively new, the funda-
mental principles of this category of metal-forming processes extend back
to artisan works in ancient times [6]. It appeared as repetitive folding and
forging as early as B500 BC for the manufacturing of high-quality steel
swords. This method was developed and maturated around AD 220 280,
termed in Chinese “Bai-Lian” (meaning numerous repetitive processing).
The steel fabricated by this process is called Bai-Lian steel. A comprehen-
sive survey of the historical aspects of SPD processing shows that the gen-
eral concept dates back at least to the ancient Chinese Han dynasty in
B200 BC [7]. At that time, artisans introduced a new technique for the
processing of steel for use in swords to form the famous Bai-Lian steels,
the metal being repetitively forged and folded. The important feature of
this technique was that repeated forging and folding of the metal resulted
Fundamentals of Severe Plastic Deformation 21

in substantial hardening. Many archeological artifacts are now available


which often have inscriptions that provide a historical record of the pro-
cessing method. A high-strength sword from 50-Lian steel was prepared
by 50 separate forging and folding operations. The fundamentals of this
technology spread to Japan and it was used for the processing of samurai
swords. In India, it led to the development of Wootz steel, a special form
of ultra high carbon steel, which was developed between approximately
300 BC and AD 300 [6]. It is informative to note that Wootz steel has
been specifically designated as an advanced material of the ancient world
because of its high superplastic properties at elevated temperatures and
hardness [8]. Further spreading of this technology to the Middle East led
to the development of the famous Damascus steel manufactured in
ancient Syria [9].
Although these developments extended readily across Asia, the princi-
ples of the processing method lacked scientific rigor, and the knowledge
of this technique was lost in the middle of the 18th century [5].

1.2.2 The Scientific Age


The modern history of SPD processing has its beginnings in the funda-
mental work of Bridgman at Harvard University who developed the sci-
entific backgrounds and techniques for materials processing through an
application of shear deformation under high hydrostatic pressure [10,11],
which today is the core of SPD methods [12]. The news of his finding
was press released with the exciting title “Scientist Makes Miniature
Earthquake,” as shown in Fig. 1.1, indicating the severe strains under high
pressure in the Earth which result in deep-seated earthquakes (this photo
was provided by the archives of Harvard University) [13]. At the begin-
ning of the 1930s, Bridgman conducted a remarkably comprehensive
series of experiments on the application of high pressures to bulk solids
[10,14], and, in 1952, the results from these many experiments were suc-
cinctly summarized in a book [15]. It is worth mentioning that Bridgman
received the Nobel Prize in Physics in 1946 for the invention of “an
apparatus to produce extremely high pressures,” and for discoveries in the
field of high-pressure physics. As mentioned, Bridgman proposed the pro-
cessing of metals through a combination of torsional straining and com-
pression. This approach was later further developed in the former Soviet
Union [16] and finally evolved into the process known as high-pressure
torsion (HPT).
22 Severe Plastic Deformation

Figure 1.1 (A) P.W. Bridgman receiving the Nobel Prize from King Gustav V of
Sweden, December 11, 1946, and (B) the document that was released after
Bridgman’s success [13].

The second major influence on SPD processing can be traced to the


classic work of Dr. Segal and his colleagues at Minsk in the former Soviet
Union (now the capital of Belarus) in the 1980s. Segal and coworkers
developed the process known as equal channel angular pressing (ECAP)
(or equal channel angular extrusion) [17]. This technique is now the most
important and the most useful SPD processing method.
However, a deficit of the earlier studies was the lack of detailed micro-
structural analysis. Also, the detailed microstructural information became
possible only with the later development of sophisticated analytical tools
such as electron backscatter diffraction, high-resolution transmission elec-
tron microscopy, modern X-ray techniques, and orientation imaging
microscopy [5].

1.2.3 The Microstructural Age


The development of modern microscopic and analytical tools has led to
the detailed examination of microstructural features of metals and alloys
processed by SPD methods. In the 1980s, Professor Valiev and his collea-
gues from Ufa, Russia, developed the approach [18] in which processing
by SPD shows remarkable grain refinement with the average grain size in
the submicrometer or nanometer range even in conventional commercial
alloys. As an example, in the classic first article in 1988, it has been shown
Fundamentals of Severe Plastic Deformation 23

that a grain size of B0.3 µm in an Al 4% Cu 0.5% Zr alloy can be pro-


duced by HPT processing [19]. This grain size is much smaller than the
grain size of B3 5 µm obtained by the use of conventional thermome-
chanical processes. Valiev et al. published a follow-up report in the 1990s
in Ufa on nanomaterials by SPD, especially on ECAP and HPT [20,21].
This report presented the advanced properties of SPD processed metals to
scientists outside of Russia, and thereby encouraged a great deal of
research and development of SPD processing in laboratories around the
world. It is informative to note that the first Western scientific paper in
the field of SPD processing appeared in 1993 with the title “An
Investigation of Ductility and Microstructural Evolution in an Al-3% Mg
Alloy with Submicron Grain Size” [22]. These investigations have had a
major impact on the more recent literature in materials science, as docu-
mented in a more recent report [23].

1.3 BASIC PRINCIPLES OF SEVERE PLASTIC DEFORMATION


METHODS
At present, the processing of metals and alloys through the application of
SPD has attracted wide attention in many research laboratories around
the world. The processing by SPD is defined as [24]:
A metal forming process under high hydrostatic pressure to apply a very high
strain on a bulk metal without any significant change in the dimensions of the
sample and has the ability to achieve an exceptional grain refinement.
Imparting large plastic strains to a workpiece by conventional forming
methods is a challenging task because of the change in dimensions and
formation of internal defects and cracks [25]. It requires a considerable
investment in tool design, which should be durable enough to sustain
repeated high stresses during material forming under high hydrostatic
stresses. A distinctive feature of SPD processing that meets these require-
ments is that a very large strain is imposed to the workpiece without any
significant change in the overall dimensions. This is achieved due to spe-
cial tool geometries that prevent the free flow of the material and thereby
produce significant hydrostatic compressive stresses. The presence of this
hydrostatic pressure is a clue for achieving the high strains required for
exceptional grain refinement. Many crystalline materials, including those
which are brittle under normal conditions (e.g., tungsten oxide, B2O3
glasses, and amorphous materials), gain substantial ductility under high
24 Severe Plastic Deformation

Figure 1.2 Simulation of the SPD process with hammer impact on glass.

hydrostatic pressure and can be deformed to large strains before failure.


Nowadays many variants of SPD techniques, which explicitly or implic-
itly employ this generic feature of high hydrostatic pressure, are readily
available for fabrication of a great variety of UFG materials [12].
The SPD process can be explained with simplicity by analogy with a
hammer impact on glass. In Fig. 1.2, the SPD process is simulated by
analogy with a hammer impacting on a window glass. The glass can be
related to the microstructure of the material, and the window frame that
holds the glass can be related to the role of hydrostatic pressure in the
SPD process. The glass is crushed when the hammer impacts the glass.
The impact generated by the hammer is similar to the high strain intro-
duced on the material by the SPD process, and the crushed glass can be
related to microstructure refinement from coarse grain material to the
UFG and NS materials. The comprehensive investigation of grain refine-
ment mechanisms in the SPD process are explained in Section 1.4.
Thus, SPD is an effective method for obtaining a very fine crystalline
structure in different crystalline materials having different crystal struc-
tures of face-centered cubic (fcc), bcc, and hexagonal close-packed (hcp)
(e.g., aluminum, iron, and magnesium). SPD causes the formation of
micrometer and submicrometer-sized grains within the original coarse
grains of the material. The structural changes caused by SPD are reflected
in improved mechanical properties [26]. In general, the features of SPD
methods are expressed as follows:
• Imposing higher strains to the sample;
• Applying high hydrostatic pressure;
• Prevention of the free flow of the material during the process;
• No significant change to the sample dimensions after the process;
• The ability to produce microstructure with high-angle grain boundaries;
Fundamentals of Severe Plastic Deformation 25

• The ability to produce a homogeneous microstructure to achieve uni-


form properties;
• No mechanical defects, cracks, or porosity in the final sample.
In the next chapters, we will explain the most important factors of
SPD methods that influence the metallurgical and mechanical properties.
There are several factors involved in determining the evolution of micro-
structure during the SPD process. These factors are related to the inher-
ent characteristics of the SPD methods which have been already
introduced by researchers. Some of these factors, such as equivalent plastic
strain, shear strain, hydrostatic pressure, strain rate, process temperature,
and post heat treatment are explained in the following chapters.

1.4 DIFFERENCE BETWEEN SEVERE PLASTIC DEFORMATION


AND CONVENTIONAL METAL-FORMING PROCESSES
One of the most attractive features of SPD processing is the ability to
refine the grains to sizes that cannot be attained with conventional ther-
momechanical treatments. Thus, SPD processing is capable of producing
grains having sizes within the submicrometer and nanometer ranges,
whereas conventional treatments may refine the grains to sizes of several
micrometers. The grains produced in SPD processing are designated
UFGs, and UFG solids are defined as “bulk solids having fairly homoge-
neous equiaxed microstructures with average grain sizes less than B1 µm
and high angle grain boundaries (HAGBs)” [27,28]. Within the range of
UFG materials, submicrometer grain sizes refer to average grain size in
the range of 100 1000 nm, and nanometer grain sizes refer to an average
grain size of less than 100 nm [24].
It is well known that heavy plastic straining by conventional methods,
such as drawing, extrusion or cold rolling, can lead to grain refinement at
room temperature. However, the microstructures formed are usually cel-
lular type substructures with low angle misorientation boundaries. On
the other hand, the structures formed by SPD are ultrafine-grained or
nanograined structures with predominantly HAGBs. SPD processed
materials often have a mean grain size of less than 1 µm, and the grain
interiors usually possess a substructure with high dislocation density due
to highly distorted crystal lattices. The formation of such nanostructures
can be achieved by SPD methods providing very large deformations at
relatively low temperatures under high hydrostatic compressive stresses
[21,29,30].
26 Severe Plastic Deformation

It is well recognized that the interrelationship between the four


components of composition, processing, structure, and property make
up the core of materials science and engineering. The conventional
process has been dominating in the control of the microstructure and
properties of the material, and they rely on the manipulation of the
thermally activated process, such as phase transformation based on the
thermodynamics, and in some cases with the help of mechanical pro-
cesses. However, in recent years, significant progress and achievements
around the world in the research and development of SPD and related
materials claim the dominant role of high strain in the control of
materials microstructure and attract the ever-increasing attention of
materials experts worldwide [7]. The development of the SPD method
must meet specific requirements to produce nanostructured materials
from billets and bulk samples. These requirements are listed as follows
[31,32]. Firstly, it is essential to obtain UFG or NG structures with
dominant HAGBs since the qualitative change in material properties
occurs only in this case. Secondly, the formation of uniform nanostruc-
tures in the whole volume of a specimen is essential to achieve the
uniform properties of processed metals and alloys. Thirdly, the samples
should be free of mechanical cracks, or damage, after substantial plastic
deformation. It is not possible to meet the above-mentioned requirements
in conventional plastic deformation methods such as drawing, rolling, or
extrusion. It is impossible to form nanostructures in bulk samples without
utilization of a particular mechanical deformation procedure for providing
large deformations at relatively low temperatures and also without designa-
tion of optimal material processing [33].
These parameters constitute a clear demarcation line between nano-
SPD materials and conventional materials with low angle subgrain struc-
tures produced by cold rolling or other conventional metal-forming
techniques. Nevertheless, the works on the SPD process have opened
the gates for microstructure refinement by deformation to remarkable
strains [12].

1.5 GRAIN REFINEMENT MECHANISMS UNDER SEVERE


PLASTIC DEFORMATION CONDITIONS
Why are the coarser grains refined under SPD conditions? This is one of
the most important questions that gets much attention from SPD researchers.
Though it has not apparently been answered, in recent years, researchers have
Fundamentals of Severe Plastic Deformation 27

presented different models to explain the grain refinement mechanisms in


SPD processing. The correct identification of the grain refinement mecha-
nism is not only important from a theoretical viewpoint but also improves the
design of SPD methods. The nature of the structure formation is determined
not only by the material itself (initial microstructure, composition, crystal
structure, stacking fault energy (SFE), etc.) but also by the severe deformation
conditions (temperature, strain rate, and hydrostatic pressure, etc.). In general,
decreases in temperature, increases in pressure, and the addition of alloying
elements contribute to grain refinement and obtaining a minimum grain size
[33]. In particular, there is no generally accepted scenario of grain fragmenta-
tion by subdivision of grains, and the underlying processes remain a riddle for
researchers to the present day [12]. Researchers have presented several models
under specific conditions, which we introduce in the following sections. The
important point is that particular SPD methods can represent a combination
of these models. The following are the most well-known grain refinement
models for fcc and hcp metals.

1.5.1 Face-Centered Cubic (fcc) Metals


The processing of fcc metals by SPD methods is relatively easy because of
the multiplicity of active slip systems. It has been well documented that
for those materials with medium or high SFE, such as Ni (150 mJ/m2),
Al (200 mJ/m2), and Cu (80 mJ/m2), coarser grains are refined upon con-
tinued straining by multiplication and migration of dislocations [34]. As is
known, metals with high SFE tend to form a cell structure in which the
cell walls are formed from dislocation networks [35]. The multiplication
and migration of dislocations (MMDs) mechanism is the most common
model that was observed in the many studies [36 40]. Since it needs the
high mobility of dislocations, the MMDs mechanism is usually more
active in metals with high SFE. In the SPD methods, the application of
strain with hydrostatic pressure causes the dislocation density to increase
in the material. The presence of dislocation leads to refining material
microstructure in the following steps:
1. At first, at the onset of straining, a random dislocation distribution is
formed as observed in Fig. 1.3.
2. As shown in Fig. 1.3B, the dislocation density increases due to the
applied strain and results in the tangling of dislocations with the regu-
lar arrangement, and consequently this leads to the formation of elon-
gated dislocation cells (D2cD1).
28 Severe Plastic Deformation

Figure 1.3 Schematic of the grain refinement mechanism in ultrafine grain copper: (A)
random dislocation distribution, (B) elongated dislocation cells, (C) the dislocations
rearrange to form finer subgrain boundaries, (D) accumulation of dislocations in the cell
walls and forming the subgrains, and (E) formation of the high angle grain boundaries.

3. As the deformation continues, the dislocations rearrange to form the


boundaries of fine grains (Fig. 1.3C).
4. At higher strains and SPD cycles, accumulation of dislocations
increases at the dislocation walls, which causes the formation of sub-
grains with low angle boundaries. With further deformation, the
number of subgrains increases until the material converts to a UFG
structure (Fig. 1.3D).
5. As can be seen in the Fig. 1.3E, the shear strain at higher SPD cycles
causes rotation of the subgrains relative to each other. Accordingly,
the subgrains tend to disorient, and the HAGBs are formed [36,41].
Although many researchers verified the MMD model, this model can-
not justify the presence of different sizes of grains in ultrafine-grained
materials. To justify the bimodal (or other) distribution of grain size,
another model based on intersection of microshear bands (IMSBs) has been
proposed [42]. This mechanism is observed in materials with a cubic crys-
tal structure, and the microshear bands convert to the kink bands in hcp
materials [38].
Since it needs high mobility of dislocations, the MMD mechanism is
usually more active in high SFE materials such as aluminum alloys.
However, the IMSB mechanism can occur in various metals and alloys
Fundamentals of Severe Plastic Deformation 29

New grain
Microshear bands

HAGBs
LAGBs

Figure 1.4 Schematic illustration of the IMSB mechanism.

[38]. Also, other parameters can affect the activity of these mechanisms.
For example, the lower deformation temperature and the higher concen-
trations of alloying elements can result in the superiority of the IMSB
mechanism and deactivation of the MMD mechanism as illustrated by
Sitdikov et al. [43]. It is speculated that the occurrence of multidirectional
shear planes and the concentration of strains in these planes in MAF can
accelerate the IMSB mechanism.
As shown in Fig. 1.4, low angle grain boundaries (LAGBs) appeared
in the initial grains due to the incidence of microshear bands (MSBs) by
plastic deformation. With further deformation, the IMSBs occurs and
causes the LAGBs to convert to HAGBs, which results in the formation
of new grains [42,44].

1.5.2 Hexagonal Close-Packed (hcp) Metals


It is known that the straining of hcp materials at room temperature is
more difficult than that of fcc metals due to the limited number of active
slip systems. This problem was shown in recent investigations of ECAP
processing of pure magnesium and its alloy [45]. It is challenging to pro-
duce magnesium alloys with a UFG structure because they are normally
processed at elevated temperatures where dynamic recrystallization occurs
[46,47]. Thus the refinement process in hcp materials occurs through the
nucleation of finer grains along preexisting twins or grain boundaries
owing to the development of stress concentrations and the activation of
both basal and nonbasal slip [48]. This mechanism is based on the princi-
ples of dynamic recrystallization and is often observed at high temperature
(420 600K) [49 52]. It leads to a necklace-like array of new grains, and
it means in practice that a critical grain size dc is needed in order to
achieve an array of equiaxed UFGs. As any model for grain refinement
30 Severe Plastic Deformation

must satisfy the experimental observations, it is important to summarize


the most common features of microstructural evolution in magnesium-
based alloys processed by ECAP [53,54]. The experimental observations
are expressed as follows:
• The new fine grains that form along grain boundaries depend on
process conditions such as strain rate and temperature.
• Homogeneous grain refinement is observed throughout the billet in some
alloys, primarily when the material is subjected to a prior extrusion.
• Bimodal grain size distributions may be observed after one pass of
ECAP, and they may evolve into homogeneous distributions of fine
grains after multiple passes.
• Bimodal or even trimodal grain size distributions may exist following
multiple passes of ECAP.
Fig. 1.5A denotes a coarse initial structure with an average grain size,
d, which is much larger than the critical size so that dcdc , Fig. 1.5D and
Initial After ECAP

d >> dc

(A) (B) (C)

d > dc

(D) (E) (F)

d > dc

(G) (H) (I)

d < dc

(J) (K)
Figure 1.5 Model for the grain refinement process of Mg alloys processed by SPD in
which the left column shows the initial condition and the second column shows the
structure after one pass, and the third column shows the structure after multiple SPD
passes; the upper two rows show the same initial structure with different processing
parameters and the third and fourth rows show different initial structures [52].
Fundamentals of Severe Plastic Deformation 31

G denote coarse initial structures with d . dc , and Fig. 1.5J denotes an


initial structure finer than the critical size so that d , dc . The second col-
umn in Fig. 1.5 denotes the structure after a single pass of ECAP and the
third column denotes the structure after multiple passes. If the starting
grains are coarser than a critical size, dc , it is anticipated that these new
refined grains will not impinge on the refined grains formed along the
opposite grain boundary, thereby leaving a central core of the original
grain remains unrefined. Examples where this may occur are given in
Fig. 1.5A, D, and G where the grain size is consistently larger than dc and
the structure after one pass, shown in Fig. 1.5B, E, and H, contains some
areas of coarser grains. It should be noted that Fig. 1.5B shows an exam-
ple where twinning within the grain leads to grain refinement along the
twin. In practice, the size and area fraction of the unrefined cores of the
coarser grains will depend upon the size of the initial grains. In Fig. 1.5,
the areas occupied by the newly formed grains are highlighted in gray to
distinguish them from the cores of the original grains, which are shown
in white. Processing by ECAP with additional passes can lead to homoge-
neous arrays as shown in Fig. 1.5F and I. In some cases, the processing
may be insufficient to remove the larger grains as shown in Fig. 1.5C,
where there is a multimodal grain size distribution even after multiple
passes. By contrast, under some conditions the multimodal distribution of
grain sizes observed in the early passes of ECAP evolves into a homoge-
neous distribution of fine grains after multiple passes through the contin-
uous refinement of the original grains as shown in Fig. 1.5I. Conversely,
if the initial grains are finer than the critical size as shown in Fig. 1.5J,
refinement occurs along the original boundaries and the new grains
impinge on the refined grains formed at the adjacent boundaries, creating
a homogeneous distribution of grains in a single pass. This is schemati-
cally depicted in Fig. 1.5K [54].
In the second row of Fig 1.5, the initial structure is similar to the first
row, but the process conditions, such as temperature and strain rate, are
different. This leads to newly formed grains that are larger and occupy a
large volume fraction after one pass, as shown in Fig. 1.5E, and to a
homogeneous structure after multiple passes as shown in Fig. 1.5F.
Therefore, it can be stated that even if the initial grain size is larger than
dc , it is still possible to achieve fine grain structure with the homogeneous
distribution of grain size.
In the third row, Fig. 1.5G illustrates the situation where the initial
structure is reasonably fine, but the grains are sufficiently coarse that a
32 Severe Plastic Deformation

bimodal or multimodal grain size distribution develops after one pass of


ECAP, but the area fraction of newly formed grains is larger than the area
occupied by the remaining cores of the initial grains. Further refinement
occurs in subsequent passes to give a homogeneous distribution of very
fine grains after multiple passes as shown in Fig. 1.5I.
Finally, if the initial grain size is small and d , dc as in Fig. 1.5J, a
homogeneous array of fine grains can be produced in the first pass as illus-
trated in Fig. 1.5K, and this structure remains homogeneous in subse-
quent passes [52].
The important conclusion from this model is that the bimodal or
multimodal grain size distributions often reported after ECAP are transi-
tional and may be removed, or at least significantly changed, if the press-
ing is continued through a sufficiently large number of passes [52].
Hcp metals have limited slip systems (especially at room temperature),
so twinning is considered to play an important role in plastic deformation.
The major role of the twinning of hcp metals during deformation is to
convert the substantial portions of grains into a twin orientation when-
ever the applied stress is directed normal to the basal plane [55,56]. The
reorientation of grains helps slip to occur by activating the slip systems
that are favorably oriented. Slip induced by reorientation is important in
plastic deformation of hcp metals because twinning cannot accommodate
a huge amount of deformation since the atomic displacement by twinning
is less than one interatomic distance [57]. The twinning mechanism
refines the microstructure during the SPD process as follows [36]:
1. Formation of microtwins with high density and conversion of the first
coarse grains into a lamellar structure;
2. Development of dislocation walls inside twin lamellae that further
subdivide the lamellar structure into equiaxed nanosized blocks;
3. Formation of the preferentially oriented blocks into randomly ori-
ented nanograins.
Fig. 1.6 is an example of the twinning mechanism that shows the
TEM (transmission electron microscopy) microstructure of a pure Ti
sample after a single pass of ECAP. The microtwin bands are obvious to
constitute the lamellar structure. Dislocation walls are observed inside
twin bands that are marked with a arrows.
The formation of many twins is quite common in the deformation
process of materials with medium to low SFEs, especially at high strain
rates and/or low temperatures as the dislocation slips are effectively sup-
pressed. The twinning mechanism can also be observed in fcc metals, in
Fundamentals of Severe Plastic Deformation 33

Figure 1.6 TEM micrograph of ECAPed pure titanium showing twins and the interior
dislocation walls [57].

which case the tendency to substitute twinning for the slip is more sensi-
tive to the strain rate [58]. For example, nanosized thick mechanical twins
are frequently formed in the Cu samples processed by high strain rate
processes such as shock loading [59], ball milling, [60] and dynamic plastic
deformation [61].

REFERENCES
[1] Hall EO. The deformation and ageing of mild steel: III Discussion of results. Proc
Phys Soc Section B 1951;64:747.
[2] Langdon TG. Creep at low stresses: an evaluation of diffusion creep and Harper-
Dorn creep as viable creep mechanisms. Metall Mater Trans A 2002;33:249 59.
[3] Langdon TG. Identifying creep mechanisms in plastic flow. Z Metall
2005;96:522 31.
[4] Tavakkoli V, Afrasiab M, Faraji G, Mashhadi M. Severe mechanical anisotropy of
high-strength ultrafine grained Cu Zn tubes processed by parallel tubular channel
angular pressing (PTCAP). Mater Sci Eng: A 2015;625:50 5.
[5] Langdon TG. Twenty-five years of ultrafine-grained materials: Achieving excep-
tional properties through grain refinement. Acta Mater 2013;61:7035 59.
[6] Valiev RZ, Zhilyaev AP, Langdon TG. Bulk nanostructured materials: fundamentals
and applications. United States: John Wiley & Sons; 2013.
[7] Wang JT. Historic retrospection and present status of severe plastic deformation in
China. Mater Sci Forum 2006;503 504:363 70.
[8] Srinivasan SRS. India’s legendary wootz steel: an advanced material of the ancient
world. Bangalore: National Institute of Advanced Studies and Indian Institute of
Science; 2004.
[9] Sherby OD, Wadsworth J. Ancient blacksmiths, the Iron Age, Damascus steels, and
modern metallurgy. J Mater Process Technol 2001;117:347 53.
[10] Bridgman PW. On torsion combined with compression. J Appl Phys
1943;14:273 83.
[11] Bridgman PW. The effect of hydrostatic pressure on plastic flow under shearing
stress. J Appl Phys 1946;17:692 8.
34 Severe Plastic Deformation

[12] Estrin Y, Vinogradov A. Extreme grain refinement by severe plastic deformation: a


wealth of challenging science. Acta Mater 2013;61:782 817.
[13] Edalati K, Horita Z. A review on high-pressure torsion (HPT) from 1935 to 1988.
Mater Sci Eng: A 2016;652:325 52.
[14] Bridgman PW. Effects of high shearing stress combined with high hydrostatic pres-
sure. Phys Rev 1935;48:825 47.
[15] Bridgman PW. Studies in large plastic flow and fracture with special emphasis on the
effects of hydrostatic pressure. New York, NY: McGraw; 1952.
[16] Smirnova NA, Levit VI, Pilyugin VI, Kuznetsov RI, Davydova LS, Sazonova VA.
Evolution of the structure of f.c.c. single crystal subjected to strong plastic deforma-
tion. Fiz Metal Metalloved 1986;61:1170.
[17] Segal VM, Reznikov VI, Drobyshevskiy AE, Kopylov VI. Plastic working of metals
by simple shear. Russian Metall 1981;99 105.
[18] Langdon TG. Research on bulk nanostructured materials in Ufa: twenty years of
scientific achievements. Mater Sci Eng: A 2009;503:6 9.
[19] Valiev RZ, Kuznetsov RI, Musalimov RSH, Tsenev NK. Low-temperature super-
plasticity of metallic materials. Doklady Akad Nauk SSSR 1988;301:864.
[20] Valiev RZ, Krasilnikov NA, Tsenev NK. Plastic deformation of alloys with
submicron-grained structure. Mater Sci Eng: A 1991;137:35 40.
[21] Valiev RZ, Korznikov AV, Mulyukov RR. Structure and properties of ultrafine-
grained materials produced by severe plastic deformation. Mater Sci Eng: A
1993;168:141 8.
[22] Wang J, Horita Z, Furukawa M, Nemoto M, Tsenev NK, Valiev RZ, et al. An
investigation of ductility and microstructural evolution in an Al 2 3% Mg alloy with
submicron grain size. J Mater Res 2011;8:2810 18.
[23] Langdon TG. The current status of bulk nanostructured materials. Rev Adv Mater
Sci 2012;31:1 4.
[24] Langdon TG. Processing by severe plastic deformation: historical developments and
current impact. Mater Sci Forum 2011;667 669:9 14.
[25] Faraji G, Kim H. Review of principles and methods of severe plastic deformation
for producing ultrafine-grained tubes. Mater Sci Technol 2017;33:905 23.
[26] Rosochowski A. Processing metals by severe plastic deformation. Solid State
Phenom 2005;101 102:13 22.
[27] Azimi A, Tutunchilar S, Faraji G, Givi MB. Mechanical properties and microstruc-
tural evolution during multi-pass ECAR of Al 1100 O alloy. Mater Design
2012;42:388 94.
[28] Babaei A, Faraji G, Mashhadi M, Hamdi M. Repetitive forging (RF) using inclined
punches as a new bulk severe plastic deformation method. Mater Sci Eng: A
2012;558:150 7.
[29] Valiev R. Ultrafine-grained materials produced by severe plastic deformation: an
introduction. Ann Chim, Sci Mater 1996;369 78.
[30] Valiev RZ, Alexandrov I, Islamgaliev R. Processing and properties of nanostructured
materials prepared by severe plastic deformation. Nanostructured materials. United
States: Springer; 1998. p. 121 42.
[31] Mahmoodian R, Annuar NSM, Faraji G, Bahar ND, Razak BA, Sparham M.
Severe plastic deformation of commercial pure titanium (CP-Ti) for biomedical
applications: a brief review. JOM 2017. Available from: https://doi.org/10.1007/
s11837-017-2672-4.
[32] Manafi B, Saeidi M, Shatermashhadi V, Abrinia K, Faraji G. Study on the deforma-
tion behavior of polyamide under the backward extrusion process. J Polym Eng
2015;35:675 87.
Fundamentals of Severe Plastic Deformation 35

[33] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nanostructured materials from
severe plastic deformation. Prog Mater Sci 2000;45:103 89.
[34] Mishra A, Richard V, Grégori F, Asaro RJ, Meyers MA. Microstructural evolution in cop-
per processed by severe plastic deformation. Mater Sci Eng: A 2005;410 411:290 8.
[35] Faraji G, Mashhadi M, Bushroa A, Babaei A. TEM analysis and determination of
dislocation densities in nanostructured copper tube produced via parallel tubular
channel angular pressing process. Mater Sci Eng: A 2013;563:193 8.
[36] Wang K, Tao NR, Liu G, Lu J, Lu K. Plastic strain-induced grain refinement at the
nanometer scale in copper. Acta Mater 2006;54:5281 91.
[37] Mishra A, Kad BK, Gregori F, Meyers MA. Microstructural evolution in copper
subjected to severe plastic deformation: experiments and analysis. Acta Mater
2007;55:13 28.
[38] Sakai T, Belyakov A, Kaibyshev R, Miura H, Jonas JJ. Dynamic and post-dynamic
recrystallization under hot, cold and severe plastic deformation conditions. Prog
Mater Sci 2014;60:130 207.
[39] Kawasaki M, Horita Z, Langdon TG. Microstructural evolution in high purity alu-
minum processed by ECAP. Mater Sci Eng: A 2009;524:143 50.
[40] Xu C, Horita Z, Langdon TG. Microstructural evolution in an aluminum solid
solution alloy processed by ECAP. Mater Sci Eng: A 2011;528:6059 65.
[41] Torabzadeh Kashi H, Bahrami M, Shahbazi Karami J, Faraji G. Microstructure and
mechanical properties of the ultrafine-grained copper tube produced by severe plas-
tic deformation. Iran J Mater Sci Eng 2017;14:32 40.
[42] Xue Q, Beyerlein IJ, Alexander DJ, Gray GT. Mechanisms for initial grain refine-
ment in OFHC copper during equal channel angular pressing. Acta Mater
2007;55:655 68.
[43] Sitdikov O, Avtokratova E, Sakai T, Tsuzaki K. Ultrafine-grain structure formation
in an Al-Mg-Sc alloy during warm ECAP. Metall Mater Trans A
2013;44:1087 100.
[44] Mesbah M, Faraji G, Bushroa A. Characterization of nanostructured pure aluminum
tubes produced by tubular channel angular pressing (TCAP). Mater Sci Eng: A
2014;590:289 94.
[45] Yamashita A, Horita Z, Langdon TG. Improving the mechanical properties of mag-
nesium and a magnesium alloy through severe plastic deformation. Mater Sci Eng: A
2001;300:142 7.
[46] Amani S, Faraji G, Abrinia K. Microstructure and hardness inhomogeneity of fine-
grained AM60 magnesium alloy subjected to cyclic expansion extrusion (CEE). J
Manuf Process 2017;28:197 208.
[47] Amani S, Faraji G, Kazemi Mehrabadi H, Abrinia K, Ghanbari H. A combined
method for producing high strength and ductility magnesium microtubes for biode-
gradable vascular stents application. J Alloys Compd 2017;723:467 76.
[48] Galiyev A, Kaibyshev R, Gottstein G. Correlation of plastic deformation and
dynamic recrystallization in magnesium alloy ZK60. Acta Mater 2001;49:1199 207.
[49] Faraji G, Mashhadi M, Kim H. Microstructure inhomogeneity in ultra-fine grained
bulk AZ91 produced by accumulative back extrusion (ABE). Mater Sci Eng: A
2011;528:4312 17.
[50] Faraji G, Mashhadi MM, Kim HS. Tubular channel angular pressing (TCAP) as a
novel severe plastic deformation method for cylindrical tubes. Mater Lett
2011;65:3009 12.
[51] Faraji G, Mashhadi MM, Kim HS. Microstructural evolution of UFG magnesium
alloy produced by accumulative back extrusion (ABE). Mater Manuf Process
2012;27:267 72.
36 Severe Plastic Deformation

[52] Figueiredo RB, Langdon TG. Grain refinement and mechanical behavior of a mag-
nesium alloy processed by ECAP. J Mater Sci 2010;45:4827 36.
[53] Figueiredo RB, Langdon TG. Principles of grain refinement in magnesium alloys
processed by equal-channel angular pressing. J Mater Sci 2009;44:4758 62.
[54] Figueiredo RB, Langdon TG. The nature of grain refinement in equal-channel
angular pressing: a comparison of representative fcc and hcp metals. Int J Mater Res
2009;100:1638 46.
[55] Ishiyama S, Hanada S, Izumi O. Orientation dependence of twinning in commer-
cially pure titanium. J Jpn Inst Metals 1990;54:976 84.
[56] Paton NE, Backofen WA. Plastic deformation of titanium at elevated temperatures.
Metall Trans 1970;1:2839 47.
[57] Kim I, Jeong W-S, Kim J, Park K-T, Shin DH. Deformation structures of pure Ti
produced by equal channel angular pressing. Scr Mater 2001;45:575 80.
[58] Christian JW, Mahajan S. Deformation twinning. Prog Mater Sci 1995;39:1 157.
[59] Murr LE, Niou CS, Garcia EP, Ferreyra E, Rivas TJM, Sanchez JC. Comparison of
jetting-related microstructures associated with hypervelocity impact crater formation
in copper targets and copper shaped charges. Mater Sci Eng: A 1997;222:118 32.
[60] Huang JY, Wu YK, Ye HQ. Deformation structures in ball milled copper. Acta
Mater 1996;44:1211 21.
[61] Zhao WS, Tao NR, Guo JY, Lu QH, Lu K. High density nano-scale twins in Cu
induced by dynamic plastic deformation. Scr Mater 2005;53:745 9.
CHAPTER 2

Severe Plastic Deformation


Methods for Bulk Samples
2.1 INTRODUCTION
This chapter aims to explain the principles of ultrafine grained (UFG)
and nanostructured bulk metal production processes developed during the
past two decades. Various methods of severe plastic deformation (SPD)
are used to produce bulk materials, sheets, and tubes. In this chapter, we
will only refer to methods suitable for processing bulk-shaped specimens.
Furthermore, we will consider the features of each procedure in detail.
The techniques introduced in this chapter are all top-down approaches in
which materials with fine or nanostructured grains are produced from
coarse-grained materials by SPD.
The first attempt to introduce scientific principles into SPD processing
lies unambiguously in the classic work of Professor Bridgman at Harvard
University [1]. Professor Bridgman carried out a comprehensive series of
experiments on the application of high pressure to bulk solids in 1930
and won the Nobel Prize for Physics in 1946 [2]. This approach was later
further developed by the former Soviet Union and ultimately led to the
procedure known as high-pressure torsion (HPT) [1]. The second major
influence on SPD processing may be traced to the classic work of Dr.
Segal and his colleagues at Minsk in 1980, which led to the development
of the most widely used SPD method: equal-channel angular pressing/
extrusion (ECAP/ECAE) [3]. Over the last three decades, many other
SPD methods have been developed around the world. A brief account of
the history of various SPD developments can be very insightful.
Developed SPD processes suitable for deforming bulk metals include, but
are not limited to, HPT [4], incremental high-pressure torsion (IHPT)
[5], single-task incremental high-pressure torsion (SIHPT) [6],
high-pressure torsion extrusion (HPTE) [7], ECAP [3], rotary-die ECAP,
side-extrusion, multipass ECAP, torsional-equal channel angular pressing
(T-ECAP), expansion ECAP (Exp-ECAP), ECAP with parallel channels,
incremental ECAP (IECAP), dual equal channel lateral extrusion

Severe Plastic Deformation © 2018 Elsevier Inc.


DOI: https://doi.org/10.1016/B978-0-12-813518-1.00002-3 All rights reserved. 37
38 Severe Plastic Deformation

(DECLE), channel angular pressing with converging billets, torsion extru-


sion (TE), multiple direct extrusion (MDE), accumulated extrusion (AE),
pure shear extrusion (PSE), equal channel forward extrusion (ECFE),
C-shape equal channel reciprocating extrusion (CECRE), twist extrusion
(TE), elliptical cross-section spiral equal channel extrusion (ECSEE),
planar twist extrusion (PTE), axisymmetric forward spiral extrusion
(AFSE), multidirectional forging (MDF), cyclic closed-die forging
(CCDF), multiaxial incremental forging and shearing (MAIFS), repetitive
forging (RF), repetitive upsetting (RU), cylinder covered compression (CCC),
repetitive upsetting and extrusion (RUE), cyclic extrusioncompression
(CEC), cyclic expansionextrusion (CEE), accumulative back extrusion
(ABE), cyclic forwardbackward extrusion (CFBE), half-channel angular
extrusion (HCAE), accumulative channel-die compression bonding (ACCB),
machining, twist channel angular pressing (TCAP), twist channel multiangular
pressing (TCMAP), and cyclic extrusion compression angular pressing
(CECAP) [8].
The methods mentioned above are discussed in detail in the following
sections. Later in this chapter, the functional structure of each method is
discussed in detail to specify the development of SPD on the material. In
each of the presented methods, various aspects of the deformation
behavior are considered to compare and determine the efficiency of the
methods.

2.2 HIGH-PRESSURE TORSION


Although the fundamental principles of the HPT process were primarily
investigated in 1930, processing by this approach became of great signifi-
cance only within the last 25 years when it was realized that this plastic
deformation process provides an opportunity to produce UFG and nanos-
tructured (NS) materials with enhanced mechanical properties. The scien-
tific origin of processing by HPT may be found in the classic paper entitled
“On Torsion Combined with Compression” [9]. In this early report,
Bridgman proved that a bar could bear angular displacement much greater
in the presence of longitudinal compression than without the compressive
load. For example, in the case of cast iron, the rupture occurred at an
angular displacement of 20 degrees with a zero compressive load, whereas
the angular displacement was increased to 85 degrees in the presence of a
compressive load with a corresponding decrease in the torque to 80% of its
maximum value [10]. This fundamental concept formed the basis of some
Severe Plastic Deformation Methods for Bulk Samples 39

research done in this field. These researchers focused on the grain refine-
ment [11], phase transitions [12,13], recrystallization [14], consolidation of
powders [15], and ability of high elongation [16].
The principles of the HPT method are schematically illustrated in
Fig. 2.1. The sample, which is a thin disk located between a plunger and
an anvil, is subjected to a high compressive load. The compressive load
decreases the thickness of the disc, and the side surfaces stick to the inner
surface of the anvil’s (shaded areas). The rotation of the anvils and the sur-
face frictional forces deform the disk by torsional straining, and the shear
deformation occurs under a hydrostatic pressure [4].
In this process, for an infinitely small rotation (dθ) and a displacement
(dl 5 rdθ, where r is the radius of the disk), the incremental shear strain
(dγ) can be calculated by [17]:
dl rdθ
dγ 5 5 (2.1)
h h
where h is the disk thickness. Assuming that the thickness of the disk is
independent of the rotation angle (θ), θ 5 2πN , the shear strain (γ) is
given by:
2πNr
γ5 (2.2)
h

Plunger

Sample

Anvil

Figure 2.1 Schematic illustration of HPT processing.


40 Severe Plastic Deformation

where N is the number of revolutions. Finally, in many investigations the


equivalent Von Mises strain is calculated using the following relationship
[1820]:
γ 2 πNr
ε 5 pffiffiffi 5 pffiffiffi (2.3)
3 3 h
The use of Eq. (2.3) is correct for small imposed shear strains but for
large strains, where γc0:8, the equivalent strain is given by [21]:
  h i
2  1=2
ε5 p ffiffi
ffi ln 11γ2 =4 1 γ=2 (2.4)
3
Another relationship has been developed to incorporate the decrease
in thickness of the disk due to the applied pressure (P). For this condi-
tion, the true strain is given by [22]:
h ϕ:r  i1=2  
h0
ε 5 ln 11 2 1 ln (2.5)
h h
where h0 and h denote the initial and final thickness of the sample,
respectively. In practice, Eq. (2.5) may be further simplified because, since
ϕ:r=hc1 and ϕ 5 2πN , it follows that [23]:
ϕr       
h0 ϕr h0 2πNrh0
ε 5 ln 1 ln 5 ln 5 ln (2.6)
h h h2 h2
The two types of this process are designated as unconstrained and con-
strained HPT, as illustrated schematically in Fig. 2.2A and B, respectively.
In unconstrained HPT, the sample is located between the lower and
upper anvils, and it is subjected to an applied pressure load and torsional
straining. Hence, the sample is not constrained on its edges, and the

(A) (B) (C)


Figure 2.2 Schematic illustration of HPT for (A) unconstrained, (B) constrained, (C)
quasiconstrained conditions.
Severe Plastic Deformation Methods for Bulk Samples 41

applied pressure can cause the material flow. In this case, only a minor
hydrostatic pressure is imported to the system due to the frictional forces
acting between the sample and the anvils [4]. On the other hand, in con-
strained HPT, the sample is fitted into a hollow cavity of the lower anvil.
This restricts the outward flow of material when the pressure load and
torsional straining are applied. In this condition, the HPT process is con-
ducted in the presence of an effective hydrostatic compressive stress.
However, it is difficult to achieve a perfect constrained mode, and experi-
ments are often conducted under some limited outward flow as shown in
Fig. 2.2C. In this case, the sample is quasiconstrained in the cavity as
some reduction of the sample’s thickness occurs, but are usually limited to
5%10% and can be considered negligible [24,25]. The limited material
flow refrains the contact of both anvils and creates a hydrostatic pressure
that restricts the free flow of materials out of HPT tool.
An important point should be considered to enhance a desirable
homogeneous ultrafine-grained microstructure in the HPT-processed spe-
cimens. The ratio of the thickness ðhÞ of the sample to the diameter ðdÞ,
h=d, should not exceed a certain limit and this depends on the specimen
material [25]. If the aspect ratio (h/d) of the disc is too large, a significant
axial inhomogeneity across the thickness has been reported in the HPT
processed disc [26].
In general, the HPT method is a powerful method for the fundamen-
tal comprehension of SPDs. The equivalent strain in this technique could
be quite high [26]. The HPT technique has several advantages, including
the following [27]:
• HPT permits a defined continuous variation of strain, while most
SPD processes apply strain to the samples in terms of the cycle by
cycle.
• In the HPT method, the extremely high shear strain can be achieved
in a very simple way.
• Because of the unique nature of higher hydrostatic pressure, relatively
hard and brittle materials can be severely deformed in the HPT
method, which is often impossible by other SPD processes. This is
because the higher hydrostatic pressure increases the workability of the
metals.
• The total torque vs. angle of rotation can be measured comfortably.
This permits an estimation of the development of the flow stress [28].
• A change of the direction of rotation also creates SPD, which is typical
for many other SPD processes [29].
42 Severe Plastic Deformation

It is very important to enhance the material properties homogeneously


for industrial applications. As already mentioned, the applied strain has
inherent radius dependency in the HPT-processed samples and cause
inhomogeneity in the mechanical properties. Also, the other problem
with this technique is that the sample size is small and it is limited to lab-
oratory scale, which can’t be used for industrial and large-scale applica-
tions [27].
There have been several attempts to develop HPT-based methods for
scaling up and processing of relatively larger samples. These methods will
be presented in Chapter 5, Severe Plastic Deformation for Industrial
Applications, and may be considered as a semiscaled-up process. For
example, Sakai et al. [30] developed an HPT process for scaling-up and
use with bulk samples. The facility is illustrated schematically in Fig. 2.3.
It consists of upper and lower anvils and dies that form a hexagon cavity
shape when the load and torsional straining are applied. Since the die has
edges inclined at 5 degrees to the perpendicular, the sample shape con-
verts from the initial shape to the final shape when the load is applied.
Thus, the total height of the cylinder is reduced, and the sample expands
outwards into a barrel-shaped configuration to fill the die [30]. Applying
this model of HPT method, the critical aspect ratio (h=d) increases and
also increases the size of the HPT-processed sample.

Load

Upper anvil

Case Die

Sample

Lower anvil

Rotation

(A) (B)
Figure 2.3 (A) Schematic illustration of the HPT facility and (B) in operation with the
load.
Severe Plastic Deformation Methods for Bulk Samples 43

2.2.1 Incremental High-Pressure Torsion


HPT is primarily used for processing of thin and small disks and cannot be
used for processing of rod-shaped samples. To overcome this problem,
Hohenwarter invented a method called IHPT that can produce relatively
large specimens with ultrafine or nanocrystalline microstructures [5]. The
processing steps of IHPT are depicted in Fig. 2.4. The cylindrical sample is
inserted between the anvils and is confined by the height of two support
cylinders. Each anvil also has a small opening to apply shear deformation in
the deformation zone, and this opening also ensures that excess material
can flow out. When the sample is pressed, the entire specimen assumes the
shape of the anvils, and the free material between the anvils bulges and
the excess material flows out of the openings as illustrated in Fig. 2.4B.
The excess material is bound between the anvils and supplies the hydro-
static pressure and prevents abrasion from the adjoining anvils. Then, one
anvil is rotated against the other as shown in Fig. 2.4B. The tangential
friction forces are created by cone-shaped anvils and the sample with the
compressive load. The sample is therefore fixed to the anvil, allowing free

Upper anvil

Support Openings Excess New support


cylinders Gap
cylinders material

Sample
Deformation
zone

(A) Lower anvil (B) (C)

Deformed Deformed Deformed


volume volume volume

New
deformation
zone

(D) (E) (F)


Figure 2.4 Schematic of IHPT illustrating: (A) the initial setup, (B) applying the load
and rotating the anvils relative to each other, (C) changing of the support cylinders,
(D) next deformation step, (E) deformed volume after second deformation step, and
(F) final stage of the deformed specimen after several deformation steps.
44 Severe Plastic Deformation

shear deformation only in the deformation zone during the process. After a
certain number of rotations, the sample is unloaded. Then, the different
support cylinders are used, where the lower one has a larger height and the
top support cylinder is for the same amount small height. As a result, a
small gap is produced between the support cylinder and the sample as
shown in Fig. 2.4C. The lower supporting cylinder then pushes the sample
upwards to fill the gap. The new undeformed material is therefore trans-
formed to the deformation zone (see Fig. 2.4D). However, the excess
material from the first deformation step cannot move and remains in the
openings. Then, one anvil is again rotated against the other for a certain
number of rotations as shown in Fig. 2.4D, and the new undeformed
material placed in the deformation zone gets deformed. This incremental
process increases the deformed volume (Fig. 2.4E). This process can be
repeated multiple times and the volume of the deformed material increases
continuously as shown in Fig. 2.4F. The growth of the deformed material
zone is controlled by the height change of the support cylinders between
consecutive deformation steps and defines a certain step size [5]. The main
disadvantage of the IHPT process is that it takes a long time and the die
needs to be disassembled in each shifting step [6].
Similar to the conventional HPT, the equivalent plastic strain of the
IHPT process is calculated using the following equation [31]:
γ 2 πNr
ε 5 pffiffiffi 5 pffiffiffi (2.7)
3 3 h
where N is the number of rotations, r denotes the sample radius, and h is
the approximate thickness of the deformation zone in the each step.
IHPT can successfully process relatively larger rod-shaped samples
instead of the small disk and can be used in industrial applications.
However, there are still several difficulties in the process, including it
being time-consuming, expensive, and the inhomogeneous microstruc-
ture of the final product. The following methods may be better than
IHPT due to simplicity in the processing steps.

2.2.2 Single-Task Incremental High-Pressure Torsion


A new IHPT process called SIHPT has been developed to improve the
sample size limitations of the HPT method, and it is more convenient to
produce large NS and UFG metallic rods. Also, SIHPT is less time-
consuming and doesn’t require disassembling of the die in each shifting
Severe Plastic Deformation Methods for Bulk Samples 45

step [6]. Fig. 2.5 shows the different steps of the SIHPT process. Using
multipiece die instead of one-piece die is an innovation of the SIHPT
design compared to the IHPT process. Each piece of die is called a step-
per, and the thickness of the stepper determines the deformation zone in
each step. Steppers have a hole in their center to insert a rod-shaped sam-
ple. Because the sample length is high, then different sections of the sam-
ple are deformed step by step. The processing of the sample begins at the
bottom and finishes at the top of the sample. At the initial step, only the
lowest stepper rotates (Fig. 2.5A). This causes a small region of the sample
to be plastically sheared. In the next step, all steppers are fixed except two
at the bottom of the sample (Fig. 2.5B). Then, the two bottom steppers
rotate at the same time. This causes the other small lengths of the sample
to be processed and the deformed zone increases. This algorithm is
repeated until the entire length of the sample is plastically sheared
(Fig. 2.5C and D). The application of a high-pressure load establishes the
desired contact between the sample and steppers and prevents the sample
from slipping during rotation. The equivalent plastic strain of the SIHPT
process can be computed similarly to the IHPT relationship [6].

(A) Axial load (B)

Fixed
steppers

Deformation
Fixed zone
(C) (D)
Unprocessed zone

Deformation zone

Fixed
steppers
Rotating
steppers

Rotating
steppers

Figure 2.5 Schematic of different steps of the SIHPT method [6].


46 Severe Plastic Deformation

2.2.3 High-Pressure Torsion Extrusion


The other method based on HPT is high-pressure torsion extrusion
(HPTE), which was presented in 2016 to improve the limitations in sample
size. In other words, the process may increase the efficiency of the cyclic
expansion extrusion (CEE) process [32,33] for achieving UFG and NG
microstructures. When the fraction of shear to total strain increases more
grain refinement is achieved. Higher shear strains can be obtained in HPTE
processes in comparison with the conventional CEE process. A schematic of
this process is illustrated in Fig. 2.6. At the beginning of the process, the
channels of all containers are filled with the material. This aim can be
accomplished by either using a specially shaped initial sample or by closing
the outlet with a plunger and then pressing. In the case of using a plunger,
the plunger is removed once the containers are filled. The punch (moving at
velocity V ) acts to extrude the material, while the two containers remain
fully filled during the extrusion. The deformation zone consists of three sec-
tions: expansion, shear, and extrusion zones. During the process, one of the
containers rotates with velocity ω, creating the shear zone. The reduction of
the channel diameter in the extrusion section and friction act in the direc-
tion opposite to the extrusion direction and produce hydrostatic pressure in
the deformation zone. Once the sample is completely extruded for the full
length of the punch, the process is stopped and the punch is pulled out, and
the next sample is loaded into the container. D0 and D2 are the input and
output diameters of the container channels, respectively, and D1 is the diam-
eter of die where they meet. The value of these parameters can be changed
to control the level of hydrostatic pressure in the shear zone [7].

Translational motion (V)

Punch
Upper
Deformation zone

container

Extension zone
D.
Shear zone Sample
Extrusion zone D1 L1
L2
Lower container

D2
Rotational motion (ω)

Figure 2.6 Schematic diagram of the HPTE process.


Severe Plastic Deformation Methods for Bulk Samples 47

The accumulated strain after one pass of the process can be calculated
from the following equation [7]:
D1 D1 1 ωR D1
ε 5 2ln 1 2ln 1 pffiffiffi (2.8)
D0 D2 3 V D2
where ðD1 =D2 Þ is the specimen diameter ratio, V is the translational
velocities, ω is the rotational speed of the lower container, and R is the
sample radius. The HPTE method allows implementing simple shear con-
ditions and higher hydrostatic pressure into a rod-shaped sample. In addi-
tion, it can provide the desired condition to produce UFG rods with a
larger length in comparison with the IHPT method. This feature is very
important for commercialization of the HPTE process [7].

2.3 EQUAL-CHANNEL ANGULAR PRESSING


Equal-channel angular passing (ECAP), also known as equal-channel
angular extrusion (ECAE), was first developed by Segal et al. in the 1980s
in the Soviet Union [17,3440]. Bridgeman introduced extraordinary
properties for materials that endured shear stresses under higher hydro-
static pressure [10]. Initially, the ECAP method received little attention,
however, in the 1990s, the outstanding capability of the ECAP method in
the production of UFG and NG metals was realized through further
investigations [41,42]. This section discusses the various ECAP methods
in terms of functional structure.

2.3.1 Conventional ECAP


The ECAP method was first invented by Segal in 1977 and introduced in
the former Soviet Union [43], and its results became publicly available in
international scientific papers [44,45]. After that, Prof. Segal presented
further research on the underlying concepts of the conventional ECAP
method and determination of the applied strain calculations [4650].
The schematic of a conventional ECAP method is shown in Fig. 2.7. It
consists of a die with two equal cross-section channels placed with at an
channel angle of ϕ. In this method, ϕ is usually considered to be 90
degrees, but other angles such as 60 or 120 degrees are also used. The
sample is machined slightly smaller than the channel size for easy place-
ment into the inlet channel. Then the sample is pressed into the deforma-
tion zone using a plunger. As a result, a large shear plastic strain is
imposed on the sample in the deformation zone.
48 Severe Plastic Deformation

Plunger

Die Sample

ϕ
ψ

Figure 2.7 Schematic of a conventional ECAP process for processing of bulk


samples.

The equivalent strain after N passes, εN , may be expressed in a general


form by the relationship [51]:
N

εN 5 pffiffiffi 2cotðφ= 2 1 ψ= 2Þ 1 ψcosecðφ= 2 1 ψ= 2Þ : (2.9)


3
As can be seen, the Von Mises strain depends on the amount of chan-
nel angle (ϕ) and the angle of curvature (ψ), which are typically chosen
as 90 and 20 degrees, respectively. This leads to an equivalent plastic strain
of about 1 for each pass. Eq. (2.9) is consistent with an earlier estimate of
the strain where a die was analyzed with ψ 5 0degree, the strain after N
passes was estimated as [45]:
2N
εN 5 pffiffiffi cotðφ= 2Þ (2.10)
3
Though a significant amount of strain is imposed on the sample in the
ECAP method, the cross-section of the sample remains constant. Thus,
this method can be conducted periodically on the sample, and a large
cumulative strain can be achieved [52]. This is one of the important char-
acteristics of SPD processing over conventional metal forming processes
[53]. The mechanism of shear stress on the shear plane during the ECAP
process is shown in Fig. 2.8. For clarity, the channel angle is considered
to be 90 degrees. A simple shear on the shear plane converts element 1 to
element 2. Therefore, it is believed that the deformation occurs locally at
the intersection of two channels, and the small region is called the shear

You might also like