You are on page 1of 43

Metal Forming

Fundamentals and
Applications

Ramy A. Mohamed
6th March 2018
Contents

Preface iii

1 Metal Forming Processes 1

1.1 Characteristics of forming methods . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Forming processes as system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2.1 Independent variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2.2 Dependent variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.3 Independent-dependent variable interrelations . . . . . . . . . . . . . . . 3

1.3 General parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4 Classification of forming methods . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4.1 Hot or cold working . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.4.2 Forming methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.5 Forming limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Flow Curves 13

2.1 Tensile test geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2 Measured variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.1 Engineering variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2.2 True variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Analysis of work hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4 Necking, or the end of uniform elongation . . . . . . . . . . . . . . . . . . . . . 23

2.5 Strain-rate sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

i
ii  CONTENTS

2.6 Physical significance of m and n . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 Rolling 31

3.1 Work of Deformation method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.2 Slab method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

References 37
Preface

Before we proceed to the details of the subject of the book it might be helpful to introduce the
subject of study and how it evolves naturally from our human curiosity to understand the world
around us.

If we consider a loaded structure such as a building or a machine element, such as rod or beam.
One of the questions that we ponder for these and more complex structures is the following: what
is the mechanism of transmission of load? The general answer to this question is: deformation. It
took millennia of empirical familiarity with natural and human-made structures before this simple
answer could be arrived at. Indeed, the majestic Egyptian pyramids, the beautiful Greek temples,
the imposing Roman arches, the overwhelming Gothic cathedrals and many other such structures
were conceived, built and utilized without any awareness of the fact that their deformation, small as
it might be, plays a crucial role in the process of transmission of load from one part of the structure
to another.

In an intuitive picture, one may say that the deformation of a continuous medium is the manifestation
of the change in atomic distances at a deeper level, a change that results in the development of
internal forces in response to the applied external loads. although this simple model should not be
pushed too far, it certainly contains enough physical motivation to get the general picture and to
be useful in many applications.

Once the role of the deformation has been recognized, we need to organize the understanding in a
threefold activities around the following questions [3]:

1. How is the deformation of a continuous medium described mathematically?


2. What are the physical laws applicable?
3. How do different materials respond to various external loads?

This subdivision of the discipline is not only useful for learning purposes, but also meaningful in
the scientific knowledge sense. The answers to the three questions just formulated are included,
respectively, under the following three disciplines:

1. continuum kinematics;
2. mechanical balance laws;
3. constitutive theory.

From a mathematical standpoint, continuum kinematics is a direct application of a branch of


mathematics known as differential geometry.

The physical balance laws that apply to all continuous media, regardless of their material constitution,
are mechanical (balance of mass, linear momentum, angular momentum) and thermodynamical

iii
iv  Preface

(balance of energy, entropy production). In some applications, electromagnetical, chemical and


other laws may be required. The fact that all these laws are formulated over a continuous entity,
rather than over a discrete collection of particles, is an essential feature of continuum hypothesis.

Finally, by not directly incorporating the more fundamental levels of physical details (molecular,
atomic, subatomic). Solid mechanics must introduce phenomenological descriptors of material
behavior. Thus, connecting the loads to deformations. In other words, geometrically identical
pieces of different materials will undergo very different deformations under the application of the
same loads. One may think that the only considerations to be kept in mind are purely experimental.
Nevertheless, there are some principles that can be established a priori on theoretical grounds,
thus justifying the name of constitutive theory for this fundamental third pillar of the discipline. In
particular, the introduction of ideal material models, such as elasticity, viscoelasticity and plasticity,
has proven historically useful in terms of proposing material responses that can be characterized by
means of a relatively small number of parameters to be determined experimentally [3].

When a solid body is subjected to a force of small magnitude it deforms elastically such that the
strain is directly proportional to the stress and when relieved of the stress it eventually returns to
its original dimensions. Elastic deformation is therefore a reversible or recoverable process. The
well known theory of elasticity is concerned with the mathematical study of stress and strain in
elastically deformed solids.

Under the influence of a substantial force a solid body may experience inelastic, plastic deformation
which is an irreversible or irrecoverable process and the body is permanently deformed. Actually,
the so-called elastic body is an idealization because all solid bodies exhibit more or less plastic
behavior even when subjected to small forces. However, this permanent deformation is so small as
to be practically immeasurable [8].

The present book is concerned with the later, the plastic deformation of metals. The chapters
are organized in a manner that corresponds to the three divisions mentioned earlier. The first
part covers the basics of the plasticity theory. It starts with the mathematical description of the
deformation in three dimensions (continuum kinematics). Then we proceed to the way mechanical
conversation laws applies to the internal forces that arises in response to the deformation.

The generalization of the experimentally measured values to the more complex loading conditions
are the subject of the rest of the part concerned with the theory of plasticity, this particular facet
of the onset of the yielding is the first step to distinguish different materials subject to the same
loading condition but responding differently (i.e. constitutive theory).

In the second part of the book, we apply what was learned in the first part to specific metal forming
processes, the emphasis is on the underlying assumptions and their corresponding limitations when
applied to a specific process.

Each chapter ends with a set of problems. Solving those problems is a very important learning
activity. I cannot stress enough on the importance of the solving the problems.
1
Metal Forming Processes

Metal processing is the branch of engineering that deals with the manufacture of parts, machines,
and structures by the processes of forming, machining, welding and casting among others.

A more general classification system of the metal processing and manufacture techniques are:

Group I-Primary forming: Original creation of shape from the molten or gaseous state or from solid
particles of undefined shape, that is, creating cohesion between particles of the material.
Group II-Deforming: Converting a given shape of a solid body to another shape without change in mass
or material composition, that is maintaining cohesion.
Group III-Separation: Machining or removal of material, that is, destroying cohesion.
Group IV-Joining: Uniting of the individual workpieces to form sub-assemblies, filling and impregnation of
workpiece, and so on, that is, increasing cohesion between several workpieces.
Group V-Coating: Application of thin layers to a workpiece, for example, galvanization, painting, coating
with plastic foils, that is, creating cohesion between substrate and coating.
Group VI-Changing the material properties: Deliberate modification of the workpiece properties in order
to achieve optimum characteristics at a particular point in the manufacturing process.

Metal forming is used synonymously with deformation or deforming and comprises the methods in
group II of the manufacturing process classification introduced earlier.

The term metal forming refers to a group of manufacturing methods by which the given shape of a
workpiece is converted to another shape without change in the mass or composition of the material.

Nearly all metal products undergo metal deformation at some stage of their manufacture. By rolling,
cast ingots, strands, and slabs are reduced in size and converted into basic forms such as sheets,
rods, and plates. These forms can then undergo further deformation to produce wire or the myriad
of finished products formed by processes such as forging, extrusion, sheet metal forming, and others.
The deformation may be bulk flow in three dimensions, simple shearing, simple or compound
bending, or complex combinations of these. The stresses induced by these deformations can be
tension, compression, shear, or any of the other varieties. For most of these processes, a wide range
of speeds, temperatures, tolerances, surface finishes, and amounts of deformation are possible.

1
2 Metal Forming Processes

1.1 Characteristics of forming methods

1. The loads and stresses required for deformation are very high. The stresses vary between
50 and 2500 N/mm2 , depending on the method and material used. In forging process, for
example, the load may reach 750 MN [6]. By comparison, a heavy planning machine produces
only a few kiloNewtons for cutting force.
2. The majority of the parts are completely deformed. Because of the high loads involved, the
tools are generally very large, heavy, and as such expensive. The manufacture of metal-forming
tools requires well equipped workshops and highly skilled workers since the tolerances required
approach those of the precision engineering and gage making.
3. Because of the high cost of machinery and tools, certain minimum quantities are prerequisite
for production to become economical. When minimum quantities are assured, the advantages
of the forming methods are a) high productivity and short production times, b) high accuracy,
withing particular tolerances, with regard to dimension and shape, and c) good mechanical
properties of the manufactured components.

1.2 Forming processes as system

Forming processes tend to be complex systems consisting of independent variables, dependent


variables, and independent-dependent interrelations. Independent variables are those aspects of
the process over which the engineer or operator has direct control, and they are generally selected
or specified when setting up a process.

1.2.1 Independent variables

Consider some of the independent variables in a typical forming process:

1. Starting material When specifying the starting material, we may define not only the chemistry of
that material but also its condition. In so doing, we define the initial properties and characteristics.
These may be chosen entirely for ease of fabrication, or they may be restricted by the desire to achieve
the required final properties upon completion of the deformation process.
2. Starting geometry of the workpiece The starting geometry may be dictated by previous processing,
or it may be selected from a variety of available shapes. Economic considerations often influence the
choice.
3. Tool or die geometry This is an area of major significance and has many aspects, such as the
diameter and profile of a rolling mill roll, the bend radius in a sheet-forming operation, the die angle
in wire drawing or extrusion, and the cavity details when forging. Since the tooling will induce and
control the metal flow as the material goes from starting shape to finished product, success or failure
of a process often depends on tool geometry.
4. Lubrication It is not uncommon for friction between the tool and the workpiece to account for more
than 50% of the power supplied to a deformation process. Lubricants can also act as coolants, thermal
barriers, corrosion inhibitors, and parting compounds. Hence, their selection is an important aspect
in the success of a forming operation. Specification includes type of lubricant, amount to be applied,
and method of application.
5. Starting temperature Since material properties can vary greatly with temperature, temperature
selection and control are often key to the success or failure of a metal-forming operation. Specification
Forming processes as system 3

of starting temperatures may include the temperatures of both the workpiece and the tooling.
6. Speed of operation Most deformation processing equipment can be operated over a range of speeds.
Since speed can directly influence the forces required for deformation, the lubricant effectiveness, and
the time available for heat transfer, its selection affects far more than the production rate.
7. Amount of deformation While some processes control this variable through the design of tooling,
others, such as rolling, may permit its adjustment at the discretion of the operator.

1.2.2 Dependent variables

After specification of the independent variables, the process in turn determines the nature and values
of a second set of features. Known as dependent variables, these, in essence, are the consequences
of the independent variable selection. Examples of dependent variables include:

1. Force or power requirements A certain amount of force or power is required to convert a selected
material from a starting shape to a final shape, with a specified lubricant, tooling geometry, speed,
and starting temperature. A change in any of the independent variables will result in a change in the
force or power required, but the effect is indirect. We cannot directly specify the force or power; we
can only specify the independent variables and then experience the consequences of that selection. It
is extremely important, however, that we be able to predict the forces or powers that will be required
for any forming operation. Without a reasonable estimate of forces or power, we would be unable to
specify the equipment for the process, select appropriate tool or die materials, compare various die
designs or deformation methods, and ultimately optimize the process.
2. Material properties of the product While we can easily specify the properties of the starting material,
the combined effects of deformation and the temperatures experienced during forming will certainly
change them. The starting properties of the material may be of interest to the manufacturer, but the
customer is far more concerned with receiving the desired final shape with the desired final properties.
It is important to know, therefore, how the initial properties will be altered by the shape-producing
process.
3. Exit (or final) temperature Deformation generates heat within the material. Hot workpieces cool
when in contact with colder tooling. Lubricants can break down or decompose when overheated or
may react with the workpiece. The properties of an engineering material can be altered by both the
mechanical and thermal aspects of a deformation process. Therefore, if we are to control a process
and produce quality products, it is important to know and control the temperature of the material
throughout the deformation. (Note: The fact that temperature may vary from location to location
within the product further adds to the complexity of this variable.)
4. Surface finish and precision The surface finish and dimensional precision of the resultant product
depend on the specific details of the forming process.
5. Nature of the material flow In deformation processes, dies and tooling generally exert forces or
pressures and control the movement of external surfaces of the workpiece. While the objective of an
operation is the production of a desired shape, the internal flow of material may actually be of equal
importance. As will be shown later in this book, product properties can be significantly affected by the
details of material flow, and that flow depends on all the details of a process. Customer satisfaction
requires not only the production of a desired geometric shape but also that the shape possess the right
set of properties, without any surface or internal defects.

1.2.3 Independent-dependent variable interrelations

Figure 1.1 serves to illustrate the major problem facing metal-forming personnel. On the left side
are the independent variables-those aspects of the process for which control is direct and immediate.
4 Metal Forming Processes

On the right side are the dependent variables-those aspects for which control is entirely indirect.

Figure 1.1L
Schematic representation of a
metal-forming system showing
independent variables, depen-
dent variables, and the various
means of linking the two.

Unfortunately, it is the dependent variables that we want to control, but their values are determined
by the process, as complex consequences of the independent variable selection. If we want to
change a dependent variable, we must determine which independent variable (or combination of
independent variables) is to be changed, in what manner, and by how much. To make appropriate
decisions, therefore, it is important for us to develop an understanding of the independent variable-
dependent variable interrelations.

Understanding the links between independent and dependent variables is truly the most important
area of knowledge for a person in metal-forming. Unfortunately, this knowledge is often difficult to
obtain. Metal-forming processes are complex systems composed of the material being deformed,
the tooling performing the deformation, lubrication at surfaces and interfaces, and various other
process parameters such as temperature and speed.

The number of different forming processes (and variations thereof) is quite large. In addition,
different materials often behave differently in the same process, and there are multitudes of available
lubricants. Some processes are sufficiently complex that they may have 15 or more interacting
independent variables. We can gain information on the interdependencies of independent and
dependent variables in three distinct ways:

1. Experience Unfortunately, this generally requires long-time exposure to a process and is often limited
to the specific materials, equipment, and products encountered during past contact.Younger employees
may not have the experience necessary to solve production problems. Moreover, a single change in
an area such as material, temperature, speed, or lubricant may make the bulk of past experience
irrelevant.
2. Experiment While possibly the least likely to be in error, direct experiment can be both time consuming
and costly. Size and speed of deformation are often reduced when conducting laboratory studies.
Unfortunately, lubricant performance and heat transfer behave differently at different speeds and
sizes, and their effects are generally altered.The most valid experiment, therefore, is one conducted
under full-size and full-speed production conditions-generally too costly to consider to any great
degree. While laboratory experiments can provide valuable insight, caution should be exercised when
extrapolating lab-scale results to more realistic production conditions.
3. Process modeling Here one approaches the process through high-speed computing and one or
more mathematical models. Numerical values are selected for the various independent variables,
and the models are used to compute predictions for the dependent outcomes. Most techniques rely
on the applied theory of plasticity with various simplifying assumptions. Alternatives vary from
General parameters 5

crude, first-order approximations to sophisticated, computer-based methods, such as finite element


analysis. Various models may incorporate strain hardening, thermal softening, heat transfer, and other
phenomena. Solutions may be algebraic relations that describe the process and reveal trends and
relations between the variables or simply numerical values based on the specific input features.

1.3 General parameters

While much metal-forming knowledge is specific to a given process, there are certain features
that are common to all processes, and these will be presented here. It is extremely important to
characterize the material being deformed. What is its strength or resistance to deformation at the
relevant conditions of temperature, speed of deformation, and amount of prior straining? What
are the formability limits and conditions of anticipated fracture? What is the effect of temperature
or variations in temperature? To what extent does the material strain-harden? What are the
recrystallization kinetics? Will the material react with various environments or lubricants? These
and many other questions must be answered to assess the suitability of a material to a given
deformation process. Since the properties of engineering materials vary widely, the details will be
presented in the next chapter.

Another general parameter is the speed of deformation and the various related effects. Some rate-
sensitive materials may shatter or crack if impacted but will deform plastically when subjected to
slow-speed loadings. Other materials appear to be stronger when deformed at higher speeds. For
these speed-sensitive materials, more energy is needed to produce the same result if we wish to do
it faster, and stronger tools may be required. Mechanical data obtained from slow strain rates in
tensile tests may be totally useless if the deformation process operates at a significantly greater rate
of deformation. Speed sensitivity is also greatest when the material is at elevated temperature, a
condition that is frequently encountered in metalforming operations. The selection of hammer or
press for the hot forging of a small product may well depend on the speed sensitivity of the material
being forged.

In addition to the changes in mechanical properties, faster deformation speeds tend to promote
improved lubricant efficiency. Faster speeds also reduce the time for heat transfer and cooling.
During hot working, workpieces stay hotter and less heat is transferred to the tools. Other general
parameters include friction, lubrication and temperature. Both of these are of sufficient importance
that they will be discussed in some detail in chapter ??.

1.4 Classification of forming methods

Many different metal forming processes are utilized in the manufacturing industry and there is
difficulty in formulating a generally accepted classification of these processes. The following are
some possible classifications: (a) cold, warm or hot forming characterized by the homologous
temperature; (b) chip forming or chipless forming characterized by whether or not metal removal is
involved; (c) state of stress in the workpiece, that is, simple or complex; (d) type of stress involved,
for example, tensile, compressive or shear; (e) size of the plastically deforming zone which may be
local or bulk; (f) steady or non-steady state deformation; (g) low or high strain-rate. Thus it may be
6 Metal Forming Processes

appreciated that a metal forming process may not easily be encompassed by a single classification.
Nearly all metal forming processes involve the workpiece being subjected to complex stress states
which can vary from, say, tri-axial compression to biaxial tension. However, shear stresses need not
be considered unless they constitute major stresses to which the workpiece is subjected and are
thus influential in contributing to plastic deformation. Most of the complex stress states can then
be approximated by their principal stress components, that is, by the normal stresses acting on
planes on which shear stresses are absent.

1.4.1 Hot or cold working

Hot working is defined as the plastic deformation of metals at a temperature above the recrys-
tallization temperature. It is important to note, however, that the recrystallization temperature
varies greatly with different materials. Tin is near hot-working conditions at room temperature,
steels require temperatures near 1093 °C, and tungsten does not enter the hot-working regime until
about 2204 °C. Thus the term hot working does not necessarily correlate with high or elevated
temperature, although such is usually the case.

Elevated temperatures bring about a decrease in the yield strength of a metal and an increase
in ductility. At the temperatures of hot working, recrystallization eliminates the effects of strain
hardening, so there is no significant increase in yield strength or hardness, or corresponding
decrease in ductility. The true stress-true strain curve is essentially flat once we exceed the yield
point, and deformation can be used to drastically alter the shape of a metal without fear of fracture
and without the requirement of excessively high forces. In addition, the elevated temperatures
promote diffusion that can remove or reduce chemical inhomogeneities, pores can be welded shut
or reduced in size during the deformation, and the metallurgical structure can often be altered
through recrystallization to improve the final properties. An added benefit is observed for steels,
where hot working involves the deformation of the weak, ductile, face-centered-cubic austenite
structure, which then cools and transforms to the stronger body-centered-cubic ferrite or much
stronger non-equilibrium structures, such as martensite.

From a negative perspective, the high temperatures of hot working may promote undesirable
reactions between the metal and its surroundings. Tolerances are poorer due to thermal contractions,
and warping or distortion can occur due to nonuniform cooling. The metallurgical structure may
also be nonuniform, since the final grain size depends on the amount of deformation, the temperature
of the last deformation/recrystallization, the cooling history after the deformation, and other factors,
all of which may vary throughout a workpiece. While recrystallization sets the minimum temperature
for hot working, the upper limit for hot working is usually determined by factors such as excess
oxidation, grain growth, or undesirable phase transformations. To keep the forming forces as low as
possible and enable hot deformation to be performed for a reasonable amount of time, the starting
temperature of the workpiece is usually set at or near the highest temperature for hot working.

The plastic deformation of metals below the recrystallization temperature is known as cold working.
Here, the deformation is usually performed at room temperature, but mildly elevated temperatures
may be used to provide increased ductility and reduced strength. From a manufacturing viewpoint,
cold working has a number of distinct advantages, and the various cold-working processes have
Classification of forming methods 7

become quite prominent. Recent advances have expanded their capabilities, and a trend toward
increased cold working appears likely to continue. When compared to hot working, the advantages
of cold working include the following:

1. No heating is required.
2. Better surface finish is obtained.
3. Superior dimensional control is achieved since the tooling sets dimensions at room temperature.
As a result, little, if any, secondary machining is required.
4. Products possess better reproducibility and interchangeability.
5. Strength, fatigue, and wear properties are all improved through strain hardening.
6. Directional properties can be imparted.
7. Contamination problems are minimized.

Some disadvantages associated with cold-working processes include the following:

1. Higher forces are required to initiate and complete the deformation.


2. Heavier and more powerful equipment and stronger tooling are required.
3. Less ductility is available.
4. Metal surfaces must be clean and scale-free.
5. Intermediate anneals may be required to compensate for the loss of ductility that accompanies
strain hardening.
6. The imparted directional properties may be detrimental.
7. Undesirable residual stresses may be produced.

The strength levels induced by strain hardening are often comparable to those produced by the
strengthening heat treatments. Even when the precision and surface finish of cold working are not
required, it may be cheaper to produce a product by cold working a less expensive alloy (achieving
the strength by strain hardening) than by heat treating parts that have been hot formed from a
heat-treatable alloy. In addition, better and more ductile metals and an improved understanding of
plastic flow have done much to reduce the difficulties often experienced during cold forming. As an
added benefit, most cold-working processes eliminate or minimize the production of waste material
and the need for subsequent machining-a significant feature with today’s emphasis on conservation
and materials recycling.

Because the cold-forming processes require powerful equipment and product-specific tools or
dies, they are best suited for large-volume production of precision parts where the quantity of
products can justify the cost of the equipment and tooling. Considerable effort has been devoted to
developing and improving cold-forming machinery along with methods to enable these processes
to be economically attractive for modest production quantities. By grouping products made from
the same starting material and using quick-change tooling, cold-forming processes can often be
adapted to small-quantity or just-in-time manufacture.

Other ways of classification, are:

• shape of workpiece or finished product (sheet metal, bar stock, massive).


• size of deformation zone (bulk).
• steady and non-steady state forming.
8 Metal Forming Processes

• low or high strain rate.


• primary or secondary forming processes.

1.4.2 Forming methods

Nearly all metal forming processes involve the workpiece being subjected to complex stress states
which can vary from, tri-axial compression to biaxial tension. However, shear stresses need not be
considered unless they constitute major stresses to which the workpiece is subjected and are thus
influential in contributing to plastic deformation. Most of the complex stress states can then be
approximated by their principal stress components, that is, by the normal stresses acting on planes
on which shear stresses are absent.

Thomsen, Yang and Kobayashi [9] have suggested, based on a scheme originally proposed by
Kienzle, that the kind of stress involved may be the best choice, thus dividing the major industrial
metal forming processes into four main groups:

1. Squeezing group in which the workpiece is subjected principally to a compressive stress state.
The processes in this group normally involve bulk plastic deformation producing considerable
change in the shape of the workpiece. They include forging (upsetting, closed die forging and
coining), forward and backward extrusion, rolling, swaging, spin forging and rotary forging.
2. Drawing group in which the workpiece is subjected principally to a tensile stress state; thus
pulling instead of pushing is implied. The processes in this group are generally limited in the
extent of plastic deformation of the workpiece which can be achieved in a single operation and
are therefore restricted to changes in shape of the workpiece rather than changes in thickness.
The workpiece is thus usually in the form of metal sheet, plate or thin walled tubing. This
group includes sheet, wire and bar drawing, tube drawing, the deep drawing of cylindrical
cup and box shapes and stretch forming.
3. Bending group in which the workpiece is subjected to couples thereby inducing tensile stresses
on one side of the workpiece and compressive stresses on the other with a stress gradient
throughout the thickness of the workpiece. This group of processes are again restricted to
change of shape rather than change in thickness and include straight flanging, stretch flanging
(concave flanges), shrink flanging (convex flanges) and the seaming of sheet or plate.
4. Cutting group are those processes which either separate excess metal from the workpiece in
a single operation or by incremental metal removal. The first category are shearing processes
such as bar cropping, and piercing and blanking of sheet metal which are regarded as chipless
forming processes. The second category are chip forming and include the conventional
machining processes such as turning, drilling, milling, grinding, sawing, broaching and
shaving.

The state of stress in some common forming processes are illustrated in Table 1.1. The corresponding
approximate stress state in the plastically deforming zone and the extent of the deformation zone are
shown. The deformation is classified as being steady or non-steady state for each of the processes.
Classification of forming methods 9

Table 1.1: Some metal forming processes and their stress state.

Process Schematic Diagram State of Stress in Main Part

Rolling

Forging

Extrusion

Shear Spinning

Tube Spinning
10  Metal Forming Processes

Process Schematic Diagram State of Stress in Main Part

Wire and tube drawing

Deep Drawing

Stretching

Swagging and kneading

Straight Bending

1.5 Forming limits

The amount of plastic deformation that may be achieved during metal forming processes is
profoundly influenced by the stress state developed in the workpiece. Plastic deformation can only
occur if shear stresses are present. Thus, a state of uniform tri-axial tensile stress or uniform tri-axial
compressive stress, referred to as a hydrostatic stress state, does not produce plastic deformation
irrespective of the magnitude of the stresses since shear stresses are absent on any arbitrarily
chosen plane. However, if the normal stresses are unbalanced, that is, they are not all of the same
Forming limits  11

magnitude, shear stresses can exist and when a critical magnitude of shear stress is attained the
onset of plastic flow occurs known as the yield condition. It is not important whether yielding occurs
due to the development of shear stresses resulting from an unbalanced tri-axial tensile stress state
or an unbalanced tri-axial compressive stress state. Nevertheless, the extent of plastic deformation
that can be achieved is highly dependent on the nature of the stress state induced.

Plastic deformation is limited by (a) necking which is a phenomenon due to an instability condition
in tension when uniform plastic flow ceases and becomes localized resulting in local thinning of the
workpiece, or (b) buckling which is associated with a transition phenomenon between elastic and
plastic stress states, or (c) fracture which is a separation process, or possibly by a combination of
these three limits. It follows that a limit to forming is imposed when uniform plastic flow ceases and
the forming limit is determined by whichever defect occurs first. For example, in uniaxial tension
the fracture of ductile metals is usually preceded by local plastic flow or necking. Consequently,
the forming limit is determined by necking rather than by fracture. A thin-walled tube subjected
to torsion fails first by buckling followed by fracture, whilst the torsion of a solid bar is limited by
fracture. The forming limits therefore depend on the state of stress induced in the workpiece.

The condition of instability in uniaxial tension producing the onset of necking occurs when the
maximum axial force is attained and at a relatively low strain. The deformation of the workpiece
that can be achieved by the processes in the drawing group is relatively low and is especially so if
the workpiece is subjected to a stretching operation as in stretch forming.

Since buckling is a transition phenomenon between the elastic and plastic stress states, if it occurs
it does so at an early stage in the forming process. A typical example of this is the buckling or
wrinkling of the flange at the commencement of the deep drawing of a cylindrical cup. This limit to
forming is usually suppressed by a change in the stress state induced in the flange by means of the
blank holder pressure.

The extent of the plastic deformation which can be achieved by those processes comprising the
squeezing group is relatively high. This can be attributed to the stress state in the workpiece being
predominantly tri-axial compressive which inhibits necking and fracture. In upsetting a cylinder of
ductile metal, such as copper, at ambient temperature a dimensional change of 20:1 can easily be
obtained and even greater strains can be induced locally when a metal billet is extruded.

In bending of sheet or plate, the outer fibers of the metal are stretched whilst the inner fibers are
compressed. Because of the stress and strain gradients induced in the bent portion of the workpiece,
the onset of necking may be delayed or perhaps eliminated. However, fracture occurs in the outer
fibers if the radius of the bend is too small.

Considerable plastic deformation can be achieved for ductile metals before separation by fracture
occurs in the shearing processes where the intention is to separate the required part from the
workpiece by chipless cutting. In the blanking process the sheared edge profile of the blanked
product is thus composed of a smooth portion due to plastic flow and an irregular portion due to
fracture. Fracture is initiated at the punch and die comer profiles where high stress concentrations
are developed.

If a hydrostatic pressure is superimposed on the normal stress state for a forming process, necking
12  Metal Forming Processes

can be eliminated and provided precautions are taken to ensure that buckling does not occur the
extent of plastic deformation can be extremely high. By increasing the magnitude of the hydrostatic
pressure the initiation of fracture is suppressed. With the advent of high pressure technology, higher
hydrostatic pressures are being utilized in a number of processes, for example, hydrostatic extrusion,
which enables otherwise difficult to form materials such as high speed tool steel and beryllium to
be successfully formed.

1.6 Conclusion

Since plastic deformation is the principal mechanism of transformation of the workpiece from
one shape to another, a sound understanding of the theory of plasticity is needed. The theory of
plasticity is the mathematical formulation of relations of stress and strain in plastically deforming
solids. The first part of this book is devoted to this subject and to methods of problem solving.
The subject is concerned principally with large plastic deformations as they occur in many forming
processes. Elasto-plastic problems in which elastic and plastic deformations are of the same magnitude
are outside the scope of this book, except when needed in spring-back analysis.

The second part will discuss the details of practice, and calculations of the major forming processes
used in industry.
2
Flow Curves

Mechanical testing has been one of the most important tools to collect relevant information about
engineering materials. Tensile properties are used in selecting materials for different applications.
Material specifications often include minimum tensile properties to ensure quality so tests must
be made to guarantee that materials meet these specifications. Tensile properties are also used in
research and development to compare new materials or processes. With plasticity theory, tensile
data can be used to predict a material’s behavior under forms of loading other than uniaxial tension.

Fundamental and practical studies of metal mechanical behavior usually originate with the uniaxial
tension test. Apparently simple and one-dimensional, a great deal of hidden information may be
obtained by careful observation and measurements of the tensile test. On the other hand, the
underlying physical complexity means that interpretation must be quite careful (along with the
procedure followed to conduct the test) if meaningful results are to be realized.

2.1 Tensile test geometry

Standard tensile test analysis is based on an ideal view of the physical problem. A long, thin rod
is subjected to an extension (usually at a constant extension rate) and the corresponding load is
measured. The basic assumptions are that the loading is purely axial and the deformation takes
place uniformly, both along the length of the specimen and throughout the cross-section. Under
these conditions, it is sufficient to measure just two macroscopic quantities for much of the desired
information: extension and load.

Two kinds of tensile specimens are used for standard tests: a round bar for bulk material (plates,
beams, etc.), and a flat specimen for sheet products. Each is subject to ASTM specifications and
has a nominal gage length of 2 inches. The gage length refers to the distance between ends of
an extension gage put on the specimen to measure extension between these points. The reduced
length also known as the deforming length is the length of the specimen that undergoes plastic
deformation during the test. This length may change but should always be significantly longer than
the gage length in order to ensure that the stress state is uniaxial and deformation is quasi uniform
over the gage length. Figure 2.1 shows the general geometry of the round specimen.

13
14  Flow Curves

Figure 2.1L
Standard tensile specimen
shapes [2].

2.2 Measured variables

A standard tensile test is carried out by moving one end of the specimen (via a machine crosshead)
at a constant speed, v , while holding the other end fixed. The primary variables recorded are load
(P ) and extension (∆l ). Note that the extension could be obtained by multiplying v times t (time).
This is done in some cases but usually the "lash" (looseness) in the system necessitates use of an
extension gage-for accuracy [11].

L
Figure 2.2
Engineering stress-strain diagram. After the maximum of the stress-strain curve, deformation localizes to
form a neck [2].

Note that the load-extension variables depend on specimen size, if, for example, the specimen
were twice as large in each direction, the load would be four times as great, and the extension
would be-twice as great. Since we want to measure material properties, we normalize the measured
variables to account for specimen size. The simplest way to do this is to normalize to the original
Measured variables  15

specimen geometry.

2.2.1 Engineering variables

The variables may be defined as:

σe = P/Ao Engineering stress.


e = ∆l/lo Engineering strain. Also called elongation.
Ao Initial cross-sectional area.
lo Initial length (gage length).
∆l = l − lo Change in gage length, extension

Engineering stress has units of force per area, and engineering strain is dimensionless (mm/mm for
example). Engineering strain is often presented as a percentage by multiplying by 100. A typical
engineering stress-strain curve is shown in Fig. 2.2.

The basic segments of the curve are as follows:

Linear segment of the curve

From the origin, 0, the initial straight-line portion is the elastic region, where stress is linearly pro-
portional to strain. When the stress is removed, if the strain disappears, the specimen is considered
completely elastic. The point at which the curve departs from the straight-line proportionality, is
the proportional limit.
σe
Other basic engineering quantities may be derived from this information as follows. E =
e
Young’s modulus, in the elastic range. The elastic limit, P, on Fig. 2.3, may coincide with the
proportionality limit, or it may occur at some greater stress. The elastic limit is the maximum
stress that can be applied without permanent deformation to the specimen. Some curves exhibit a
definite yield point, while others do not. When the stress exceeds a value corresponding to the yield
strength, the specimen undergoes gross plastic deformation. If the load is subsequently reduced to
0, the specimen will remain permanently deformed [7].

Non-linear segment of the curve (uniform strain)

The stress at which plastic deformation or yielding is observed to begin depends on the sensitivity
of the strain measurements. With most materials, there is a gradual transition from elastic to plastic
behavior, and the point at which plastic deformation begins is difficult to define with precision. In
tests of materials under uniaxial loading, three criteria for the initiation of yielding have been used:
the elastic limit, the proportional limit, and the yield strength.
16  Flow Curves

Figure 2.3 L
(a) Typical stress- strain behavior for a metal showing elastic and plastic deformations, the proportional
limit P, and the yield strength σy , as determined using the 0.002 strain offset method. (b) Representative
stress-strain behavior found for some steels demonstrating the yield point phenomenon. Discontinuous
yielding with an upper yield point A and a relatively constant yielding stress B to C [2, 7]. (Right) Lüders
bands or stretcher strains that form when this material is stretched to an amount less than the yield-point
runout (point C) [1].

• Elastic limit shown at point P in Fig. 2.3, is the greatest stress the material can withstand without any
measurable permanent strain remaining after the complete release of load. With increasing sensitivity
of strain measurement, the value of the elastic limit is decreased until it equals the true elastic limit
determined from microstrain measurements. With the sensitivity of strain typically used in engineering
studies (10-4 mm/mm), the elastic limit is greater than the proportional limit. Determination of the
elastic limit requires a tedious incremental loading-unloading test procedure. For this reason, it is
often replaced by the proportional limit.
• Yield strength shown at point σy in Fig. 2.3, is the stress required to produce a small specified amount
of plastic deformation. The usual definition of this property is the offset yield strength determined
by the stress corresponding to the intersection of the stress-strain curve offset by a specified strain.
In the United States, the offset is usually specified as a strain of 0.2% or 0.1% (e = 0.002 or 0.001).
Offset yield strength determination requires a specimen that has been loaded to its 0.2% offset yield
strength and unloaded so that it is 0.2% longer than before the test. The offset yield strength is referred
to in ISO Standards as the proof stress (R p 0.1 or R p 0.2 )- In the EN standards for materials that do
not have a yield phenomenon present, the 0.2% proof strength (R p 0.2 ) or 0.5% (R p 0.5 ) is determined.
The non-proportional elongation is either 0.1%, 0.2%, or 0.5%. The yield strength obtained by an
offset method is commonly used for design and specification purposes, because it avoids the practical
difficulties of measuring the elastic limit or proportional limit [7].
Measured variables  17

Many metals, particularly annealed low-carbon steel, show a localized, heterogeneous type of
transition from elastic to plastic deformation that produces a yield point in the stress-strain curve.
Rather than having a flow curve with a gradual transition from elastic to plastic behavior, such as
Fig. 2.3 (a), metals with a yield point produce a flow curve or a load-elongation diagram similar to
Fig. 2.3 (b). The load increases steadily with elastic strain, drops suddenly, fluctuates about some
approximately constant value of load, and then rises with further strain.

Typical yield point behavior of low-carbon steel is shown in Fig. 2.3 (b). The slope of the initial
linear portion of the stress-strain curve, designated by E , is the modulus of elasticity. The load at
which the sudden drop occurs is called the upper yield point. The constant load is called the lower
yield point, and the elongation that occurs at constant load is called the yield-point elongation. The
deformation occurring throughout the yield-point elongation is heterogeneous. At the upper yield
point, a discrete band of deformed metal, often readily visible, appears at a stress concentration
such as a fillet. Coincident with the formation of the band, the load drops to the lower yield point.
The band then propagates along the length of the specimen, causing the yield-point elongation.

In typical cases, several bands form at several points of stress concentration. These bands are
generally at approximately 45° to the tensile axis. They are usually called Lüders bands, Hartmann
lines, or stretcher strains, and this type of deformation is sometimes referred to as the Piobert effect.
They are visible and can be aesthetically undesirable. When several Lüders bands are formed, the
flow curve during the yield-point elongation is irregular, each jog corresponding to the formation
of a new Lüders band. After the Lüders bands have propagated to cover the entire length of the
specimen test section, the flow will increase with strain in the typical manner. This marks the end
of the yield-point elongation. These bands are also formed in certain aluminum-magnesium alloys.

The stress required to produce continued plastic deformation increases with increasing plastic
strain; that is, the metal strain hardens. As the specimen elongates, its cross-sectional area decreases
uniformly along the gage length. Initially, the strain hardening more than compensates for this
decrease in area, and the engineering stress (proportional to load P ) continues to rise with
increasing strain. Eventually, a point is reached where the decrease in specimen cross-sectional
area is greater than the increase in deformation load arising from strain hardening. This condition
will be reached first at some point in the specimen that is slightly weaker than the rest. All further
plastic deformation is concentrated in this region, and the specimen begins to neck or thin down
locally.

Non-linear segment of the curve (post-uniform strain)

The strain up to the point of necking has been uniform, as indicated on Fig. 2.2. Because the
cross-sectional area is now decreasing far more rapidly than the ability to resist the deformation by
strain hardening, the actual load required to deform the specimen decreases and the engineering
stress continues to decrease until fracture occurs.

The point shown in curve as maximum load σult , i.e. corresponds to the maximum load sustainable
by the specimen and is defined as:
Pmax
σult = (2.1)
Ao
18  Flow Curves

and it also used to define e u the uniform elongation (elongation before necking begins).

The breaking strength defines:

• e t total elongation or e f the fracture strain.


• e pu = e t − e u post uniform elongation
de
The engineering strain rate eÛ is defined as dt , and is the rate at which strain increases. This quantity
can be obtained simply by noting that all the strain takes place in the deformed reduced length L ,
so that the crosshead speed, v is the same as the extension rate of L (See Fig. 2.1). That is,
de dL/Lo v crosshead speed
eÛ = = = = (2.2)
dt dt Lo deforming length
The third equality is correct because the region outside of L is rigid; that is it does not deform, so
that the velocity of all points outside of DL is the same.

2.2.2 True variables

Assume the original tensile test shown in Figure 2.4(a) is stopped at point B, and specimen is
unloaded to point E. If the tensile test is then restarted, the line (B-E) will be followed approximately,
and the specimen will behave as if no interruption occurred. If, instead, we remove the specimen
and hand it to a new person to test, as in Figure 2.4(b), the result will be quite different. The second
person will measure the cross-sectional area and find a new number, Ao0 , because the previous
deformation reduced the width and thickness while increasing the length. The same load will be

Figure 2.4L
(a) (b) Interrupted tensile test.

required to deform the specimen, but the engineering stress will be different: σe = P/Ao0 . Obviously,
if yield stress is to have a real material meaning, the yield strength should be the same σy = σy0 ,
independent of who tests it. Similarly a small extension at point B will produce different measured
engineering strains for the same reason:
∆l ∆l
ea = , eb = (2.3)
lo lo0
Clearly from the previous case, the engineering stress-strain curve does not give a true indication of
the deformation characteristics of a metal, because it is based entirely on the original dimensions
of the specimen and these dimensions change continuously during the test. Also, a ductile metal
Measured variables  19

that is pulled in tension becomes unstable and necks down during the course of the test. Because
the cross-sectional area of the specimen is decreasing rapidly at this stage in the test, the load
required to continue deformation lessens. The average stress based on the original area likewise
decreases, and this produces the downturn in the engineering stress-strain curve beyond the point
of maximum load. Actually, the metal continues to strain harden to fracture, so that the stress
required to produce further deformation should also increase.

If the true stress, based on the actual cross-sectional area of the specimen, is used, the stress-strain
curve increases continuously to fracture. If the strain measurement is also based on instantaneous
measurement, the curve that is obtained is known as true-stress/ true-strain curve. The true stress-
strain curve is also known as a flow curve, because it represents the basic plastic-flow characteristics
of the material. Any point on the flow curve can be considered the yield stress for a metal strained
in tension by the amount shown on the curve. Thus, if the load is removed at this point and then
reapplied, the material will behave elastically throughout the entire range of reloading.

To be able to calculate true stress-strain values, we need real or true strain, an increment of which
refers to an infinitesimal extension per unit of current length. We limit ourselves to a small extension
to insure that the current length is constant and well known. By assuming that the incremental
strain over the current gage length is uniform we can write mathematically that the true strain
increment is d = dl /l (not dl /lo ). We can express the total true strain as a simple integral:
∫ t ∫l
dl l
= d = ⇒ ln (2.4)
l lo
 t =0 lo

Similarly, the real or true stress refers to the load divided by the current cross-sectional area:
P P
σt = not (2.5)
A Ao
Exactly analogous to the definition of engineering-strain rate, the true. strain rate is defined as
d/dt .
As in § 2.2.1, this rate is simply related to the crosshead speed:
d dL/L v crosshead speed
Û = = = = (2.6)
dt dt L current deforming length
l −l
Since e = l o = ll − 1 we can relate the true strain  as:
o o

l
 = ln = ln(e + 1) and e = exp() − 1 (2.7)
lo
The relation between engineering and true stresses cannot be solved until a relationship between the
original and current cross-sectional area (Ao and A ) is known. A material assumption is required
-namely that plastic deformation produces no net change in volume. This condition is called plastic
incompressibility and is a very accurate assumption for metals and for most other liquids and dense
solids [11].

So, consider the original volume Vo = lo Ao and the current volume V = Al , since there is no
20  Flow Curves

change in the volume V = Vo :


lo 1 Ao
A = Ao = Ao =
l exp() (e + 1)
and since Ao = P σe , we can calculate the true stress as:
P P
σt = = A = σe (e + 1) = σe exp() (2.8)
A o
(e +1)

And inversely
σt σt
σe = = = σt exp(−) (2.9)
(e + 1) exp()
These equations is applicable only to the onset of necking for the reasons discussed above. Beyond

Figure 2.5L
Comparison of engineering and true stress-strain curves. Before necking, a point on the true stress-strain
curve (σt −  ) can be constructed from a point on the engineering stress-strain curve (σe − e ) with equations
(2.7) and (2.8). After necking, the cross-sectional area at the neck must be measured to find the true stress
and strain [4].

maximum load, the true strain should be based on actual area or diameter, D , measurements:
  
Ao Do
 = ln = 2 ln (2.10)
A D
Figure 2.5 compares the true-stress/true-strain curve with its corresponding engineering stress-strain
curve. In agreement with equations (2.7) and (2.8), the true-stress/true-strain curve is always to the
left of the engineering curve until the maximum load is reached.

Example 2.1. In a tensile test, a material fractured before necking. The true stress and strain
at fracture were 630 MPa and 0.18, respectively. What is the tensile strength of the material?
Analysis of work hardening  21

Solution: The engineering strain at fracture was e = exp(0.18) − 1 = 0.197. Because


σe = σt /(1 + e ), the tensile strength = 630/1.197 = 526 MPa.

Example 2.2. When a tensile specimen with a diameter of 0.505 in. and a gage length of 2.0
in. was loaded to 10,000 lbs., it was found that the gage length was 2.519 in. Assuming that
the deformation was uniform,

a) compute the true stress and true strain.


b) find the diameter.

Solution:

a) With do = 0.505 in, Ao = 0.200 in2


σe = 10000/0.2 = 50000 psi;
e = (2.519−2.000)/2.000 = 0.259.
σt = σe (1 + e ) = 50000(1.259) = 62950 psi,
 = ln(1 + 0.259) = 0.230.
b) since the plastic deformation is constant volume A/Ao = lo/l = 1/exp( ), and then
d = √ do = √ 0.505 = 0.491 in.
exp() exp(0.23)

2.3 Analysis of work hardening

J. H. Hollomon discovered in 1945 that many engineering alloys, particularly ferrous alloys, obey a
simple true stress-strain relationship in the plastic regime. His equation states that:

σt = K  n (2.11)

where K and n are constants known as the strength coefficient and work-hardening rate, or work-
hardening exponent respectively. This is the mathematical expression of the flow curve. The
constants K and n are determined from the true stress-strain curve by taking logarithms of both
sides of Equation (2.11).

log(σt ) = log(K ) + n log() (2.12)

Note that Equation (2.12) is the equation of a line whose slope is n and intersect the y axis at
log(K ). By taking logarithms of experimental true stress-strain pairs and plotting as a straight line,
K and n can be obtained.

Figure 2.6 shows such a plot for an aluminum alloy. Note that there are three zones. Zone 1 is the
elastic region where σ = E . Zone II is the region of transition between elastic and fully plastic
behavior and the material in Zone III is fully plastic. Strictly equations (2.11) and (2.12) apply only
to the plastic part of the strain, but since the elastic strain is small relative to the plastic strain after
a few percent, that distinction can be ignored.
22  Flow Curves

Figure 2.6L
True stress strain curve of alu-
minum 1100-O plotted on log-
arithmic coordinates.

Example 2.3. The plastic behavior of a metal can be expressed as σt = 500 0.50 MPa. Estimate
the yield strength if a bar of this material is uniformly cold worked to a reduction of r = 0.3.

Solution: The reduction is a measure of ductility, and is defined as:


Ao − A A 1
r = =1− =1−
Ao Ao exp()
So, we can invert the relation to get:
1
= 1−r
exp()
1
exp() =
 1 − r
1
 = ln
1−r

With r = 0.3 we have  = ln( 1−10.3 ) = 0.357, we then substitute in the Hollomon equation to
get the σ = 500(0.357)0.50 = 198.6 MPa. Here σ  t can be interpreted as the new yield strength
1
after a cold reduction corresponding to  = ln .
1−r
Necking, or the end of uniform elongation  23

2.4 Necking, or the end of uniform elongation

Necking is the localization of strain that occurs near the end of a tensile test. Once it has begun,
none of the equations we have developed is applicable because the strains and stresses are no longer
uniform over our length of measurement, the gage length. That is, the current deforming length l
is less than our gage length G . It is therefore necessary to know the limit of uniform elongation.
This limit is also a measure of formability, because it approximates the end of the tensile test for
most metals at room temperature.

Note that the necking phenomenon is a result of the competition between work hardening (σ
increases with increasing  ) and the reduction of cross-sectional area because of continuing extension.
At the start of a tensile test, the strain hardening dominates the geometric softening and the load
increases. Eventually; the reduction of cross-section dominates and the load decreases. These effects
are just balanced when dP = 0. The maximum load point is readily available on load-elongation
curves or engineering stress-strain curves, but also may be found purely by knowing the true
work-hardening law.
Example 2.4. Show that the maximum load in a tension test starts when  = n .

Solution: Since P = σA , and the load is maximum when dP = 0.


So, σdA + Adσ = 0. Rearranging, dσ σ = − A . Substituting d = − A , the maximum load
dA dA

d = σ .
corresponds to dσ
With the power law, σ = K  n and dσ n−1 .
d = nK 
Equating and simplifying, K  n = nK  n−1 , leads to  = n .
Thus maximum load and necking start when the strain equals strain-hardening exponent.

This surprising result, known as the Considére criterion says that the onset of necking and the end
of uniform elongation occur when the true work-hardening rate n exactly equals the true strain.
This criterion is applicable even if the material does not obey the Hollomon equation (2.11). But in
this case we need to interpret n differently. n must be interpreted as the quantity (d ln σ/d ln  ), a
quantity that may vary with strain. In fact, the general form of Considére’s criterion states that:
dσ d ln σ
= σ or = (2.13)
d d ln 
at the onset of instability (at the limit of uniform elongation). Note that dP = 0 is not in general
the proper condition for plastic instability, although this condition yields an identical result for
rate-sensitive materials.
Example 2.5. Show that the maximum load in a tension test starts when  = n .

Solution:

P = σA = K  n Ao exp(−)
taking the natural logarithm

ln P = ln(K ) + n ln() + ln(Ao ) − 


24  Flow Curves

We can now set the slope to zero to obtain


d ln P n
0= = −1
d 

It might be interesting to check the value of the true stress at the onset of necking. The true stress
at maximum load can be expressed as

σu = K  un = K n n (2.14)
A A
Since σult Ao = σu Au = (K n n )Au , σult = (K n n ) Au . Substituting Au = exp(− u ), we arrive at
o o
 n n
σult = K (2.15)
e
where e here is the natural number.
Example 2.6. In the tension test for Figure 2.6, the tensile strength was experimentally
measured as 28000 psi. Is this consistent with the values of n = 0.25 and K = 50000?

Solution: Using equation (2.15),

σult = 50000(0.25/e )0.25 = 27535 psi


This is within 2% so it is reasonable in view of errors in establishing n and K .

2.5 Strain-rate sensitivity

Suppose that we perform two tensile tests on identical specimens but at two different cross-head
velocities v 1 and v 2 where v 2 > v 1 . Many materials are more difficult to deform at higher rates,
so that the superimposed tensile tests will look like the one in Figure 2.7. This effect, strain-rate

L
Figure 2.7
The flow curve of copper in cold forming was
recorded at two different strain rates.

sensitivity, is often described by a power law exactly analogous to the Hollomon equation (2.11) for
Strain-rate sensitivity  25

strain hardening:

σ = K 0Ûm (2.16)

In order to analyze tensile data to obtain m value we write the equations for two cross-head speeds:

σ1 = σ@v 1 = K 0Û1 m
σ2 = σ@v 2 = K 0Û2 m
This equations could be rearranged:
 m
σ1 Û1 σ1 Û1 ln(σ1/σ2 )
  
= ⇒ ln = m ln ⇒m= (2.17)
σ2 Û2 σ2 Û2 ln(Û1/Û2 )
Equation (2.17) shows how to obtain m from two tensile tests conducted at different extension rates.
Since only the ratio of the true strain rates is required, the ratio of known cross-head velocities will
suffice: v 1/v 2 = Û1/Û2 provided they both have the same reduced length. The same is true for the
stresses:
σt1 σe 1 P 1
= =
σt2 σe 2 P 2
The strain-rate sensitivity index may therefore be obtained very simply, without reference to stresses
or strains:
ln(σ1/σ2 ) ln(σe 1/σe 2 ) ln(P1/P2 )
m= = = (2.18)
ln(Û1/Û2 ) ln(Û1/Û2 ) ln(v 1/v 2 )

Example 2.7. Derive the plastic instability point (limit of uniform elongation) for a material
with power-law hardening and strain-rate sensitivity:

σ = K  n Ûm

Solution: We first need to relate specimen length, area, strain and their corresponding strain
rates:
dl dA
d = =−
l A
d 1 dl lÛ AÛ
Û = = = =−
dt dt l  l A  
2 2
Ü
d Û ll − l l lÛ Û Ü Û
l AÜ AÛ
Ü = = = − =− +
dt l2 l l A A
 2  2
Û AÛ
but, for a constant velocity tensile test, lÜ = 0, so Ü = − ll = A = −Û2 . Recalling the rule
of complete derivatives of a general function f = f (x, y):
∂f ∂f
   
df dx dy
= fÛ = +
dt ∂x y dt ∂y x dt
So, if we write generally σ as a function of x and y , we can write the time derivative of the
26  Flow Curves

true stress as:


∂σ dx ∂σ dy
   

σ
Û = = +
dt ∂x y dt ∂y x dt
dy
Then if we replace x by  , dx
dt = Û and y by Û, dt = Ü, we obtain:

∂σ ∂σ ∂σ ∂σ
       

σ
Û = = Û + Ü = Û − Û2 (2.19)
dt ∂ Û ∂Û  ∂ Û ∂Û 
where the quantities in parentheses-are material properties at an instant in time whether the
function σ = σ(, )
Û is known explicitly or not.
For the Considére Criterion, we find the point at which PÛ = 0:

Û + Aσ σ
Û AÛ σ
Û
PÛ = Aσ Û =0 = − , or = Û

σ A σ
For a rate-sensitive material, we then substitute Eq. (2.19)for σ Û :
a

σ ∂ ln σ ∂ ln σ
   
Û
= Û − Û2 = σ
Û (2.20)
σ ∂ Û ∂Û 
Dividing by Û yields the Considére condition (ie. where PÛ = 0) for a strain-rate sensitive
material;
∂ ln σ ∂ ln σ
   
− Û = 1 (2.21)
∂ Û ∂Û 
For  of a power-law material (for strain and strain-rate hardening) n =
 the particular case
∂ ln σ ∂ ln σ
and m = sob :
∂ ln  Û ∂ ln Û 
n n
− m = 1, or  u = (2.22)
 1+m

a where we have used the standard relationship χ = d (l n χ).
b where we utilized ∂ =  ∂ ln  , and similarly for ∂Û = Û∂ ln Û

2.6 Physical significance of m and n

As a simple rule; the work-hardening rate affects the stress-strain curve primarily up to the uniform
strain and the strain-rate sensitivity index affects behavior primarily in the post-uniform or necking
region. Increasing n and m increases the total Strain to failure and therefore increases the formability
of the material [11].

To sum up, we rephrase the condition for plastic stability, (i.e. for plastic deformation to be stable
or uniform for material with m > 0):
n n
+ m > 1 or m  (2.23)
 
In this case the tensile specimen undergoes stable deformation in the region of the Considére
Physical significance of m and n  27

Figure 2.8L
Considére construction for necking in ten-
sion. The critical condition for necking is
dσ/de = σ/1+e [5].

construction (Fig. 2.8) up to the point P on the stress-strain curve, characterized by the condition:
dσ dσ σ dσe dP
= σ or = or = =0 (2.24)
d de 1+e de dl
Further strain occurs in the necked-down region which the specimen ultimately fractures. As shown

n =0
n =0.3
n =0.35
n =0.4
n =0.45
n =0.5
1 n =1
σ/K

0
Figure 2.9L 0 1 1.5 2
Various forms of power curve σ = K  n . ǫ

in Fig. 2.9, the strain-hardening exponent may have values from n = 0 (perfectly plastic or ideal
solid) to n = 1 (elastic solid). For most metals, n has values between 0.10 and 0.50 (see Table 2.1).
28  Flow Curves

Table 2.1: Values for n and K for metals at room temperature [7].

Metal Condition n K MPa


0.05% carbon steel Annealed 0.26 530
SAE 4340 steel Annealed 0.15 641
0.6% carbon steel Quenched and
tempered at 540 °C 0.10 1572
0.6% carbon steel Quenched and
tempered at 705 °C 0.19 1227
Copper Annealed 0.54 320
70/30 brass Annealed 0.49 896

Problems

Problem 2.1. The following plot of load versus extension was obtained using a specimen (shown
in the figure) of an alloy remarkably similar to the aluminum killed steel found in automotive
fenders, hoods, and so forth. The cross-head speed, v, was 3.3 × 10−4 inch/second. The extension
was measured using an extensometer with a gage length of two inches, as shown (G). Eight points
on the plastic part of the curve have been digitized for you. Use these points to help answer the
following questions.

a) Determine the following quantities. Do not neglect to include proper units in your answer.
yield stress Young’s modulus
ultimate tensile strength total elongation
uniform elongation post-uniform elongation
engineering strain rate
b) Construct a table with the following headings, left to-right: extension, load, engineering strain,
engineering stress, true strain, true stress. Fill in for the eight points on graph. What is the
percentage difference between true and engineering strains for the first point? What is the
percentage difference between true and engineering strains for the last point?
Problems  29

c) Plot the engineering and true stress-strain curves on a single graph using the same units.
d) Calculate the work-hardening rate graphically and provide the ln-ln plot along with the value
of n . How does n compare with the uniform elongation in part a? Why?
e) A second tensile test was carried out on an identical specimen of this material, this time using
a cross-head speed of 3.3 × 10−2 inch/second. The load at an extension of 0.30 inch was 763.4 lb.
What is the strain-rate sensitivity index, m , for this material?
Problem 2.2. Consider a steel plate with a yield strength of 40 ksi, Young’s modulus of 30 × 106 psi,
and a Poisson’s ratio of 0.30 loaded under balanced biaxial tension. What is the volume change,
∆V /V , just before yielding?
Problem 2.3. The strain-hardening behavior of an annealed low-carbon steel is approximated by
σ = 700 0.20 MPa.

a) Estimate the yield strength after the bar is cold worked 50%.
b) Suppose another bar of this same steel was cold worked an unknown amount and then cold
worked 15% more and found to have a yield strength of 525 MPa. What was the unknown
amount of cold work?
Problem 2.4. When a brass tensile specimen, initially 0.505 in. in diameter, is tested, the maximum
load of 12,000 lbs was recorded at an elongation of 40%. What would the load be on an identical
tensile specimen when the elongation is 20%?
Problem 2.5. During a tension test the tensile strength was found to be 340 MPa. This strength
was recorded at an elongation of 30%. Determine n and K .
Problem 2.6. * Starting from the basic idea that tensile necking begins at the maximum load point,
find the true strain and engineering strain where necking begins for the following material laws.
Derive a general expression for the form and find the actual strains.

a) Ideal plastic material σ = σo , find the actual strain value for σ = 500.
b) σ = K ( +  o )n , find the actual strain value for σ = 500( + 0.005)0.25 .
c) σ = σo + K  , find the actual strain value for σ = 250 + 350 .
d) σ = K sin(B ), find the actual strain value for, σ = 500 sin(2π ).
e) σ = A[1 − exp(−B )], find the actual strain value for σ = 500[1 − exp(−3)].
Problem 2.7. * Express the tensile strength, in terms of material parameters for the material laws
in Problem 2.6.
Problem 2.8. * For each of the explicit hardening laws presented in Problem 2.6, calculate the true
stress at  = 0.05, 0.10, 0.15, 0.20, 0.25 and plot the results on a ln σ − ln  figure. Use the figure to
calculate a best-fit n value for each material and compare this with the uniform strain calculated in
Problem 2.6. Why are they different, in view of  u = n ?
Problem 2.9. Tensile tests at two cross-head speeds (1 mm/sec and 10 mm/sec) can be fit to the following
hardening laws:
σ = 500( + 0.05)0.25 at V1 = 1 mm/sec
σ = 520( + 0.05)0.25 at V2 = 10 mm/sec,
What is the strain-rate sensitivity index for these two materials? Does it vary with strain? What is
the uniform strain of each, according to the Considére criterion?
30  Flow Curves

Problem 2.10. Repeat Problem 2.9 with two other stress-strain curves;
at V1 = 1 mm/sec, σ = 550 0.25
at V2 = 10 mm/sec, σ = 500 0.20
Plot the stress-strain curves and find the strain-rate sensitivity index at strains of 0.05, 0.15, and
0.25. In view of these results, does Equation (2.16) apply to this material?

Problem 2.11. Consider the engineering stress-strain curves for three materials labeled A, B, and
C in the following illustration.

Qualitatively, put the materials in order in terms of largest-to-smallest strain hardening (n -value),
strain-rate sensitivity (m -value), and total ductility (formability).
3
Rolling

Rolling is the process of reducing the cross-section of the workpiece by passing it between two
rotating rolls. The reduction in cross-section is accompanied by elongation in the direction of
rolling and there may also be lateral spread of the workpiece. Large reductions in cross-section as,
for example, in the rolling of ingots and billets are achieved at elevated temperatures, that is, above
the recrystallization temperature such that the homologous temperature Th > 0.5. This process is
therefore usually referred to as hot rolling. Hot rolling is one of the major industrial methods of
producing bars of rectangular cross-section and the hot rolling mill consists of two large parallel
cylindrical rolls mounted vertically one above the other. Vertical edging rolls may be provided to
control the width of the workpiece during rolling. The mill is consequently described as a 2-high
mill as distinct from a 4-high or more complex mills which are used in cold-rolling. Rolls having
special profiles are used in the hot-rolling of other cross-sections including round, hexagon, channel,
angle and I section.

During rolling, the plastically deforming region is restricted to a zone of small volume. It is thus
possible to process large ingots using mills of moderate capacity. For steel sheet manufacture, ingots
may weigh more than 0.2 MN and have a cross-section of 600 mm square. The rolling operation is
fast and is more economical than forging. A limiting factor in speed of manufacture is the time
required to transport the slab back to the entry side of the rolls. However, this limitation is obviated
in most cases by using reversing mills. This is important because in hot working processes the
workpiece hardens rapidly on cooling. Hot rolling improves the mechanical properties of the cast
metal by homogenizing and refining the structure producing greater strength and toughness.

3.1 Work of Deformation method

The work of deformation was the first method developed to give an approximate magnitude of the
separating force between the rolls. It gives no information on the stress distribution in the deformed
metal. The assumptions made in this analysis:

1. spreading in width direction is negligible, hence plane strain conditions are assumed.
2. deformation is uniform in both the length and thickness direction.
3. the rolls are assumed to remain rigid through the whole process.
4. roll friction is zero.

31
32  Rolling

The work of deformation or the specific internal energy in rolling with notations shown in fig. 3.1,
can be expressed as:
∫ ¯
ui = σ̄d ,
¯ (3.1)
0

y
θd

T/2

F Lo
Li
s

x
h i ld h o

F s

T/2

L
Figure 3.1
Schematic of a strip rolling showing roll
torque T and separating force Fs .

The infinitesimal effective strain can be evaluated from the equation:


r
2 1
d ¯ = √ (d x − d y )2 + (d y − d z )2 + (d z − d x )2 (3.2)

3 2
under plane strain conditions; if the width of the plate is taken in the z direction, and the spread is
neglected, then d z = 0, from which it follows:

d y = −d x (3.3)

upon substitution in Eq.(3.2):


2
d ¯ = √ d y (3.4)
3
For the assumed condition of uniform deformation, d y = dh/h , constant principal stress ratios,
and no work hardening, i.e. σ̄ = σy . With the additional assumption of ideal plasticity we can
integrate Eq.(3.1) to yield:
2 ho
u i = √ σy ln (3.5)
3 hi
the same equation can also be used for work hardening material if σy is considered the mean flow
Slab method  33

stress. The rolling internal specific energy can be more realistic by introducing a shear correction
factor C , i.e.
2 ho
u i = √ C σy ln (3.6)
3 hi
Now, we consider the external energy for the deformation of the workpiece. The roll torque T ,
shown in Fig. 3.1, supplies the required energy given by:
T θd
ue = (3.7)
(hi +ho )ld/2

where θd is the angle of rotation of the rolls corresponding to the sweeping of the deformed volume.
Equating external specific energy to the internal specific energy assuming no losses we obtain:
C (hi + ho )ld σy ln ho/hi
T = √ (3.8)
3 θd
The roll separating force Fs acts at a distance a from the centers of rotation of the rolls. Assuming
that this force is in equilibrium with the roll torque:
T
Fs = (3.9)
2a
For small reduction it is reasonable to assume that a = ld/2. It must be pointed out, however, that
the roll separating force calculated this way is often not even approximately equal to the actual one,
because the roll torque must include the horizontal component of the force acting on the rolls.

3.2 Slab method

Consider what happens when a strip of material of initial thickness ho enters the rolls. As the
strip passes through the roll gap, it first experiences an elastic compression until it yields, then is
subjected to plastic deformation (work-hardening with increasing strain), and on leaving the roll
gap there is elastic recovery to the final thickness hi . In the theory, it will be assumed that the
material is rigid plastic (work hardening only). This means that the contribution of the elastic arcs
(at entry and exit) to separating force and torque will be neglected.

Now, let us consider the constant volume condition, since we are neglecting the spread in the lateral
direction. The width could be assumed to be the same during the whole process. So,

bv o ho = bvh = bv i hi
where b is the width, v is the velocity, and h is the instantaneous thickness of the plate at arbitrary
location in the deformation zone. While ho , v o , hi , and v i are the entry and exit thickness and
velocity respectively.

This mean that, at entry of the strip into the roll gap , the velocity of the strip will ve smaller than
the radial velocity of the roll. Then at some point the relative velocity between the roll and the
strip should become zero.

For the solution of the rolling problems by the slab method, two dimensional assumptions are
34  Rolling

added: 1) the stresses in the x, y, z directions are principal stresses and are uniformly distributed
over any transverse section. 2) the friction between the rolls and the workpiece is of Coulomb type
and sticking is ruled out. With reference to the geometry of Fig. 3.2(a), the stress analysis can be

y
θd

T/2

θ θ

o
N
h x
h i
i o

x
dx
xb

T/2

(a) (b)

L
Figure 3.2
Schematic of rolling stresses under the arc of contact.

done by investigating equilibrium of the forces acting on an infinitesimal slab shown in Fig. 3.2(b)
in the x− direction. This leads to the equation:

h dσx + σx dh + 2p dx(tan θ + µ) = 0 (3.10)

where p = −σy under the assumption that θ is small. Introduction of the yield criterion gives the
relation of p and σx for the present plane strain problem.
2
σx + p = √ σ̄ = 2k (3.11)
3
where σ̄ is again the effective stress or mean effective stress, depending on the the assumed
hardening behavior of the metal being rolled. Introduction of Eq. (3.11) into Eq. (3.10) results in

− hdp + (2k − p)dh + 2pdx(tan θ + µ) = 0 (3.12)


Slab method  35

The further substitution of dx = dh/2 tan θ into the equations yield the the differential equation:
 µ 
hdp − p + 2k dh = 0 (3.13)
tan θ
References

[1] J T. Black and Ronald A. Kohser. DeGarmo’s Materials and Processes in Manufacturing. John
Wiley & Sons, 2008.

[2] William D. Callister and David G. Rethwisch. materials science and engineering: an introduction.
John Wiley & Sons, 9th edition, 2014.

[3] Marcelo Epstein. The Elements of Continuum Biomechanics. Wiley, 2012.

[4] William F. Hosford. Mechanical Behavious of Materials. Cambridge University Press, 2nd edition,
2010.

[5] William F. Hosford and Robert M. Caddell. Metal Forming: Mechanics and Metallurgy. Cambridge
University Press, 4th, 2011.

[6] Kurt Lange, editor. Metal Forming Handbook. SME (Society of manufacturing Engineers), 1985.

[7] Charles Moosbrugger. Representation of stress-strain behavior. In Charles Moosbrugger,


editor, Atlas Of Stress-Strain Curves. ASM (American society of metals), 2nd edition, 2002.

[8] R. A. C. Slater. Engineering Plasticity: Theory and Application to Metal Forming Processes. THE
MACMILLAN PRESS LTD, 1977.

[9] Erich G. Thomsen, Charles T. Yang, and Shiro Kobayashi. Mechanics of Plastic Deformation In
Metal Processing. McMillan, 1965.

[10] Ansel C. Ugural and Saul K. Fenster. Advanced Mechanics of Materials and Applied Elasticity.
Pearson Education inc., 5th edition, 2012.

[11] R. H. Wagoner and J. -L. Chenot. Fundamentals of Metal Forming. John Wiley & Sons, 1997.

37

You might also like