Vous êtes sur la page 1sur 43

COMPARITIVE STUDY OF MECHANICAL

PROPERTIES OF ENGINEERING MATERIAL BY


FOLLOWING SUITABLE TESTING PROCEDURES

Submitted by

Sayantan Mukherjee (071160132020)


Rashmi Sharma (071160132048)
Debarghya Mukherjee (071160132002)
Sayak Sen (071160132027)
Puranjoy Banerjee (071160132002)
Md. Sadique Gazi (071160132015)

Under the supervision of


Mr. Manik Chandra Das

REPORT SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENT FOR


THE DEGREE OF BACHELOR OF TECHNOLOGY IN AUTOMOBILE ENGINEERING OF
WEST BENGAL UNIVERSITY OF TECHNOLOGY
2010-2011 SESSION

DEPARTMENT OF AUTOMOBILE ENGINEERING


MCKV INSTITUTE OF ENGINEERING
243 G.T. ROAD (NORTH) LILUAH
HOWRAH-711204
CONTENTS

Topics Page No.

1) Introduction 03

2) Literature Review 04

3) Progress Made So Far 10

A) Defining the Properties In Details 11

B) A Datasheet on Auto Components 40A

4) Future Plan of the Project 41

5) Bibliography 42

2
INTRODUCTION

As the name suggests the main aim of the project is to study the different mechanical properties
of different engineering materials. By mechanical properties we mean, the properties of a
material that reveal its elastic and inelastic (plastic) behavior when force is applied, thereby
indicating its suitability for mechanical (load-bearing) application, fatigue limit, hardness,
modulus of elasticity, tensile strength, yield strength etc. The mechanical properties of
engineering materials determine the range of usefulness of the material and establish the service
that can be expected. Mechanical properties are also used to help specify and identify
materials. The project has an outer aim to find and determine the applicability of different
engineering materials in automobile. A need for durable components and maintenance-free
operation is pushing materials used in automobiles to their mechanical and physical limits. This
is especially true for specialty alloys in today's hotter-running engines, sensors, solenoids,
computers, and controls. Such applications need strong metals that are nearly impervious to heat
and corrosion, with special electrical or magnetic properties. Additionally, advances such as
camless valve trains, and continuously variable-transmissions will demand much of alloys. Steel
makers continue to keep pace with a lot of new ideas. The main plan of the project proceedings
include discussion of the various mechanical properties of different engineering materials, find
their applicability in automobiles, reasons to be the best and then determine the measurement of
some of the most important mechanical properties by some suitable testing. An automotive body
uses various engineering materials for its different components. The project deals with the
comparative study of the physical properties of these materials. It is important to understand why
a particular component is made up of a particular material. When a particular material is chosen
over all other there must be some particular property associated with that material that allows
smooth functioning of the corresponding component. Any further innovation in the field of
component development can be done only after understanding the advantage of one material
over the other mostly in terms of mechanical properties.

3
LITERATURE REVIEW

For many years researchers kept themselves busy in the search of materials giving the optimum
performance. Studying the grain structure, alloying properties, mechanical properties under
various temperature conditions has always been the first steps towards search for new and
advanced materials. Application of material research is found in all kind of major discoveries,
innovations and modifications. Some of the research works done has been entitled in the
following:-
Ø Title:- Volume Fraction Effects of Silicon Carbide on the Wear Behavior of SiCp-
Reinforced Magnesium Matrix Composites
Journal: - Advanced Materials Research (Volumes 152 - 153)
Author(s): - Song Jeng Huang, Yo Zhi Dai, Yeau Ren Jeng
Abstract: - This paper investigated the volume fraction effect of micro-sized SiC on the
tribological behavior of SiCp reinforced AZ91D Mg-based metal-matrix composites
(MMCs). The Mg MMCs were prepared by the melt-stirring technique for wear tests.
The hardness and coefficient of friction of Mg MMCs increase as increasing volume
fraction of SiC particle in MMCs. The SiCp/AZ91D MMCs exhibit superior wear
resistance under lower and moderate sliding condition. However, the effects of the SiC
particle reinforcements on wear resistance are not as conclusive under severe sliding
condition (50N-1500 rpm for all vol.% of MMCs, 50N-1000rpm for 3 vol. % MMCs),
since the matrix of MMCs were softened at elevated temperature under such severe
condition. However, the incorporation of SiC particles could enhance the wear resistance
of AZ91D matrix alloy for most of the sliding conditions.

Ø Title: - Indentation Size Effect and the Hall-Petch ‘Law’


Journal: - Materials Science Forum (Volume 662)
Author(s):- L.M. Brown
Abstract: - The flow of material out from under regions in compression must occur by
the operation of many slip systems, which together produce rotational flow. Such flow
requires the accumulation of geometrically necessary dislocations, and leads to the
indentation size effect: smaller indents produce higher hardness, a component of the

4
hardness being inversely proportional to the square-root of the indenter size. A pattern of
flow in polycrystals which satisfies both continuity of normal stress and continuity of
matter at boundaries can be achieved by rotational flow, and it leads to a grain-size effect.
Under most circumstances, the flow stress has a component which is inversely
proportional to the square-root of the grain size, the Hall-Petch law. The flow is
accompanied by the build-up of internal stress which can be relieved by intercrystalline
cracking, thereby limiting the cohesive strength of polycrystals. The relationship between
these ideas and traditional views is briefly explained, and an analysis is given of recent
experimental results.

Ø Title:- Residual Microstress of Austenitic Stainless Steel Due to Tensile Deformation


Journal: - Materials Science Forum (Volume 652)
Author(s):- Kenji Suzuki, Takahisa Shobu
Abstract: - Material of the specimen was austenitic stainless steel (SUS316L). The
specimens were given tensile plastic strains from 0% to 55%. The Vickers hardness of
the specimen corresponded to the plastic strain. The residual macrostress was measured
by Mn-Kα radiations. The residual macrostress of the annealed specimen had a small
compression and changed into a tension after ten- sile plastic deformation. The specimen
with 1% plastic strain showed the maximum tensile residual stress. To examine the
dependency of the residual stress on the lattice plane, the residual microstress for each
lattice plane was measured by hard synchrotron X-rays. The residual microstress was
related with Young’s modulus which was calculated by Kro¨ ner model. A new method,
2θ-cos2 χ method, was proposed to solve the problem of coarse grains and it was
excellent in comparison with the sin2 ψ method.

Ø Title:- A Comparative Study on the Microstructures and Mechanical Properties of Al


6061 Alloy and the MMC Al 6061/TiB2/12P
Journal: - Journal of Minerals & Materials Characterization & Engineering, Vol. 9, No.1
Author: - T.V. Christy1, N. Murugan and S. Kumar

5
Ø Title:- A comparative study of the mechanical properties of TiN coatings using the non-
destructive surface acoustic wave method, scratch test and four-point bending test
Journal: - Surface and Coatings Technology, Volume 84, Issues 1-3, October 1996
Author(s):- H. Ollendorf, D. Schneider, Th. Schwarz, G. Kirchhoff and A. Mucha

Ø Title:- Residual Stress Evaluation of Carburized Transmission Steel Gear Using Neutron
and Synchrotron X-Ray Diffraction and Finite Element Methods
Journal: - Materials Science Forum (Volume 652)
Author(s):- Yoshihisa Sakaida, Takanori Serizawa, M. Kawauchi, M. Manzanka
Abstract: - A motorcycle transmission gear of chromium-molybdenum steel with 0.2%C
was carburized in carrier gas. Carburizing process including heating, carburizing,
diffusing and quenching was simulated using elastoplastic finite element method. The
carbon content, hardness, residual strain and residual stress fields of gear were analyzed.
The unstressed lattice plane spacing and residual strains of the interior near the internal
spline of gear were experimentally measured by synchrotron x-ray and neutron
diffraction methods. As a result, the analyzed carbon content and hardness gradients of
gear accorded with the experimental results. The radial, hoop and axial directions of
cylindrical gear were found to be not always principal axes of residual stress field. On the
other hand, the analyzed residual strains in the radial, hoop and axial directions of gear
slightly discorded with the experimental results. Although correlation between the
measured three strains was similar to that of the weighted average of analyzed strains,
residual strain and stress fields of motorcycle transmission gear could not be accurately
predicted at the present finite element analysis. It was concluded that carbon diffusion
phenomenon and resultant hardening could be analyzed by the finite element method, and
the actual interior residual strain and stress fields should be nondestructively measured by
neutron diffraction method.

Ø Title:- Fracture and Strength of Solids VII (ISBN-13 978-0-87849-210-7)


Journal: - Key Engineering Materials, Volumes 462 - 463 , peer reviewed papers of the
Eight International Conference on Fracture and Strength of Solids (FEOFS) 2010, Kuala
Lumpur, Malaysia, June 7-9, 2010

6
Author(s):- Ahmad Kamal Ariffin, Shahrum Abdullah, Aidy Ali, Andanastuti Muchtar,
Mariyam Jameelah Ghazali and Zainuddin Sajuri
Ø Title:- Fatigue Failure under Varying Loading within Fatigue Limit Diagram
Journal: - Materials Science Forum (Volumes 567 - 568), Materials Structure &
Micromechanics of Fracture V
Author(s):- Yoshiyuki Kondo, H. Eda, Masanobu Kubota
Abstract: - Although fatigue limit diagram is defined in principle for constant stress
amplitude condition, it is often considered that fatigue failure would not occur even in
varying loading if applied stresses were kept within the fatigue limit diagram. However,
it was shown in the case of small-notched specimen and fretting fatigue that fatigue
failure occurred in some special case of variable amplitude loading condition even when
all stress amplitudes were kept within the fatigue limit diagram. The cause of this
phenomenon was examined using two-step and repeated two-step stress patterns in which
the first step stress was with zero mean stress and the second step stress had a high mean
stress. A non-propagating crack was formed by the first step stress. This crack functioned
as a pre-crack for the second step stress with high mean stress. Consequently, fatigue
failure occurred even when all stress amplitudes were kept within the fatigue limit
diagram. It was an unexpected fracture caused by the interference effect of non-
propagating crack and mean stress change.

Ø Title:- Overview of Fatigue Behaviour of Ultrafine-Grained Copper Produced by Severe


Plastic Deformation
Journal: - Materials Science Forum (Volumes 567 - 568)
Author(s):- Petr Lukáš, Ludvík Kunz, Milan Svoboda
Abstract: - Fatigue behaviour of ultrafine-grained copper of purity 99.9 % produced by
ECAP technique was studied in a broad region of stress amplitudes. Fatigue strength is
by a factor of about 2 higher than that of conventional-grain-size copper in the broad
region of fatigue lives from 6x103 to 2x1010 cycles. The grain structure is stable and
undergoes only very marginal changes during cycling. Fatigue slip markings on specimen
surface follow the trace of the shear plane of the last ECAP pass. Fatigue notch
sensitivity is also higher than that of conventional-grain-size copper, but not dramatically.

7
The cyclic stress-strain curve of studied copper is temperature insensitive, while its S-N
curve is temperature dependent.
Ø Title: - Modeling of Cracks Crossing an Interface between Dissimilar Elastic Anisotropic
Materials
Journal: - Materials Structure & Micromechanics of Fracture V
Author(s):- Michal Kotoul, Tomáš Profant, Oldrich Sevecek, Martin Krejcir
Abstract: - Matched asymptotic procedure is used to analyze crack crossing a sharp
interface between dissimilar elastic anisotropic materials. The link to the configurational
forces approach is suggested.

Ø Title: - Potential Automotive Uses of Wrought Magnesium Alloys


Conference: - Automotive Technology Development, Detroit, Michigan, October 28 -
November 1, 1996
Author(s):- L. Gaines, R. Cuenca, F. Stodolsky and S. Wu
Abstract: - Vehicle weight reduction is one of the major means available to improve
automotive fuel efficiency. High-strength steels, aluminum (Al), and polymers are
already being used to reduce weight significantly, but substantial additional reductions
could be achieved by greater use of low-density magnesium (Mg) and its alloys. Mg
alloys are currently used in relatively small quantities for auto parts, generally limited to
die castings (e.g., housings). Argonne National Laboratory's Center for Transportation
Research has performed a study for the Lightweight Materials Program within DOE's
Office of Transportation Materials to evaluate the suitability of wrought Mg and its alloys
to replace steel/aluminum for automotive structural and sheet applications. Mg sheet
could be used in body nonstructural and semi-structural applications, while extrusions
could be used in such structural applications as space frames. This study identifies high
cost as the major barrier to greatly increased Mg use in autos. Two technical R&D areas,
novel reduction technology and better hot-forming technology, could enable major cost
reductions.

8
Ø Title: - Dynamic Mechanical Properties of Automotive Thin Sheet Steel in
Tension, Compression and Shear
Journal:- J pm N FRANCE 7 (1 997)
colloque C3, Supplkment au Journal de Physique 111 d'aoiit 1997
Author(s):- M, Quik, K. Labibes, C. Albertini, T. Valentin and P. Magain
Abstract: - Thin sheet steel has been tested at different strain rates ranging under
different deformation modes. The specimens were cut from thin sheet at 0 and 90 degrees
with respect to the rolling direction. It was found that pre-strained specimens in tension
are characterized by higher initial yielding and a strong decrease of the strain hardening.
Specimens at 90 degrees show at increasing strain rate a higher yielding stress with
respect to 0 degrees specimens emphasizing the importance of anisotropic effects in sheet
metals. Furthermore, the equivalent flow curve at relatively large strain values do not
coincide at different deformation modes (shear, compression and tension).

Ø Title: - Performance of Composite Construction Products in Reaction-to-Fire Test


Conference: - COMPOSITES 2006 Convention and Trade Show American Composites
Manufacturers Association
Author(s):- J. Huczek, M. Mehrafza and M. Janssens
Abstract: - A glass-fiber reinforced plastic panel construction product was tested at
SwRI in the ASTM E 84 Steiner tunnel, the ISO 9705 room and the EN 13823 single
burning item (SBI) test apparatus. Most of the tests were conducted in duplicate or
triplicate to assess the repeatability of the measurements. The same product was also
tested in the Cone Calorimeter (ASTM E 1354 and ISO 5660) at different heat flux levels
to obtain ignition, heat release and smoke production properties. This paper summarizes
the results from the tests and compares the corresponding reaction-to-fire classification in
the United States, Australia, Europe and Japan. In addition to the data acquired in the
SwRI project, additional available literature data are provided for similar materials (fiber
reinforced plastic panels) to illustrate the range of product performance in a given
standard fire test.

9
PROGRESS MADE SO FAR

We undertook the project with a specific plan. According to the topic of our project we are
supposed to study some of the mechanical properties of materials. For this reason we are
supposed to go into details of the various mechanical properties a material possess alongwith the
result of such possession, e.g. we all know that in general engine cylinder blocks are made up of
grey cast iron or aluminum alloys although in the late 1990s engine blocks made from plastic and
other experimental materials were being used in prototype cars with the hope of developing more
lightweight, efficient vehicles. This is because grey cast iron or aluminum alloys give great
resistance to wear and tear, are corrosion resistant etc. Therefore, we can conclude that the
property of being more corrosion resistant makes aluminum the best material for engine cylinder
blocks over other materials. The aim of the project is explained by this. We are to first find out
the properties that separate one material from the other and the advantages of such properties.
Some of the common mechanical properties we aim to study are as follows:-
1. Hardness
2. Toughness
3. Ultimate Tensile strength
4. Compressive strength
5. Yield strength
6. Ductility
7. Fatigue strength
8. Brittleness
9. Corrosion resistance
10. Creep
11. Resilience
Note: - The reason why these mechanical properties are considered over many others is that in
the use of different engineering materials in automobiles these properties find their maximum
applicability.

10
Defining The Properties In Details
1) Hardness: - It is the measure of how resistant solid matter is to various kinds of permanent
shape change when a force is applied. Macroscopic hardness is generally characterized by
strong intermolecular bonds, however the behavior of solid materials under force is complex,
therefore there are different measurements of hardness: scratch hardness, indentation hardness,
and rebound hardness. It is dependent on ductility, elasticity, plasticity, strain, strength,
toughness, viscoelasticity, and viscosity. The Metals Handbook defines hardness as "Resistance
of metal to plastic deformation, usually by indentation. However, the term may also refer to
stiffness or temper, or to resistance to scratching, abrasion, or cutting. It is the property of a
metal, which gives it the ability to resist being permanently, deformed (bent, broken, or have its
shape changed), when a load is applied. The greater the hardness of the metal, the greater
resistance it has to deformation. In mineralogy the property of matter commonly described as the
resistance of a substance to being scratched by another substance. In metallurgy hardness is
defined as the ability of a material to resist plastic deformation. The dictionary of Metallurgy
defines the indentation hardness as the resistance of a material to indentation. This is the usual
type of hardness test, in which a pointed or rounded indenter is pressed into a surface under a
substantially static load. Hardness covers several properties: resistance to deformation, resistance
to friction and abrasion. The well known correlation links hardness with tensile strength, while
resistance to deformation is dependent on modulus of elasticity. The frictional resistance may be
divided in two equally important parts: the chemical affinity of materials in contact, and the
hardness itself. So it is easy to understand that surface treatments modify frictional coefficients
and behaviour of the parts in contact. The abrasion resistance is partially related to hardness
(between 2 metallic parts in frictional contact, the less hard one will be the more rapidly worn),
but experiments carried out at Centre de Recherches PECHINEY in Voreppe (CRV), with
TABER test show that the correlation resistance against wear/ hardness presents some
inversions. A correlation may be established between hardness and some other material property
such as tensile strength. Then the other property (such as strength) may be estimated based on
hardness test results, which are much simpler to obtain. This correlation depends upon specific
test data and cannot be extrapolated to include other materials not tested.

11
Hardness Measurement and Its Types

There are three main types of hardness measurements: scratch, indentation, and rebound. Within
each of these classes of measurement there are individual measurement scales. For practical
reasons conversion tables are used to convert between one scale and another.

Scratch hardness is the measure of how resistant a sample is to fracture or plastic


(permanent) deformation due to friction from a sharp object. The principle is that an object made
of a hard material will scratch an object made of a softer material. The most common test
is Mohs scale, which is used in mineralogy. One tool to make this measurement is
the sclerometer.

Indentation hardness measures the resistance of a sample to permanent plastic deformation due
to a constant compression load from a sharp object; they are primarily used in engineering and
metallurgy fields. The tests work on the basic premise of measuring the critical dimensions of an
indentation left by a specifically dimensioned and loaded indenter.

Common indentation hardness scales are Rockwell, Vickers, and Brinell.

Rebound hardness, also known as dynamic hardness, measures the height of the "bounce" of a
diamond-tipped hammer dropped from a fixed height onto a material. This type of hardness is
related to elasticity. The device used to take this measurement is known as a scleroscope. Two
scales that measures rebound hardness are the Leeb rebound hardness test and Bennett hardness
scale.

Hardness measurement can be defined as macro-, micro- or nano- scale according to the forces
applied and displacements obtained. Measurement of the macro-hardness of materials is a quick
and simple method of obtaining mechanical property data for the bulk material from a small
sample. It is also widely used for the quality control of surface treatments processes. However,
when concerned with coatings and surface properties of importance to friction and wear
processes for instance, the macro-indentation depth would be too large relative to the surface-
scale features. Where materials have a fine microstructure, are multi-phase, non-homogeneous or
prone to cracking, macro-hardness measurements will be highly variable and will not identify
individual surface features. It is here that micro-hardness measurements are appropriate.

12
Microhardness is the hardness of a material as determined by forcing an indenter such as a
Vickers or Knoop indenter into the surface of the material under 15 to 1000 gf load; usually, the
indentations are so small that they must be measured with a microscope. Capable of determining
hardness of different microconstituents within a structure, or measuring steep hardness gradients
such as those encountered in casehardening. Conversions from microhardness values to tensile
strength and other hardness scales (e.g. Rockwell) are available for many metals and alloys.
Micro-indenters works by pressing a tip into a sample and continuously measuring: applied load,
penetration depth and cycle time.

Hardness Measurement Methods

There are three types of tests used with accuracy by the metals industry; they are the Brinell
hardness test, the Rockwell hardness test, and the Vickers hardness test. Since the definitions of
metallurgic ultimate strength and hardness are rather similar, it can generally be assumed that a
strong metal is also a hard metal. The way the three of these hardness tests measure a metal's
hardness is to determine the metal's resistance to the penetration of a non-deformable ball or
cone. The tests determine the depth which such a ball or cone will sink into the metal, under a
given load, within a specific period of time.

Rockwell Hardness Test

The Rockwell Hardness test is a hardness measurement based on the net increase in depth of
impression as a load is applied. Hardness numbers have no units and are commonly given in the
R, L, M, E and K scales. The higher the number in each of the scales means the harder the
material. Hardness has been variously defined as resistance to local penetration, scratching,
machining, wear or abrasion, and yielding. The multiplicity of definitions, and corresponding
multiplicity of hardness measuring instruments, together with the lack of a fundamental
definition, indicates that hardness may not be a fundamental property of a material, but rather a
composite one including yield strength, work hardening, true tensile strength, modulus of
elasticity, and others. In the Rockwell method of hardness testing, the depth of penetration of an
indenter under certain arbitrary test conditions is determined. The indenter may either be a steel
ball of some specified diameter or a spherical diamond-tipped cone of 120° angle and 0.2 mm tip
radius, called Brale. The type of indenter and the test load determine the hardness scale(A, B, C,
etc). The hardness of ceramic substrates can be determined by the Rockwell hardness test,

13
according to the specifications of ASTM E-18. This test measures the difference in depth caused
by two different forces, using a dial gauge. Using standard hardness conversion tables, the
Rockwell hardness value is determined for the load applied, the diameter of the indentor, and the
indentation depth. The Rockwell hardness tester to measure the hardness of metal measures
resistance to penetration like the Brinell test, but in the Rockwell case, the depth of the
impression is measured rather than the diametric area. With the Rockwell tester, the hardness is
indicated directly on the scale attached to the machine. This dial like scale is really a depth
gauge, graduated in special units. The Rockwell hardness test is the most used and versatile of
the hardness tests. For soft materials such as copper alloys, soft steel, and aluminum alloys a
1/16" diameter steel ball is used with a 100-kilogram load and the hardness is read on the "B"
scale. In testing harder materials, hard cast iron and many steel alloys, a 120 degrees diamond
cone is used with up to a 150 kilogram load and the hardness is read on the "C" scale. The
Rockwell test uses two loads, one applied directly after the other. The first load, known as the
"minor", load of 10 kilograms is applied to the specimen to help seat the indenter and remove the
effects, in the test, of any surface irregularities. In essence, the minor load creates a uniformly
shaped surface for the major load to be applied to. The difference in the depth of the indentation
between the minor and major loads provides the Rockwell hardness number. There are several
Rockwell scales other than the "B" & "C" scales, (which are called the common scales). The
other scales also use a letter for the scale symbol prefix, and many use a different sized steel ball
indenter. A properly used Rockwell designation will have the hardness number followed by
"HR" (Hardness Rockwell), which will be followed by another letter which indicates the specific
Rockwell scale. An example is 60 HRB, which indicates that the specimen has a hardness
reading of 60 on the B scale. There is a second Rockwell tester referred to as the "Rockwell
Superficial Hardness Tester". This machine works the same as the standard Rockwell tester, but
is used to test thin strip, or lightly carburized surfaces, small parts or parts that might collapse
under the conditions of the regular test. The Superficial tester uses a reduced minor load, just 3
kilograms, and has the major load reduced to either 15 or 45 kilograms depending on the
indenter, which are the same ones used for the common scales. Using the 1/16" diameter, steel
ball indenter, a "T" is added (meaning thin sheet testing) to the superficial hardness designation.
An example of a superficial Rockwell hardness is 15T-22, which indicates the superficial

14
hardness as 22, with a load of 15 kilograms using the steel ball. If the 120¡ diamond cone were
used instead, the "T" would be replaced with "N".
The Rockwell Hardness Scale
A -Cemented carbides, thin steel and shallow case hardened steel
B -Copper alloys, soft steels, aluminum alloys, malleable iron, etc.
C -Steel, hard cast irons, pearlitic malleable iron, titanium, deep case hardened
steel and other materials harder than B 100
D -Thin steel and medium case hardened steel and pearlitic malleable iron
E -Cast iron, aluminum and magnesium alloys, bearing metals
F -Annealed copper alloys, thin soft sheet metals
G -Phosphor bronze, beryllium copper, malleable irons
H -Aluminum, zinc, lead
K, L, M, P, R, S, V -Bearing metals and other very soft or thin materials,
including plastics.
Brinell Hardness Test
Brinell hardness is determined by forcing a hard steel or carbide sphere of a specified diameter
under a specified load into the surface of a material and measuring the diameter of the
indentation left after the test. The Brinell hardness number, or simply the Brinell number, is
obtained by dividing the load used, in kilograms, by the actual surface area of the indentation, in
square millimeters. The result is a pressure measurement, but the units are rarely stated. The
Brinell hardness test method consists of indenting the test material with a 10 mm diameter
hardened steel or carbide ball subjected to a load of 3000 kg. For softer materials the load can be
reduced to 1500 kg or 500 kg to avoid excessive indentation. The full load is normally applied
for 10 to 15 seconds in the case of iron and steel and for at least 30 seconds in the case of other
metals. The diameter of the indentation left in the test material is measured with a low powered
microscope. The Brinell harness number is calculated by dividing the load applied by the surface
area of the indentation. The diameter of the impression is the average of two readings at right
angles and the use of a Brinell hardness number table can simplify the determination of the
Brinell hardness. A well structured Brinell hardness number reveals the test conditions, and
looks like this, "75 HB 10/500/30" which means that a Brinell Hardness of 75 was obtained
using a 10mm diameter hardened steel with a 500 kilogram load applied for a period of 30

15
seconds. On tests of extremely hard metals a tungsten carbide ball is substituted for the steel ball.
Compared to the other hardness test methods, the Brinell ball makes the deepest and widest
indentation, so the test averages the hardness over a wider amount of material, which will more
accurately account for multiple grain structures and any irregularities in the uniformity of the
material. This method is the best for achieving the bulk or macro-hardness of a material,
particularly those materials with heterogeneous structures.

Vickers Hardness Test


It is the standard method for measuring the hardness of metals, particularly those with extremely
hard surfaces: the surface is subjected to a standard pressure for a standard length of time by
means of a pyramid-shaped diamond. The diagonal of the resulting indention is measured under
a microscope and the Vickers Hardness value read from a conversion table. The indenter
employed in the Vickers test is a square-based pyramid whose opposite sides meet at the apex at
an angle of 136º. The diamond is pressed into the surface of the material at loads ranging up to
approximately 120 kilograms-force, and the size of the impression (usually no more than 0.5
mm) is measured with the aid of a calibrated microscope. The Vickers number (HV) is calculated
using the following formula:
HV = 1.854(F/D2),
with F being the applied load (measured in kilograms-force) and D2 the area of the indentation
(measured in square millimetres). The applied load is usually specified when HV is cited.
The Vickers hardness test method consists of indenting the test material with a diamond indenter,
in the form of a right pyramid with a square base and an angle of 136 degrees between opposite
faces subjected to a load of 1 to 100 kgf. The full load is normally applied for 10 to 15 seconds.
The two diagonals of the indentation left in the surface of the material after removal of the load
are measured using a microscope and their average calculated. The area of the sloping surface of

16
the indentation is calculated. The Vickers hardness is the quotient obtained by dividing the kgf
load by the square mm area of indentation.
F= Load in kgf, d = Arithmetic mean of the two diagonals, d1 and d2 in mm, HV = Vickers
hardness

When the mean diagonal of the indentation has been determined the Vickers hardness may be
calculated from the formula, but is more convenient to use conversion tables. The Vickers
hardness should be reported like 800 HV/10, which means a Vickers hardness of 800, was
obtained using a 10 kgf force. Several different loading settings give practically identical
hardness numbers on uniform material, which is much better than the arbitrary changing of scale
with the other hardness testing methods. The advantages of the Vickers hardness test are that
extremely accurate readings can be taken, and just one type of indenter is used for all types of
metals and surface treatments. Although thoroughly adaptable and very precise for testing the
softest and hardest of materials, under varying loads, the Vickers machine is a floor standing unit
that is more expensive than the Brinell or Rockwell machines.

2) Toughness- Toughness, in materials science and metallurgy, is the resistance to fracture of a


material when stressed. It is defined as the amount of energy per volume that a material can
absorb before rupturing. The ability of a metal to deform plastically and to absorb energy in the
process before fracture is termed toughness. The emphasis of this definition should be placed on
the ability to absorb energy before fracture. Recall that ductility is a measure of how much
something deforms plastically before fracture, but just because a material is ductile does not
make it tough. The key to toughness is a good combination of strength and ductility. A material
with high strength and high ductility will have more toughness than a material with low strength
and high ductility. Therefore, one way to measure toughness is by calculating the area under the
stress strain curve from a tensile test. This value is simply called “material toughness” and it has
units of energy per volume. Material toughness equates to a slow absorption of energy by the
material. There are several variables that have a profound influence on the toughness of a
material. These variables are:

17
• Strain rate (rate of loading)
• Temperature
• Notch effect
A metal may possess satisfactory toughness under static loads but may fail under dynamic loads
or impact. As a rule ductility and, therefore, toughness decrease as the rate of loading increases.
Temperature is the second variable to have a major influence on its toughness. As temperature is
lowered, the ductility and toughness also decrease. The third variable is termed notch effect, has
to due with the distribution of stress. A material might display good toughness when the applied
stress is uniaxial; but when a multiaxial stress state is produced due to the presence of a notch,
the material might not withstand the simultaneous elastic and plastic deformation in the various
directions. There are several standard types of toughness test that generate data for specific
loading conditions and/or component design approaches. Three of the toughness properties that
will be discussed in more detail are 1) impact toughness, 2) notch toughness and 3) fracture
toughness.
Impact Toughness- The impact toughness (AKA Impact strength) of a material can be
determined with a Charpy or Izod test. These tests are named after their inventors and were
developed in the early 1900’s before fracture mechanics theory was available. Impact properties
are not directly used in fracture mechanics calculations, but the economical impact tests continue
to be used as a quality control method to assess notch sensitivity and for comparing the relative
toughness of engineering materials. The two tests use different specimens and methods of
holding the specimens, but both tests make use of a pendulum-testing machine. For both tests,
the specimen is broken by a single overload event due to the impact of the pendulum. A stop
pointer is used to record how far the pendulum swings back up after fracturing the specimen. The
impact toughness of a metal is determined by measuring the energy absorbed in the fracture of
the specimen. This is simply obtained by noting the height at which the pendulum is released and
the height to which the pendulum swings after it has struck the specimen . The height of the
pendulum times the weight of the pendulum produces the potential energy and the difference in
potential energy of the pendulum at the start and the end of the test is equal to the absorbed
energy. Since toughness is greatly affected by temperature, a Charpy or Izod test is often
repeated numerous times with each specimen tested at a different temperature. This produces a
graph of impact toughness for the material as a function of temperature. An impact toughness

18
versus temperature graph for a steel is shown in the image. It can be seen that at low
temperatures the material is more brittle and impact toughness is low. At high temperatures the
material is more ductile and impact toughness is higher. The transition temperature is the
boundary between brittle and ductile behavior and this temperature is often an extremely
important consideration in the selection of a material.

Fig1. Explaining Impact Testing

Notch-Toughness- Notch toughness is the ability that a material possesses to absorb energy in
the presence of a flaw. As mentioned previously, in the presence of a flaw, such as a notch or
crack, a material will likely exhibit a lower level of toughness. When a flaw is present in a
material, loading induces a triaxial tension stress state adjacent to the flaw. The material
develops plastic strains as the yield stress is exceeded in the region near the crack tip. However,
the amount of plastic deformation is restricted by the surrounding material, which remains
elastic. When a material is prevented from deforming plastically, it fails in a brittle manner.
Notch-toughness is measured with a variety of specimens such as the Charpy V-notch impact
specimen or the dynamic tear test specimen. As with regular impact testing the tests are often
repeated numerous times with specimens tested at a different temperature. With these specimens
and by varying the loading speed and the temperature, it is possible to generate curves such as
those shown in the graph. Typically only static and impact testing is conducted but it should be
recognized that many components in service see intermediate loading rates in the range of the
dashed red line.

19
Fracture Toughness- Fracture toughness is an indication of the amount of stress required to
propagate a preexisting flaw. It is a very important material property since the occurrence of
flaws is not completely avoidable in the processing, fabrication, or service of a
material/component. Flaws may appear as cracks, voids, metallurgical inclusions, weld defects,
design discontinuities, or some combination thereof. Since engineers can never be totally sure
that a material is flaw free, it is common practice to assume that a flaw of some chosen size will
be present in some number of components and use the linear elastic fracture mechanics (LEFM)
approach to design critical components. This approach uses the flaw size and features,
component geometry, loading conditions and the material property called fracture toughness to
evaluate the ability of a component containing a flaw to resist fracture. A parameter called the
stress-intensity factor (K) is used to determine the fracture toughness of most materials. A
Roman numeral subscript indicates the mode of fracture and the three modes of fracture are
illustrated in the image to the right. Mode I fracture is the condition in which the crack plane is
normal to the direction of largest tensile loading. This is the most commonly encountered mode
and, therefore, for the remainder of the material we will consider KI
The stress intensity factor is a function of loading, crack size, and structural geometry. The stress
intensity factor may be represented by the following equation:

Where: KI is the fracture toughness in


s is the applied stress in MPa or psi
a is the crack length in meters or inches
is a crack length and component geometry factor that is different for each
B
specimen and is dimensionless.

3) Ultimate Tensile Strength- The ultimate tensile strength (UTS) or, more simply, the tensile
strength, is the maximum engineering stress level reached in a tension test. The ultimate tensile
strength (UTS) is the maximum resistance to fracture. It is equivalent to the maximum load that
can be carried by one square inch of cross-sectional area when the load is applied as simple
tension. It is the maximum stress that a material can withstand while being stretched or pulled
before necking, which is when the specimen's cross-section starts to significantly contract.
Mechanically, tensile strength is opposite of compressive strength, although in many materials

20
the magnitudes of these two strengths are quite different. The strength of a material is its ability
to withstand external forces without breaking. In brittle materials, the UTS will be at the end of
the linear-elastic portion of the stress-strain curve or close to the elastic limit. In ductile
materials, the UTS will be well outside of the elastic portion into the plastic portion of the stress-
strain curve. On the stress-strain curve above, the UTS is the highest point where the line is
momentarily flat. Since the UTS is based on the engineering stress, it is often not the same as the
breaking strength. In ductile materials strain hardening occurs and the stress will continue to
increase until fracture occurs, but the engineering stress-strain curve may show a decline in the
stress level before fracture occurs. This is the result of engineering stress being based on the
original cross-section area and not accounting for the necking that commonly occurs in the test
specimen. The UTS may not be completely representative of the highest level of stress that a
material can support, but the value is not typically used in the design of components anyway. For
ductile metals the current design practice is to use the yield strength for sizing static components.
However, since the UTS is easy to determine and quite reproducible, it is useful for the purposes
of specifying a material and for quality control purposes. On the other hand, for brittle materials
the design of a component may be based on the tensile strength of the material.

Fig.2 Explaining the Stress-Strain Curve

21
4) Compressive Strength- Compressive strength is the capacity of a material to withstand
axially directed pushing forces. When the limit of compressive strength is reached, materials are
crushed. Concrete can be made to have high compressive strength, e.g. many concrete structures
have compressive strengths in excess of 50 MPa, whereas a material such as soft sandstone may
have a compressive strength as low as 5 or 10 MPa. By definition, the compressive strength of a
material is that value of uniaxial compressive stress reached when the material fails completely.
The compressive strength is usually obtained experimentally by means of a compressive test.

The compressive strength of the material would correspond to the stress at the red point shown
on the curve. Even in a compression test, there is a linear region where the material
follows Hooke's Law. Hence for this region σ = Eε where this time E refers to the Young's
Modulus for compression. In theory, the compression test is simply the opposite of the tension
test with respect to the direction of loading. In compression testing the sample is squeezed while
the load and the displacement are recorded. Compression tests result in mechanical properties
that include the compressive yield stress, compressive ultimate stress, and compressive modulus
of elasticity. Compressive yield stress is measured in a manner identical to that done for tensile
yield strength. When testing metals, it is defined as the stress corresponding to 0.002 in./in.
plastic strain. For plastics, the compressive yield stress is measured at the point of permanent
yield on the stress-strain curve. Moduli are generally greater in compression for most of the
commonly used structural materials. Ultimate compressive strength is the stress required to
rupture a specimen. This value is much harder to determine for a compression test than it is for a
tensile test since many material do not exhibit rapid fracture in compression. Materials such as
most plastics that do not rupture can have their results reported as the compressive strength at a
specific deformation such as 1%, 5%, or 10% of the sample's original height. For some materials,

22
such as concrete, the compressive strength is the most important material property that engineers
use when designing and building a structure. Compressive strength is also commonly used to
determine whether a concrete mixture meets the requirements of the job specifications. It is the
maximum stress a material can sustain under crush loading. The compressive strength of a
material that fails by shattering fracture can be defined within fairly narrow limits as an
independent property. However, the compressive strength of materials that do not shatter in
compression must be defined as the amount of stress required to distort the material an arbitrary
amount. Compressive strength is calculated by dividing the maximum load by the original cross-
sectional area of a specimen in a compression test.

5) Yield strength- The yield strength or yield point of a material is defined


in engineering and materials science as the stress at which a material begins to deform
plastically. Prior to the yield point the material will deform elastically and will return to its
original shape when the applied stress is removed. Once the yield point is passed some fraction
of the deformation will be permanent and non-reversible. In the three-dimensional space of the
principal stresses (σ1,σ2,σ3), an infinite number of yield points form together a yield surface.
Knowledge of the yield point is vital when designing a component since it generally represents
an upper limit to the load that can be applied. It is also important for the control of many
materials production techniques such as forging, rolling, or pressing. In structural engineering,
this is a soft failure mode which does not normally cause catastrophic failure or ultimate
failure unless it accelerates buckling. Yield strength is one of the types of tensile strength. Yield
strength is defined as the yield stress, which is actually the stress level at which a permanent
deformation of 0.2% of the original dimension of the material happens, and is defined as the
stress level at which a material can withstand the stress before it is deformed permanently.
Yield strength is the amount of stress at which plastic deformation becomes noticeable and
significant. The figure is an engineering stress-strain diagram in tensile test. Because there is no
definite point on the curve where elastic strain ends and plastic strain begins, the yield strength is
chosen to be that strength when a definite amount of plastic strain has occurred. For the general
engineering structural design, the yield strength is chosen when 0,2 percent plastic strain has
taken place. The 0.2% yield strength or the 0.2% offset yield strength is calculated at 0.2% offset
from the original cross-sectional area of the sample (s=P/A). During yielding stage, the material

23
deforms without an increase in applied load, but during the strain hardening stage, the material
undergoes changes in its atomic and crystalline structure, resulting in increased resistance of
material to further deformation. Yield strength is a very important value for use in engineering
structural design. If we are designing a component that must support a force during use, we must
be sure that the component does not plastically deform. We must therefore select a material that
has high yield strength, or we must make the component large enough so that the applied force
produces a stress that is below the yield strength. In contrast, the tensile strength is relatively
unimportant for ductile materials selection and application since too much plastic deformation
takes place before it is reached. However, the tensile strength can give some indication of the
materials, such as hardness and material defects.

The composition affects the yield strength only through the variation in alloying elements such
as carbon (C), chromium (Cr), manganese (Mn), molybdenum (Mo), nickel (Ni) and silicon (Si).
Because of the mechanical property degradation experienced by the nuclear power plant
components, it is important to evaluate the evolution of the mechanical properties during service,
such as the yield strength. A lack of materials available from nuclear power plants has led to the
development of more and more small-sized specimen tests such as the small-punch test that has
the advantage of using very small amounts of material with very simple working conditions
(machining and test). A whole new method of measurement of yield strength has been proposed
as Punch test. This method is based on the elastic deformation energy method and is compared
with the current methods which include the “two tangents” method and the offset method. This
new method consists of the determination of the yield strength by measuring the energy beneath

24
the load-displacement curve in the elastic deformation domain. The elastic deformation domain
is defined by measuring the elastic reverse displacement during an unloading phase.
Subsequently, the elastic displacement reverse is applied at the beginning of the load-
displacement curve. Good agreement has been found between the elastic deformation energy and
the square of the tensile yield strength. This method allows a better evaluation of the tensile yield
strength and with a reduced scatter compared to the current conventional methods.

6) Ductility- Ductility is a mechanical property that describes the extent in which solid materials
can be plastically deformed without fracture. In materials science, ductility specifically refers to
a material's ability to deform under tensile stress; this is often characterized by the material's
ability to be stretched into a wire. It is the physical property of being capable of sustaining large
permanent changes in shape without breaking (in metals, such as being drawn into a wire). It is
characterized by the internal structure of the material flowing under shear stress. In Earth
science the brittle-ductile transition zone is a zone, at an approximate depth of 15 km (9 mi) in
continental, at which rock becomes less likely to fracture and more likely to deform ductility. In
glacial ice this zone is at approximately 30 m (100 ft) depth. It is still possible for material above
a brittle-ductile transition zone to deform ductility, and possible for material below to deform
brittle. The zone exists because as depth increases confining pressure increases, and brittle
strength increases with confining pressure whilst ductile strength decreases with increasing
temperature. The transition zone occurs at the point where brittle strength exceeds ductile
strength. Ductility is especially important in metalworking, as materials that crack or break under
stress cannot be manipulated using metal forming processes, such as hammering, rolling,
and drawing. Malleable materials can be formed using stamping or pressing, whereas brittle
metals and plastics must be molded. High degrees of ductility occur due to metallic bonds, which
are found predominantly in metals and leads to the common perception that metals are ductile in
general. In metallic bonds valence shell electrons are delocalized and shared between many
atoms. The delocalized electrons allow metal atoms to slide past one another without being
subjected to strong repulsive forces that would cause other materials to shatter. Ductility can be
quantified by the fracture strain , which is the engineering strain at which a test specimen
fractures during a uniaxial tensile test. Another commonly used measure is the reduction of area
at fracture q. The following list ranks some of the metals according to their mechanical property

25
measured as the greatest ductility to
least: gold, silver, platinum, iron, nickel, copper, aluminum, zinc, tin, and lead. The malleability
of the same metals are then ranked from greatest to least: gold, silver, lead, copper, aluminum,
tin, platinum, zinc, iron, and nickel. The ductility of steel varies depending on the alloying
constituents. Increasing levels of carbon decreases ductility. Many plastics and amorphous
solids, such as Play-Doh, are also malleable. The ductile-brittle transition temperature (DBTT),
nil ductility temperature (NDT), or nil ductility transition temperature of a metal represents the
point at which the fracture energy passes below a pre-determined point (for steels typically 40
J for a standard Charpy impact test). DBTT is important since, once a material is cooled below
the DBTT, it has a much greater tendency to shatter on impact instead of bending or deforming.
For example, zamak 3 exhibits good ductility at room temperature but shatters at sub-zero
temperatures when impacted. DBTT is a very important consideration in materials selection
when the material in question is subject to mechanical stresses. A similar phenomenon, the glass
transition temperature, occurs with glasses and polymers, although the mechanism is different in
these amorphous materials. In some materials this transition is sharper than others. For example,
the transition is generally sharper in materials with a body-centered cubic (BCC) lattice than
those with a face-centered cubic (FCC) lattice. DBTT can also be influenced by external factors
such as neutron radiation, which leads to an increase in internal lattice defects and a
corresponding decrease in ductility and increase in DBTT. The most accurate method of
measuring the BDT or DBT temperature of a material is by fracture testing. Typically, four point
bend testing at a range of temperatures is performed on pre-cracked bars of polished material.
For experiments conducted at higher temperatures, dislocation activity increases. At a certain
temperature, dislocations shield the crack tip to such an extent the applied deformation rate is not
sufficient for the stress intensity at the crack-tip to reach the critical value for fracture (KiC). The
temperature at which this occurs is the ductile-brittle transition temperature. If experiments are
performed at a higher strain rate, more dislocation shielding is required to prevent brittle fracture
and the transition temperature is raised. Ductility is strictly defined as the ability of a material to
be drawn into a wire. More generally, it is used to refer to the amount of plastic deformation a
material can endure before failure. Ductility is not a uniquely defined material property,
quantifiable by a single test. Instead, there are a number of tests that are used to give a measure

26
of ductility. The test chosen will usually depend on the application in which the alloy is to be
used.

Explaining Ductility Before Breaking

7) Fatigue Strength- Fatigue limit, endurance limit, and fatigue strength are all expressions used
to describe a property of materials: the amplitude (or range) of cyclic stress that can be applied to
the material without causing fatigue failure. Ferrous alloys and titanium alloys have a distinct
limit, amplitude below which there appears to be no number of cycles that will cause failure.
Other structural metals such as aluminum and copper, do not have a distinct limit and will
eventually fail even from small stress amplitudes. In these cases, a number of cycles (usually
107) are chosen to represent the fatigue life of the material. The ASTM defines fatigue
strength, SNf, as the value of stress at which failure occurs after Nf cycles, and fatigue limit, Sf, as
the limiting value of stress at which failure occurs as Nf becomes very large. ASTM does not
define endurance limit, the stress value below which the material will withstand many number of
load cycles, but implies that it is similar to fatigue limit. Some authors use endurance limit, Se,
for the stress below which failure never occurs, even for an indefinitely large number of loading

27
cycles, as in the case of steel; and fatigue limit or fatigue strength, Sf, for the stress at which
failure occurs after a specified number of loading cycles, such as 500 million, as in the case of
aluminium. Other authors do not differentiate between the expressions even if they do
differentiate between the two types of materials. Typical values of the limit (Se) for steels are 1/2
the ultimate tensile strength, to a maximum of 100 ksi (690 MPa). For iron, aluminium, and
copper alloys, Se is typically 0.4 times the ultimate tensile strength. Maximum typical values for
irons are 24 ksi (165 MPa), aluminums 19 ksi (131 MPa), and coppers 14 ksi (96.5 MPa).
Fatigue considerations are important because the consequent failure is generally sudden and at a
stress level much lower than the ultimate stress. Fatigue properties of materials are generally
determined by producing Wohler /S-N Plots. These are simply plots with stress as the vertical
axis and log (number of complete stress reversals) as the horizontal axis. A number of material
specimens are tested and the points at which they break are plotted on the S-N curve. It is a
useful property of steel (and titanium) that when the stress level fall below a certain value the
specimen is effectively never likely to fail. Generally other materials do not exhibit this effect.
The fatigue strength is the maximum completely reversed stress under which a material will fail
after it has experienced the stress for a specified number of cycles. (The strength is accompanied
by the number of cycles). ..Fatigue Strength (fixed number of cycles) = Sn The Fatigue limit is
the maximum completely reversed stress for which it is assumed that the material will never fail
regardless of the number of cycles. Fatigue Limit = S'n Experiments have shown little direct
relationship between the fatigue limit and the yield strength, ductility etc. However some
relationship between the fatigue limit and the tensile strength Su has been established for
unotched polished specimens tested using the rotating beam method. This method loads the
specimens by reversed bending. The following table gives some important information regarding
fatigue limit of some of the important engineering materials:-
S'e = 0,5 Su for Wrought Steels where Su < 1400mpa
S'e = 690MPa for Wrought Steels where Su > 1400MPa
S'e = 0,5 Su for Titanium
S'e = 0,4 Su for cast steel and cast iron
for magnesium casting and wrought alloys (based on
S'n = 0,38 Su
10 6 cycle life)

28
S'n = 0,35 Su ->0,5 Su for nickel alloys (based on 10 8 cycle life)
S'n = 0,25 Su ->0,5 Su for copper based alloys (based on 10 8 cycle life)
for for wrought aluminium alloys up to a strength of
S'n = 0,38 Su
280 MPa (based on 5 x 10 8 cycle life)
for for cast aluminium alloys up to a strength of 350
S'n = 0,16 Su
MPa (based on 5 x 10 8 cycle life)
Note: The fatigue limit S'n a is pseudo limit based on a number of stress cycles this applies to the
engineering metals which will eventually fail at some time if subject to continuous reversing
/repeated stress cycles. Ferrous metals and titanium can operate continuously without failure at
stress levels at or below the stress limit S'e.
Note; All of the above relationships are based on a 50% survival life.
The fatigue limit for reversed axial load of a polished, unnotched specimen is about 15% lower
than that for reversed bending. The fatigue limit for torsional testing of polished unnotched
specimens is
• S'es is about 0,58 x the fatigue limit in reversed bending for steel.
• S'es is about 0,8 x the fatigue limit in reversed bending for cast iron.
• S'ns is about 0,48 x the fatigue limit in reversed bending for copper.
The above values are all experimentally derived under relatively ideal conditions.
These values should be modified using factors that take into account actual operating conditions.
Approximations for endurance limits for three types of loading for steel are as follows
• Bending S'e is about 0,5 Su
• Axial S'e is about 0,45 Su
• Torsion S'e is about 0,29 Su

12) Brittleness- A material is brittle if, when subjected to stress, it fails without significant
deformation (strain). Brittle materials absorb relatively little energy prior to fracture, even those
of high strength. Failure is often accompanied by a snapping sound. Brittle materials include
most ceramics and glasses (which do not deform plastically) and some polymers, such
as PMMA and polystyrene. Many steels become brittle at low temperatures (see ductile-brittle
transition temperature), depending on their composition and processing. When used in materials

29
science, it is generally applied to materials that fail in tension rather than shear, or when there is
little or no evidence of plastic deformation before failure. When a material has reached the limit
of its strength, it usually has the option of either deformation or fracture. A
naturally malleable metal can be made stronger by impeding the mechanisms of plastic
deformation (reducing grain size, dispersion strengthening, work hardening, etc.), but if this is
taken to an extreme, fracture becomes the more likely outcome, and the material can become
brittle. Improving material toughness is therefore a balancing act. The term brittleness is used
differently by different authors. The various phenomena associated with brittle substances, e.g.
low value of elongation, fracture failure, formation of fines, higher ratio of compressive to
tensile strength, higher angle of internal friction, formation of cracks in indentation, etc., can be
used for measuring brittleness. Experimental results based on the measurement of some of the
above quantities to represent brittleness have been presented. A discussion is made to compare
the results of different formulation. It appears that there is no uniformity in different formulation
of brittleness, e.g. brittleness based on strain ratio, brittleness based on energy ratio, brittleness
from Mohr's envelope, brittleness from compressive and tensile strength, brittleness from
Protodyakonov Impact Test, brittleness from macro- and micro-hardness, etc. Each concept of
brittleness should, therefore, be treated and used separately with reference to its practical utility.
Naturally brittle materials, such as glass, are not difficult to toughen effectively. Most such
techniques involve one of two mechanisms: to deflect or absorb the tip of a propagating crack, or
to create carefully-controlled residual stresses so that cracks from certain predictable sources will
be forced closed. The first principle is used in laminated glass where two sheets of glass are
separated by an interlayer of polyvinyl butyral, which as a viscoelastic polymer absorbs the
growing crack. The second method is used in toughened glass and pre-stressed concrete. A
demonstration of glass toughening is provided by Prince Rupert's Drop. Brittle polymers can be
toughened by using rubber particles to initiate crazes when a sample is stressed, a good example
being high impact polystyrene or HIPS. The least brittle structural ceramics are silicon
carbide (mainly by virtue of its high strength) and transformation-toughened zirconia. Generally,
the brittle strength of a material can be increased by pressure. This happens as an example in
the brittle-ductile transition zone at an approximate depth of 10 kilometres (6.2 mi) in the Earth's
crust, at which rock becomes less likely to fracture, and more likely to deform ductility. Failure
of brittle materials can be determined using several approaches:

30
§ Phenomenological failure criteria
§ Linear elastic fracture mechanics
§ Energy-based methods
Phenomenological failure criteria- The failure criteria that were developed for brittle solids
were the maximum stress/strain criteria. The maximum stress criterion assumes that a material
fails when the maximum principal stress σ1 in a material element exceeds the uniaxial tensile
strength of the material. Alternatively, the material will fail if the minimum principal stress σ3 is
less than the uniaxial compressive strength of the material. If the uniaxial tensile strength of the
material is σt and the uniaxial compressive strength is σc, then the safe region for the material is
assumed to be
σc < σ3 < σ1 < σt
Note that the convention that tension is positive has been used in the above expression.
The maximum strain criterion has a similar form except that the principal strains are compared
with experiment all determined uniaxial strains at failure, i.e.,

The maximum principal stress and strain criteria continue to be widely used in spite of severe
shortcomings. Numerous other phenomenological failure criteria can be found in the engineering
literature. The degree of success of these criteria in predicting failure has been limited. For brittle
materials, some popular failure criteria are
§ criteria based on invariants of the Cauchy stress tensor
§ the Tresca or maximum shear stress failure criterion
§ the von Mises or maximum elastic distortional energy criterion
§ the Mohr-Coulomb failure criterion for cohesive-frictional solids
§ the Drucker-Prager failure criterion for pressure-dependent solids
§ the Bresler-Pister failure criterion for concrete
§ the Willam-Warnke failure criterion for concrete
§ the Hankinson criterion, an empirical failure criterion that is used for orthotropic
materials such as wood.
§ the Hill yield criteria for anisotropic solids
§ the Tsai-Wu failure criterion for anisotropic composites

31
§ the Johnson-Holmquist damage model for high-rate deformations of isotropic
solids
§ the Hoek-Brown failure criterion for rock masses
Linear elastic fracture mechanics- The approach taken in linear elastic fracture mechanics is to
estimate the amount of energy needed to grow a preexisting crack in a brittle material. The
earliest fracture mechanics approach for unstable crack growth is Griffiths' theory [3]. When
applied to the mode I opening of a crack, Griffiths' theory predicts that the critical stress (σ)
needed to propagate the crack is given by

where E is the Young's modulus of the material, γ is the surface energy per unit area of the crack,

and 2a is the crack length. The quantity is postulated as a material parameter called the
'fracture toughness. The mode I fracture toughness is defined as

and is determined experimentally. Similar quantities KIIc and KIIIc can be determined for mode
II and model III loading conditions. The state of stress around cracks of various shapes can be
expressed in terms of their stress intensity factors. Linear elastic fracture mechanics predicts that
a crack will extend when the stress intensity factor at the crack tip is greater than the fracture
toughness of the material. Therefore the critical applied stress can also be determined once the
stress intensity factor at a crack tip is known.
Energy-based methods- The linear elastic fracture mechanics method is difficult to apply for
anisotropic materials (such as composites) or for situations where the loading or the geometry are
complex. The strain energy release rate approach has proved quite useful for such situations.
The strain energy release rate for a mode I crack which runs through the thickness of a plate is
defined as

where P is the applied load, t is the thickness of the plate, u is the displacement at the point of
application of the load due to crack growth, and 2a is the length of the crack. The crack is

32
expected to propagate when the strain energy release rate exceeds a critical value GIc - called
the critical strain energy release rate.
The fracture toughness and the critical stain energy release rate are related by

where E is the Young's modulus. If an initial crack size is known, then a critical stress can be
determined using the strain energy release rate criterion.

8) Corrosion Resistance- Corrosion is the disintegration of an engineered material into its


constituent atoms due to chemical reactions with its surroundings. In the most common use of
the word, this means electrochemical oxidation of metals in reaction with an oxidant such
as oxygen. Formation of an oxide of iron due to oxidation of the iron atoms in solid solution is a
well-known example of electrochemical corrosion, commonly known as rusting. This type of
damage typically produces oxide(s) and/or salt(s) of the original metal. Corrosion can also refer
to other materials than metals, such as ceramics or polymers, although in this context, the term
degradation is more common. In other words, corrosion is the wearing away of metals due to a
chemical reaction. Many structural alloys corrode merely from exposure to moisture in the air,
but the process can be strongly affected by exposure to certain substances. Corrosion can be
concentrated locally to form a pit or crack, or it can extend across a wide area more or less
uniformly corroding the surface. Because corrosion is a diffusion controlled process, it occurs on
exposed surfaces. Some metals are more intrinsically resistant to corrosion than others, either
due to the fundamental nature of the electrochemical processes involved or due to the details of
how reaction products form. If a more susceptible material is used, many techniques can be
applied during an item's manufacture and use to protect its materials from damage. The materials
most resistant to corrosion are those for which corrosion is thermodynamically unfavorable.
More common "base" metals can only be protected by more temporary means. Some metals have
naturally slow reaction kinetics, even though their corrosion is thermodynamically favorable.
These include such metals as zinc, magnesium, and cadmium. While corrosion of these metals is
continuous and ongoing, it happens at an acceptably slow rate. An extreme example is graphite,
which releases large amounts of energy upon oxidation, but has such slow kinetics that it is
effectively immune to electrochemical corrosion under normal conditions. Iron, in its various

33
forms, is exposed to all kinds of environments. It tends to be highly reactive with most of them
because of its natural tendency to form iron oxide. When it does resist corrosion it is due to the
formation of a thin film of protective iron oxide on its surface by reaction with oxygen of the air.
This film can prevent rusting in air at 99% RH, but a contaminant such as acid rain may destroy
the effectiveness of the film and permit continued corrosion. Thicker films of iron oxide may act
as protective coatings, and after the first year or so, could reduce the corrosion rate significantly
as shown in the following figure.

While the corrosion rate of bare steel tends to decrease with time in most cases, the difference in
corrosivity of different atmospheres for a particular product is tremendous. Similar ranges in
corrosivity were determined by the ISO 9223 corrosion rates for steel. In a few cases, the
corrosion rates of ferrous metals have been reported as increasing with time, and careful analysis
of the exposure conditions generally reveals that an accumulation of contaminating corrosive
agents has occurred, thus changing the severity of the exposure. It is generally conceded that
steels containing low amounts of copper are particularly susceptible to severe atmospheric
corrosion. In one test over a 3 1/2-year period in both a marine and an industrial atmosphere, a
steel containing 0.01 % copper corroded at a rate of 80 µm/y, whereas increasing the copper
content by a factor of five reduced the corrosion rate to only 35 µm/y. Other tests comparing
gray cast iron, malleable iron, and low-alloy steels indicated that their corrosion resistances were
approximately the same. Plain cast iron appears to have a corrosion rate about one half that
of 0.2% copper steel in a marine atmosphere. One has to be careful in citing such differences to

34
stipulate the composition of the carbon steel because corrosion behavior of carbon steels is
influenced so markedly by small variations in copper and phosphorus content. In an industrial
atmosphere, structural carbon steel showed a penetration of about 20 µm, a copper structural
steel about 10 µm, and low-alloy steel about 4 µm after five years of exposure. As indicated
in ISO 9223 , it is impossible to give a corrosion rate for steel in the atmosphere without
specifying the location, composition, and certain other factors. If one can relate exposure
conditions to those described in the literature, a fairly good estimate can be made of the probable
corrosion behavior of a selected material. However, all aspects of the exposure of the metal
surface must be considered. A high-strength, low-alloy (HSLA) steel may show an advantage in
corrosion resistance of 12:1 over carbon steel when freely exposed in a mild environment. As the
severity or the physical conditions of exposure change, the HSLA steel will show less
superiority, until in crevices or the backside of structural forms in a corrosive atmosphere, it will
be no better than carbon steel. Cast Iron has, for hundreds of years, been the preferred piping
material throughout the world for drain, waste, and vent plumbing applications and water
distribution. Gray iron can be cast in the form of pipe at low cost and has excellent strength
properties. Unique corrosion resistance characteristics make cast iron soil pipe ideally suited for
plumbing applications. Galvanic corrosion is self-generating and occurs on the surface of a metal
exposed to an electrolyte (such as moist, salt-laden soil). The action is similar to that which
occurs in a wet, or dry, cell battery. Differences in electrical potential between locations on the
surface of the metal (pipe) in contact with such soil may occur for a variety of reasons, including
the joining of different metals (iron and copper or brass for example). Potential differences also
may be due to the characteristics of the soil in contact with the pipe surface, e.g., pH, soluble
salt, oxygen and moisture content, soil resistivity, temperature and presence of certain bacteria.
Any one or a combination of these factors may cause a small amount of electrical current to flow
through the soil between areas on the pipe or metal surface. Where this current discharge into the
soil from such an area, metal is removed from the pipe surface and corrosion occurs.
Electrolytic corrosion occurs when direct current from outside sources enters and then leaves an
underground metal surface to return to its source through the soil; metal is removed and in this
process and corrosion occurs.

35
10) Creep- Creep is a time-dependent deformation of a material while under an applied load that
is below its yield strength. It is most often occurs at elevated temperature, but some materials
creep at room temperature. In materials science, creep is the tendency of a solid material to
slowly move or deform permanently under the influence of stresses. It occurs as a result of long
term exposure to high levels of stress that are below the yield strength of the material. Creep is
more severe in materials that are subjected to heat for long periods, and near melting point.
Creep always increases with temperature. The rate of this deformation is a function of the
material properties, exposure time, exposure temperature and the applied structural load.
Depending on the magnitude of the applied stress and its duration, the deformation may become
so large that a component can no longer perform its function — for example creep of a turbine
blade will cause the blade to contact the casing, resulting in the failure of the blade. Creep is
usually of concern to engineers and metallurgists when evaluating components that operate under
high stresses or high temperatures. Creep is a deformation mechanism that may or may not
constitute a failure mode. Moderate creep in concrete is sometimes welcomed because it relieves
tensile stresses that might otherwise lead to cracking. Unlike brittle fracture, creep deformation
does not occur suddenly upon the application of stress. Instead, strain accumulates as a result of
long-term stress. Creep deformation is "time-dependent" deformation. The temperature range in
which creep deformation may occur differs in various materials. For example, tungsten requires
a temperature in the thousands of degrees before creep deformation can occur while ice will
creep near 0 °C (32 °F). As a rule of thumb, the effects of creep deformation generally become
noticeable at approximately 30% of the melting point for metals and 40–50% of melting point
for ceramics. Virtually any material will creep upon approaching its melting temperature. Since
the minimum temperature is relative to melting point, creep can be seen at relatively low
temperatures for some materials. Plastics and low-melting-temperature metals, including many
solders, creep at room temperature as can be seen markedly in old lead hot-water
pipes. Glacier flow is an example of creep processes in ice. Creep terminates in rupture if steps
are not taken to bring to a halt. Creep data for general design use are usually obtained under
conditions of constant uniaxial loading and constant temperature. Results of tests are usually
plotted as strain versus time up to rupture. As indicated in the image, creep often takes place in
three stages. In the initial stage, strain occurs at a relatively rapid rate but the rate gradually
decreases until it becomes approximately constant during the second stage. This constant creep

36
rate is called the minimum creep rate or steady-state creep rate since it is the slowest creep rate
during the test. In the third stage, the strain rate increases until failure occurs. Creep in service is
usually affected by changing conditions of loading and temperature and the number of possible
stress-temperature-time combinations is infinite. While most materials are subject to creep, the
creep mechanism is often different between metals, plastics, rubber, and concrete.

Stress rupture testing is similar to creep testing except that the stresses are higher than those used
in a creep testing. Stress rupture tests are used to determine the time necessary to produce failure
so stress rupture testing is always done until failure. A straight line or best fit curve is usually
obtained at each temperature of interest. This information can then be used to extrapolate time to
failure for longer times. A typical set of stress rupture curves is shown below.

In the initial stage, or primary creep, the strain rate is relatively high, but slows with increasing
strain. This is due to work hardening. The strain rate eventually reaches a minimum and becomes
near constant. This is due to the balance between work hardening and annealing (thermal

37
softening). This stage is known as secondary or steady-state creep. This stage is the most
understood. The characterized "creep strain rate" typically refers to the rate in this secondary
stage. Stress dependence of this rate depends on the creep mechanism. In tertiary creep, the strain
rate exponentially increases with strain because of necking phenomena. Though mostly due to
the reduced yield stress at higher temperatures, the Collapse of the World Trade Center was
due in part to creep from increased temperature operation. The creep rate of hot pressure-loaded
components in a nuclear reactor at power can be a significant design-constraint, since the creep
rate is enhanced by the flux of energetic particles. Creep was blamed for the Big Dig tunnel
ceiling collapse in Boston, Massachusetts that occurred in July 2006. An example of an
application involving creep deformation is the design of tungsten light bulb filaments. Sagging of
the filament coil between its supports increases with time due to creep deformation caused by the
weight of the filament itself. If too much deformation occurs, the adjacent turns of the coil touch
one another, causing an electrical short and local overheating, which quickly leads to failure of
the filament. The coil geometry and supports are therefore designed to limit the stresses caused
by the weight of the filament, and a special tungsten alloy with small amounts of oxygen trapped
in the crystallite grain boundaries is used to slow the rate of Coble creep. In steam turbine power
plants, pipes carry steam at high temperatures (566 °C or 1050 °F) and pressures (above 24.1
MPa or 3500 psi). In jet engines, temperatures can reach up to 1400 °C (2550 °F) and initiate
creep deformation in even advanced-coated turbine blades. Hence, it is crucial for correct
functionality to understand the creep deformation behavior of materials. Creep deformation is
important not only in systems where high temperatures are endured such as nuclear power
plants, jet engines and heat exchangers, but also in the design of many everyday objects. For
example, metal paper clips are stronger than plastic ones because plastics creep at room
temperatures. Aging glass windows are often erroneously used as an example of this
phenomenon: measurable creep would only occur at temperatures above the glass transition
temperature around 500 °C (900 °F). While glass does exhibit creep under the right conditions,
apparent sagging in old windows may instead be a consequence of obsolete manufacturing
processes, such as that used to create crown glass, which resulted in inconsistent thickness.

38
11) Resilience- Resilience is the property of a material to absorb energy when it is
deformed elastically and then, upon unloading to have this energy recovered. In other words, it is
the maximum energy per unit volume that can be elastically stored. It is represented by the area
under the curve in the elastic region in the stress-strain curve. Modulus of resilience, Ur, can be
calculated using the following formula:

, where
σy is yield stress, E is Young's modulus, and ε is strain. Material’s resilience is the ratio of a
material’s yield strength to its elastic modulus. It can be used to predict the comparative
performance of different materials. A highly resilient material is able to withstand greater
deflections and generates higher forces than a less resilient material. Resilience is also non
dimensional. This material property will have the same value in any system of units, which
makes it easy to compare specifications of different materials, even if their respective data sheets
consist of different units. Resilience does have some drawbacks. A low strength material with a
very low modulus will have a high resilience. In some cases, a lower resilience material will
create a higher force. For example, let us compare the performance of four different materials
used in the same cantilever beam. Let us set the length as 1.0 inch, the width as 0.5 inches, and
the thickness as 0.010 inches. We will compare 1/2 H temper 510 phosphor bronze, T6 temper
7075 aluminum, 25 ½ HT temper copper beryllium, and 301 H temper stainless steel. the
phosphor bronze and stainless steel have nearly equal resilience. However, the steel alloy will
create a higher force due to its greater elastic modulus. Resilience is generally best used to
compare alloys with similar elastic moduli. Fortunately, the elastic moduli of copper alloys do
not vary greatly. Note that the copper beryllium has the highest resilience, and the greatest
maximum deflection and contact force.

However, there is a related parameter that probably gives a better indication of the performance
of a springing material. Take another look at the equations for the maximum allowable force and
deflection of a cantilever beam. Notice that the deflection at yield depends on the resilience and
the force at yield depends on the yield strength. For the best performance, we want high

39
allowable deflection and high force. Therefore, we would like the product of these two terms to
be as high as possible. We know that a force multiplied by a distance results in work, which is
equivalent to energy. Therefore, the product of the force and deflection gives the energy required
to deflect the beam. The equation below derives the maximum strain energy absorbed by the
beam before it yields:

Notice that the twos in the numerator and denominator have not been cancelled by each other.
The reason is that the term on the far right of the equation is a special quantity, known as the
modulus of resilience. This is a quantity that can be derived from a material’s stress-strain curve.
The area under the curve at any point is the strain energy density (strain energy per unit volume)
required to stress the material to that point on the curve. The modulus of resilience is the area
under the curve up to the yield strength. Therefore, the modulus of resilience is the strain energy
density required to stress the material to its yield strength. Let us re-examine the equation for the
energy absorbed by a cantilever beam deflected up to its yield point. The term on the far right is
the strain energy per unit volume at the yield point. The center term is the volume. These two
multiplied together give the total strain energy. The 2/9 term is merely a shape factor for the
cantilever beam. (In this case, it is a straight cantilever beam with a rectangular cross section.).

This strain energy is entirely elastic, which means that it will want to recover when the load is
removed. Until the load is removed, all of the strain energy works to maintain a normal force at
the point of contact. Higher strain energy means better contact force. Therefore, the modulus of
resilience gives an indication of the ability of the material to perform in a contact spring under
load.
We also studied about the different automotive parts, how they work, what are the different types
of stresses/forces they are to undergo. Thus we can make an assessment of which component is
made up of which material that makes it work smoothly. A datasheet is hereby produced.

40
FUTURE PLAN OF THE PROJECT

Now, we are in a position where we have studied most of the important mechanical properties of
an engineering material. We know by now that how the mechanical properties of an engineering
material almost define its applicability in different industries. Thus we can figure out the next
phase of action for the ongoing project. We are to decide and choose some of the most important
mechanical properties of the ones mentioned already. Then we are going to bring some market
available engineering materials for our testing purpose. We are going to perform some testing
(mostly destructive) upon these engineering materials. For this we require to make out suitable
specimens out of the samples obtained from the market. While undergoing the suitable testing
procedures for each of the chosen mechanical properties we compare the obtained values for
different engineering materials. The obtained results help us to understand why a particular
component assigned for one kind of job is made of a particular material. This always gives an
opportunity to commence further research in the search of a material giving the optimum
performance at suitable conditions. The world today is in a search for the best material in all
assigned jobs. That may be automotive design, plant design, and construction purpose and water
tank design. The best material is required not only for cost effectiveness but for endurance and
durability as well. Thus optimization is highly required in all fields of material manufacturing
and development. Our project may be the first step to a bright opening.

41
BIBLIOGRAPHY

All the information we gathered in course of our ongoing project were of immense importance.
The following are some of the references:-
1) http://www.sciencedirect.com/science?_ob=ArticleURL&_udi=B6TVV-3YRNTPD-
V&_user=10&_coverDate=10/31/1996&_rdoc=1&_fmt=high&_orig=search&_origin=se
arch&_sort=d&_docanchor=&view=c&_acct=C000050221&_version=1&_urlVersion=0
&_userid=10&md5=289ea2c053c28ffa31224007c3e52b71&searchtype=a
2) http://www.sciencedirect.com/science?_ob=ArticleURL&_udi=B6TXD-4HDX6WX-
4&_user=10&_coverDate=01/15/2006&_rdoc=1&_fmt=high&_orig=search&_origin=se
arch&_sort=d&_docanchor=&view=c&_acct=C000050221&_version=1&_urlVersion=0
&_userid=10&md5=14473df17831e5ea0c92d514154991cb&searchtype=a
3) http://www.springerlink.com/content/d78361850m127154/
4) http://www.springerlink.com/content/v28t491512103583/
5) http://www.scientific.net/MSF.584-586.734
6) http://www.scientific.net/AMR.59.116
7) http://www.scientific.net/AMR.152-153.11
8) http://www.ttp.net
9) http://www.gordonengland.co.uk/hardness/brinell.htm
10) http://www.gordonengland.co.uk/hardness/vickers.htm
11) http://www.ndt-
ed.org/EducationResources/CommunityCollege/Materials/Mechanical/ImpactToughness.
htm
12) http://www.tpub.com/content/doe/h1017v1/css/h1017v1_73.htm
13) http://en.wikipedia.org/wiki/Compressive_strength
14) http://en.wikipedia.org/wiki/Strength_of_materials
15) http://www.instron.in/wa/glossary/Compressive-Strength.aspx
16) http://en.wikipedia.org/wiki/Yield_(engineering)
17) http://www.engineersedge.com/material_science/yield_strength.htm
18) http://www.tppinfo.com/defect_analysis/yield_strength.html

42
19) http://www.astm.org/DIGITAL_LIBRARY/JOURNALS/TESTEVAL/PAGES/JTE1016
57.htm
20) http://www.differencebetween.net/science/difference-between-yield-strength-and-tensile-
strength/
21) http://en.wikipedia.org/wiki/Ductility
22) http://www.engineersedge.com/material_science/ductility.htm
23) http://www.engineersedge.com/material_science/ductility.htm
24) http://aluminium.matter.org.uk/content/html/eng/default.asp?catid=167&pageid=214441
6578
25) http://en.wikipedia.org/wiki/Fatigue_limit
26) http://www.roymech.co.uk/Useful_Tables/Fatigue/Fatigue.html
27) http://en.wikipedia.org/wiki/Brittleness
28) http://www.sciencedirect.com/science?_ob=ArticleURL&_udi=B6V4V-482HJNH-
G1&_user=10&_coverDate=10/31/1974&_rdoc=1&_fmt=high&_orig=search&_origin=s
earch&_sort=d&_docanchor=&view=c&_acct=C000050221&_version=1&_urlVersion=
0&_userid=10&md5=a2dd7125b86e5282cfe73c78b9e3856a&searchtype=a
29) http://en.wikipedia.org/wiki/Corrosion
30) http://en.wikipedia.org/wiki/Failure_theory_(material)
31) http://corrosion-doctors.org/Corrosion-Atmospheric/Corrosion-resistance.htm
32) http://en.wikipedia.org/wiki/Corrosion
33) http://www.corrosion-doctors.org/MatSelect/CorrResistance.htm
34) http://www.cispi.org/corrosion_resistance.htm
35) http://www.ndt-
ed.org/EducationResources/CommunityCollege/Materials/Mechanical/Creep.htm
36) http://en.wikipedia.org/wiki/Creep_(deformation)
37) http://en.wikipedia.org/wiki/Resilience

--------------------------------------------------------------------------------------

43

Vous aimerez peut-être aussi