You are on page 1of 104

Online Support Module

MECHANICAL ENGINEERING

CONTENTS
Learning Objectives 2 Investigation of Engineering Failures 58
M.1 Introduction 2
M.17 Reasons for Failure 58
M.18 Root Causes 59
Materials Selection for a Torsionally
Stressed Cylindrical Shaft (Case Study) 2 The Failure Analysis 60
M.2 Strength ConsiderationsTorsionally M.19 What Exactly Is the Failure
Stressed Shaft 3 Problem? 60
M.3 Other Property Considerations and the M.20 What Is the Root Cause of the Failure
Final Decision 8 Problem? 61
M.21 What Are the Possible Solutions? 75
Fracture 9 M.22 Which of These Is the Best
Solution? 75
M.4 Principles of Fracture Mechanics 9 M.23 Effectiveness Evaluation of Corrective
M.5 Flaw Detection Using Nondestructive Actions 76
Testing Techniques 22 M.24 The Final Report 76
M.6 Fracture Toughness Testing 26
M.7 Impact Fracture Testing 30
Failure of an Automobile Rear Axle
(Case Study) 77
Fatigue 35
M.25 Introduction 77
M.8 Cyclic Stresses 35 M.26 Testing Procedure and Results 79
M.9 The S-N Curve 37 M.27 Discussion 86
M.10 Crack Initiation and Propagation 39
M.11 Crack Propagation Rate 42 Summary 86
M.12 Factors That Affect Fatigue Life 47 Equation Summary 91
M.13 Environmental Effects 50 Important Terms and Concepts 93
References 93
Questions and Problems 94
Automobile Valve Spring (Case Study) 51
Design Problems 98
M.14 Mechanics of Spring Deformation 51 Problem Conversion Guide 103
M.15 Valve Spring Design and Material Glossary 105
Requirements 52 Answers to Selected Problems 106
M.16 One Commonly Employed Steel Alloy 55 Index for Support Module 107
1
Learning Objectives
After studying this web module, you should be able to do the following:
1. Describe the manner in which materials 7. Define fatigue and specify the conditions
selection charts are employed in the materials under which it occurs.
selection process. 8. From a fatigue plot for some material,
2. Explain why the strengths of brittle materials determine (a) the fatigue lifetime (at a speci-
are much lower than predicted by theoretical fied stress level), and (b) the fatigue strength
calculations. (at a specified number of cycles).
3. Define fracture toughness in terms of (a) a 9. Cite five measures that may be taken to
brief statement, and (b) an equation; define improve the fatigue resistance of a metal.
all parameters in this equation. 10. Briefly describe the steps that are used to
4. Make distinctions between stress intensity ascertain whether a particular metal alloy is
factor, fracture toughness, and plane strain suitable for use in an automobile valve spring.
fracture toughness. 11. List and briefly explain the three root causes of
5. In a qualitative manner, describe how a condi- failure.
tional value of plane-strain fracture toughness 12. List the four questions that a typical failure
is determined using ASTM Standard E 399-09. investigation seeks to answer.
6. Name and briefly describe the two techniques 13. Make a list of procedures/analyses that were used
that are used to measure impact energy (or to determine the cause of failure described in the
notch toughness) of a material. Failure of an Automobile Rear Axle case study.

M.1 INTRODUCTION
Due to constraints on book length, several topics especially suited to the discipline
of mechanical engineering were either not discussed in sufficient detail or omitted
from the print textbook. Therefore, it was decided to provide this online web mod-
ule supplement, which includes the following:
A materials selection case studyMaterials Selection for a Torsionally
Stressed Cylindrical Shaft.
Alternative (and more detailed) versions of Sections 8.5 and 8.6 of Introduction
(Sections 9.5 and 9.8 of Fundamentals)Principles of Fracture Mechanics and
Fracture Toughness/Impact Fracture Testing.
An alternative (and more detailed) treatment of the topic of fatigueSections
8.7 through 8.11 of Introduction (Sections 9.9 through 9.14 of Fundamentals).
A case study on constraints and materials used for an automobile valve spring.
A submodule Investigation of Engineering Failures that outlines a protocol
that may be used to analyze the failure of engineered components.
Another case study that details an investigation that was conducted to deter-
mine the cause of failure of an automobile rear axle.

M a t e r i a l s S e l e c t i o n f o r a To r s i o n a l l y
St r e s s e d C y l i n d r i c a l S h a f t ( C a s e St u d y )
We begin this web module for mechanical engineers by presenting a case study on
materials selection. This process of materials selection involves, for some specified
application, choosing a material having a desirable or optimum property or com-
2
M.2 Strength ConsiderationsTorsionally Stressed Shaft 3

bination of properties. Selection of the proper material can reduce costs and im-
prove performance. Elements of this materials selection process involve deciding
on the constraints of the problem and, from these, establishing criteria that can be
used in materials selection to maximize performance.
The component or structural element we have chosen to discuss is one that has
relevance to a mechanical engineer: a solid cylindrical shaft that is subjected to a
torsional stress. Strength of the shaft will be considered in detail, and criteria will
be developed for maximizing strength with respect to both minimum material mass
and minimum cost. Other parameters and properties that may be important in this
selection process are also discussed briefly.

M.2 STRENGTH CONSIDERATIONSTORSIONALLY


STRESSED SHAFT
For this portion of the design problem, we will establish a criterion for selection of
light and strong materials for this shaft. We will assume that the twisting moment
and length of the shaft are specified, whereas the radius (or cross-sectional area)
may be varied. We develop an expression for the mass of material required in terms
of twisting moment, shaft length, and density and strength of the material. Using
this expression, it will be possible to evaluate the performancethat is, maximize
the strength of this torsionally stressed shaft with respect to mass and, in addition,
relative to material cost.
Consider the cylindrical shaft of length L and radius r, as shown in Figure M.1.
The application of twisting moment (or torque), Mt produces an angle of twist f.
Shear stress t at radius r is defined by the equation
Mt r
t (M.1)
J
Here, J is the polar moment of inertia, which for a solid cylinder is
pr 4
J (M.2)
2
Thus,
2Mt
t (M.3)
pr 3
A safe design calls for the shaft to be able to sustain some twisting moment with-
out fracture. In order to establish a materials selection criterion for a light and strong
material, we replace the shear stress in Equation M.3 with the shear strength of the
material tf divided by a factor of safety N, as1

Figure M.1 A solid cylindrical shaft that


experiences an angle of twist f in response to
r the application of a twisting moment Mt.
Mt

1
The factor of safety concept as well as guidelines for selecting values are discussed in
Section 6.12 of Introduction (Section 7.20 of Fundamentals).
4 Online Support Module: Mechanical Engineering

tf 2M t
 (M.4)
N pr 3
It is now necessary to take into consideration material mass. The mass m of any
given quantity of material is just the product of its density (r) and volume. Since
the volume of a cylinder is pr2L, then
m  pr2Lr (M.5)
or, the radius of the shaft in terms of its mass is
m
r (M.6)
B pLr
Substituting this r expression into Equation M.4 leads to
tf 2Mt

m 3
pa b
N
A pLr

pL3r3
 2Mt (M.7)
B m3
For a cylindrical Solving this expression for the mass m yields
shaft of length L and

m  12NMt 2 23 1p1 3L2 a 2 3 b


radius r that is r
stressed in torsion, (M.8)
expression for mass tf
in terms of density
and shear strength of The parameters on the right-hand side of this equation are grouped into three sets of
the shaft material parentheses. Those contained within the first set (i.e., N and Mt) relate to the safe func-
tioning of the shaft. Within the second parentheses is L, a geometric parameter. Finally,
the material properties of density and strength are contained within the last set.
The upshot of Equation M.8 is that the best materials to be used for a light
shaft that can safely sustain a specified twisting moment are those having low rt2f 3
ratios. In terms of material suitability, it is sometimes preferable to work with what
performance index is termed a performance index, P, which is just the reciprocal of this ratio; that is

Strength tf23
performance index P r (M.9)
expression for a
torsionally stressed
cylindrical shaft In this context we want to use a material having a large performance index.
At this point it becomes necessary to examine the performance indices of a
materials selection variety of potential materials. This procedure is expedited by the use of materials
chart selection charts.2 These are plots of the values of one material property versus those
of another property. Both axes are scaled logarithmically and usually span about
five orders of magnitude, so as to include the properties of virtually all materials.
For example, for our problem, the chart of interest is logarithm of strength versus
logarithm of density, which is shown in Figure M.2.3 It may be noted on this plot
that materials of a particular type (e.g., woods, and engineering polymers) cluster

2
A comprehensive collection of these charts may be found in M. F. Ashby, Materials Selec-
tion in Mechanical Design, 2nd edition, Butterworth-Heinemann, Woburn, UK, 2002.
3
Strength for metals and polymers is taken as yield strength; for ceramics and glasses, com-
pressive strength; for elastomers, tear strength; and for composites, tensile failure strength.
M.2 Strength ConsiderationsTorsionally Stressed Shaft 5

10,000
Diamond
Engineering SiC
Si3N4
ceramics Sialons Engineering
Al2O3
ZrO2 alloys
Cermets
B MgO
Glasses Si Ge

CFRP
1000 GFRP
UNIPLY Steels
Engineering KFRP Pottery Ti W Alloys
CFRP Be Alloys
composites Mo Alloys
GFRP
Laminates Cast
Irons Ni Alloys
KFRP Al Alloys
Cu Alloys
Mg Stone,
Rock Zn
Alloys Alloys
Ash Nylons
100
P = 100 Oak
Pine
PMMA Engineering
Fir PP alloys
Strength (MPa)

MEL
Parallel PVC
to Grain Wood PS Lead
Epoxies Alloys
Products Cement
Polyesters
Balsa Concrete
HDPE
Ash PTFE
Woods Oak
Pine Porous
P = 30 Fir PU ceramics
10 Perpendicular
to Grain
Engineering
LDPE polymers
Silicone
Soft
Balsa Butyl

Elastomers
P = 10
Polymer
foams
Cork
1

P=3

0.1
0.1 0.3 1 3 10 30
Density (Mg /m3)

Figure M.2 Strength versus density materials selection chart. Design guidelines for
performance indices of 3, 10, 30, and 100 (MPa)23m3/Mg have been constructed, all having
a slope of 32. (Adapted from M. F. Ashby, Materials Selection in Mechanical Design.
Copyright 1992. Reprinted by permission of Butterworth-Heinemann Ltd.)

together and are enclosed within an envelope delineated with a bold line. Subclasses
within these clusters are enclosed using finer lines.
Now, taking the logarithm of both sides of Equation M.9 and rearranging yields
3 3
log tf  log r  log P (M.10)
2 2
This expression tells us that a plot of log tf versus log r will yield a family of straight
and parallel lines all having a slope of 32; each line in the family corresponds to a
different performance index, P. These lines are termed design guidelines, and four
6 Online Support Module: Mechanical Engineering

have been included in Figure M.2 for P values of 3, 10, 30, and 100 (MPa)2 3m3/Mg.
All materials that lie on one of these lines will perform equally well in terms of
strength-per-mass basis; materials whose positions lie above a particular line will
have higher performance indices, whereas those lying below will exhibit poorer per-
formances. For example, a material on the P  30 line will yield the same strength
with one-third the mass as another material that lies along the P  10 line.
The selection process now involves choosing one of these lines, a selection
line that includes some subset of these materials; for the sake of argument let us

10,000
Diamond
Engineering SiC
Si3N4
ceramics Sialons Engineering
Al2O3
ZrO2 alloys
Cermets
B MgO
Glasses Si Ge

CFRP
1000 GFRP
UNIPLY Steels
Engineering
KFRP Pottery Ti W Alloys
composites CFRP Be Alloys
Mo Alloys
GFRP
300 MPa Laminates Cast
Irons Ni Alloys
KFRP Al Alloys
Cu Alloys
Mg Stone,
Rock Zn
Alloys Alloys
Ash Nylons
100 Oak PMMA Engineering
Pine
Fir PP alloys
Strength (MPa)

MEL
Parallel PVC
to Grain Wood PS Lead
Epoxies Alloys
Products Cement
Polyesters
Balsa Concrete
HDPE
Ash PTFE
Woods Oak
Pine Porous
Fir PU ceramics
10 P = 10 Perpendicular
to Grain Engineering
(MPa)2/3 m3/Mg LDPE
Silicone polymers
Soft
Balsa Butyl
Elastomers

Polymer
foams
Cork
1

0.1
0.1 0.3 1 3 10 30
Density (Mg /m3)

Figure M.3 Strength versus density materials selection chart. Materials within the shaded
region are acceptable candidates for a solid cylindrical shaft that has a mass-strength
performance index in excess of 10 (MPa)23m3/Mg and a strength of at least 300 MPa
(43,500 psi). (Adapted from M. F. Ashby, Materials Selection in Mechanical Design.
Copyright 1992. Reprinted by permission of Butterworth-Heinemann Ltd.)
M.2 Strength ConsiderationsTorsionally Stressed Shaft 7

pick P  10 1MPa2 23m3/Mg, which is represented in Figure M.3. Materials lying


along this line or above it are in the search region of the diagram and are possi-
ble candidates for this rotating shaft. These include wood products, some plastics, a
number of engineering alloys, the engineering composites, and glasses and engi-
neering ceramics. On the basis of fracture toughness considerations, the engineer-
ing ceramics and glasses are ruled out as possibilities.
Let us now impose a further constraint on the problemnamely that the
strength of the shaft must equal or exceed 300 MPa (43,500 psi). This may be rep-
resented on the materials selection chart by a horizontal line constructed at 300
MPa, see Figure M.3. Now the search region is further restricted to the area above
both of these lines. Thus, all wood products, all engineering polymers, other engi-
neering alloys (viz., Mg and some Al alloys), and some engineering composites are
eliminated as candidates; steels, titanium alloys, high-strength aluminum alloys, and
the engineering composites remain as possibilities.
At this point we are in a position to evaluate and compare the strength per-
formance behavior of specific materials. Table M.1 presents the density, strength,
and strength performance index for three engineering alloys and two engineering
composites, which were deemed acceptable candidates from the analysis using the
materials selection chart. In this table, strength was taken as 0.6 times the tensile
yield strength (for the alloys) and 0.6 times the tensile strength (for the compos-
ites); these approximations were necessary because we are concerned with strength
in torsion, and torsional strengths are not readily available. Furthermore, for the
two engineering composites, it is assumed that the continuous and aligned glass and
carbon fibers are wound in a helical fashion (Figure 16.15 of Introduction, Figure
15.15 of Fundamentals) and at a 45 angle referenced to the shaft axis. The five ma-
terials in Table M.1 are ranked according to strength performance index, from high-
est to lowest: carbon fiber-reinforced and glass fiber-reinforced composites, followed
by aluminum, titanium, and 4340 steel alloys.
Materials cost is another important consideration in the selection process. In
real-life engineering situations, economics of the application often is the overriding
issue and normally will dictate the material of choice. One way to determine ma-
terials cost is by taking the product of the price (on a per-unit mass basis) and the
required mass of material.
Cost considerations for these five remaining candidate materialssteel, alu-
minum, and titanium alloys, and two engineering compositesare presented in

Table M.1 Density (), Strength (ttf), and Strength Performance Index (P)
for Five Engineering Materials
f f23/r  P
Material (Mg/m3) (MPa) [(MPa)23m3/Mg]
Carbon fiber-reinforced composite 1.5 1140 72.8
(0.65 fiber fraction)a
Glass fiber-reinforced composite 2.0 1060 52.0
(0.65 fiber fraction)a
Aluminum alloy (2024-T6) 2.8 300 16.0
Titanium alloy (Ti-6Al-4V) 4.4 525 14.8
4340 Steel (oil-quenched 7.8 780 10.9
and tempered)
a
The fibers in these composites are continuous, aligned, and wound in a helical fashion at
a 45 angle relative to the shaft axis.
8 Online Support Module: Mechanical Engineering

Table M.2 Tabulation of the /f2/3 Ratio, Relative Cost (c



), and Product of /f2/3
a
and c for Five Engineering Materials
f23 c c ( 23 )
f
2 2
Material {10 [Mg(MPa)23m3]} ($$) {10 ($$)[Mg(MPa)23m3]}
4340 Steel (oil-quenched 9.2 3.0 27
and tempered)
Glass fiber-reinforced com- 1.9 28.3 54
posite (0.65 fiber fraction)b
Carbon fiber-reinforced com- 1.4 43.1 60
posite (0.65 fiber fraction)b
Aluminum alloy (2024-T6) 6.2 12.4 77
Titanium alloy (Ti-6Al-4V) 6.8 94.2 641
a
The relative cost is the ratio of the price per unit mass of the material and low-carbon steel.
b
The fibers in these composites are continuous, aligned, and wound in a helical fashion at a 45 angle relative to
the shaft axis.

Table M.2. In the first column is tabulated rt2f 3. The next column lists the ap-
proximate relative cost, denoted as c; this parameter is simply the per-unit mass
cost of material divided by the per-unit mass cost for low-carbon steel, one of the
common engineering materials. The underlying rationale for using c is that although
the price of a specific material will vary over time, the price ratio between that
material and another will, most likely, change more slowly.
Finally, the right-hand column of Table M.2 shows the product of c and rt2f 3.
This product provides a comparison of these materials on the basis of the cost of ma-
terials for a cylindrical shaft that would not fracture in response to the twisting
moment Mt. We use this product inasmuch as r/t2/3 f is proportional to the mass of
material required (Equation M.8) and c is the relative cost on a per-unit mass basis.
Now the most economical is the 4340 steel, followed by the glass fiber-reinforced com-
posite, the carbon fiber-reinforced composite, 2024-T6 aluminum, and the titanium
alloy. Thus, when the issue of economics is considered, there is a significant alteration
within the ranking scheme. For example, inasmuch as the carbon fiber-reinforced com-
posite is relatively expensive, it is significantly less desirable; in other words, the higher
cost of this material may not outweigh the enhanced strength it provides.

M.3 OTHER PROPERTY CONSIDERATIONS


AND THE FINAL DECISION
To this point in our materials selection process we have considered only the strength
of materials. Other properties relative to the performance of the cylindrical shaft
may be importantfor example, stiffness, and, if the shaft rotates, fatigue behavior
(Sections M.9 through M.13). Furthermore, fabrication costs should also be con-
sidered; in our analysis they have been neglected.
Relative to stiffness, a stiffness-to-mass performance analysis similar to the one
just discussed could be conducted. For this case, the stiffness performance index Ps is
1G
Ps  (M.11)
r
where G is the shear modulus. The appropriate materials selection chart (log G
versus log r) would be used in the preliminary selection process. Subsequently,
performance index and per-unit-mass cost data would be collected on specific
M.4 Principles of Fracture Mechanics 9

candidate materials; from these analyses the materials would be ranked on the
basis of stiffness performance and cost.
In deciding on the best material, it may be worthwhile to make a table employ-
ing the results of the various criteria that were used. The tabulation would include, for
all candidate materials, performance index, cost, and so forth for each criterion, as well
as comments relative to any other important considerations. This table puts in per-
spective the important issues and facilitates the final decision process.

Fr a c t u r e
M.4 PRINCIPLES OF FRACTURE MECHANICS 4
Brittle fracture of normally ductile materials, such as that shown in the chapter-
opening photograph (of the oil barge) for Chapter 8 of Introduction (Chapter 9 of
Fundamentals) has demonstrated the need for a better understanding of the mech-
anisms of fracture. Extensive research endeavors over the past century have led to
fracture mechanics the evolution of the field of fracture mechanics. This subject allows quantification
of the relationships between material properties, stress level, the presence of crack-
producing flaws, and crack propagation mechanisms. Design engineers are now
better equipped to anticipate, and thus prevent, structural failures. The present dis-
cussion centers on some of the fundamental principles of the mechanics of fracture.

Stress Concentration
The fracture strength of a solid material is a function of the cohesive forces that ex-
ist between atoms. On this basis, the theoretical cohesive strength of a brittle elastic
solid has been estimated to be approximately E/10, where E is the modulus of elas-
ticity. The experimental fracture strengths of most engineering materials normally
lie between 10 and 1000 times below this theoretical value. In the 1920s, A. A. Grif-
fith proposed that this discrepancy between theoretical cohesive strength and ob-
served fracture strength could be explained by the presence of microscopic flaws or
cracks that always exist under normal conditions at the surface and within the inte-
rior of a body of material. These flaws are a detriment to the fracture strength be-
cause an applied stress may be amplified or concentrated at the tip, the magnitude
of this amplification depending on crack orientation and geometry. This phenome-
non is demonstrated in Figure M.4, a stress profile across a cross section containing
an internal crack. As indicated by this profile, the magnitude of this localized stress
diminishes with distance away from the crack tip. At positions far removed, the stress
is equal to the nominal stress s0, or the applied load divided by the specimen cross-
sectional area (perpendicular to this load). Because of their ability to amplify an
stress raiser applied stress in their locale, these flaws are sometimes called stress raisers.
If it is assumed that a crack has an elliptical shape (or is circular) and is oriented
perpendicular to the applied stress, the maximum stress at the crack tip, sm, is equal to
For tensile loading,
a 1 2
computation of sm  s0 c 1  2 a b d (M.12a)
maximum stress at rt
a crack tip
where s0 is the magnitude of the nominal applied tensile stress, rt is the radius of cur-
vature of the crack tip (Figure M.4a), and a represents the length of a surface crack,

4
This section is an expanded and more detailed version of Section 8.5 of Introduction and
Section 9.5 of Fundamentals.
10 Online Support Module: Mechanical Engineering

s0

sm

rt

Stress
a

X X'

2a s0
x x'

x x'

Position along XX'


(a) (b)
s0

Figure M.4 (a) The geometry of surface and internal cracks. (b) Schematic stress profile
along the line X-X in (a), demonstrating stress amplification at crack tip positions.

or half of the length of an internal crack. For a relatively long microcrack that has a
small tip radius of curvature, the factor 1a/rt 2 1 2 may be very large (certainly much
greater than unity); under these circumstances Equation M.12a takes the form

a 1 2
sm 2s0 a b (M.12b)
rt

Furthermore, sm will be many times the value of s0.


Sometimes the ratio sm0 is denoted as the stress concentration factor Kt:
a 1 2
 2a b
sm
Kt  (M.13)
s0 rt
which is simply a measure of the degree to which an external stress is amplified at
the tip of a crack.
By way of comment, it should be said that stress amplification is not restricted
to these microscopic defects; it may occur at macroscopic internal discontinuities
(e.g., voids or inclusions), at sharp corners, scratches, and notches. Figure M.5 shows
theoretical stress concentration factor curves for several simple and common macro-
scopic discontinuities.
The effect of a stress raiser is more significant in brittle than in ductile materi-
als. For a ductile material, plastic deformation ensues when the maximum stress ex-
ceeds the yield strength. This leads to a more uniform distribution of stress in the
vicinity of the stress raiser and to the development of a maximum stress concen-
tration factor less than the theoretical value. Such yielding and stress redistribution
do not occur to any appreciable extent around flaws and discontinuities in brittle
materials; therefore, essentially the theoretical stress concentration will result.
Griffith then went on to propose that all brittle materials contain a population
of small cracks and flaws that have a variety of sizes, geometries, and orientations.
Fracture will result when, upon application of a tensile stress, the theoretical cohe-
M.4 Principles of Fracture Mechanics 11

Figure M.5 3.4

Stress concentration
Theoretical stress
3.0
concentration factor

factor Kt
curves for three 2.6
w
simple geometrical
d
shapes. (From G. H. 2.2
Neugebauer, Prod.
Eng. (NY), Vol. 14, 1.8
0 0.2 0.4 0.6 0.8 1.0
pp. 8287, 1943.) d
w
(a)

3.8

Stress concentration factor Kt


3.4

b 3.0 b = 4
r
2.6
b = 1
w h r
2.2

b 1.8
b = 1
r 2
1.4
2r
1.0
0 0.2 0.4 0.6 0.8
r
(b) h

3.2

3.0

2.8
Stress concentration factor Kt

2.6
w w
2.4 = 2.00
h
2.2

2.0 w
= 1.25
+ h
r 1.8
h 1.6

1.4
w
1.2 = 1.10
h

1.0
0 0.2 0.4 0.6 0.8 1.0
r
(c) h

sive strength of the material is exceeded at the tip of one of these flaws. This leads
to the formation of a crack that then rapidly propagates. If no flaws were present,
the fracture strength would be equal to the cohesive strength of the material. Very
small and virtually defect-free metallic and ceramic whiskers have been grown with
fracture strengths that approach their theoretical values.

Griffith Theor y of Brittle Fracture


During the propagation of a crack, there is a release of what is termed the elastic strain
energy, some of the energy that is stored in the material as it is elastically deformed.
Furthermore, during the crack extension process, new free surfaces are created at the
12 Online Support Module: Mechanical Engineering

faces of a crack, which give rise to an increase in surface energy of the system. Griffith
developed a criterion for crack propagation of an elliptical crack (Figure M.4a) by
performing an energy balance using these two energies. He demonstrated that the crit-
ical stress sc required for crack propagation in a brittle material is described by
Critical stress for
2Egs 1 2
sc  a b
crack propagation in
a brittle material (M.14)
pa
where
E  modulus of elasticity
gs  specific surface energy
a  one-half the length of an internal crack
Worth noting is that this expression does not involve the crack tip radius rt, as does
the stress concentration equations (Equations M.12a and M.12b); however, it is as-
sumed that the radius is sufficiently sharp (on the order of the interatomic spacing) so
as to raise the local stress at the tip above the cohesive strength of the material.
The previous development applies only to completely brittle materials, for which
there is no plastic deformation. Most metals and many polymers do experience some
plastic deformation during fracture; thus, crack extension involves more than pro-
ducing just an increase in the surface energy. This complication may be accommo-
dated by replacing gs in Equation M.14 by gs  gp, where gp represents a plastic
deformation energy associated with crack extension. Thus,
2E1gs  gp 2 1 2
sc  c d (M.15a)
pa
For highly ductile materials, it may be the case that gp >> gs such that
2Egp 1 2
sc  a b (M.15b)
pa
In the 1950s, G. R. Irwin chose to incorporate both gs and gp into a single term, Gc, as
Gc  21gs  gp 2 (M.16)
Gc is known as the critical strain energy release rate. Incorporation of Equation M.16
into Equation M.15a after some rearrangement leads to another expression for the
Griffith cracking criterion as
ps2a
Gc  (M.17)
E
Thus, crack extension occurs when ps2aE exceeds the value of Gc for the partic-
ular material under consideration.

EXAMPLE PROBLEM M.1

Maximum Flaw Length Computation


A relatively large plate of a glass is subjected to a tensile stress of 40 MPa. If
the specific surface energy and modulus of elasticity for this glass are 0.3 J/m2
and 69 GPa, respectively, determine the maximum length of a surface flaw
that is possible without fracture.
M.4 Principles of Fracture Mechanics 13

Solution
To solve this problem it is necessary to employ Equation M.14. Rearrange-
ment of this expression such that a is the dependent variable, and realizing
that s  40 MPa, gs  0.3 J/m2, and E  69 GPa leads to
2Egs
a
ps2
122169 109 N/m2 210.3 N/m2

p 140 106 N/m2 2 2
 8.2 106 m  0.0082 mm  8.2 mm

Stress Analysis of Cracks


As we continue to explore the development of fracture mechanics, it is worthwhile
to examine the stress distributions in the vicinity of the tip of an advancing crack.
There are three fundamental ways, or modes, by which a load can operate on a
crack, and each will affect a different crack surface displacement; these are illus-
trated in Figure M.6. Mode I is an opening (or tensile) mode, whereas modes II and
III are sliding and tearing modes, respectively. Mode I is encountered most fre-
quently, and only it will be treated in the ensuing discussion on fracture mechanics.
For this mode I configuration, the stresses acting on an element of material are
shown in Figure M.7. Using elastic theory principles and the notation indicated, ten-
sile (sx and sy)5 and shear (txy) stresses are functions of both radial distance r and
the angle u as follows:6
fx 1u2
K
sx  (M.18a)
22pr

Figure M.6 The


three modes of
crack surface
displacement.
(a) Mode I, opening
or tensile mode;
(b) mode II, sliding
mode; and (c) mode
III, tearing mode.
(a) (b) (c)

5
This sy denotes a tensile stress parallel to the y-direction and should not be confused
with the yield strength (Section 6.6 of Introduction, Section 7.6 of Fundamentals), which
uses the same symbol.
6
The f(u) functions are as follows:

fx 1u2  cos a1  sin sin b


u u 3u
2 2 2

fy 1u2  cos a1  sin sin b


u u 3u
2 2 2

fxy 1u2  sin cos cos


u u 3u
2 2 2
14 Online Support Module: Mechanical Engineering

Figure M.7 The stresses acting


in front of a crack that is loaded
in a tensile mode I configuration.
txy sy
y

sx
sz
r

q
x

fy 1u2
K
sy (M.18b)
22pr

fxy 1u2
K
txy (M.18c)
22pr
If the plate is thin relative to the dimensions of the crack, then sz 0, or a condi-
tion of plane stress is said to exist. At the other extreme (a relatively thick plate),
plane strain sz n(sx sy), and the state is referred to as plane strain (since z 0); n in this
expression is Poissons ratio.
stress intensity factor In Equations M.18, the parameter K is termed the stress intensity factor; its use
provides for a convenient specification of the stress distribution around a flaw. It
should be noted that this stress intensity factor and the stress concentration factor
Kt in Equation M.13, although similar, are not equivalent.
The stress intensity factor is related to the applied stress and the crack length
by the following equation:
Stress intensity
factordependence
K Ys 1pa (M.19)
on applied stress and
crack length
Here Y is a dimensionless parameter or function that depends on both the crack
and specimen sizes and geometries, as well as the manner of load application. More
will be said about Y in the discussion that follows. Moreover, it should be noted
that K has the unusual units of MPa 1m 1psi1in. [alternatively ksi1in.]).

Fracture Toughness
In the previous discussion, a criterion was developed for the crack propagation in a
brittle material containing a flaw; fracture occurs when the applied stress level exceeds
some critical value sc (Equation M.14). Similarly, since the stresses in the vicinity of
fracture toughness a crack tip can be defined in terms of the stress intensity factor, a critical value of K
exists that may be used to specify the conditions for brittle fracture; this critical value
Fracture toughness
dependence on is termed the fracture toughness Kc, and, from Equation M.19, is defined by
critical stress for
crack propagation Kc Y1a W2 sc 1pa (M.20)
and crack length
M.4 Principles of Fracture Mechanics 15

Figure M.8 Schematic


representations of (a) an
interior crack in a plate
of infinite width, and
(b) an edge crack in a
plate of semi-infinite
width.

2a a

(a) (b)

Here, sc again is the critical stress for crack propagation, and we now represent Y as
a function of both crack length (a) and component width (W)i.e., as Y(aW).
Relative to this Y(aW) function, as the aW ratio approaches zero (i.e., for very
wide planes and short cracks), Y(aW) approaches a value of unity. For example, for
a plate of infinite width having a through-thickness crack, Figure M.8a, Y(aW)  1.0;
whereas for a plate of semi-infinite width containing an edge crack of length a
(Figure M.8b), Y(aW) 1.1. Mathematical expressions for Y(aW) (often relatively
complex) in terms of a/W are required for components of finite dimensions. For
example, for a center-cracked (through-thickness) plate of width W (Figure M.9),
pa 12
Y1a/W2  a b
W
tan (M.21)
pa W
Here the pa/W argument for the tangent is expressed in radians, not degrees. It is
often the case for some specific component-crack configuration that Y(a/W)
is plotted versus a/W (or some variation of a/W). Several of these plots are shown
in Figures M.10a, b, and c; included in the figures are equations that are used to
determine Kcs.
By definition, fracture toughness is a property that is the measure of a mater-
ials resistance to brittle fracture when a crack is present. Its units are the same as
for the stress intensity factor (i.e., MPa 1m or psi 1in.2.

s Figure M.9 Schematic representation of a flat


plate of finite width having a through-thickness
center crack.
W
W
2

2a

B
s
16 Online Support Module: Mechanical Engineering

F
4.00

YF
Kc  a
W WB
3.00

a Y

2.00

B
1.00
F 0.0 0.1 0.2 0.3 0.4 0.5 0.6
a/W
(a)

F 1.22

1.20
YF
Kc  a
W WB
1.18

a a Y 1.16

1.14

1.12

B
1.10
F 0.0 0.1 0.2 0.3 0.4 0.5 0.6
2a/W
(b)

1.9

1.7
F 3FSY
Kc  a
4W2B

B 1.5

Y
W S/W  8
1.3
a

S 1.1 S/W  4

0.9
0.0 0.1 0.2 0.3 0.4 0.5 0.6

a/W
(c)

Figure M.10 Y calibration curves for three simple crack-plate geometries. (Copyright
ASTM. Reprinted with permission.)
M.4 Principles of Fracture Mechanics 17

Figure M.11 Schematic


representation showing the
effect of plate thickness on
fracture toughness.

Fracture toughness Kc
KIc

Plane stress Plane strain


behavior behavior

Thickness B

For relatively thin specimens, the value of Kc will depend on and decrease with
increasing specimen thickness B, as indicated in Figure M.11. Eventually, Kc becomes
independent of B, at which time the condition of plane strain is said to exist.7
plane strain fracture The constant Kc value for thicker specimens is known as the plane strain fracture
toughness toughness KIc, which is also defined by8

Plane strain fracture KIc  Ys1pa (M.22)


toughness for mode
I crack surface
It is the fracture toughness normally cited since its value is always less than Kc. The
displacement
I subscript for KIc denotes that this critical value of K is for mode I crack dis-
placement, as illustrated in Figure M.6a. Brittle materials, for which appreciable
plastic deformation is not possible in front of an advancing crack, have low KIc val-
ues and are vulnerable to catastrophic failure. On the other hand, KIc values are
relatively large for ductile materials. Fracture mechanics is especially useful in pre-
dicting catastrophic failure in materials having intermediate ductilities. Plane strain
fracture toughness values for a number of different materials are presented in Table
M.3; a more extensive list of KIc values is contained in Table B.5 of Appendix B of
both Introduction and Fundamentals.
The stress intensity factor K in Equations M.18 and the plane strain fracture tough-
ness KIc are related to one another in the same sense as are stress and yield strength.
A material may be subjected to many values of stress; however, there is a specific stress
level at which the material plastically deformsthat is, the yield strength. Likewise, a
variety of Ks are possible, whereas KIc is unique for a particular material, and indi-
cates the conditions of flaw size and stress necessary for brittle fracture.
Several different testing techniques are used to measure KIc; one of these is de-
scribed later in Section M.6. Virtually any specimen size and shape consistent with
mode I crack displacement may be utilized, and accurate values will be realized pro-
vided that the Y scale parameter in Equation M.22 has been properly determined.

7
Experimentally, it has been verified that for plane strain conditions

Minimum specimen KIc 2


thickness for a B
2.5a b (M.23)
condition of plane sy
strain
where sy is the 0.002 strain offset yield strength of the material.
8
In the ensuing discussion we will use Y to designate Y(a/W) in order to simplify the form
of the equations.
18 Online Support Module: Mechanical Engineering

Table M.3 Room-Temperature Yield Strength and Plane Strain Fracture


Toughness Data for Selected Engineering Materials
Yield Strength KIc
Material MPa ksi MPa 1m ksi 1in .
Metals
Aluminum alloya 495 72 24 22
(7075-T651)
Aluminum alloya 345 50 44 40
(2024-T3)
Titanium alloya 910 132 55 50
(Ti-6Al-4V)
Alloy steela 1640 238 50.0 45.8
(4340 tempered @ 260C)
Alloy steela 1420 206 87.4 80.0
(4340 tempered @ 425C)
Ceramics
Concrete 0.21.4 0.181.27
Soda-lime glass 0.70.8 0.640.73
Aluminum oxide 2.75.0 2.54.6
Polymers
Polystyrene 25.069.0 3.6310.0 0.71.1 0.641.0
(PS)
Poly(methyl methacrylate) 53.873.1 7.810.6 0.71.6 0.641.5
(PMMA)
Polycarbonate 62.1 9.0 2.2 2.0
(PC)
a
Source: Reprinted with permission, Advanced Materials and Processes, ASM Interna-
tional, 1990.

The plane strain fracture toughness KIc is a fundamental material property that
depends on many factors, the most influential of which are temperature, strain rate,
and microstructure. The magnitude of KIc diminishes with increasing strain rate and
decreasing temperature. Furthermore, an enhancement in yield strength wrought
by solid solution or dispersion additions or by strain hardening generally produces
a corresponding decrease in KIc. Furthermore, KIc normally increases with reduction
in grain size as composition and other microstructural variables are maintained con-
stant. Yield strengths are included for some of the materials listed in Table M.3.

Design Using Fracture Mechanics


According to Equations M.20 and M.22, three variables must be considered relative
to the possibility for fracture of some structural componentviz. the fracture tough-
ness (Kc) or plane strain fracture toughness (KIc), the imposed stress (s), and the flaw
size (a), assuming, of course, that Y has been determined. When designing a compo-
nent, it is first important to decide which of these variables are constrained by the
application and which are subject to design control. For example, material selection
(and hence Kc or KIc) is often dictated by factors such as density (for lightweight
applications) or the corrosion characteristics of the environment. Or, the allowable
flaw size is either measured or specified by the limitations of available flaw detection
techniques. It is important to realize, however, that once any combination of two of
the preceding parameters is prescribed, the third becomes fixed (Equations M.20 and
M.22). For example, assume that KIc and the magnitude of a are specified by applica-
tion constraints; therefore, the design (or critical) stress sc must be
M.4 Principles of Fracture Mechanics 19

KIc
Computation of sc  (M.24)
design stress Y1pa
On the other hand, if stress level and plane strain fracture toughness are fixed by
the design situation, then the maximum allowable flaw size ac is
Computation of 1 KIc 2
maximum allowable ac  a b (M.25)
flaw length p sY

EXAMPLE PROBLEM M.2

Determination of the Possibility of Critical Flaw Detection


A structural component in the form of a very wide plate, as shown in Fig-
ure M.8a, is to be fabricated from a 4340 steel. Two sheets of this alloy, each
having a different heat treatment and thus different mechanical properties,
are available. One, denoted material A, has a yield strength of 860 MPa
(125,000 psi) and a plane strain fracture toughness of 98.9 MPa 1m
190 ksi1in.2 . For the other, material Z, sy and KIc values are 1515 MPa
(220,000 psi) and 60.4 MPa1m 155 ksi1in.2, respectively.
(a) For each alloy, determine whether or not plane strain conditions prevail
if the plate is 10 mm (0.39 in.) thick.
(b) It is not possible to detect flaw sizes less than 3 mm, which is the resolu-
tion limit of the flaw detection apparatus. If the plate thickness is sufficient
such that the KIc value may be used, determine whether or not a critical flaw
is subject to detection. Assume that the design stress level is one half of the
yield strength; also, for this configuration, the value of Y is 1.0.
Solution
(a) Plane strain is established by Equation M.23. For material A,
KIc 2 98.9 MPa1m 2
B  2.5 a b  2.5 a b
sy 860 MPa
 0.033 m  33 mm (1.30 in.)
Thus, plane strain conditions do not hold for material A because this value of
B is greater than 10 mm, the actual plate thickness; the situation is one of
plane stress and must be treated as such.
And for material Z,
60.4 MPa 1m 2
B  2.5 a b  0.004 m  4.0 mm 10.16 in.2
1515 MPa
which is less than the actual plate thickness, and therefore the situation is one
of plane strain.
(b) We need only to determine the critical flaw size for material Z because
the situation for material A is not plane strain, and KIc may not be used. Em-
ploying Equation M.25 and taking s to be sy/2,
60.4 MPa1m 2
a b
1
ac 
p 11211515/22 MPa
 0.002 m  2.0 mm (0.079 in.)
Therefore, the critical flaw size for material Z is not subject to detection since
it is less than 3 mm.
20 Online Support Module: Mechanical Engineering

DESIGN EXAMPLE M.1

Material Specification for a Pressurized Spherical Tank


Consider the thin-walled spherical tank of radius r and thickness t (Figure M.12)
that may be used as a pressure vessel.
(a) One design of such a tank calls for yielding of the wall material prior to fail-
ure as a result of the formation of a crack of critical size and its subsequent rapid
propagation. Thus, plastic distortion of the wall may be observed and the pres-
sure within the tank released before the occurrence of catastrophic failure. Con-
sequently, materials having large critical crack lengths are desired. On the basis of
this criterion, rank the metal alloys listed in Table B.5, Appendix B (of both
Introduction and Fundamentals), as to critical crack size, from longest to shortest.

s Figure M.12
Schematic diagram
showing the cross
section of a spherical
2a tank that is subjected
p
to an internal pressure
t p p p, and that has a radial
r crack of length 2a in
p p its wall.

p p s
p

(b) An alternative design that is also often utilized with pressure vessels is termed
leak-before-break. Using principles of fracture mechanics, allowance is made for the
growth of a crack through the thickness of the vessel wall prior to the occurrence
of rapid crack propagation (Figure M.12). Thus, the crack will completely penetrate
the wall without catastrophic failure, allowing for its detection by the leaking of
pressurized fluid. With this criterion the critical crack length ac (i.e., one-half of the
total internal crack length) is taken to be equal to the pressure vessel thickness t.
Allowance for ac  t instead of ac  t/2 assures that fluid leakage will occur prior
to the buildup of dangerously high pressures. Using this criterion, rank the metal
alloys in Table B.5, Appendix B as to the maximum allowable pressure.
For this spherical pressure vessel, the circumferential wall stress s is a function
of the pressure p in the vessel and the radius r and wall thickness t according to
pr
s (M.26)
2t
For both parts (a) and (b) assume a condition of plane strain.

Solution
(a) For the first design criterion, it is desired that the circumferential wall stress
be less than the yield strength of the material. Substitution of sy for s in Equa-
tion M.22, and incorporation of a factor of safety N leads to
M.4 Principles of Fracture Mechanics 21

Table M.4 Ranking of Several Metal Alloys Relative to Critical Crack Length
(Yielding Criterion) for a Thin-Walled Spherical Pressure Vessel
KIc 2
a b (mm)
Material Sy
Medium carbon (1040) steel 43.1
AZ31B magnesium 19.6
2024 aluminum (T3) 16.3
Ti-5Al-2.5Sn titanium 6.6
4140 steel (tempered @ 482C) 5.3
4340 steel (tempered @ 425C) 3.8
Ti-6Al-4V titanium 3.7
17-7PH stainless steel 3.4
7075 aluminum (T651) 2.4
4140 steel (tempered @ 370C) 1.6
4340 steel (tempered @ 260C) 0.93

sy
KIc  Y a b 1pac (M.27)
N
where ac is the critical crack length. Solving for ac yields the following expression:

N 2 KIc 2
ac  a b (M.28)
Y 2p sy

Therefore, the critical crack length is proportional to the square of the KIc-sy
ratio, which is the basis for the ranking of the metal alloys in Table B.5. The rank-
ing is provided in Table M.4, where it may be seen that the medium carbon (1040)
steel with the largest ratio has the longest critical crack length, and, therefore, is
the most desirable material on the basis of this criterion.
(b) As stated previously, the leak-before-break criterion is just met when one-
half of the internal crack length is equal to the thickness of the pressure vessel
(i.e., when a  t). Substitution of a  t into Equation M.22 gives

KIc  Ys1pt (M.29)


And, from Equation M.26,
pr
t (M.30)
2s
The stress is replaced by the yield strength, inasmuch as the tank should be designed
to contain the pressure without yielding; furthermore, substitution of Equation M.30
into Equation M.29, after some rearrangement, yields the following expression:
K 2Ic
a b
2
p (M.31)
Y 2pr sy
Hence, for some given spherical vessel of radius r, the maximum allowable pres-
sure consistent with this leak-before-break criterion is proportional to K 2Ic sy.
The same several materials are ranked according to this ratio in Table M.5; as
may be noted, the medium carbon steel will contain the greatest pressures.
22 Online Support Module: Mechanical Engineering

Of the eleven metal alloys that are listed in Table B.5, the medium carbon
steel ranks first according to both yielding and leak-before-break criteria. For
these reasons, many pressure vessels are constructed of medium carbon steels,
when temperature extremes and corrosion need not be considered.

Table M.5 Ranking of Several Metal Alloys Relative to Maximum Allowable


Pressure (Leak-Before-Break Criterion) for a Thin-Walled
Spherical Pressure Vessel
KIc2
Material Sy (MPa 1m)

Medium carbon (1040) steel 11.2


4140 steel (tempered @ 482C) 6.1
Ti-5Al-2.5Sn titanium 5.8
2024 aluminum (T3) 5.6
4340 steel (tempered @ 425C) 5.4
17-7PH stainless steel 4.4
AZ31B magnesium 3.9
Ti-6Al-4V titanium 3.3
4140 steel (tempered @ 370C) 2.4
4340 steel (tempered @ 260C) 1.5
7075 aluminum (T651) 1.2

M.5 FLAW DETECTION USING NONDESTRUCTIVE


TESTING TECHNIQUES
nondestructive testing A number of nondestructive testing (NDT) techniques have been developed that per-
mit detection and measurement of both internal and surface flaws.9 Such techniques
are used to examine structural components currently in service for defects and flaws
that could lead to premature failure; in addition, NDTs are used as a means of qual-
ity control for manufacturing processes. As the name implies, these techniques must
not destroy or damage the material/structure being examined nor impair its future
serviceability. Some testing techniques are capable of detecting only surface defects,
some only subsurface (interior) defects, while other tests detect defects at both sur-
face and subsurface sites. Furthermore, in some instances location of testing is
important. Some testing methods must be conducted in a laboratory setting; others
may be adapted for use in the field. Several commonly employed NDT techniques
are visual inspection, optical microscopy, scanning electron microscopy, dye (or
liquid) penetrant, magnetic particle, radiographic, ultrasonic, and acoustic emission.
A listing of these techniques and their characteristics are presented in Table M.6.
The following discussions on some of these techniques are very brief and con-
densed. More detailed treatments are provided in the end-of-module references.

Visual Inspection
Visual inspection is probably the most common detection technique and the easi-
est to conduct; of course, only cracks and defects found on surfaces may be ob-
served visually. Only relatively large cracks/defects are subject to detection with the

9
Sometimes the terms nondestructive evaluation (NDE) and nondestructive inspection
(NDI) are also used for these techniques.
M.5 Flaw Detection Using Nondestructive Testing Techniques 23

Table M.6 A List of Several Common Nondestructive Testing (NDT) Techniques


Defect Size
Technique Defect Location Sensitivity (mm) Testing Location
Visual inspection Surface 0.1 Laboratory/in-field
Optical microscopy Surface 0.10.5 Laboratory
Scanning electron Surface 0.001 Laboratory
microscopy (SEM)
Dye penetrant Surface 0.0350.25 Laboratory/in-field
Magnetic particle Surface 0.5 Laboratory/in-field
Radiography Subsurface 2% of Laboratory/in-field
(x-ray/gamma ray) specimen thickness
Ultrasonic Subsurface 0.50 Laboratory/in-field
Acoustic emission Surface/ 0.1 Laboratory/in-field
subsurface
Source: Adapted from ASM Handbook, Vol. 19, Fatigue and Fracture. Reprinted with per-
mission of ASM International. All rights reserved. www.asminternational.org.

unaided eye or a magnifying glass. For inspection of inaccessible/remote regions,


mirrors, fiberscopes, and borescopes may be used. Fiberscopes and borescopes are
optical devices composed of an eyepiece on the inspection end and a lens on the
observation end, which are linked by either a rigid or flexible tube (normally of an
optical fiber bundle and a protective outer sheath), which acts as the optical relay
system, and in some cases is used to illuminate the remote object. An imaging
device (e.g., video camera) may also be incorporated.
Portable video inspection camera systems are used to inspect the interiors of
large structures (e.g., containers, railroad tank cars, sewer lines) that are inaccessi-
ble and/or hazardous. A video camera (with a zoom lens) is mounted on a pole,
cable, or trolley for deployment into the structure that is to be inspected.
Visual inspection of some confined (and normally horizontal) and long struc-
tures (such as pipelines, air ducts, and reactors) is possible using robotic crawlers.
These devices typically include a sensor (or video camera) mounted on a mobile
support carriage that is capable of traveling through the system to be inspected.
The crawler must be steerable, capable of both forward and reverse motions, as well
as acceleration, deceleration, and stopping; in addition, it must have the ability
to negotiate bends in piping and to pass through different diameters of pipes. An
illumination system is also provided, and the camera must deliver full-directional
viewing as well as have a remote adjustable focus.
Optical and Scanning Electron Microscopy Inspections
For detection of small surface cracks (less than about 0.5 mm in size), employment
of optical and/or scanning electron microscopic techniques is necessary. Normally
inspections of these types are conducted in a laboratory setting (as opposed to in
the field). Brief discussions of these two techniques are presented in Section 4.10
of Introduction and Section 5.12 of Fundamentals.

Dye (Liquid) Penetrant Inspection


This common and low-cost technique is used to detect surface cracks in nonporous
materials. In essence, a liquid is used to enhance the visual contrast between a
defect and the bulk solid material. This liquid must have high surface wetting char-
acteristics (i.e., a low surface tension) and is applied (by spraying, brushing, or dip-
ping) onto the surface of the part to be inspected. After adequate time has been
24 Online Support Module: Mechanical Engineering

Figure M.13 Photograph


showing a surface crack in an
automobile engine connecting rod
that was made visible using a dye
penetrant. (Photograph courtesy
of Center for NDE, Iowa State
University.)

allowed for this liquid to penetrate (by capillary action) into any surface-breaking
defects that are present, excess liquid is removed, and a powder developer is
applied that draws the penetrant out of any defects and to the surface, making pos-
sible their observation. Visual inspection is performed using a white light. The pen-
etrant may also be loaded with a fluorescent dye to increase detection sensitivity;
under these circumstances an ultraviolet (or black) light (in a darkened environ-
ment) is used to reveal the defects.
Figure M.13 shows a surface crack that was exposed using this technique.

Magnetic Particle Inspection


A variation of the dye penetrant technique is used to detect both surface and near-
surface defects in ferrous alloys that may be magnetized. For such a material that has
been magnetized, in the vicinity of a surface flaw (or discontinuity) there is a distor-
tion of the magnetic flux (or field) such that the flux leaks from (or passes out of)
the solid. This inspection technique utilizes fine iron particles that are suspended in
a suitable liquid carrier (e.g., kerosene); these particles are often coated with a fluo-
rescent dye. The article to be inspected is first magnetized, and when such a suspen-
sion is applied (sprayed or painted) onto its surface, iron particles become attracted
to regions where any leakage fields are presenti.e., they cluster in the vicinity of
any surface defects. Visual detection of these clusters (and defects) is possible under
proper lighting conditions (e.g., a fluorescent light and a dark ambient). A surface
crack that was made visible using magnetic particles is presented in Figure M.14.
After observation, demagnetization of the inspected object is possible.

The final three nondestructive testing techniques (radiographic, ultrasonic [pulse


echo], and acoustic emission) are utilized to detect subsurface (interior) defects. The
first two techniques (radiographic and ultrasonic testing) employ some type of signal
or energy source (e.g., ultrasonic waves, x-rays) to probe the object to be examined.
An interaction between the incoming signal and a defect or crack causes some type
of signal disruption, and a response in the form of an image or signature that may be
sensed and recorded (such as a photograph or a blip on a screen).

Radiographic Testing
The radiographic testing technique utilizes x-rays or gamma radiation (from a ra-
dioactive isotope such as iridium-192 or cobalt-60) as a signal source. This radiation
is directed upon, penetrates, and passes through the object to be inspected. An
M.5 Flaw Detection Using Nondestructive Testing Techniques 25

Figure M.14
Photograph
showing a surface
crack in a crane
hook that was
revealed using the
magnetic particle
inspection
technique.
(Photograph
courtesy of Center
for NDE, Iowa
State University.)

image is generated on the remote side by radiation that is transmitted through the
object. Photographic film (that is sensitive to the type of radiation) is most
frequently employed as the detection device to record this image (Figure M.15);
fluoroscopic screens and digitized systems with video monitors may also be used.
An image results from variations in the transmitted radiation intensity over the
cross-section of the object. Defects and cracks will appear as part of this image inas-
much as transmission intensity will be different through regions containing defects
than those regions that are defect-free. This technique is commonly used to assess
the integrity of welds.
It should be noted that health hazards are associated with radiographic testing;
exposure to both x-rays and gamma rays must be avoided since both are forms of
ionizing radiation.

Figure M.15
Photograph
generated using
x-radiation that
shows an internal
defect (dark
wedge-shaped
region) in an
automobile motor
mount casting.
(Photograph
courtesy of Center
for NDE, Iowa
State University.)
26 Online Support Module: Mechanical Engineering

Ultrasonic (Pulse-Echo) Inspection


For this inspection technique, the input energy (or signal) is in the form of ultra-
sonic wavesi.e., sound waves having high frequencies, normally in the range of
0.1 to 50 MHz. These waves are emitted from a transducer as intermittent pulses,
which are introduced into and propagate through the object that is being inspected.
Normally both transducer and test object are immersed in a coupling medium (a
liquid such as water or oil) so as to promote the transfer of ultrasonic waves. These
waves experience reflection (or echo) whenever they encounter some type of in-
terface or discontinuity, such as the back face of the test object or the surface of an
interior defect. A reflected wave is received by the same transducer, which then
converts the wave signal into an electrical signal. Test results are displayed on a
screen as reflected signal strength amplitude versus travel time (i.e., the time be-
tween when the pulsed signal was sent and when the reflected signal was received).
A high intensity peak of signal strength represents the time at which a reflected
wave was received. Signal travel time may be converted into travel distance inas-
much as the velocity of the ultrasonic waves is known. Measurement of travel time
is important in order to distinguish between back-surface and defect echos, and, in
addition, to determine the location (depth) of a defect. A complete inspection in-
volves passing the transducer probe over the entire region of the test object.
Aerospace, aviation, and automotive industries utilize this NDT technique ex-
tensively.
Acoustic Emission
Acoustic emission testing also utilizes ultrasonic waves to detect the presence of
cracksusually in the frequency range between 30 kHz and 1MHz. However, un-
like conventional ultrasonic testing, this technique monitors sound (acoustic) waves
that are emitted during the failure process (i.e., as a crack forms and then propa-
gates) and while a structure is in servicethat is no ultrasonic signal is artificially
generated and then collected.Associated with the formation and extension of cracks
is the release of elastic strain energy, in the form of sound waves, that propagate
throughout the material and ultimately to its surface where they may be recorded
using some type of sensor (i.e., a transducer). This signal is converted into an elec-
trical signal and then displayed on a screen for analysis.
One advantage of acoustic emission testing (over other NDT techniques) is that
it monitors failure processes that are dynamic (i.e., crack formation and growth).
As such, instantaneous information relative to the status of and risk of failure is
provided. This technique is frequently used on aircraft. For example, a group of
transducers mounted in a highly stressed area can detect the presence of a crack
the instant that it forms, and, in addition, very precisely determine its origin by meas-
urement of the time it takes for the signal to reach different transducers.

M.6 FRACTURE TOUGHNESS TESTING


As noted earlier, fracture toughness is defined as a materials resistance to crack prop-
agation and ultimately to brittle fracture. In Section M.4 we used the symbol KIc to
represent the fracture toughness for the condition of plane strain (i.e., when speci-
men thickness is greater than crack length) and also when stress application is such
as to promote mode I crack surface displacement (Figure M.6). Inasmuch as KIc is
such an important material property with regard to fracture prevention, it seems rea-
sonable to explore the manner in which it is measured. A variety of standardized tests
have been devised. In the United States, these test methods are developed by ASTM;
for the international marketplace, standards are established by the International
M.6 Fracture Toughness Testing 27

Organization for Standardization (ISO). Most of these techniques are designed for
measuring fracture toughness values for metals and their alloys; in addition, some
have also been developed for ceramics, polymers, and composite materials.
In essence, a typical fracture toughness test is conducted on a standard specimen
that contains a preexisting crack. A testing apparatus loads the specimen at a pre-
scribed rate, and continuously records load magnitude and crack displacement. Re-
sulting data are analyzed and fracture toughness parameters are determined. These
parameters are then subjected to qualification procedures in order to ensure they meet
established criteria before the fracture toughness values are deemed acceptable. We
have chosen to describe one of the earliest and least complicated fracture toughness
test standards that was developed: ASTM Standard E 399-09, Standard Test Method
for Linear-Elastic Plane-Strain Fracture Toughness KIc of Metallic Materials.
First of all, as the title of this standard suggests, the test is used to measure KIc
when the crack-tip region is exposed to a condition of plane-strain upon load
application. In addition, the material being tested should exhibit linear-elastic
behaviorthat is, a plot of load versus crack displacement is linear, and virtually all
deformation to the point of fracture is elastic (i.e., the material has limited ductility).
Furthermore, it may be recalled (Section M.4), that the elastic stress field near a crack
tip can be described in terms of the stress intensity factor K; as will be seen we use
this parameter in the development of the methodology for this testing technique.
Two specimen geometries permitted by Standard E 399single-edge notched
bend and compact tensionare represented in Figures M.16a and M.16b, respec-
tively. As noted in these illustrations, a three-point loading scheme is used for the
bend specimen, whereas the compact specimen is loaded in tension. Specimen size
is not specified by this standard, which must be selected. Test validity is dependent
on specimen size, which is not subject to evaluation until after the conclusion of the
test; therefore, unless an adequate size is chosen, the test will need to be repeated.
A notch is machined in each specimen, after which a very sharp precrack of length
a is introduced at the notch root using cyclic fatigue-loading. As noted in Figures
M.16a and M.16b, initial crack length includes both notch depth as well as precrack

Precrack
P
Loading pin
a
B
W
Precrack
a

S Support pin
P
(a) (b)

Figure M.16 Configuration of (a) single-edge notched bend and (b) compact-tension
specimens used for fracture toughness tests (ASTM Standard E 399-09). (From V. J.
Colangelo and F. A. Heiser, Analysis of Metallurgical Failures, 2nd edition. Copyright 1987
by John Wiley & Sons, New York. Reprinted with permission of John Wiley & Sons, Inc.)
28 Online Support Module: Mechanical Engineering

P Figure M.17 Schematic diagram


Test showing a displacement gage that has
specimen
been installed on a single-edge notched
bend specimen in preparation for a
fracture toughness test. (Adapted with
W permission from Figure A2.1 in ASTM
E399-09 Standard Test Method for
Linear-Elastic Plane-Strain Fracture
Toughness KIc of Metallic Materials.
Copyright ASTM International, 100
Displacement Barr Harbor Drive, West Conshohocken,
gage
PA 19428. A copy of this standard may
be obtained from ASTM International,
www.astm.org.)
Test fixture

length. Details relating to specimen size selection, geometrical tolerances, notch con-
figuration, and precracking procedures are contained in the ASTM standard.
During testing, load is applied at a specified rate and measured using a load cell,
which is one component of the testing apparatus. Furthermore, a clip gage, mounted
on the test specimen across the open end of the notch (Figure M.17), monitors crack
displacement. Results are plotted as load (P) versus displacement (v). The test is con-
tinued until fracture, after which the initial crack length (a) (Figures M.16a and M.16b)
is physically measured on the broken specimen halves. From these data a conditional
load PQ is measured, from which a conditional KIc may be determined (and labeled
KQ); this KQ is then evaluated as to its validity as we explain below.
Three different types of load versus displacement curves have been observed,
which are presented in Figure M.18 (and labeled I, II, and III). The procedure for
determination of this conditional PQ value is described as follows: For each curve
type, a tangent is constructed at the initial linear portion of the curve (OA) and
its slope is determined. A straight-line segment having a slope 5% less than this

A Pmax A Pmax A

P5  PQ PQ
Pmax  PQ
P5 P5
Load, P

Type I Type II Type III

95% Secant

O O O
Displacement,

Figure M.18 Three principal types of load versus displacement curves that may be
generated during a fracture toughness test. (ASTM Standard E399-09). (Adapted with
permission from Figure 7 in ASTM E399-09 Standard Test Method for Linear-Elastic
Plane-Strain Fracture Toughness KIc of Metallic Materials. Copyright ASTM International,
100 Barr Harbor Drive, West Conshohocken, PA 19428. A copy of this standard may be
obtained from ASTM International, www.astm.org.)
M.6 Fracture Toughness Testing 29

initial tangent is then constructed from the origin; the intersection of this secant
(OP5) with the load-displacement curve is indicated by the point labeled P5 for each
of the curves shown in this plot. If, on a load-displacement curve, every force point
that precedes P5 is less than P5 (as is the case for only curve I in Figure M.18), then
PQ  P5. On the other hand, when there is a sharp drop in load just past the ter-
mination of the linear-elastic region, such that maximum load on the resultant cusp
precedes and is greater than P5 (Figure M.18, curves II and III), then this maximum
load is taken as PQ.
At this time it becomes necessary to impose the first validity criterionto
determine whether the specimen is too ductile to be tested using this technique.
This criterion is expressed quantitatively by the expression
KIc validity Pmax
criterionmaximum 1.10 (M.32)
degree of ductility PQ

where Pmax is the maximum force that the test specimen is able to sustain (Figure
M.18). If this criterion is not satisfied, then another fracture toughness testing tech-
nique must be employed.10
However, if the criterion specified by Equation M.32 is realized, the next step
is to determine a value for the conditional KQ. For the single-edge notched bend
specimen, the following equation is employed:
PQS
fa b
a
KQ  3/2
(M.33)
BW W
In this expression (and from Figure M.16a)
PQ  the conditional load value, determined as described above
S  distance between support points
B  specimen thickness
W  specimen width (or depth)
a  precrack length

In Equation M.33, f a b is a calibration function that depends on the a/W


a
W
ratio as

fa b
a
W

a 2
1.99  a ba1  b c 2.15  3.93a b  2.7a b d
a a a


a W W W W
3 (M.34)
AW a 3 2
2 c 1  2a bd c1  a bd
a
W W
Similarly, the following equation is used to compute KQ for the compact-
tension specimen configuration:
PQ
fa b
a
KQ  (M.35)
B2W W

10
For example, ASTM Standard E 1820.
30 Online Support Module: Mechanical Engineering

in which B and W are the specimen thickness and width (depth), respectively, and
a is the precrack length (Figure M.16b). In this case

fa b
a
W
a 2 a 3 a 4
a2  b c 0.866  4.64a b  13.32a b  14.72a b  5.6a b d
a a
W W W W W
(M.36)
a 3/2
a1  b
W
Again, KQ is conditional, and before it can be accepted as a valid KIc value,
verification of a condition of plane-strain must be established. Such is possible when
the following criterion is satisfied (for both specimen geometries)11:
KIc validity 2
KQ
b
criterionminimum
crack length and a and B
2.5a (M.37)
sy
minimum specimen
thickness
Here a and B are, respectively, crack length and specimen thickness, and sy is the
0.2% offset yield strength (measured in tension).

By way of summary:
When the criteria specified by Equations M.32 and M.37 are met, then KQ is
a valid value for KIc, and may be reported as such.
If the condition of Equation M.32 is not satisfied, then another testing tech-
nique must be employed.
And, finally, when the criterion of Equation M.32 is met, while at the same
time Equation M.37 is not realized, then the test must be repeated using a
thicker specimen. A new specimen thickness may be estimated by incorporat-
ing the measured value of KQ into Equation M.37.

M.7 IMPACT FRACTURE TESTING12


Prior to the advent of fracture mechanics as a scientific discipline, impact testing tech-
niques were established so as to ascertain the fracture characteristics of materials at
high loading rates. It was realized that the results of laboratory tensile tests (at low
loading rates) could not be extrapolated to predict fracture behavior. For example, un-
der some circumstances normally ductile metals fracture abruptly and with very little
plastic deformation under high loading rates. Impact test conditions were chosen to
represent those most severe relative to the potential for fracturenamely, (1) defor-
mation at a relatively low temperature, (2) a high strain rate (i.e., rate of deformation),
and (3) a triaxial stress state (which may be introduced by the presence of a notch).

Impact Testing Techniques


Charpy test, Izod test Two standardized tests,13 the Charpy and Izod, were designed and are still used to
impact energy measure the impact energy (sometimes also termed notch toughness). The Charpy
11
Note the similarity between Equations M.23 and M.37, the former of which was cited
earlier as a minimum thickness for the condition of plane strain.
12
This section is virtually identical to Section 8.6 of Introduction and Section 9.8 of
Fundamentals.
13
ASTM Standard E 23, Standard Test Methods for Notched Bar Impact Testing of
Metallic Materials.
M.7 Impact Fracture Testing 31

V-notch (CVN) technique is most commonly used in the United States. For both
Charpy and Izod, the specimen is in the shape of a bar of square cross section, into
which a V-notch is machined (Figure M.19a). The apparatus for making V-notch
impact tests is illustrated schematically in Figure M.19b. The load is applied as an
impact blow from a weighted pendulum hammer that is released from a cocked po-
sition at a fixed height h. The specimen is positioned at the base as shown. Upon
release, a knife edge mounted on the pendulum strikes and fractures the specimen
at the notch, which acts as a point of stress concentration for this high-velocity

Figure M.19 (a)


8 mm
Specimen used for (0.32 in.)
Charpy and Izod
impact tests. (b) A
schematic drawing of
an impact testing
apparatus. The
(a)
hammer is released 10 mm
from fixed height h (0.39 in.)

and strikes the


specimen; the energy
10 mm
expended in fracture (0.39 in.)
is reflected in the
difference between h
and the swing height
h. Specimen
placements for both
Charpy and Izod Notch
tests are also shown.
[Figure (b) adapted
from H. W. Hayden, Charpy
W. G. Moffatt, and J. Scale
Wulff, The Structure
and Properties of Izod
Materials, Vol. III,
Mechanical Behavior,
p. 13. Copyright Pointer Starting position
1965 by John
Wiley & Sons, New Hammer
York. Reprinted by
permission of John
Wiley & Sons, Inc.]

End of swing
h
Specimen

h'

Anvil

(b)
32 Online Support Module: Mechanical Engineering

impact blow. The pendulum continues its swing, rising to a maximum height h,
which is lower than h. The energy absorption, computed from the difference
between h and h, is a measure of the impact energy. The primary difference be-
tween the Charpy and Izod techniques lies in the manner of specimen support, as
illustrated in Figure M.19b. Furthermore, these are termed impact tests in light of
the manner of load application. Variables including specimen size and shape as well
as notch configuration and depth influence the test results.
Both plane strain fracture toughness and these impact tests determine the frac-
ture properties of materials. The former are quantitative in nature, in that a specific
property of the material is determined (i.e., KIc). The results of the impact tests, on
the other hand, are more qualitative and are of little use for design purposes. Impact
energies are of interest mainly in a relative sense and for making comparisons
absolute values are of little significance. Attempts have been made to correlate plane
strain fracture toughnesses and CVN energies, with only limited success. Plane strain
fracture toughness tests are not as simple to perform as impact tests; furthermore,
equipment and specimens are more expensive.

Ductile-to-Brittle Transition
One of the primary functions of Charpy and Izod tests is to determine whether a
ductile-to-brittle material experiences a ductile-to-brittle transition with decreasing temperature and,
transition if so, the range of temperatures over which it occurs. As may be noted in the chapter-
opening photograph (of the oil barge) for Chapter 8 of Introduction (Chapter 9 of
Fundamentals), widely used steels can exhibit this ductile-to-brittle transition with
disastrous consequences. The ductile-to-brittle transition is related to the tempera-
ture dependence of the measured impact energy absorption. This transition is rep-
resented for a steel by curve A in Figure M.20. At higher temperatures the CVN

Temperature (F)

40 0 40 80 120 160 200 240 280

100

A
80
100
Shear fracture (%)
Impact energy (J)

Impact
60 energy 80
Shear
fracture
60
40 B
40

20
20

0 0
40 20 0 20 40 60 80 100 120 140
Temperature (C)

Figure M.20 Temperature dependence of the Charpy V-notch impact energy (curve A)
and percent shear fracture (curve B) for an A283 steel. (Reprinted from Welding Journal.
Used by permission of the American Welding Society.)
M.7 Impact Fracture Testing 33

59 12 4 16 24 79

Figure M.21 Photograph of fracture surfaces of A36 steel Charpy V-notch specimens tested
at indicated temperatures (in C). (From R. W. Hertzberg, Deformation and Fracture
Mechanics of Engineering Materials, 3rd edition, Fig. 9.6, p. 329. Copyright 1989 by John
Wiley & Sons, Inc., New York. Reprinted by permission of John Wiley & Sons, Inc.)

energy is relatively large, in correlation with a ductile mode of fracture. As the tem-
perature is lowered, the impact energy drops suddenly over a relatively narrow tem-
perature range, below which the energy has a constant but small value; that is, the
mode of fracture is brittle.
Alternatively, appearance of the failure surface is indicative of the nature of frac-
ture and may be used in transition temperature determinations. For ductile fracture,
this surface appears fibrous or dull (or of shear character) as in the steel specimen of
Figure M.21 that was tested at 79C. Conversely, totally brittle surfaces have a granu-
lar (shiny) texture (or cleavage character) (the 59C specimen, Figure M.21). Over
the ductile-to-brittle transition, features of both types will exist (in Figure M.21, dis-
played by specimens tested at 12C, 4C, 16C, and 24C). Frequently, the percent
shear fracture is plotted as a function of temperaturecurve B in Figure M.20.
For many alloys there is a range of temperatures over which the ductile-to-
brittle transition occurs (Figure M.20); this presents some difficulty in specifying a
single ductile-to-brittle transition temperature. No explicit criterion has been estab-
lished, and so this temperature is often defined as that temperature at which the CVN
energy assumes some value (e.g., 20 J or 15 ft-lbf), or corresponding to some given
fracture appearance (e.g., 50% fibrous fracture). Matters are further complicated inas-
much as a different transition temperature may be realized for each of these criteria.
Perhaps the most conservative transition temperature is that at which the fracture
surface becomes 100% fibrous; on this basis, the transition temperature is approxi-
mately 110C (230F) for the steel alloy that is shown in Figure M.20.
Structures constructed from alloys that exhibit this ductile-to-brittle behavior
should be used only at temperatures above the transition temperature, to avoid brittle
and catastrophic failure. Classic examples of this type of failure occurred, with disas-
trous consequences, during World War II when a number of welded transport ships,
away from combat, suddenly and precipitously split in half.The vessels were constructed
of a steel alloy that possessed adequate ductility according to room-temperature ten-
sile tests. The brittle fractures occurred at relatively low ambient temperatures, at about
4C (40F), in the vicinity of the transition temperature of the alloy. Each fracture crack
originated at some point of stress concentration, probably a sharp corner or fabrication
defect, and then propagated around the entire girth of the ship.
In addition to the ductile-to-brittle transition represented in Figure M.20, two
other general types of impact energy-versus-temperature behavior have been
observed; these are represented schematically by the upper and lower curves of
Figure M.22. Here it may be noted that low-strength FCC metals (some aluminum
34 Online Support Module: Mechanical Engineering

Low-strength (FCC and HCP) metals Figure M.22 Schematic curves for
the three general types of impact
energy-versus-temperature
behavior.

Impact energy
Low-strength steels (BCC)

High-strength materials

Temperature

and copper alloys) and most HCP metals do not experience a ductile-to-brittle tran-
sition (corresponding to the upper curve of Figure M.22), and retain high impact
energies (i.e., remain ductile) with decreasing temperature. For high-strength
materials (e.g., high-strength steels and titanium alloys), the impact energy is also
relatively insensitive to temperature (the lower curve of Figure M.22); however,
these materials are also very brittle, as reflected by their low impact energy values.
And, of course, the characteristic ductile-to-brittle transition is represented by the
middle curve of Figure M.22. As noted, this behavior is typically found in low-
strength steels that have the BCC crystal structure.
For these low-strength steels, the transition temperature is sensitive to both al-
loy composition and microstructure. For example, decreasing the average grain size
results in a lowering of the transition temperature. Hence, refining the grain size both
strengthens (Section 7.8 of Introduction, Section 8.9 of Fundamentals) and toughens
steels. In contrast, increasing the carbon content, while increasing the strength of steels,
also raises the CVN transition of steels, as indicated in Figure M.23.
Izod or Charpy tests are also conducted to assess impact strength of poly-
meric materials. As with metals, polymers may exhibit ductile or brittle fracture
under impact loading conditions, depending on the temperature, specimen size,
strain rate, and mode of loading, as discussed in the preceding section. Both semi-

Figure M.23 Temperature (F)


Influence of carbon 200 0 200 400
content on the
240
Charpy V-notch
300
energy-versus-
temperature 200
behavior for steel.
0.01 0.11
(Reprinted with
Impact energy (ft-lbf)
Impact energy (J)

160
permission from
200 0.22
ASM International,
Metals Park, OH 120
44073-9989, USA;
J. A. Reinbolt and 0.31
0.43 80
W. J. Harris, Jr., 100
Effect of Alloying 0.53
Elements on Notch 0.63 40
Toughness of
0.67
Pearlitic Steels,
Transactions of ASM, 0 0
200 100 0 100 200
Vol. 43, 1951.) Temperature (C)
M.8 Cyclic Stresses 35

crystalline and amorphous polymers are brittle at low temperatures, and both
have relatively low impact strengths. However, they experience a ductile-to-brit-
tle transition over a relatively narrow temperature range, similar to that shown
for a steel in Figure M.20. Of course, impact strength undergoes a gradual de-
crease at still higher temperatures as the polymer begins to soften. Ordinarily,
the two impact characteristics most sought after are a high impact strength at
the ambient temperature and a ductile-to-brittle transition temperature that lies
below room temperature.
Most ceramics also experience a ductile-to-brittle transition, which occurs only
at elevated temperatures, ordinarily in excess of 1000C (1850F).

Fa t i g u e
fatigue Fatigue is a form of failure that occurs in structures subjected to dynamic and fluctu-
ating stresses (e.g., bridges, aircraft, and machine components). Under these circum-
stances it is possible for failure to occur at a stress level considerably lower than the
tensile or yield strength for a static load. The term fatigue is used because this type of
failure normally occurs after a lengthy period of repeated stress or strain cycling.
Fatigue is important inasmuch as it is the single largest cause of failure in metals, es-
timated to comprise approximately 90% of all metallic failures; polymers and ceram-
ics (except for glasses) are also susceptible to this type of failure. Furthermore, fatigue
failure is catastrophic and insidious, occurring very suddenly and without warning.
Fatigue failure is brittlelike in nature even in normally ductile metals, in that
there is very little, if any, gross plastic deformation associated with failure. The
process occurs by the initiation and propagation of cracks, and ordinarily the frac-
ture surface is perpendicular to the direction of an applied tensile stress.

M.8 CYCLIC STRESSES14


The applied stress may be axial (tension-compression), flexural (bending), or torsional
(twisting) in nature. In general, three different fluctuating stress-time modes are pos-
sible. One is represented schematically by a regular and sinusoidal time dependence
in Figure M.24a, wherein the amplitude is symmetrical about a mean zero stress level,
for example, alternating from a maximum tensile stress (smax) to a minimum com-
pressive stress (smin) of equal magnitude; this is referred to as a reversed stress cycle.
Another type, termed repeated stress cycle, is illustrated in Figure M.24b; the maxima
and minima are asymmetrical relative to the zero stress level. Finally, the stress level
may vary randomly in amplitude and frequency, as exemplified in Figure M.24c.
Also indicated in Figure M.24b are several parameters used to characterize
the fluctuating stress cycle. The stress amplitude alternates about a mean stress sm,
Mean stress for defined as the average of the maximum and minimum stresses in the cycle, or
cyclic loading
dependence on smax  smin
maximum and sm  (M.38)
2
minimum stress
levels
Furthermore, the range of stress sr is just the difference between smax and
sminnamely,
Computation of
range of stress for sr  smax  smin (M.39)
cyclic loading
14
This section is virtually identical to Section 8.7 of Introduction and Section 9.9 of
Fundamentals.
36 Online Support Module: Mechanical Engineering

Figure M.24 Variation


of stress with time that
max accounts for fatigue

Tension
failures. (a) Reversed
+
stress cycle, in which the

Stress
stress alternates from a
0

Compression
maximum tensile stress
(+) to a maximum


compressive stress ()
min of equal magnitude. (b)
Time Repeated stress cycle, in
(a) which maximum and
minimum stresses are
asymmetrical relative to
the zero stress level;
max
mean stress sm, range
a
of stress sr, and stress
amplitude sa are
Tension

r
+ indicated. (c) Random
m stress cycle.
Stress

0
Compression

min

Time
(b)
Tension

+
Stress
Compression

Time
(c)

Stress amplitude sa is just one-half of this range of stress, or


Computation of
stress amplitude for sr smax  smin
sa   (M.40)
cyclic loading 2 2

Finally, the stress ratio R is just the ratio of minimum and maximum stress amplitudes:

smin
Computation of R (M.41)
stress ratio smax

By convention, tensile stresses are positive and compressive stresses are nega-
tive. For example, for the reversed stress cycle, the value of R is 1.
M.9 The SN Curve 37

Flexible coupling

Counter
Specimen
High-speed
motor
ing
Bearing housing g hous
Bearin

Load Load

Figure M.25 Schematic diagram of fatigue-testing apparatus for making rotating-


bending tests. (From Keyser, Materials Science in Engineering, 4th Edition, 1986.
Adapted by permission of Pearson Education, Inc., Upper Saddle River, NJ.)

M.9 THE S-N CURVE15


As with other mechanical characteristics, the fatigue properties of materials can be
determined from laboratory simulation tests.16 A test apparatus should be designed
to duplicate as nearly as possible the service stress conditions (stress level, time
frequency, stress pattern, etc.). A schematic diagram of a rotating-bending test ap-
paratus, commonly used for fatigue testing, is shown in Figure M.25; the compres-
sion and tensile stresses are imposed on the specimen as it is simultaneously bent
and rotated. Tests are also frequently conducted using an alternating uniaxial ten-
sion-compression stress cycle.
A series of tests are commenced by subjecting a specimen to the stress cycling
at a relatively large maximum stress amplitude (smax), usually on the order of two
thirds of the static tensile strength; the number of cycles to failure is counted. This
procedure is repeated on other specimens at progressively decreasing maximum
stress amplitudes. Data are plotted as stress S versus the logarithm of the number N
of cycles to failure for each of the specimens. The values of S are normally taken as
stress amplitudes (sa, Equation M.40); on occasion, smax or smin values may be used.
Two distinct types of SN behavior are observed, which are represented
schematically in Figure M.26. As these plots indicate, the higher the magnitude of
the stress, the smaller the number of cycles the material is capable of sustaining
before failure. For some ferrous (iron base) and titanium alloys, the SN curve
(Figure M.26a) becomes horizontal at higher N values; or, there is a limiting stress
fatigue limit level, called the fatigue limit (also sometimes the endurance limit), below which fa-
tigue failure will not occur. This fatigue limit represents the largest value of fluctu-
ating stress that will not cause failure for essentially an infinite number of cycles.
For many steels, fatigue limits range between 35% and 60% of the tensile strength.
Most nonferrous alloys (e.g., aluminum, copper, magnesium) do not have a fa-
tigue limit, in that the SN curve continues its downward trend at increasingly
greater N values (Figure M.26b). Thus, fatigue will ultimately occur regardless of
the magnitude of the stress. For these materials, one fatigue response is specified as
fatigue strength fatigue strength, which is defined as the stress level at which failure will occur for
some specified number of cycles (e.g., 107 cycles). The determination of fatigue
strength is also demonstrated in Figure M.26b.
15
This section is virtually identical to Section 8.8 of Introduction and Section 9.10 of
Fundamentals.
16
See ASTM Standard E 466, Standard Practice for Conducting Force Controlled Con-
stant Amplitude Axial Fatigue Tests of Metallic Materials, and ASTM Standard E 468,
Standard Practice for Presentation of Constant Amplitude Fatigue Test Results for Metal-
lic Materials.
38 Online Support Module: Mechanical Engineering

Figure M.26 Stress


amplitude (S) versus
logarithm of the
number of cycles to

Stress amplitude, S
fatigue failure (N)
for (a) a material
that displays a
Fatigue
limit fatigue limit, and (b)
a material that does
not display a fatigue
limit.

103 104 105 106 107 108 109 1010


Cycles to failure, N
(logarithmic scale)
(a)
Stress amplitude, S

S1

Fatigue strength
at N1 cycles

103 104 Fatigue life 107 N1 108 109 1010


at stress S1

Cycles to failure, N
(logarithmic scale)
(b)

Another important parameter that characterizes a materials fatigue behavior


fatigue life is fatigue life Nf. It is the number of cycles to cause failure at a specified stress level,
as taken from the SN plot (Figure M.26b).
Unfortunately, there always exists considerable scatter in fatigue datathat is,
a variation in the measured N value for a number of specimens tested at the same
stress level. This variation may lead to significant design uncertainties when fatigue
life and/or fatigue limit (or strength) are being considered. The scatter in results is
a consequence of the fatigue sensitivity to a number of test and material parame-
ters that are impossible to control precisely. These parameters include specimen fab-
rication and surface preparation, metallurgical variables, specimen alignment in the
apparatus, mean stress, and test frequency.
Fatigue SN curves similar to those shown in Figure M.26 represent best fit
curves that have been drawn through average-value data points. It is a little unset-
tling to realize that approximately one-half of the specimens tested actually failed
at stress levels lying nearly 25% below the curve (as determined on the basis of sta-
tistical treatments).
Several statistical techniques have been developed to specify fatigue life and
fatigue limit in terms of probabilities. One convenient way of representing data
M.10 Crack Initiation and Propagation 39

Figure M.27 70
Fatigue SN
probability of failure 60
400 P = 0.99
curves for a 7075-T6
aluminum alloy; P P = 0.90 50

Stress (103 psi)


Stress, S (MPa)
denotes the
300 P = 0.50
probability of 40
failure. (From G. M.
P = 0.01
Sinclair and T. J. 200 30
P = 0.10
Dolan, Trans. ASME,
75, 1953, p. 867. 20
Reprinted with 100
permission of the 10
American Society of 104 105 106 107 108 109
Mechanical Cycles to failure, N
Engineers.) (logarithmic scale)

treated in this manner is with a series of constant probability curves, several of which
are plotted in Figure M.27. The P value associated with each curve represents the
probability of failure. For example, at a stress of 200 MPa (30,000 psi), we would
expect 1% of the specimens to fail at about 106 cycles and 50% to fail at about
2 107 cycles, and so on. Remember that SN curves represented in the literature
are normally average values, unless noted otherwise.
The fatigue behaviors represented in Figures M.26a and M.26b may be classi-
fied into two domains. One is associated with relatively high loads that produce not
only elastic strain but also some plastic strain during each cycle. Consequently, fa-
tigue lives are relatively short; this domain is termed low-cycle fatigue and occurs
at less than about 104 to 105 cycles. For lower stress levels wherein deformations
are totally elastic, longer lives result. This is called high-cycle fatigue inasmuch as
relatively large numbers of cycles are required to produce fatigue failure. High-
cycle fatigue is associated with fatigue lives greater than about 104 to 105 cycles.

M.10 CRACK INITIATION AND PROPAGATION 17


The process of fatigue failure is characterized by three distinct steps: (1) crack ini-
tiation, wherein a small crack forms at some point of high stress concentration; (2)
crack propagation, during which this crack advances incrementally with each stress
cycle; and (3) final failure, which occurs very rapidly once the advancing crack has
reached a critical size. The fatigue life Nf, the total number of cycles to failure, there-
fore can be taken as the sum of the number of cycles for crack initiation Ni and
crack propagation Np:
Nf  Ni  Np (M.42)
The contribution of the final failure step to the total fatigue life is insignificant since
it occurs so rapidly. Relative proportions to the total life of Ni and Np depend on
the particular material and test conditions. At low stress levels (i.e., for high-cycle
fatigue), a large fraction of the fatigue life is utilized in crack initiation. With in-
creasing stress level, Ni decreases and the cracks form more rapidly. Thus, for low-
cycle fatigue (high stress levels), the propagation step predominates (i.e., Np > Ni).

17
This section is an expanded and more detailed version of Section 8.9 of Introduction and
Section 9.12 of Fundamentals.
40 Online Support Module: Mechanical Engineering

s Figure M.28 Schematic representation showing stages


I and II of fatigue crack propagation in polycrystalline
metals. (Copyright ASTM. Reprinted with permission.)

Stage II

Stage I

Cracks associated with fatigue failure almost always initiate (or nucleate) on
the surface of a component at some point of stress concentration. Crack nucleation
sites include surface scratches, sharp fillets, keyways, threads, dents, and the like. In
addition, cyclic loading can produce microscopic surface discontinuities resulting
from dislocation slip steps which may also act as stress raisers, and therefore as
crack initiation sites.
Once a stable crack has nucleated, it then initially propagates very slowly and,
in polycrystalline metals, along crystallographic planes of high shear stress; this is
sometimes termed stage I propagation (Figure M.28). This stage may constitute
a large or small fraction of the total fatigue life depending on stress level and
the nature of the test specimen; high stresses and the presence of notches favor a
short-lived stage I. In polycrystalline metals, cracks normally extend through only
several grains during this propagation stage. The fatigue surface that is formed dur-
ing stage I propagation has a flat and featureless appearance.
Eventually, a second propagation stage (stage II ) takes over, wherein the crack
extension rate increases dramatically. Furthermore, at this point there is also a
change in propagation direction to one that is roughly perpendicular to the applied
tensile stress (see Figure M.28). During this stage of propagation, crack growth
proceeds by a repetitive plastic blunting and sharpening process at the crack tip, a
mechanism illustrated in Figure M.29. At the beginning of the stress cycle (zero or
maximum compressive load), the crack tip has the shape of a sharp double-notch
(Figure M.29a). As the tensile stress is applied (Figure M.29b), localized deforma-
tion occurs at each of these tip notches along slip planes that are oriented at 45
angles relative to the plane of the crack. With increased crack widening, the tip
advances by continued shear deformation and the assumption of a blunted config-
uration (Figure M.29c). During compression, the directions of shear deformation
at the crack tip are reversed (Figure M.29d) until, at the culmination of the cycle,
a new sharp double-notch tip has formed (Figure M.29e). Thus, the crack tip has
advanced a one-notch distance during the course of a complete cycle. This process is
repeated with each subsequent cycle until eventually some critical crack dimension is
achieved that precipitates the final failure step and catastrophic failure ensues.
The region of a fracture surface that formed during stage II propagation may
be characterized by two types of markings termed beachmarks and striations. Both
of these features indicate the position of the crack tip at some point in time and
M.10 Crack Initiation and Propagation 41

(a)
(d)

(e)
(b)

(c) (f)

Figure M.29 Fatigue crack propagation mechanism (stage II) by repetitive crack tip
plastic blunting and sharpening; (a) zero or maximum compressive load, (b) small tensile
load, (c) maximum tensile load, (d) small compressive load, (e) zero or maximum
compressive load, (f ) small tensile load. The loading axis is vertical. (Copyright ASTM.
Reprinted with permission.)

appear as concentric ridges that expand away from the crack initiation site(s),
frequently in a circular or semicircular pattern. Beachmarks (sometimes also called
clamshell marks) are of macroscopic dimensions (Figure M.30), and may be
observed with the unaided eye. These markings are found for components that
experienced interruptions during stage II propagationfor example, a machine that
operated only during normal work-shift hours. Each beachmark band represents a
period of time over which crack growth occurred.

Figure M.30 Fracture


surface of a rotating steel
shaft that experienced
fatigue failure. Beachmark
ridges are visible in the
photograph. (Reproduced
with permission from D. J.
Wulpi, Understanding How
Components Fail, American
Society for Metals, Materials
Park, OH, 1985.)
42 Online Support Module: Mechanical Engineering

Figure M.31 Transmission


electron fractograph showing
fatigue striations in
aluminum. 9000 . (From V.
J. Colangelo and F. A.
Heiser, Analysis of
Metallurgical Failures, 2nd
edition. Copyright 1987 by
John Wiley & Sons, New
York. Reprinted by
permission of John Wiley &
Sons, Inc.)

1 m

On the other hand, fatigue striations are microscopic in size and subject to
observation with the electron microscope (either TEM or SEM). Figure M.31 is an
electron fractograph which shows this feature. Each striation is thought to repre-
sent the advance distance of the crack front during a single load cycle. Striation
width depends on, and increases with, increasing stress range.
At this point it should be emphasized that although both beachmarks and stri-
ations are fatigue fracture surface features having similar appearances, they are nev-
ertheless different, both in origin and size. There may be thousands of striations
within a single beachmark.
Often the cause of failure may be deduced after examination of the failure sur-
faces. The presence of beachmarks and/or striations on a fracture surface confirms
that the cause of failure was fatigue. Nevertheless, the absence of either or both
does not exclude fatigue as the cause of failure.
One final comment regarding fatigue failure surfaces: Beachmarks and striations
will not appear on that region over which the rapid failure occurs (which region is
noted in Figure M.32). Rather, the rapid failure may be either ductile or brittle; evi-
dence of plastic deformation will be present for ductile, and absent for brittle, failure.

M.11 CRACK PROPAGATION RATE


Even though measures may be taken to minimize the possibility of fatigue failure,
cracks and crack nucleation sites will always exist in structural components. Under
the influence of cyclic stresses, cracks will inevitably form and grow; this process, if
unabated, can ultimately lead to failure. The intent of the present discussion is to
develop a criterion whereby fatigue life may be predicted on the basis of material
and stress state parameters. Principles of fracture mechanics (Section M.4) will be
employed inasmuch as the treatment involves determination of a maximum crack
length that may be tolerated without inducing failure. It should be noted that this
discussion relates to the domain of high-cycle fatigue, that is, for fatigue lives greater
than about 104 to 105 cycles.
M.11 Crack Propagation Rate 43

Region of slow Figure M.32 Fatigue


crack propagation
failure surface. A crack
formed at the top edge.
The smooth region also
2 cm
near the top corresponds
to the area over which
the crack propagated
slowly. Rapid failure
occurred over the area
having a dull and fibrous
texture (the largest area).
Approximately 0.5 .
(Reproduced by
permission from Metals
Handbook: Fractography
and Atlas of Fractographs,
Vol. 9, 8th edition, H. E.
Boyer, Editor, American
Society for Metals, 1974.)

Region of rapid failure

Results of fatigue studies have shown that the life of a structural component
may be related to the rate of crack growth. During stage II propagation, cracks may
grow from a barely perceivable size to some critical length. Experimental techniques
are available which are employed to monitor crack length during the cyclic stress-
ing. Data are recorded and then plotted as crack length a versus the number of
cycles N. A typical plot is shown in Figure M.33, where curves are included from
data generated at two different stress levels; the initial crack length a0 for both sets
of tests is the same. Crack growth rate da/dN is taken as the slope at some point of
the curve. Two important results are worth noting: (1) initially, growth rate is small,
but increases with increasing crack length; and (2) growth rate is enhanced with
increasing applied stress level and for a specific crack length (a1 in Figure M.33).

Figure M.33 Crack length


s2 > s 1 versus the number of cycles at
s1 stress levels s1 and s2 for
fatigue studies. Crack growth
rate da/dN is indicated at
s2
crack length a1 for both stress
Crack length a

da
 dN  a1, s2 levels.
da
a1  dN  a1, s1

a0

Cycles N
44 Online Support Module: Mechanical Engineering

Fatigue crack propagation rate during stage II is a function of not only stress
level and crack size but also material variables. Mathematically, this rate may be
expressed in terms of the stress intensity factor K (developed using fracture me-
chanics in Section M.4) and takes the form
Dependence of stage
II crack propagation da
rate on stress A1K2 m (M.43)
dN
intensity factor
range at a crack tip
The parameters A and m are constants for the particular material, which will also
depend on environment, frequency, and the stress ratio (R in Equation M.41). The
value of m normally ranges between 1 and 6.
Furthermore, K is the stress intensity factor range at the crack tip, that is,
K Kmax Kmin (M.44a)
or, from Equation M.19,
K Ys1pa Y 1smax smin 2 1pa (M.44b)
Since crack growth stops or is negligible for a compression portion of the stress
cycle, if smin is compressive, then Kmin and smin are taken to be zero; that is, K
Kmax and s smax. Also note that Kmax and Kmin in Equation M.44a represent
stress intensity factors, not the fracture toughness Kc nor the plane strain fracture
toughness KIc.
The typical fatigue crack growth rate behavior of materials is represented
schematically in Figure M.34 as the logarithm of crack growth rate da/dN versus
the logarithm of the stress intensity factor range K. The resulting curve has a sig-
moidal shape which may be divided into three distinct regions, labeled I, II, and III.
In region I (at low stress levels and/or small crack sizes), preexisting cracks will not
grow with cyclic loading. Furthermore, associated with region III is accelerated crack
growth, which occurs just prior to the rapid fracture.
The curve is essentially linear in region II, which is consistent with Equation M.43.
This may be confirmed by taking the logarithm of both sides of this expression, which
leads to

log a b log 3A1K2 m 4


da
(M.45a)
dN

log a b m log K log A


da
(M.45b)
dN
Indeed, according to Equation M.45b, a straight-line segment will result when
log (da/dN)-versus-log K data are plotted; the slope and intercept correspond to
the values of m and log A, respectively, which may be determined from test data
that have been represented in the manner of Figure M.34. Figure M.35 is one such
plot for a Ni-Mo-V steel alloy. The linearity of the data may be noted, which veri-
fies the power law relationship of Equation M.43.
One of the goals of failure analysis is to be able to predict fatigue life for some
component, given its service constraints and laboratory test data. We are now able
to develop an analytical expression for Nf, due to stage II, by integration of Equa-
tion M.43. Rearrangement is first necessary as follows:
da
dN (M.46)
A1K2 m
M.11 Crack Propagation Rate 45

Figure M.34 Schematic


representation of logarithm
fatigue crack propagation rate
da/dN versus logarithm stress
intensity factor range K. The
three regions of different crack
growth response (I, II, and III)
are indicated. (Reprinted with
permission from ASM
Fatigue crack growth rate, da (log scale)
da
dN
 A(K)m International, Metals Park, OH
44073-9989. W. G. Clark, Jr.,
How Fatigue Crack Initiation
dN

and Growth Properties Affect


Material Selection and Design
Criteria, Metals Engineering
Quarterly, Vol. 14, No. 3, 1974.)

Region I Region II Region III


Non- Linear relationship Unstable
propagating between crack
da
fatigue log K and log dN growth
cracks

Stress intensity factor range, K (log scale)

which may be integrated as


Nf ac


da
Nf  dN  (M.47)
0 a0
A1K2 m
The limits on the second integral are between the initial flaw length a0, which may
be measured using nondestructive examination techniques, and the critical crack
length ac determined from fracture toughness tests.
Substitution of the expression for K (Equation M.44b) leads to
ac


da
Nf 
a0
A1Ys1pa2 m
Computation of
predicted fatigue life
ac


1 da

Ap 1s2 m
m/2 m m/2
(M.48)
a0 Y a

Here it is assumed that s (or smax  smin) is constant; furthermore, in general Y will
depend on crack length a and therefore cannot be removed from within the integral.
A word of caution: Equation M.48 presumes the validity of Equation M.43 over
the entire life of the component; it ignores the time taken to initiate the crack and also
for final failure. Therefore, this expression should only be taken as an estimate of Nf.
46 Online Support Module: Mechanical Engineering

MPa m Figure M.35 Plot of


20 40 60 80 100 logarithm crack growth
10 3
rate versus logarithm
0.2% yield strength = 84,500 psi stress intensity factor
Test temp. = 24C (75F) range for a Ni-Mo-V
Test frequency = 1800 cpm (30 Hz)
Max. cyclic load, ibf steel. These data provide
10 2
experimental verification
for the linearity of
4000
region II of Figure M.34,
5000
and, in addition, for
6000
7000
Equation M.45b [i.e., a
8000
straight-line segment
9200 when log (da/dN) is
9900 plotted versus log K].
10 4 (Reprinted by

(mm/cycle)
(in./cycle)

permission of the
Society for
Experimental
Crack growth rate, dN
da

Mechanics, Inc.)

Crack growth rate, dN


da
10 3

10 5

10 4

10 6
10 20 40 60 80 100
Stress intensity factor range, K
(10 3 psi in.)

DESIGN EXAMPLE M.2

Fatigue Life Prediction


A relatively large sheet of steel is to be exposed to cyclic tensile and compres-
sive stresses of magnitudes 100 MPa and 50 MPa, respectively. Prior to testing, it
has been determined that the length of the largest surface crack is 2.0 mm (2
103 m). Estimate the fatigue life of this sheet if its plane strain fracture tough-
ness is 25 MPa 1m and the values of m and A in Equation M.43 are 3.0 and 1.0
1012, respectively, for s in MPa and a in m. Assume that the parameter Y is
independent of crack length and has a value of 1.0.
M.12 Factors That Affect Fatigue Life 47

Solution
It first becomes necessary to compute the critical crack length ac, the integration up-
per limit in Equation M.48. Equation M.25 is employed for this computation, as-
suming a stress level of 100 MPa, since this is the maximum tensile stress. Therefore,
1 KIc 2
ac a b
p sY
1 25 MPa 1m 2
c d 0.02 m
p 1100 MPa2112
We now want to solve Equation M.48 using 0.002 m as the lower integration limit
a0, as stipulated in the problem. The value of s is just 100 MPa, the magnitude of
the tensile stress, since smin is compressive. Therefore, integration yields
a


1 c
da
Nf
Ap 1s2 a0 Ymam/2
m/2 m

ac


1
a 3/2da
Ap3/2 1s2 3Y3 a0
ac
122a `
1 1 2

Ap3 2 1s2 3Y3 a0

a b
2 1 1

Ap3 2 1s2 3Y3 1a0 1ac

a b
2 1 1

11 10 21p2 11002 112 10.002
12 3 2 3 3
10.02
5.49 106 cycles

M.12 FACTORS THAT AFFECT FATIGUE LIFE18


As mentioned in Section M.9, the fatigue behavior of engineering materials is highly
sensitive to a number of variables. Some of these factors include mean stress level,
geometrical design, surface effects, and metallurgical variables, as well as the environ-
ment. This section is devoted to a discussion of these factors and, in addition, to meas-
ures that may be taken to improve the fatigue resistance of structural components.

Mean Stress
The dependence of fatigue life on stress amplitude is represented on the S-N plot.
Such data are taken for a constant mean stress sm, often for the reversed cycle sit-
uation (sm 0). Mean stress, however, will also affect fatigue life; this influence
may be represented by a series of S-N curves, each measured at a different sm, as
depicted schematically in Figure M.36. As may be noted, increasing the mean stress
level leads to a decrease in fatigue life (as well as a decrease in fatigue strength).
Goodman equation Empirical equations have been developed that express the dependence of fatigue
computation of the strength on mean stress. One of these, the Goodman equation, is written as follows:
nonzero-mean-stress
fatigue limit for a
sfs sfs0 a1 b
material using tensile sm
(M.49)
strength and zero- TS
mean-stress fatigue
limit values 18
This section is a slightly modified version of Section 8.10 of Introduction and Section
9.13 of Fundamentals.
48 Online Support Module: Mechanical Engineering

Figure M.36 Demonstration of influence


of mean stress sm on SN fatigue
m3 > m2 > m1 behavior.

Stress amplitude, a
m1
m2

m3

Cycles to failure, N
(logarithmic scale)

In this expression
sm  mean stress
sfs  fatigue strength for sm  0
sfs0  fatigue strength for sm  0
TS  tensile strength

Surface Effects
For many common loading situations, the maximum stress within a component or
structure occurs at its surface. Consequently, most cracks leading to fatigue failure
originate at surface positions, specifically at stress amplification sites. Therefore, it
has been observed that fatigue life is especially sensitive to the condition and con-
figuration of the component surface. Numerous factors influence fatigue resistance,
the proper management of which will lead to an improvement in fatigue life. These
include design criteria as well as various surface treatments.

Design Factors
The design of a component can have a significant influence on its fatigue charac-
teristics. Any notch or geometrical discontinuity can act as a stress raiser and fatigue
crack initiation site; these design features include grooves, holes, keyways, threads,
and so on. The sharper the discontinuity (i.e., the smaller the radius of curvature),
the more severe the stress concentration. The probability of fatigue failure may be
reduced by avoiding (when possible) these structural irregularities, or by making de-
sign modifications whereby sudden contour changes leading to sharp corners are
eliminatedfor example, calling for rounded fillets with large radii of curvature at
the point where there is a change in diameter for a rotating shaft (Figure M.37).

Figure M.37 Demonstration of how design


Fillet can reduce stress amplification. (a) Poor
design: sharp corner. (b) Good design:
fatigue lifetime improved by incorporating
rounded fillet into a rotating shaft at the
point where there is a change in diameter.

(a) ( b)
M.12 Factors That Affect Fatigue Life 49

Surface Treatments
During machining operations, small scratches and grooves are invariably introduced
into the workpiece surface by cutting tool action. These surface markings can limit
the fatigue life. It has been observed that improving the surface finish by polishing
will enhance fatigue life significantly.
One of the most effective methods of increasing fatigue performance is by
imposing residual compressive stresses within a thin outer surface layer. Thus, a
surface tensile stress of external origin will be partially nullified and reduced in
magnitude by the residual compressive stress. The net effect is that the likelihood
of crack formation and therefore of fatigue failure is reduced.
Residual compressive stresses are commonly introduced into ductile metals me-
chanically by localized plastic deformation within the outer surface region. Com-
mercially, this is often accomplished by a process termed shot peening. Small, hard
particles (shot) having diameters within the range of 0.1 to 1.0 mm are projected
at high velocities onto the surface to be treated. The resulting deformation induces
compressive stresses to a depth of between one-quarter and one-half of the shot
diameter. The influence of shot peening on the fatigue behavior of steel is demon-
strated schematically in Figure M.38.
case hardening Case hardening is a technique by which both surface hardness and fatigue life
are enhanced for steel alloys. This is accomplished by a carburizing or nitriding
process whereby a component is exposed to a carbonaceous or nitrogenous at-
mosphere at an elevated temperature.A carbon- or nitrogen-rich outer surface layer
(or case) is introduced by atomic diffusion from the gaseous phase. The case is
normally on the order of 1 mm deep and is harder than the inner core of material.
(The influence of carbon content on hardness for Fe-C alloys is demonstrated in
Figure 10.29a of Introduction and Figure 11.30a of Fundamentals.) The improve-
ment of fatigue properties results from increased hardness within the case, as well
as from the introduction of residual compressive stresses that attends the carburiz-
ing or nitriding process. A carbon-rich outer case may be observed for the gear
shown in the top chapter-opening photograph for Chapter 5 for Introduction (Chap-
ter 6 of Fundamentals); it appears as a dark outer rim within the sectioned segment.
The increase in case hardness is demonstrated in the photomicrograph appearing
in Figure M.39. The dark and elongated diamond shapes are Knoop microhardness
indentations. The upper indentation, lying within the carburized layer, is smaller
than the core indentation.

Figure M.38 Schematic SN fatigue curves


for normal and shot-peened steel.

Shot peened
Stress amplitude

Normal

Cycles to failure
(logarithmic scale)
50 Online Support Module: Mechanical Engineering

Figure M.39 Photomicrograph


showing both core (bottom) and
carburized outer case (top) regions of a
Case case-hardened steel. The case is harder
as attested by the smaller
microhardness indentation. 100 .
(From R. W. Hertzberg, Deformation
and Fracture Mechanics of Engineering
Materials, 3rd edition. Copyright
1989 by John Wiley & Sons, New
York. Reprinted by permission of
Core
region John Wiley & Sons, Inc.)

M.13 ENVIRONMENTAL EFFECTS 19


Environmental factors may also affect the fatigue behavior of materials. A few brief
comments will be given relative to two types of environment-assisted fatigue fail-
ure: thermal fatigue and corrosion fatigue.
thermal fatigue Thermal fatigue is normally induced at elevated temperatures by fluctuating
thermal stresses; mechanical stresses from an external source need not be present.
The origin of these thermal stresses is the restraint to the dimensional expansion
and/or contraction that would normally occur in a structural member with varia-
tions in temperature. The magnitude of a thermal stress developed by a tempera-
ture change T is dependent on the coefficient of thermal expansion al and the
Thermal stress modulus of elasticity E according to
dependence on
coefficient of
thermal expansion, s  alET (M.50)
modulus of elasticity,
and temperature [The topics of thermal expansion and thermal stresses are discussed in Sections 19.3
change and 19.5 of Introduction (Sections 17.3 and 17.5 of Fundamentals).] Of course,
thermal stresses will not arise if this mechanical restraint is absent. Therefore, one
obvious way to prevent this type of fatigue is to eliminate, or at least reduce,
the restraint source, thus allowing unhindered dimensional changes with tempera-
ture variations, or to choose materials with appropriate physical properties.
Failure that occurs by the simultaneous action of a cyclic stress and chemical
corrosion fatigue attack is termed corrosion fatigue. Corrosive environments have a deleterious in-
fluence and produce shorter fatigue lives. Even the normal ambient atmosphere will
affect the fatigue behavior of some materials. Small pits may form as a result of
chemical reactions between the environment and material, which serve as points of
stress concentration, and therefore as crack nucleation sites. In addition, crack prop-
agation rate is enhanced as a result of the corrosive environment. The nature of the
stress cycles will influence the fatigue behavior; for example, lowering the load ap-
plication frequency leads to longer periods during which the opened crack is in con-
tact with the environment and to a reduction in the fatigue life.

19
This section is virtually identical to Section 8.11 of Introduction and Section 9.14 of
Fundamentals.
M.14 Mechanics of Spring Deformation 51

Several approaches to corrosion fatigue prevention exist. On one hand, we can


take measures to reduce the rate of corrosion by some of the techniques discussed
in Chapter 17 of Introduction (Chapter 16 of Fundamentals)for example, apply
protective surface coatings, select a more corrosion-resistant material, and reduce
the corrosiveness of the environment. And/or it might be advisable to take actions
to minimize the probability of normal fatigue failure, as outlined previouslyfor
example, reduce the applied tensile stress level and impose residual compressive
stresses on the surface of the member.

A u t o m o b i l e Va l ve S p r i n g ( C a s e St u d y )
The following submodule is a case study that discusses the valve spring found in a
typical automobile engine. Issues addressed include mechanics of the deformation
of helical springs, constraints imposed on the deformation of a typical valve spring,
and, in addition, one of the steel alloys that is commonly used for these springs and
the rational for its use.

M.14 MECHANICS OF SPRING DEFORMATION


The basic function of a spring is to store mechanical energy as it is initially elasti-
cally deformed and then recoup this energy at a later time as the spring recoils. In
this section helical springs that are used in mattresses and in retractable pens and
as suspension springs in automobiles are discussed. A stress analysis will be con-
ducted on this type of spring, and the results will then be applied to a valve spring
that is used in automobile engines.
Consider the helical spring shown in Figure M.40, which has been constructed
of wire having a circular cross section of diameter d; the coil center-to-center di-
ameter is denoted as D. The application of a compressive force F causes a twisting
force, or moment, denoted T, as shown in the figure. A combination of shear stresses
result, the sum of which, t, is
8FD
t Kw (M.51)
pd3
where Kw is a force-independent constant that is a function of the D/d ratio:
D 0.140
Kw  1.60a b (M.52)
d
F Figure M.40 Schematic
diagram of a helical spring
showing the twisting moment T
that results from the
compressive force F. (Adapted
from K. Edwards and P.
McKee, Fundamentals of
Mechanical Component Design.
Copyright 1991 by McGraw-
d Hill, Inc. Reproduced with
T permission of The McGraw-Hill
Companies.)

D
F
52 Online Support Module: Mechanical Engineering

D D D
2 2 2

(a) (b)

Figure M.41 Schematic diagrams of one coil of a helical spring, (a) prior to being
compressed, and (b) showing the deflection dc produced from the compressive force F.
(Adapted from K. Edwards and P. McKee, Fundamentals of Mechanical Component
Design. Copyright 1991 by McGraw-Hill, Inc. Reproduced with permission of The
McGraw-Hill Companies.)

In response to the force F, the coiled spring will experience deflection, which
will be assumed to be totally elastic. The amount of deflection per coil of spring, dc,
as indicated in Figure M.41, is given by the expression
8FD3
dc  (M.53)
d4G
where G is the shear modulus of the material from which the spring is constructed.
Furthermore, dc may be computed from the total spring deflection, ds, and the num-
ber of effective spring coils, Nc, as
ds
dc  (M.54)
Nc
Now, solving for F in Equation M.53 gives
d4dcG
F (M.55)
8D3
and substituting for F in Equation M.51 leads to
dcGd
t Kw (M.56)
pD2

Condition for Under normal circumstances, it is desired that a spring experiences no per-
nonpermanent spring manent deformation upon loading; this means that the right-hand side of Equa-
deformationshear tion M.56 must be less than the shear yield strength ty of the spring material, or
yield strength and its that
relationship to shear
modulus, number of
effective coils, and
dcGd
ty 7 Kw (M.57)
spring and wire pD2
diameters
M.15 VALVE SPRING DESIGN AND MATERIAL
REQUIREMENTS
We shall now apply the results of the preceding section to an automobile valve spring.
A cut-away schematic diagram of an automobile engine showing these springs is pre-
sented in Figure M.42. Functionally, springs of this type permit both intake and ex-
haust valves to alternately open and close as the engine is in operation. Rotation of
the camshaft causes a valve to open and its spring to be compressed, so that the load
M.15 Valve Spring Design and Material Requirements 53

Figure M.42 Cutaway drawing of a


Cam section of an automobile engine in
which various components including
Camshaft valves and valve springs are shown.

Valve
spring

Exhaust
valve

Intake
valve

Piston

Crankshaft

on the spring is increased. The stored energy in the spring then forces the valve to
close as the camshaft continues its rotation. This process occurs for each valve for
each engine cycle, and over the lifetime of the engine it occurs many millions of times.
Furthermore, during normal engine operation, the temperature of the springs is
approximately 80C (175F).
A photograph of a typical valve spring is shown in Figure M.43. The spring has
a total length of 1.67 in. (42 mm), is constructed of wire having a diameter d of 0.170

Figure M.43 Photograph of a typical automobile


valve spring.
54 Online Support Module: Mechanical Engineering

in. (4.3 mm), has six coils (only four of which are active), and has a center-to-
center diameter D of 1.062 in. (27 mm). Furthermore, when installed and when a
valve is completely closed, its spring is compressed a total of 0.24 in. (6.1 mm),
which, from Equation M.54, gives an installed deflection per coil dic of

 0.060 in./coil 11.5 mm/coil2


0.24 in.
dic 
4 coils
The cam lift is 0.30 in. (7.6 mm), which means that when the cam completely opens
a valve, the spring experiences a maximum total deflection equal to the sum of the
valve lift and the compressed deflection, namely, 0.30 in.  0.24 in.  0.54 in. (13.7
mm). Hence, the maximum deflection per coil, dmc, is

 0.135 in./coil 13.4 mm/coil2


0.54 in.
dmc 
4 coils
Thus, we have available all of the parameters in Equation M.57 (taking dc  dmc),
except for ty, the required shear yield strength of the spring material.
However, the material parameter of interest is really not ty inasmuch as the
spring is continually stress cycled as the valve opens and closes during engine op-
eration; this necessitates designing against the possibility of failure by fatigue rather
than against the possibility of yielding. This fatigue complication is handled by
choosing a metal alloy that has a fatigue limit (Figure M.26a) that is greater than
the cyclic stress amplitude to which the spring will be subjected. For this reason,
steel alloys, which have fatigue limits, are normally employed for valve springs.
When using steel alloys in spring design, two assumptions may be made if the
stress cycle is reversed (if tm  0, where tm is the mean stress, or, equivalently, if
tmax  tmin, in accordance with Equation M.38 and as noted in Figure M.44). The
first of these assumptions is that the fatigue limit of the alloy (expressed as stress
amplitude) is 45,000 psi (310 MPa), the threshold of which occurs at about 106 cycles.
Secondly, for torsion and on the basis of experimental data, it has been found that
the fatigue strength at 103 cycles is 0.67TS, where TS is the tensile strength of the
material (as measured from a pure tension test). The SN fatigue diagram (i.e., stress
amplitude versus logarithm of the number of cycles to failure) for these alloys is
shown in Figure M.45.
Now let us estimate the number of cycles to which a typical valve spring may
be subjected in order to determine whether it is permissible to operate within the
fatigue limit regime of Figure M.45 (i.e., if the number of cycles exceeds 106). For
the sake of argument, assume that the automobile in which the spring is mounted
travels a minimum of 100,000 miles (161,000 km) at an average speed of 40 mph (64.4

Figure M.44 Stress versus time


for a reversed cycle in shear.
max
Stress

min

Time
M.16 One Commonly Employed Steel Alloy 55

Figure M.45 Shear


0.67TS stress amplitude versus
logarithm of the number
of cycles to fatigue failure

Stress amplitude, S
for typical ferrous alloys.

45,000 psi

103 105 107 109


Cycles to failure, N
(logarithmic scale)

km/h), with an average engine speed of 3000 rpm (rev/min). The total time it takes
the automobile to travel this distance is 2500 h (100,000 mi/40 mph), or 150,000 min.
At 3000 rpm, the total number of revolutions is (3000 rev/min)(150,000 min)  4.5
108 rev, and since there are 2 rev/cycle, the total number of cycles is 2.25 108. This
result means that we may use the fatigue limit as the design stress inasmuch as the
limit cycle threshold has been exceeded for the 100,000-mile distance of travel (i.e.,
since 2.25 108 cycles 106 cycles).
Furthermore, this problem is complicated by the fact that the stress cycle is not
completely reversed (i.e., tm  0) inasmuch as between minimum and maximum de-
flections the spring remains in compression; thus, the 45,000-psi (310-MPa) fatigue
limit is not valid. What we would now like to do is to make an appropriate extrap-
olation of the fatigue limit for this tm  0 case and then compute and compare with
this limit the actual stress amplitude for the spring; if the stress amplitude is signif-
icantly below the extrapolated limit, then the spring design is satisfactory.
A reasonable extrapolation of the fatigue limit for this tm  0 situation may
be made using the Goodman equation (Equation M.49) modified to take into ac-
count the application of shear (rather than tensile) stresses. This modified expres-
sion takes the form

tfl  tfl0 a1  b
tm
(M.58)
0.67TS
where tfl is the fatigue limit for the mean stress tm; tfl0 is the fatigue limit for tm 
0 [i.e., 45,000 psi (310 MPa)]; and, again, TS is the tensile strength of the alloy]. To
determine the new fatigue limit tfl from this expression necessitates the computa-
tion of both the tensile strength of the alloy and the mean stress for the spring.

M.16 ONE COMMONLY EMPLOYED STEEL ALLOY


One common spring alloy is an ASTM 232 chromevanadium steel, having a composi-
tion of 0.480.53 wt% C, 0.801.10 wt% Cr, a minimum of 0.15 wt% V, and the balance
being Fe. Spring wire is normally cold drawn [Section 11.4 of Introduction (Section 14.2
of Fundamentals)] to the desired diameter; consequently, tensile strength will increase
For an ASTM 232 with the amount of drawing (i.e., with decreasing diameter). For this alloy it has been
steel wire, experimentally verified that, for the diameter d in inches, the tensile strength is
dependence of
tensile strength on TS 1psi2  169,0001d2 0.167 (M.59)
drawn wire diameter
56 Online Support Module: Mechanical Engineering

Since d  0.170 in. for this spring,


TS 1psi2  169,00010.170 in.2 0.167
 227,200 psi 11579 MPa2
Computation of the mean stress tm is made using Equation M.38 modified to
the shear stress situation as follows:
tmin  tmax
tm  (M.60)
2
It now becomes necessary to determine the minimum and maximum shear stresses
for the spring, using Equation M.56. The value of tmin may be calculated from Equa-
tions M.56 and M.52 inasmuch as the minimum dc is known (i.e., dic  0.060 in.). A
shear modulus of 11.5 106 psi (79 GPa) will be assumed for the steel; this is the room-
temperature value, which is also valid at the 80C service temperature. Thus, tmin is just
dicGd
tmin  Kw (M.61a)
pD2
D 0.140
c 1.60a b d
dicGd

pD2 d
10.060 in.2111.5 106 psi210.170 in.2 1.062 in. 0.140
 c d c 1.60a b d
p11.062 in.2 2 0.170 in.
 41,000 psi 1280 MPa2
Now tmax may be determined taking dc  dmc  0.135 in. as follows:
D 0.140
c 1.60 a b d
dmcGd
tmax  (M.61b)
pD2 d
10.135 in.2111.5 106 psi210.170 in.2 1.062 in. 0.140
 c d c 1.60a b d
p11.062 in.2 2 0.170 in.
 92,200 psi 1635 MPa2
Now, from Equation M.60,
tmin  tmax
tm 
2
41,000 psi  92,200 psi
  66,600 psi 1460 MPa2
2
The variation of shear stress with time for this valve spring is noted in Figure M.46;
the time axis is not scaled, inasmuch as the time scale will depend on engine speed.
Our next objective is to determine the fatigue limit 1tfl 2 for this tm  66,600 psi
(460 MPa) using Equation M.58 and for tfl0 and TS values of 45,000 psi (310 MPa)
and 227,200 psi (1570 MPa), respectively. Thus,

tfl  tfl0 a1  b
tm
0.67TS

 145,000 psi2 c 1  d
66,600 psi
10.6721227,200 psi2
 25,300 psi 1175 MPa2
M.16 One Commonly Employed Steel Alloy 57

100 Figure M.46 Shear stress


versus time for an automobile
valve spring.
80

Stress (103 psi)


60 aa = 25,600 psi

max = 92,200 psi


40
m = 66,600 psi

20 min = 41,000 psi

0
Time

Now let us determine the actual stress amplitude taa for the valve spring using
Equation M.40 modified to the shear stress condition:
tmax  tmin
taa 
2
92,200 psi  41,000 psi
  25,600 psi 1177 MPa2
2
Thus, the actual stress amplitude is slightly greater than the fatigue limit [i.e.,
taa 125,600 psi2 7 tfl 125,300 psi2 ], which means that this spring design is marginal.
The fatigue limit of this alloy may be increased to greater than 25,300 psi (175
MPa) by shot peening, a procedure described in Section M.12. Shot peening in-
volves the introduction of residual compressive surface stresses by plastically de-
forming outer surface regions; small and very hard particles are projected onto the
surface at high velocities. This is an automated procedure commonly used to im-
prove the fatigue resistance of valve springs; in fact, the spring shown in Figure M.43
has been shot peened, which accounts for its rough surface texture. Shot peening
has been observed to increase the fatigue limit of steel alloys in excess of 50% and,
in addition, to reduce significantly the degree of scatter of fatigue data.
This spring design, including shot peening, may be satisfactory; however, its ad-
equacy should be verified by experimental testing. The testing procedure is rela-
tively complicated and, consequently, will not be discussed in detail. In essence, it
involves performing a relatively large number of fatigue tests (on the order of 1000)
on this shot-peened ASTM 232 steel, in shear, using a mean stress of 66,600 psi
(460 MPa) and a stress amplitude of 25,600 psi (177 MPa), and for 106 cycles. On
the basis of the number of failures, an estimate of the survival probability can
be made. For the sake of argument, let us assume that this probability turns out to
be 0.99999; this means that one spring in 100,000 produced will fail.
Suppose that you are employed by one of the large automobile companies that
manufactures on the order of 1 million cars per year, and that the engine powering
each automobile is a six-cylinder one. Since for each cylinder there are two valves,
and thus two valve springs, a total of 12 million springs would be produced every
year. For the preceding survival probability rate, the total number of spring failures
would be approximately 120, which also corresponds to 120 engine failures. As a
practical matter, one would have to weigh the cost of replacing these 120 engines
against the cost of a spring redesign.
58 Online Support Module: Mechanical Engineering

Redesign options would involve taking measures to reduce the shear stresses
on the spring, by altering the parameters in Equations M.52 and M.56. This would
include either (1) increasing the coil diameter D, which would also necessitate
increasing the wire diameter d, or (2) increasing the number of coils Nc.

I n ve s t i ga t i o n o f E n g i n e e r i n g Fa i l u r e s
The entirety of Chapter 8 of Introduction (Chapter 9 of Fundamentals) was devoted
to discussions of the various forms of failure that materials experience, of failure
mechanisms, and, in some instances, of measures that may be taken to prevent, or
at least, minimize the possibility of failure. However, once an unexpected failure
has occurred an investigation may be conducted in order to determine the causes
or factors that led to the failure, and to recommend courses of action that, if taken,
will prevent or at least reduce the likelihood of future events. In some instances the
primary purpose of organizing a failure investigation is to assign legal responsibil-
ity for the consequences of the failure incidentwho is to be held accountable: the
company/individual that manufactured the failed component, or the company/indi-
vidual that was operating the component when it failed? Thus, the term forensic
engineering is sometimes used in the context of failure investigations and analyses.
Inasmuch as some engineers will be expected to conduct failure investigations,
we have included this submodule as a guide for planning and conducting effective
and organized investigations. The discussion that follows addresses the following
topics: causes and kinds of failure, planning a failure investigation, types of failure
mechanisms, procedures that may be used to ascertain root causes, and how to de-
termine corrective actions. More detailed treatments of the whys and hows of fail-
ure investigations are contained in the reference list at the end of this module.

M.17 REASONS FOR FAILURE


At the outset of such an investigation, one of the first issues to be addressed is why
the failure occurred. And as we shall see below, such a failure analysis is just one
aspect of the overall failure investigation. There are many possible reasons for en-
gineering failures, and one way of classifying the various types is as follows:
Design errors
Fabrication/manufacturing defects
Assembly errors
Misuse during operation
Improper maintenance

Design Errors
Several aspects of design determine a products overall reliability. The shape, size, and
configuration of a component are important so that it will (1) perform the function
intended, (2) withstand any applied loads without deforming excessively or fractur-
ing, and (3) not fail as a result of unanticipated stress levels that result from the pres-
ence of stress raiserssharp corners, configurational discontinuities, etc. Selecting ma-
terials that have an appropriate combination of properties (mechanical, electrical,
etc.) is also an important aspect of design; this also includes the specification of any
treatments to which the materials are to be subjected (e.g., heat treatments, cold work-
ing, etc.). Designation of manufacturing and assembly procedures is also part of the
design process, which also have an influence on the lifetime of a product.
M.18 Root Causes 59

Fabrication/Manufacturing Defects
There are many possible types of fabrication/manufacturing defects, which
normally are relatively easy to identify as causes of failure. Virtually all of the
fabrication techniques discussed in Chapters 11, 13, and 15 of Introduction (Chap-
ter 14 of Fundamentals) are prone to the introduction of defects. Some of the
more common fabrication/manufacturing defects include welding defects (poros-
ity, lack of penetration), improper heat treating, machining/grinding defects
(gouges, burns, tears, scratches, cracks), decarburization, and casting defects
(porosity, shrinkage cavities).

Assembly Errors
During a manufacturing process, the various components must be assembled to-
gether to form the desired product. In todays world, in order to be economically
competitive, industries have to devise faster and cheaper assembly processes. This,
coupled with increasingly more complicated products, leads to a greater likelihood
that components wont be assembled correctly. Furthermore, automated inspection
techniques often do not detect misassembled products. And, of course, a misas-
sembled product has a greater probability of failing prematurely than one that was
assembled correctly.

Misuse During Operation


Most products and machines are designed to have a reasonable lifetime expectancy;
this life expectancy is often expressed in terms of a warranty by the manufacturer.
A failure occurs when the component/machine wears out sooner than expected.
Many times this type of failure results when the component/machine is operated im-
properly or is abused during servicei.e., when operating procedures recommended
by the manufacturer are not observed. For example, the radiator in an automobile
may fail if the appropriate water level is not maintained. This type of failure is one
of the most common, and should be one of the first suspects in the investigation.

Improper Maintenance
In order to function properly, many products require periodic maintenancefor ex-
ample, automobiles (engine oil changes, tire rotation), lawn mowers (lubrication),
aircraft (inspection/replacement of high-stress structures), computers (virus checks),
etc. Improper maintenance can result in a premature failure of a component, struc-
ture, or machine, and may be intentional or unintentional. Corrosion failures often
result from maintenance neglect.

M.18 ROOT CAUSES


root cause We sometimes refer to the actual and true cause of failure as the root cause. This
root cause will most likely be related to one or more of the reasons for failure dis-
cussed in the previous section (e.g., design errors, fabrication/manufacturing defects,
etc.). Furthermore, there are really three levels or classifications of root causes
viz. physical, human, and latentwhich are described as follows:
PhysicalFor physical, the primary cause for the failure of a component/struc-
ture is related to one of the failure types or mechanisms discussed previously
viz. fracture due to overload, fatigue, creep, etc.
HumanFor human, a physical cause may be of secondary importance, in that
the actions of an individual led up to the failurefor example, a poorly writ-
ten set of instructions on how to use or properly maintain a product.
60 Online Support Module: Mechanical Engineering

Figure M.47 Root causes of failure


Interrelationships
among the root
causes of failure and
Physical Human Latent
reasons for failure.

Reasons for failure

Design Fabrication/ Assembly Misuse Improper


errors manufacturing errors during maintenance
defects operation

LatentA latent root cause relates to failures resulting from organizational


policye.g., company cost-reduction measures with the elimination of critical
testing procedures.

The relationships among reasons for failure and root causes are presented in Fig-
ure M.47.
Unfortunately, some failure investigations never discern the actual root cause.
Whereas the real root cause may be human or latent, the investigation stops at the
physical cause level. It is essential that the failure investigation be conducted so as
to include the possibility of involvment of human and latent factors.
Another complicating issue is that a series of events may lead up to the even-
tual failure. For example, a failure-producing crack is initiated by stress corrosion;
this crack then propagates in response to cyclic stresses (i.e., it becomes a fatigue
crack); and final failure results from a mechanical overload condition (when this
crack reaches some critical length). Thus, three physical causes are involved in this
failure. Determination of the real root cause (i.e., the crack induced by stress cor-
rosion) becomes a complex problem for the failure investigator.

T h e Fa i l u r e A n a l ys i s
A failure investigation is essentially an exercise in problem solving, which can be
broken down into finding the answers to the following four questions:
1. What exactly is the failure problem?
2. What is the root cause of the failure problem?
3. What are possible solutions?
4. Which of these is the best solution?
These four steps are also commonly used by engineers to solve most general engi-
neering problems.
A schematic diagram that outlines those procedures used to answer the above
four questions is shown in Figure M.48. The sections that follow discuss the imple-
mentation of these protocols.

M.19 WHAT EXACTLY IS THE FAILURE PROBLEM?


The first question that should be asked in any failure investigation is: What event
precipitated the malfunction of a component, machine, or process? The answer,
in essence, defines the purpose of the investigation. It will help also to determine
M.20 What Is the Root Cause of the Failure Problem? 61

Figure M.48 What is the failure What is the


Schematic outline of problem? root cause?
procedures used to
answer the four
Identification of
failure investigation possible root causes
questions.

Implementation of appropriate
analytical procedures
(NDT, microscopic examination,
mechanical testing, etc.)

Construction of
fault tree

Prove or disapprove
each possible root cause

Identification of What are possible


root cause solutions?

Construction of Which solution


corrective action tree is best?

what kind(s) of expertise is (are) needed, as well as the time and resources
required.

M.20 WHAT IS THE ROOT CAUSE


OF THE FAILURE PROBLEM?
Ascertaining the root cause of the failure is one of the primary goals of a failure
investigation. It is at this point that planning and organization of the investigation
take place. This includes the formation of an investigation team, which will be com-
posed of technical experts who have appropriate expertise and experience.
As noted in Figure M.48 (the column below What is the root cause?), it is
first necessary to identify all possible root causes.This is accomplished by conducting
a variety of analytical procedures so as to gain the best possible understanding of
the failure. One tool that may be implemented to help organize the investigation
and discern the actual root cause from the list of possibilities is a fault tree.

Failure Analysis Procedures


A well organized failure analysis will involve a number of procedures; some of the
common ones are provided in the following list. The sequence followed in an ac-
tual analysis need not be as given; furthermore, not all procedures are included in
every investigation.
1. Collection of background data and selection of samples for examination
2. Preliminary visual examination of the failed part
3. Nondestructive testing
4. Mechanical testing (e.g., tensile, hardness, impact)
62 Online Support Module: Mechanical Engineering

5. Selection, identification, preservation, and/or cleaning of critical specimens


6. Macroscopic examination and analysis of fracture surfaces, secondary cracks,
and other important surface features
7. Microscopic examination and analysis of fracture surfaces
8. Selection, preparation, examination, and analysis of metallographic sections
9. Determination of the actual stress state of the failed component
10. Determination of the failure mode
11. Chemical analyses (bulk, local, surface corrosion products, and deposits or
coatings)
12. Application of fracture mechanics
We now present some discussion for each of these procedures.

Collection of Background Data and Samples


Background data should include, when available, information pertaining to the orig-
inal design (including all underlying assumptions), manufacture, processing, fabrica-
tion, and service history of the failed component. Details regarding abnormal and
unusual conditions such as loading excursions, variations in temperature, the pres-
ence of a corrosive environment, and any accidental events are part of the service
record. Photographs of the failed component and its surrounding environment are
also essential background information. It may be necessary to select samples for both
macroscopic and microscopic examinations.These specimens should be carefully cho-
sen so as to include not only the region that encompasses the failure, but also other
locations both adjacent to and far removed from the failure site. Care should be ex-
ercised so as to preserve any debris or oxide materials that are present.

Preliminary Visual Examination


The next step is to perform an examination, using the unaided eye, of the part that
failed as well as all of its broken fragments. Of particular interest are the features
of and changes in texture across the fracture surface, any evidence of corrosion, sur-
face marks, and angle of fracture. Details of this examination should be documented
both in writing and with photographs. When taking photographs, direction of light
illumination may be important so as to reveal critical surface characteristics. Ex-
amination of fine features of the failure surface may be necessary using a magni-
fying glass.

Nondestructive Inspection
Some of the nondestructive testing techniques discussed in Section M.5 (Table M.6)
for detecting flaws in structural components may also be utilized in failure
analysesto detect small surface cracks and discontinuities in failed parts. Dye-
penetrant, ultrasonic, and radiographic are those most commonly used.

Mechanical Testing
Mechanical tests on failed parts are conducted for several reasons: to determine if
the material conforms to specifications; to determine the heat treatment; to detect
any alteration of mechanical properties due to cold working or overheating; and to
detect decarburization or any increase in carbon and/or nitrogen concentration.
Hardness tests are easiest to conduct, but tensile and impact tests are also possible
provided that adequate material is available for the fabrication of test specimens.
M.20 What Is the Root Cause of the Failure Problem? 63

Specimen Preservation and Selection


This stage is important in order that evidence critical to the investigation is not de-
stroyed, obscured, or altered. Fracture surfaces may be susceptible to damage from
mechanical forces or some chemical environments, and, therefore, should be pro-
tected during the investigation. The investigator should not try to fit back together
broken sections, and touching and rubbing fracture surfaces should be avoided. The
best way to prevent chemical damage is to place the fracture specimen in a desic-
cator, or pack it with a desiccant material (one that removes water vapor from the
air). In some cases it may be necessary to dry the specimen, which may be accom-
plished using a jet of dry air (which will of course blow away any surface foreign
residue that may be important to the investigation).
In order to perform some tests and examinations (e.g., hardness, electron mi-
crographic, photomicrographic), it may be necessary to remove a portion of the
fracture specimen of convenient size. This is normally done using a cutting or sec-
tioning procedure. Measures to protect the area of fracture are necessary, and the
location of any cutting action should be chosen such that the fracture region itself
as well as adjacent areas are not damaged or altered. The cutting action attendant
to sectioning will necessarily heat neighboring regions with possible alteration of
microstructure and properties; it is essential that microstructural elements and prop-
erties of critical areas be preserved.

Macroscopic Examination
Macroscopic examinations are conducted with the unaided eye, and/or using a hand-
held magnifying glass, a low-power stereoscopic microscope, and/or a scanning elec-
tron microscope (SEM) (at low magnifications). In general magnifications range
between 1 and 50. Reasons for conducting this type of examination include: to locate
the crack origin, to determine its shape and size as well as the path of crack propaga-
tion, to characterize the texture of the fracture surface, and to note possible points of
stress concentration (e.g., drilled holes, hammer marks, accidental dents, etc.) as well
as any other gross features that may provide clues as to the mode of failure. In addi-
tion, an attempt should be made to determine if there is more than one crack origin.

Surface Topography. For failures that result from overload conditions, the
topography of the fracture surface depends on whether the material was ductile or
brittle, as well as the manner of loading (i.e., tensile, shear, torsional, bending, or
combinations of these loading modes). Figures M.49a and M.49b show schematic

Tension Torsion

Ductile Brittle Ductile Brittle


(a) (b) (c) (d)

Figure M.49 Characteristic fracture surface contours for (a) ductile and (b) brittle
materials that are stressed in uniaxial tension, and (c) ductile and (d) brittle materials that
are stressed in torsion. (Adapted from D. J. Wulpi, Understanding How Components Fail,
ASM International, 1985, p. 30. Reprinted with permission of ASM International. All
rights reserved. www.asminternational.org.)
64 Online Support Module: Mechanical Engineering

(a) (b)

Figure M.50 (a) Cup-and-cone fracture in aluminum. (b) Brittle fracture in a mild steel.

representations of fracture surfaces for cylindrical specimens of both ductile and


brittle materials that failed from overloading in uniaxial tension.The fracture surface
for the ductile material (normally a metal) has the typical cup-and-cone configuration
(per the photograph of Figure M.50a)i.e., central regions of both mating pieces
are relatively flat, oriented perpendicular to the stress direction, and have a rough
and fibrous texture, whereas the plane of the outer-periphery shear lips makes a
45 angle with the stress direction. (The mechanism of crack formation and
propagation for this situation is represented in Figure M.51.) By way of contrast,
for the brittle material (Figure M.49b), once formed, a crack propagates within a
plane that is oriented perpendicular to the stress axis, and yields a flat failure surface.
A photograph of a specimen that failed in this manner is shown in Figure M.50b.
Consider now the situation in which the overload stress is torsional in nature.
For cylindrical specimens of ductile and brittle materials, schematic failure profiles
are shown, respectively, in Figures M.49c and M.49d. For the ductile material, the
fracture surface is flat and oriented parallel to the direction of the applied torsional
stress. And a helical fracture surface results when the material is brittle.

Figure M.51 Stages in the cup-and-cone


fracture. (a) Initial necking. (b) Small
cavity formation. (c) Coalescence of
cavities to form a crack. (d) Crack
propagation. (e) Final shear fracture at a
45 angle relative to the tensile direction.
(From K. M. Ralls, T. H. Courtney, and J.
Wulff, Introduction to Materials Science
and Engineering, p. 468. Copyright 1976
(a) (b) (c) by John Wiley & Sons, New York.
Reprinted by permission of John Wiley &
Sons, Inc.)

Shear
Fibrous

(d) (e)
M.20 What Is the Root Cause of the Failure Problem? 65

Region of slow Figure M.52 Fatigue


crack propagation
failure surface. A crack
formed at the top edge.
The smooth region also
2 cm
near the top corresponds
to the area over which
the crack propagated
slowly. Rapid failure
occurred over the area
having a dull and fibrous
texture (the largest area).
Approximately 0.5 .
(Reproduced by
permission from Metals
Handbook: Fractography
and Atlas of
Fractographs, Vol. 9, 8th
edition, H. E. Boyer,
Editor, American Society
for Metals, 1974.)

Region of rapid failure

Failures resulting from other mechanisms may have yet other surface config-
urations. For example, Figure M.52 is a photograph of the surface of a shaft that
failed by fatigue. Important features shown here include the crack origin [on the
outer surface (near the top edge)], the region of slow crack propagation during cy-
cling (that appears light and has a smooth texture), and the area of rapid failure
(having a dull and fibrous texturecorresponding to the region of largest cross-
sectional area).

At this time, it is appropriate to make a distinction between brittle materials


and brittle fractures. A brittle fracture is one in which there is little or no gross plas-
tic deformation on a macroscale. Of course, when brittle materials are overloaded,
they fracture in a brittle manner. On the other hand, under some circumstances,
there may be very little evidence of any macroscale deformation on the failure
surface of a ductile metali.e., the mode of fracture is brittle. For example, for a
failure mechanism in which cracks form and then propagate relatively slowly (i.e.,
as with fatigue or stress-corrosion cracks), crack growth proceeds until the intact
cross-sectional area of the part reaches a state of overloading, at which time rapid
crack propagation and sudden failure occur. This type of failure may be observed
in the photograph of Figure M.52the rotating steel shaft that experienced a fa-
tigue failure. In this case, a condition of overload was achieved once the crack had
propagated through that cross-sectional region that appears light in the photograph.
Furthermore, the area of rapid failure (the largest cross-sectional area that appears
dark) has a dull and fibrous texture (and no evidence of plastic deformation). These
features are characteristic of a brittle fracture, in spite of the fact that this steel is
a ductile material.
66 Online Support Module: Mechanical Engineering

(a)

(b)
10 mm

Figure M.53 (a) Photograph showing V-shaped chevron markings characteristic of


brittle fracture. Arrows indicate origin of crack. Approximately actual size. (b) Photograph
of a brittle fracture surface showing radial fan-shaped ridges. Arrow indicates origin of
crack. Approximately 2. [(a) From R. W. Hertzberg, Deformation and Fracture Mechanics
of Engineering Materials, 3rd edition. Copyright 1989 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc. Photograph courtesy of Roger Slutter,
Lehigh University. (b) Reproduced with permission from D. J. Wulpi, Understanding How
Components Fail, American Society for Metals, Materials Park, OH, 1985.]

Surface Features. In addition to failure surface topography, surface features


that are present may also provide valuable information about the failure mode. For
example, Figure M.53a is a photograph (unmagnified) that shows matching cross
sections of a structure that failed in a brittle manner. The fracture surfaces are rel-
atively flat (indicative of a brittle fracture), and V-shaped chevron markings may
be observed that point back toward the crack origin.
Another type of brittle fracture surface is presented in the photograph of Figure
M.53b; here fan-shaped ridges may be observed that radiate from the crack origin.
The photograph of Figure M.54 shows the fracture surface of a rotating steel
shaft that experienced fatigue failure. In addition to the point of crack origin
M.20 What Is the Root Cause of the Failure Problem? 67

Figure M.54 Fracture


surface of a rotating steel
shaft that experienced
fatigue failure. Beachmark
ridges are visible in the
photograph. (Reproduced
with permission from D. J.
Wulpi, Understanding
How Components Fail,
American Society for
Metals, Materials Park,
OH, 1985.)

(at the corner of a keyway on the shaft) and site of final rupture, beachmark ridges
may be observed; beachmarks, features found on some fatigue failure surfaces, are
discussed briefly in Section M.10.

In this discussion, we have treated some of the more common types of macro-
scopic configurations and features found on surfaces of failures; of course, others are
also possible. It should be noted that accurate interpretation of them is a skill acquired
only through experience; and space does not allow a more thorough treatment here.

Microscopic Examination
Microscopic (or fractographic) examinations are conducted at higher magnifications
than macroscopic ones; normally a scanning electron microscope is used. Magnifi-
cations as high as 200,000 times are possible, as are also large depths of field. Depth
of field is important in order to adequately observe topographical features of the
failure surface at these magnifications. Some SEMs are equipped with energy-
dispersive x-ray spectroscopes, which permit semiquantative and quantitative chem-
ical analyses of selected areas. This capability is useful in ascertaining the chemistry
of microstructural features. The most significant limitation of SEM analysis is spec-
imen size; in order to fit within the examination chamber, a specimen must have a
diameter less than about 200 mm (8 in.)consequently, it is necessary to section
pieces that are larger than this.
A microscopic examination may also provide valuable evidence regarding the
mechanism of failure. For example, an SEM micrograph for a ductile metal that
failed due to overloading in tension will appear as that shown in Figure M.55a; that
is, spherical dimples will be present. Whereas, for a shear overloading failure (also
of a ductile metal), dimples will have a parabolic shape (Figure M.55b).
As noted in Section 8.4 of Introduction (Section 9.4 of Fundamentals), brittle
failures of metals may be transgranular (crack propagation is through interiors of
68 Online Support Module: Mechanical Engineering

(a) 5 m (b) 4 m

Figure M.55 (a) Scanning electron fractograph showing spherical dimples characteristic
of ductile fracture resulting from uniaxial tensile loads. 3300. (b) Scanning electron
fractograph showing parabolic-shaped dimples characteristic of ductile fracture resulting
from shear loading. 5000. (From R. W. Hertzberg, Deformation and Fracture Mechanics
of Engineering Materials, 3rd edition. Copyright 1989 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc.)

grains) or intergranular (crack propagation is along grain boundaries). For trans-


granular, an SEM micrograph will reveal cleavage facets, Figure M.56a, whereas
a grainy or faceted texture (characteristic of the three-dimensional nature of
the grains) will exist when the failure is intergranular (Figure M.56b). Intergranu-
lar fracture is caused by some process that weakens or embrittles grain boundary

(b)
200 m

Figure M.56 (a) Scanning electron fractograph of ductile cast iron showing a transgranular
fracture surface. Magnification unknown. (b) Scanning electron fractograph showing an
intergranular fracture surface. 50. [(a) From V. J. Colangelo and F. A. Heiser, Analysis of
Metallurgical Failures, 2nd edition. Copyright 1987 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc. (b) Reproduced with permission from
ASM Handbook, Vol. 12, Fractography, ASM International, Materials Park, OH, 1987.]
M.20 What Is the Root Cause of the Failure Problem? 69

Figure M.57 Scanning electron


micrograph of a mixed-mode
fracture surface, which is
composed of both cleavage and
dimpled regions. 570.
(Reproduced with permission
from Handbook of Case Studies
in Failure Analysis, Vol. 1 (1992),
ASM International, Materials
Park, OH, 44073-0002.)

regionse.g., segregation of embrittling species (viz., hydrogen, liquid metals) along


grain boundaries, intergranular stress-corrosion cracking, etc.
On occasion, more than one fracture mode is involved in a failure process. For
example, a mixed-mode fracture is shown in the SEM fractograph of Figure M.57.
Here may be observed both circular dimples (characteristic of ductile fracture, Fig-
ure M.55a), as well as cleavage facets (characteristic of transgranular brittle frac-
ture, Figure M.56a).
For some (but not all) failures resulting from fatigue, electron fractographs will
reveal the presence of striations (Figure M.58)closely spaced and parallel lines

Figure M.58 Transmission


electron fractograph
showing fatigue striations in
aluminum. 9000. (From
V. J. Colangelo and F. A.
Heiser, Analysis of
Metallurgical Failures, 2nd
edition. Copyright 1987
by John Wiley & Sons, New
York. Reprinted by
permission of John Wiley &
Sons, Inc.)

1 m
70 Online Support Module: Mechanical Engineering

Mist region Hackle region

Origin Mirror region

Figure M.59 Photomicrograph of the fracture surface of a 6-mm-diameter fused silica rod
that was fractured in four-point bending. Features typical of this kind of fracture are
notedi.e., the origin, as well as mirror, mist, and hackle regions. 500. (Courtesy of
George Quinn, National Institute of Standards and Technology, Gaithersburg, MD.)

or curves. These striations are oriented perpendicular to the direction of crack prop-
agation, and each striation represents the advance distance of the crack front dur-
ing a single loading cycle.
Fractographs of brittle ceramic materials have their own distinctive features.
Figure M.59, an optical photomicrograph that shows the fracture surface of a fused
silica rod that was fractured in four-point bending, represents one possibility. As
noted, a flat, smooth, and highly reflective mirror region is present; it corresponds
to the fracture surface that formed during the initial stage of crack propagation.
The outer perimeter of this region is circular, with the crack origin located near its
centerin this case on the surface of the rod. The measured radius of this circle
may be used to approximate the stress level that caused fracture; fracture stress
level is inversely proportional to the square root of radius (Equation 12.6 of Intro-
duction, Equation 9.14 of Fundamentals). Also shown in Figure M.59 are the mist
and hackle, annular areas that reside outside the mirror region. The advancing crack
began to branch and change propagation direction within the mist and hackle
regions, which gave rise to rougher surface textures.
M.20 What Is the Root Cause of the Failure Problem? 71

Metallographic Examination
For metallographic examination, a specimen is first removed from the failed piece
by sectioning; this specimen is then polished and etched so as to reveal details of
its microstructuree.g., grains, the various phases present, etc. Microstructural ob-
servations are normally conducted using an optical microscope.
A wealth of microstructural data regarding probable causes of failure may be
gleaned from a metallographic examination. Types of information that are available
include the following:
Grain size and shape provide some indication as to thermal and mechnical
history. A coarse grain structure (i.e., large grains) indicates that the material
was likely subjected to an annealing heat treatment or perhaps overheating
during service (with grain growth). Whereas an elongated grain structure
means that the specimen was deformed by some mechanical process (e.g.,
forging, rolling, drawing). Also, deformation direction may be deduced from
the grain-elongation direction.
Identification of the microconstituents present is helpful in determining
whether the material was propertly heat treated, as well as ascertaining other
aspects of its thermal historye.g., was the microstructure produced inten-
tionally during manufacturing, or did it result inadvertently while in service?
The presence of a grain boundary phase may explain intergranular brittle
fracture. Also, determination of heat treatment deficiencies, such as surface
decarburization is possible.
Internal imperfections such as pores, inclusions, and welding defects, that may
have played a role in the failure process, may also be revealed.
Other effects resulting from service conditions may be investigated. These in-
clude the occurrence of oxidation, corrosion, and severe surface strain hardening.
And, determination of the mode of crack propagation (viz. transgranular or
intergranular) is possible when both crack and grain structure are shown in a
photomicrograph.
In some investigations it is imperative to ascertain whether or not the microstruc-
ture of the failure region is indeed representative of the microstructure of the com-
ponent in its as-manufactured condition (i.e., did the service conditions alter the
microstructure, and, if so, what was the nature of the alteration?). This determina-
tion is possible by comparing photomicrographs of specimens taken both from a
region adjacent to the failure and from a remote area.

Stress Analyses
When an excessively high load level (i.e., a condition of overload) is suspected as
the direct cause of failure, a stress analysis is warranted so as to ascertain if it is in-
deed the root cause. Determinations of stress magnitude as well as type (tension,
torsion, bending, static, fluctuating) are the goals of such an analysis. The failure
analyst should also endeavor to obtain records from the manufacturer relating to
allowable stress levels that were predicted during the original design of the part.
Experimental verification of these data is possible by taking strain gauge measure-
ments on other identical (or similar) components during exposure to in-service
stresses. The above information is useful in determining whether the failed part was
sized properly, if properties of the materials called for in the design of the part met
specifications, or if failure was a result of overloading.
72 Online Support Module: Mechanical Engineering

Determination of Failure Mode


Of course, one critical element of an investigation involves determining the mecha-
nism(s) that was (were) responsible for the failure. There exist a large number of pos-
sible failure modes or mechanisms; some (but certainly not all) are listed as follows:
Ductile fracture
Brittle fracture
Fatigue (high-cycle, low-cycle, corrosion, thermal)
Corrosion (uniform, galvanic, pitting, crevice, etc.)
Stress-corrosion cracking
Distortion (elastic and plastic)
Creep and creep rupture
Liquid-metal embrittlement
Hydrogen embrittlement
Radiation damage
Preceding treatments (viz. macroscopic, microscopic, and metallographic examina-
tions) as well as sections in the printed text discuss characteristics of some of these
failure modes that may be used for making a reliable identification. References at the
end of this module provide additional instruction on the determination of failure mode.
Chemical Analyses
A chemical analysis of the failed material may also be necessary in the investiga-
tion. Normally, this is one of the last procedures conducted inasmuch as it requires
the destruction of some of the failure specimen. One reason for doing a chemical
(or compositional) analysis is to certify that the failed part was fabricated from the
correct material; in most instances, slight deviations from the specified composition
are not critical. Chemical analyses may also be performed on corrosion products
and other deposits associated with the failure. The presence of gaseous elements
(e.g., hydrogen, oxygen, nitrogen) may have deleterious effects on the mechanical
properties of some metal alloys; detection of these elements in concentrations above
acceptable limits is desirable.
Techniques utilized for chemical analyses include the following: wet chemistry
routes, emission spectrography, spectrophotometry, atomic-absorption spectroscopy,
x-ray diffraction, x-ray fluorescence spectroscopy, infrared and ultraviolet spec-
troscopy,Auger electron spectroscopy, energy and wavelength-dispersive x-ray spec-
troscopy (with SEMs), and electron microprobe analysis.
Application of Fracture Mechanics
Principles of fracture mechanics (Section M.4) are used to quantify the relationship
between the fracture toughness of a material, applied stress level, and the size of
crack-producing flaws (Equation M.20). Utilization of these principles allows the
evaluation of structural reliability and prediction of service lifetimes when there is
the possibility of ductile or brittle fracture, as well as failure by fatigue.

Identification of Possible Root Causes


Once an investigator has acquired a clear understanding of the failure (using the
techniques described in the preceding paragraphs), the next step is to clearly and
objectively identify and analyze all possible root causes. One organizational tool
that is especially useful to accomplish this goal is a fault tree. A fault tree is simply
M.20 What Is the Root Cause of the Failure Problem? 73

Why Do the Forgings Have


Penetrant Defects?

Casting Process Forging Process Heat


Defects Defects Treatment
(1) (2) Process
Defects
(3)

Shrinkage Hot Defect Cavity


Nonmetallic Burst
Porosity Short Enhancement
(1A) (2A) Quench
(1C) (2B) (2C)
Cracking
Hydrogen (3A)
Porosity
(1B)
Adiabatic Localized
Heating Chemistry
Hydrogen
(2B1) (2B2)
Content
(1B1)
Forging
Casting Temperatures
Temperature (2A1 and 2C1)
Chemical (1C1)
Composition Forging
(1A1) Ingot Drop Strain Rate
Speed (2A2 and 2C2)
(1C2)
Filtering
Processes Metal Flow Forging
(1A2) Rate Strain Direction
(1C3) (2A3 and 2C3)

Figure M.60 Fault tree diagram for a dye penetrant defect in a forging that failed.
(Adapted from D. P. Dennies, How to Organize and Run a Failure Investigation, ASM
International, 2005, p. 109. Reprinted with permission of ASM International. All rights
reserved. www.asminternational.org.)

a taxonomic chart that shows the interrelationships among a single major root cause
and all possible sources (or causes) that can lead up to (or are responsible for) this
root cause. For example, Figure M.60 is a fault tree for a dye penetrant defect
(a root cause) that was observed in a failed forging; shown here is a hierarchy of
possible causes for the root cause. There are three primary possible sources (or
causes) of this defect [in the boxes labeled (1) through (3)], and for each of these
defects there is a least one (secondary) cause, and in some instances also tertiary
and quaternary causes. To each cause (or box) is assigned a number (or numbers)
and in some cases a letter; this scheme helps to organize the failure investigation.
There are a number of approaches that may be used to design and construct
a comprehensive and appropriately organized fault tree; we suggest the following
procedure:
Define all possible root causes. Techniques that may be employed are dis-
cussed previously in this section.
Analyze each proposed root cause as to its possible causes; this will generate
a set of primary causes. Such analyses are possible only if all information
about the failed part has been gatherede.g., details of its design, the man-
ner in which it was fabricated, specifics of its operational history. On occa-
sion, experts in these areas may need to be consulted.
74 Online Support Module: Mechanical Engineering

Brainstorming sessions are also necessary when performing these analyses.


Team members should meet, question, and thoroughly discuss the reason(s)
why each primary cause could have occurred. This process will generate a set
of secondary causes, which are then entered on the fault tree. Continuation
of this procedure will produce sets of tertiary, quaternary, etc. causes, and,
hopefully, the ultimate discovery of all causing effects (and the generation of
a complete fault tree).
The fault tree is not to be treated as a static document. Rather, it should con-
tinue to change and improve as new information is uncovered throughout the
course of the investigation.
Once the construction of a fault tree has pretty much stabilized, two addi-
tional questions should still be addressed: (1) What is unique or different
about this failure? and (2) Are there any things that are still missing from
our investigation?
There are several distinct advantages for constructing and utilizing a fault tree in
failure investigations:
First and foremost, it is a problem-solving guide. A fault tree is essentially an
outline that may be used to generate a set of procedures and evaluations de-
signed and planned to either prove or disprove each possible root cause.
The fault tree organizes the complex problem of determining each root cause
of a failure into a series of simpler and more manageable components.
It is a form of documentation or a record of how the failure investigation
was organized and executed.

Identification of the Root Cause Responsible for Failure


Once the fault tree (or trees) has (have) been constructed, it next becomes neces-
sary to objectively evaluate the likelihood that each of the possible root causes was
responsible for the failure. The first step in this process is to assess the probability
of each of the root causes in the fault treeis it likely, possible, or unlikely? In this
regard, it is important to rely on the technical expertise of members of the investi-
gation team. The next step is prioritize the order in which each of the possible root
causes is to be investigated; in all likelihood, time, financial, and/or personnel re-
sources will not permit testing of them all. These investigations will undoubtedly
involve some of the analytical procedures discussed in the previous section (e.g.,
microscopic/metallographic examinations, NDT, mechanical testing, etc.). At this
time the rationale that was used to assign the probability and priority to each root
cause should be documented. It is sometimes convenient to summarize and record
the information formulated in these three steps in chart form.
It is now time to create a plan that will either prove or disprove each of the
possible root causesthat is, prepare a schedule of appropriate tests and analyses
based on the probability/prioritization scheme described in the preceding para-
graph. Then, for each root cause, it is necessary to make a list of what physical ev-
idence is required to substantiate that this root cause occurred or was present. In
preparing this list, one should keep in mind that it is sometimes easier to disprove
a root cause than to prove it. The next step is to decide which of the procedures
given under the Failure Analysis Procedures heading are to be used to provide
this physical evidence. Responsible personnel are then assigned to perform these
test/analysis procedures, and completion deadlines are set. And, finally, upon com-
pletion of the tests/analyses, results are tabulated; these results are the basis for the
M.22 Which of These Is the Best Solution? 75

final decision as to which of the possible root cause(s) is (are) most likely respon-
sible for the failure. This planning and testing phase of the investigation may take
weeks or even months to carry out.
The above discussion has dealt with the determination of the physical root
cause. As noted above, there is also the possibility that human or latent root causes
are responsible for this physical one; furthermore, it is important that the investi-
gation be designed to also ascertain whether or not either of these factors is the
real cause of failure.
At this point we have finally answered the second question that was presented
at the outset of our discussion on failure analysis: What is the root cause of the
failure problem?

M.21 WHAT ARE POSSIBLE SOLUTIONS?


Now, after the completion of this laborious and time-consuming root-cause-determi-
nation phase, the investigation is at the stage where possible corrective actions may
be ascertainedthat is, it is possible to answer the What are possible solutions? ques-
tion (Figure M.48). One approach is to identify corrective actions for the most likely
root cause(s) responsible for the failure. This is possible by creation of a corrective ac-
tion tree, which is formatted after the manner of the fault tree. For example, for the
forging penetrant defects fault tree, Figure M.60, the corresponding corrective action
tree might appear as in Figure M.61. Here it is assumed that the primary root cause
was a casting process defect (which is the reason that most of the other box lines in
the diagram are dashed). The proposed corrective actions listed in Figure M.61 would
come from brainstorming sessions involving members of the investigation team.
This action tree approach may also be used for suggesting preventative actions.
Whereas a corrective action is taken to ensure that the root cause of failure does
not happen again, the purpose of a preventative action is to reveal the presence of
a root cause so that failure does not occur. For example, in Figure M.61 there is a
fourth primary corrective action NDE Improvements (i.e., nondestructive evalu-
ation improvements), whereas there are only three primary root causes in the fault
tree of Figure M.60. Here, nondestructive evaluation improvements is really a pre-
ventative action.

M.22 WHICH OF THESE IS THE BEST SOLUTION?


Finally, the investigation has progressed to a point such that the last of our
four questions may be answeredWhich of the possible solutions is the best?
The investigation team should first objectively evaluate each of the corrective
actions in the corrective action tree of Figure M.61, with regard to its likelihood of
remedying the root cause, and, in addition, feasibility, cost effectiveness, and
implementation time. On the basis of this evaluation a decision is made as to the
optimal corrective action(s) that should be taken.
Of course, if the real root cause of failure was determined to be due to a hu-
man or latent factor (as opposed to a physical cause), the proposed solution should
include corrective actions appropriate to remedy the fundamental problem.
A set of recommendations should be carefully formulated for the purpose
of eliminating future failures. These recommendations might include some of the
following measures:
Design changes
Metallurgical alterations
76 Online Support Module: Mechanical Engineering

What Is the Corrective Action for the


Forgings with Penetrant Defects?

Casting Process Forging Process Heat Treatment NDE


Improvements Improvements Process Improvements
(1) (2) Improvements (4)
(3)

Nonmetallic Hydrogen Shrinkage Improve Add Improve


Inclusions Porosity Porosity Forging Double Forging Heat- Measure H2 in
Develop Ingot
Content Reduction Reduction Reduction Cross Up Process Prevent All Ingots and
Slice Exam
Reduction (1B) (1C) (2A) Work (2C) Hydrogen Forgings
(4A)
(1A) (2B) Pickup (4C)
(3A)
H2 Content Improved
Visual, Dye
Reduction Ultrasonic
Penetrant or
(1B1) Inspection of
Ultrasonic
Forgings
Casting Fumace Inspection
(4B)
Temperature Atmosphere (4A1)
LIMCA Forging
(1C1) (2C1 and 3A1)
Device Temperatures
(1A1) (2A1)
Ingot Drop Coat the Develop
Immersion
Speed Improve Forging Forgings Rating
Improve Tank
(1C2) Strain Rate (2C2 and 3A2) System
Filtering (4B1)
Processes (2A2) (4A2)
(1A2) Increase Metal Flow Add
Fluxing and Rate Improve Forging
Shear Wave
Settling (1C3) Strain Direction
Use New (4B2)
Time (2A3)
Filters
(1B1A)
(1A2A)

Figure M.61 Corrective action tree diagram for the fault tree diagram of Figure M.60.
(Adapted from D. P. Dennies, How to Organize and Run a Failure Investigation, ASM
International, 2005, p. 119. Reprinted with permission of ASM International. All rights
reserved. www.asminternational.org.)

Changes in manufacturing
Improvement of quality control
Improvement of repair procedures
Use of warning labels

M.23 EFFECTIVENESS EVALUATION


OF CORRECTIVE ACTIONS
Follow-up plans should be made for evaluating the effectiveness of the corrective
action(s) called for. This evaluation should come after a specified time period sub-
sequent to implementation as called for by the investigation team. Furthermore, it
may be necessary to conduct periodic evaluations.

M.24 THE FINAL REPORT


After completion of the failure investigation, it is imperative that a final report be
prepared.This should be done as soon as possible so that important details regarding
the investigation are not forgotten and, therefore, omitted. Furthermore, the cus-
tomer is entitled to and should expect to receive a report.
M.25 Introduction 77

The first section should be a summary of the investigation, its results, conclu-
sions, and recommendations. It has been suggested that the remainder of the report
be divided into the following sections:
1. Description of the failed component
2. Service conditions at the time of failure
3. Prior service history
4. Manufacturing and processing history of the component
5. Mechanical structural analysis of the failed part
6. Metallurgical study of the failure
7. Metallurgical evaluation of quality
8. Summary of mechanisms that caused failure
9. Recommendations for prevention of similar failures or for correction of simi-
lar components in service
10. Appendixto include any figures, tables, etc.
This final report should be concise, clearly written, and logically organized.
Furthermore, those parties on the investigation team who were responsible for
making recommendations should be provided the opportunity to review the report
for accuracy and any omissions.

We now demonstrate the use of some of these investigative techniques and


principles with the following case study on the failure of an automobile rear axle.

Fa i l u r e o f a n A u t o m o b i l e
Re a r A x l e ( C a s e St u d y ) 2 0
M.25 INTRODUCTION
After an accident in which a light pickup truck left the road and overturned, it was
noted that one of the rear axles had failed at a point near the wheel mounting
flange. This axle was made of a steel that contained approximately 0.3 wt% C. Fur-
thermore, the other axle was intact and did not experience fracture. An investiga-
tion was carried out to determine whether the axle failure caused the accident or
whether the failure occurred as a consequence of the accident.
Figure M.62 is a schematic diagram that shows the components of a rear axle as-
sembly of the type used in this pickup truck. The fracture occurred next to the bear-
ing lock nut, as noted in this schematic. A photograph of one end of the failed axle
shaft is presented in Figure M.63a, and Figure M.63b is an enlarged view
of the other fractured piece that includes the wheel mounting flange and the stub end
of the failed axle. Here (Figure M.63b), note that a keyway was present in the area of
failure; furthermore, threads for the lock nut were also situated next to this keyway.
Upon examination of the fracture surface it was noted that the region corre-
sponding to the outside shaft perimeter [being approximately 6.4 mm (0.25 in.)
wide] was very flat; furthermore, the center region was rough in appearance.

20
This case study was taken from Lawrence Kashar, Effect of Strain Rate on the Failure
Mode of a Rear Axle, Handbook of Case Histories in Failure Analysis, Vol. 1, pp. 7478,
ASM International, Materials Park, OH, 1992.
78 Online Support Module: Mechanical Engineering

Axle shaf

Point of failure

Figure M.62 Schematic diagram showing typical components of a light truck axle and the
fracture site for the failed axle of this case study. (Reproduced from MOTOR Auto Repair
Manual, 39th Edition Copyright 1975. By permission of the Hearst Corporation.)

Let us take a moment to apply some of the failure investigation principles discussed
in Sections M.19 through M.22specifically in addressing the questions that were posed.
The answer to the first question (What exactly is the failure problem?) is
a fractured rear axle of a pickup truck.
For the second question (What is the root cause of the failure problem?)
there are really only two possible answers: either (1) the accident caused the
axle failure, or (2) the axle failure caused the accident.
Inasmuch as the purpose of this investigation is to identify the root cause, it
is not necessary to find answers to the last two questions: What are possible
solutions? and Which of these is the best solution?

(a) (b)

Figure M.63 (a) Photograph of one section of the failed axle. (b) Photograph showing
wheel mounting flange and stub end of failed axle. [Reproduced with permission from
Handbook of Case Studies in Failure Analysis, Vol. 1 (1992), ASM International, Materials
Park, OH, 44073-0002.]
M.26 Testing Procedure and Results 79

Figure M.64 Optical


micrograph of failed
section of axle that
shows the keyway
(bottom), as well as the
flat outer-perimeter and
rough core regions.
[Reproduced with
permission from
Handbook of Case
Studies in Failure
Analysis, Vol. 1 (1992),
ASM International,
Materials Park, OH,
44073-0002.]

Let us now explore the analytical procedures that were utilized to determine
the root cause of the axle failure.

M.26 TESTING PROCEDURE AND RESULTS


Details of the fracture surface in the vicinity of the keyway are shown in the pho-
tograph of Figure M.64; note that the keyway appears at the bottom of the photo-
graph. Both the flat outer perimeter and rough interior regions may be observed
in the photograph. There are chevron patterns that emanate inward from the cor-
ners of and parallel to the sides of the keyway; these are barely discernible in the
photograph but indicate the direction of crack propagation.
Fractographic analyses were also conducted on the fracture surface. Figure
M.65 shows a scanning electron micrograph taken near one of the keyway corners.

Figure M.65 Scanning electron micrograph


of failed axle outer-perimeter region near
the keyway, which shows cleavage features.
3500. [Reproduced with permission from
Handbook of Case Studies in Failure
Analysis, Vol. 1 (1992), ASM International,
Materials Park, OH, 44073-0002.]

2 m
80 Online Support Module: Mechanical Engineering

Figure M.66 Scanning electron


micrograph of the failed axle
rough core region, which is
composed of mixed cleavage and
dimpled regions. 570.
[Reproduced with permission
from Handbook of Case Studies
in Failure Analysis, Vol. 1 (1992),
ASM International, Materials
Park, OH, 44073-0002.]

Cleavage features may be noted in this micrograph, whereas any evidence of dim-
ples and fatigue striations is absent. These results indicate that the mode of frac-
ture within this outer periphery of the shaft was brittle.
An SEM micrograph taken of the rough central region (Figure M.66) revealed
the presence of both brittle cleavage features and also dimples; thus, it is apparent
that the failure mode in this central interior region was mixed; that is, it was a com-
bination of both brittle and ductile fracture.
Metallographic examinations were also performed. A transverse cross section
of the failed axle was polished, etched, and photographed using the optical micro-
scope. The microstructure of the outer periphery region, as shown in Figure M.67,
consisted of tempered martensite.21 On the other hand, the microstructure in the
central region was completely different, per the photomicrograph of Figure M.68.
It may be noted that the microconstituents are ferrite, pearlite, and possibly some
bainite.22 In addition, transverse microhardness measurements were taken along
the cross section; Figure M.69 is a plot of the resulting hardness profile. Here it may
be noted that the maximum hardness of approximately 56 HRC occurred near the
surface, and that hardness diminished with radial distance to a hardness of about
20 HRC near the center. On the basis of the observed microstructures and this hard-
ness profile, it was assumed that the axle had been induction hardened.23

21
For a discussion of tempered martensite, see Section 10.8 of Introduction (Section 11.8
of Fundamentals).
22
Ferrite, pearlite, and bainite microconstituents are discussed in Sections 10.5 and 10.7 of
Introduction (Sections 11.5 and 11.7 of Fundamentals).
23
With induction hardening, the surface of a piece of medium-carbon steel is rapidly
heated using an induction furnace. The piece is then quickly quenched so as to produce
an outer surface layer of martensite (which is subsequently tempered), with a mixture of
ferrite and pearlite at interior regions.
M.26 Testing Procedure and Results 81

20 m

Figure M.67 Optical photomicrograph of the failed axle outer-perimeter region, which is
composed of tempered martensite. 500. [Note: Although the microstructure for tempered
martensite in this figure appears to be different from that shown in Figure 10.33 of
Introduction (Figure 11.34 of Fundamentals), they are nevertheless the same. One reason for
this disparity in appearance is due to the difference in magnification of the micrographs:
Figure 10.33/11.34 is an electron micrograph with a magnification that is approximately twenty
times greater than this optical micrograph. Furthermore, the dark regions of this
photomicrograph are clusters of Fe3C particles (which are unresolved) that stand in relief
above the etched a-ferrite matrix, which appears light.] [Reproduced with permission from
Handbook of Case Studies in Failure Analysis, Vol. 1 (1992), ASM International, Materials
Park, OH, 44073-0002.]

Ferrite

Pearlite

20 m

Figure M.68 Optical photomicrograph of the failed axle core region, which
is composed of ferrite and pearlite (and possibly bainite). 500.
[Reproduced with permission from Handbook of Case Studies in Failure
Analysis, Vol. 1 (1992), ASM International, Materials Park, OH, 44073-0002.]
82 Online Support Module: Mechanical Engineering

60 Figure M.69 Transverse


hardness profile across the
axle cross section.

Hardness (converted to Rockwell C)


50 (Microhardness readings
were converted to Rockwell
C values). [Reproduced with
40 permission from Handbook
of Case Studies in Failure
Analysis, Vol. 1 (1992), ASM
30
International, Materials Park,
OH, 44073-0002.]
20

10
0.0 0.2 0.4 0.6 0.8
Distance from outer surface (in.)

The results of these fractographic/metallographic analyses and hardness tests


are summarized in Table M.7.

At this point in the investigation it was not possible to ascertain irrefutably whether
the accident caused the axle failure (scenario 1) or whether the axle fracture caused
the accident (scenario 2). The high hardness and, in addition, the evidence of cleavage
of the outer surface layer indicated that this region failed in a brittle manner as a re-
sult of being overloaded (i.e., as a result of the accident, scenario 1). On the other hand,
the evidence of a mixed ductile-brittle mode of fracture in the central region neither
supported nor refuted either of these two failure scenarios.24
Assuming the validity of the first scenario, it was hypothesized that the core re-
gion was strain-rate sensitive to fracture; that is, at high strain rates, as with the truck
rollover, the fracture mode would be brittle. By contrast, if failure resulted from
loads that were applied relatively slowly, as under normal driving conditions (the
second scenario), the mode of failure would be more ductile.
In order to explore the feasibility of scenario 1 (i.e., the strain-rate sensitivity
of the core region), it was decided to fabricate and test impact specimens taken
from both outer-perimeter and core regions; in addition, a tension test was
to be conducted on a core-region specimen. Failure surfaces for all three specim-
ens were to be subjected to SEM examinations. The following results from these

Table M.7 Tabulation of Test Results on Specimens Taken from Failed Rear
Truck Axle
Analytical Result for Result for
Technique Outer Region Core Region
Fractographic Cleavage features Cleavage features/dimples
(brittle fracture) (ductile/brittle fracture)
Metallographic Tempered martensite Ferrite, pearlite (bainite?)
Hardness tests (profile) Induction hardened
(i.e., heat treatment)

24
It is at this stage of an investigation that a fault tree may be implemented. We have
chosen not to use a fault tree due to the simplicity of this failure problem.
M.26 Testing Procedure and Results 83

Figure M.70 Fracture surface of the Charpy


impact specimen that was taken from the
outer-perimeter region. [Reproduced with
permission from Handbook of Case Studies in
Failure Analysis, Vol. 1 (1992), ASM
International, Materials Park, OH, 44073-0002.]

tests/examinations would be expected if the core region of the axle were sensitive
to the rate of straining:
The failure of the core-region specimen to be impact (high strain rate) tested
would not be totally ductile in nature.
The core-region specimen to be tensile (low strain rate) tested would display
at least a moderate degree of ductility.
The outer-perimeter specimen to be impact tested would fail in a totally brit-
tle manner.

Impact Tests
For the impact tests, small [~2.5-mm- (0.1-in.-) wide] Charpy V-notch test specimens
were prepared from both outer perimeter and interior areas. Because the hardened
outer region was very thin (6.4 mm thick), careful machining of these specimens
was required. Impact tests were conducted at room temperature, and the energy
absorbed by the surface specimen was significantly lower than for the core speci-
men [4 J (3 ft-lbf) versus 11 J (8 ft-lbf)]. Furthermore, the appearances of the frac-
ture surfaces for the two specimens were dissimilar. Very little, if any, deformation
was observed for the outer-perimeter specimen (Figure M.70); conversely, the core
specimen deformed significantly (Figure M.71).
Fracture surfaces of these impact specimens were then subjected to examina-
tion using the SEM. Figure M.72, a micrograph of the outer-perimeter specimen
that was impact tested, reveals the presence of cleavage features, which indicates
that this was a brittle fracture. Furthermore, the morphology of this fracture sur-
face is similar to that of the actual failed axle (Figure M.65).
For the impact specimen taken from the core region the fracture surface had a
much different appearance; Figures M.73a and M.73b show micrographs for this
specimen, which were taken at relatively low and high magnifications, respectively.
These micrographs reveal the details of this surface to be composed of interspersed
cleavage features and shallow dimples, being similar to the failed axle, as shown in
Figure M.66. Thus, the fracture of this specimen was of the mixed-mode type, hav-
ing both ductile and brittle components.
84 Online Support Module: Mechanical Engineering

Figure M.71 Fracture surface of the Charpy


impact specimen that was taken from the core
region. [Reproduced with permission from
Handbook of Case Studies in Failure Analysis,
Vol. 1 (1992), ASM International, Materials
Park, OH, 44073-0002.]

Tensile Tests
A tensile specimen taken from the core region was pulled in tension to failure.The frac-
tured specimen displayed the cup-and-cone configuration, which indicated at least a
moderate level of ductility.A fracture surface was examined using the SEM, and its mor-
phology is presented in the micrograph of Figure M.74. The surface was composed en-
tirely of dimples, which confirms that this material was at least moderately ductile and
that there was no evidence of brittle fracture. Thus, although this core material exhib-
ited mixed-mode fracture under impact loading conditions, when the load was applied
at a relatively slow rate (as with the tensile test), failure was highly ductile in nature.

A summary of the results of these impact and tensile tests is presented in


Table M.8.

Figure M.72 Scanning electron micrograph


of the fracture surface for the impact
specimen prepared from the outer-perimeter
region of the failed axle. 3000.
[Reproduced with permission from
Handbook of Case Studies in Failure
Analysis, Vol. 1 (1992), ASM International,
Materials Park, OH, 44073-0002.]

5 m
M.26 Testing Procedure and Results 85

(a)
100 m

Figure M.73 (a) Scanning electron micrograph of the fracture surface for the impact
specimen prepared from the core region of the failed axle. 120. (b) Scanning electron
micrograph of the fracture surface for the impact specimen prepared from the core region of
the failed axle taken at a higher magnification than (a); interspersed cleavage and dimpled
features may be noted. 3000. [Reproduced with permission from Handbook of Case Studies
in Failure Analysis, Vol. 1 (1992), ASM International, Materials Park, OH, 44073-0002.]

Figure M.74 Scanning electron micrograph


of the fracture surface for the core specimen
that was tensile tested; a completely dimpled
structure may be noted. Approximately
3500. [Reproduced with permission from
Handbook of Case Studies in Failure
Analysis, Vol. 1 (1992), ASM International,
Materials Park, OH, 44073-0002.]

2 m

Table M.8 Tabulation of Impact and Tension Test Results for Specimens Taken
from Core and Outer Regions of Failed Truck Rear Axle
Specimen/Test Fracture Mode Fractographic Features
Core region/impact Some ductility Dimples and cleavage features
Outer region/impact Brittle Cleavage features
Core region/tension Ductile Dimples
86 Online Support Module: Mechanical Engineering

M.27 DISCUSSION
In light of the previous discussion, it was supposed that the truck rollover was
responsible for the axle failure (i.e., scenario 1 was valid). Reasons for this suppo-
sition are as follows (see Tables M.7 and M.8):
1. The outer-perimeter region of the failed axle shaft failed in a brittle manner,
as did also the specimen taken from this region that was impact tested. This
conclusion was based on the fact that both fracture surfaces were very flat
and that SEM micrographs revealed the presence of cleavage features.
2. The fracture behavior of the core region was strain-rate sensitive and indi-
cated that axle failure was due to a single high strain-rate incident. Fracture
surface features for both the failed axle and impact-tested (i.e., high-strain-
rate-tested) specimens taken from this core region were similar: SEM micro-
graphs revealed the presence of features (cleavage features and dimples) that
are characteristic of mixed-mode (brittle and ductile) fracture.
In spite of evidence supporting the validity of the accident-caused-axle-failure sce-
nario, the plausibility of the other (axle-failure-caused-the-accident) scenario (sce-
nario 2) was also explored. This latter scenario assumes that a fatigue crack or some
other slow-crack propagation mechanism initiated the sequence of events that
caused the accident. In this case it is important to consider the mechanical charac-
teristics of the portion of the specimen that was last to failin this instance, the
core region. If failure was due to fatigue, then any increase in loading level of this
core region would have occurred relatively slowly, not rapidly as with impact load-
ing conditions. During this gradually increasing load level, fatigue crack propaga-
tion would have continued until a critical length was achieved (i.e., until the
remaining intact axle cross section was no longer capable of sustaining the applied
load); at this time, final failure would have occurred.
On the basis of the tensile tests (i.e., slow strain-rate tests) performed on this
core region, the appearance of the axle fracture surface would be entirely ductile
(i.e., dimpled, as per the SEM micrograph of Figure M.74). Inasmuch as this core
region of the failed shaft exhibited mixed-mode (ductile and brittle) fracture fea-
tures (both cleavage features and dimples, Figure M.66) and not exclusively dim-
ples, the axle-failure-caused-the-accident scenario was rejected.

SUMMARY
Materials Selection for a Torsionally Stressed Cylindrical Shaft
For a torsionally stressed cylindrical shaft, an expression for strength perform-
ance index was derived (Equation M.9).
Using the appropriate materials selection chart (log strength versus log density,
Figure M.2) a preliminary candidate search was conducted. From the results of
this search, several candidate engineering materials were ranked on both strength-
per-unit mass and cost bases.

Fracture
Principles of Fracture Mechanics
The significant discrepancy between actual and theoretical fracture strengths of
brittle materials is explained by the existence of small flaws that are capable of
Summary 87

amplifying an applied tensile stress in their vicinity, leading ultimately to crack


formation. Fracture ensues when the theoretical cohesive strength is exceeded at
the tip of one of these flaws.
The maximum stress that may exist at the tip of a crack (oriented as in Figure
M.4a) is dependent on crack length and tip radius, as well as the applied tensile
stress according to Equation M.12a.
An expression was developed by A. A. Griffith for a crack propagation critical
stress in brittle materials (Equation M.14), which is a function of elastic modu-
lus, specific surface energy, and crack length. Fracture ensues when this critical
stress is exceeded at the tip of a preexisting flaw or crack.
Sharp corners may also act as points of stress concentration and should be avoided
when designing structures that are subjected to stresses.
There are three different crack displacement modes (Figure M.6): opening (ten-
sile), sliding, and tearing.
A condition of plane strain exists when specimen thickness is much greater than
crack lengththat is, there is no strain component perpendicular to the speci-
men faces.
Stress distributions in front of an advancing crack may be expressed in terms of
position (as radial and angular coordinates) as well as stress intensity factor, K.
The fracture toughness of a material (or critical value of the stress intensity factor)
is indicative of its resistance to brittle fracture when a crack is present. For the plane
strain situation (and mode I loading) it is dependent on applied stress, crack length,
and the dimensionless scale parameter Y as represented in Equation M.22:
KIc Ys1pa
The units for KIc are MPa 1m 1ksi1in.2 .
Minimum specimen thickness for the condition of plane strain (B) is a function
of fracture toughness and yield strength according to Equation M.23.
KIc is the parameter normally cited for design purposes; its value is relatively
large for ductile materials (and small for brittle ones) and is a function of mi-
crostructure, strain rate, and temperature.
With regard to designing against the possibility of fracture, consideration must
be given to material (its fracture toughness), the stress level, and the flaw size de-
tection limit.
Several nondestructive testing (NDT) techniques for detecting and measuring
preexisting cracks were discussed brieflyviz.
Visual
Optical and scanning electron microscopic
Dye penetrant
Magnetic particle
Radiographic
Ultrasonic (pulse-echo)
Acoustic emission

Fracture Toughness Testing


The procedure and details of ASTM Standard E 399-09 for plane-strain fracture
toughness testing of single-edge notched bend and compact tension specimens
were discussed.
A conditional value of plane-strain fracture toughness is determined from a plot
of load versus crack displacement.
88 Online Support Module: Mechanical Engineering

Validation of this conditional value depends on the satisfaction of two criteria:


One criterion (Equation M.32) ascertains whether the material being tested
is too ductile to yield a valid value.
The second criterion (Equation M.37) determines if a condition of plane
strain is met.

Impact Fracture Testing


Three factors that may cause a metal to experience a ductile-to-brittle transition
are exposure to stresses at relatively low temperatures, high strain rates, and the
presence of a sharp notch.
Qualitatively, the fracture behavior of materials may be determined using Charpy
and Izod impact testing techniques (Figure M.19).
On the basis of the temperature dependence of measured impact energy (or ap-
pearance of the fracture surface), it is possible to ascertain whether a material
experiences a ductile-to-brittle transition and the temperature range over which
such a transition occurs.
Low-strength steel alloys typify this ductile-to-brittle behavior and, for structural
applications, should be used at temperatures in excess of the transition range.
Furthermore, low-strength FCC metals, most HCP metals, and high-strength ma-
terials do not experience this ductile-to-brittle transition.

Fatigue
Fatigue is a common type of catastrophic failure wherein the applied stress level
fluctuates with time; it may occur when the maximum stress level is considerably
lower than the static tensile or yield strength.

Cyclic Stresses
Fluctuating stresses are categorized into three general stress-versus-time cycle
modes: reversed, repeated, and random (Figure M.24). Reversed and repeated
are characterized in terms of mean stress, range of stress, and stress amplitude.

The SN Curve
In-service conditions that should be replicated in a fatigue test are stress type,
time frequency, and stress pattern.
Test data are plotted as stress (normally stress amplitude) versus the logarithm
of the number of cycles to failure.
For many metals and alloys, stress diminishes continuously with increasing num-
ber of cycles at failure; fatigue strength and fatigue life are parameters used to
characterize the fatigue behavior of these materials (Figure M.26b).
For other metals (e.g., ferrous and titanium alloys), at some point, stress ceases
to decrease with, and becomes independent of, the number of cycles; the fatigue
behavior of these materials is expressed in terms of fatigue limit (Figure M.26a).

Crack Initiation and Propagation


Fatigue cracks normally nucleate on the surface of a component at some point
of stress concentration.
Summary 89

Crack propagation proceeds in two stages (Figure M.28), which are characterized
by propagation direction and rate. The mechanism for the more rapid stage II
corresponds to a repetitive plastic blunting and sharpening process at the ad-
vancing crack tip (Figure M.29).
Two characteristic fatigue surface features are beachmarks and striations.
Beachmarks form on components that experience applied stress interrup-
tions; they normally may be observed with the naked eye.
Fatigue striations are of microscopic dimensions, and each is thought to
represent the crack tip advance distance over a single load cycle.
An analytical expression (Equation M.43) was proposed for fatigue crack
propagation rate in terms of the stress intensity factor range at the crack tip.
Integration of the expression yields an equation whereby fatigue life may be
estimated.

Factors That Affect Fatigue Life


The Goodman equation (Equation M.49) may be used to estimate fatigue life for
a nonzero mean stress.
Measures that may be taken to extend fatigue life include the following:
Reducing the mean stress level
Eliminating sharp surface discontinuities
Improving the surface finish by polishing
Imposing surface residual compressive stresses by shot peening
Case hardening by using a carburizing or nitriding process

Environmental Effects
Thermal stresses may be induced in components that are exposed to elevated
temperature fluctuations and when thermal expansion and/or contraction is
restrained; fatigue for these conditions is termed thermal fatigue.
The presence of a chemically active environment may lead to a reduction in
fatigue life for corrosion fatigue. Measures that may be taken to prevent this type
of fatigue include the following:
Application of a surface coating
Utilization of a more corrosion-resistant material
Reducing the corrosiveness of the environment
Reducing the applied tensile stress level
Imposing residual compressive stresses on the surface of the specimen

Automobile Valve Spring (Case Study)


A stress analysis was first performed on a helical spring, which was then extended
to an automobile valve spring. Since the valve spring is subjected to cyclic load-
ing, the possibility of fatigue failure is crucial to its performance.
The results of this analysis included computation of the shear stress amplitude,
the magnitude of which was almost identical to the calculated fatigue limit for a
chromevanadium steel that is commonly used for valve springs. It was noted that
the fatigue limit of valve springs is often enhanced by shot peening.
Finally, a procedure was suggested for assessing the economic feasibility of this
spring design incorporating the shot-peened chromevanadium steel.
90 Online Support Module: Mechanical Engineering

Investigation of Engineering Failures


The five possible reasons for failure were presented and explained briefly:
Design errors
Fabrication/manufacturing defects
Assembly errors
Misuse during operation
Improper maintenance
In addition, the three root (or actual) causes of failure were discussed; these root
causes are
Physicalfailure is related to a type of mechanism discussed in the module
(e.g., fracture, fatigue)
Humanthe actions of an individual are the primary causes of failure
Latentfailure is a result of organizational policy
A failure investigation seeks to find answers to four questions:
What exactly is the failure problem?
What is the root cause of the failure problem?
What are possible solutions?
Which of these is the best solution?
With regard to determination of the root cause, a number of procedures are avail-
able, including:
Nondestructive testing
Mechanical testing (tensile, hardness, impact)
Microscopic and macroscopic examination of fracture surfaces
Metallographic examinations of microstructure
Determination of the stress state
Determination of fracture mode
Chemical analyses
Application of fracture mechanics
The complex problem of determining possible root causes of failure is simplified
by using a fault treea taxonomic chart that displays the interrelationships
among a major root cause and its subordinate root causes.
Identification of the root cause may be determined by evaluating the likelihood
of each of the possible root causes in the fault tree. This is possible using the
tests/analyses detailed above.
Formulation of possible solutions (including the best solution) to a failure prob-
lem may be accomplished utilizing a corrective action treea chart constructed
from (and in the form of ) the fault tree, which consists of corrective actions for
each of the fault tree entries.

Failure of an Automobile Rear Axle


(Case Study)
Several procedures were employed in this failure investigation in order to deter-
mine whether the accident caused the axle failure; these are listed as follows:
Fracture surfaces were visually inspected
SEM fractographic inspections were made on both outer perimeter and
interior core regions of the failed axle
Metallographic examinations of both outer surface and interior regions, were
performed and the microconstituents present were identified
Summary 91

A hardness profile across the axle cross section was generated


Impact specimens taken from both surface and interior regions were prepared
and tested; fractographic examinations of both fractured specimens were also
conducted
A tensile specimen taken from the core region was prepared and tested to
fracture. Furthermore, its fracture surface was examined with an SEM.
On the basis of the outcomes of these procedures/analyses, it was concluded that
the automobile accident (i.e., truck rollover) was responsible for the rear axle
failure.

Equation Summary
Equation Page
Number Equation Solving for Number

tf23 Strength performance index


M.9 P for a torsionally stressed 4
r
cylindrical shaft

a 12
M.12a sm s0 c 1 2 a b d 9
rt Maximum stress at tip of
elliptically-shaped crack
a 12
M.12b 2s0 a b 10
rt

2Egs 12 Critical stress for crack


M.14 sc a b propagation in a brittle 12
pa
material

M.19 K Ys 1pa Stress intensity factor 14

M.20 Kc Ysc 1pa Fracture toughness 14

Plane-strain fracture
M.22 KIc Ys1pa toughness, mode I crack 17
surface displacement

KIc 2
B 2.5 a b
Minimum thickness for
M.23 condition of plane strain 17
sy

KIc
M.24 sc Design (or critical) stress 19
Y1pa

1 KIc 2
M.25 ac a b Maximum allowable 19
p sY flaw size

smax smin
M.38 sm Mean stress (fatigue tests) 35
2

M.39 sr smax smin Range of stress 35


(fatigue tests)
92 Online Support Module: Mechanical Engineering

Equation Page
Number Equation Solving for Number
smax smin Stress amplitude
M.40 sa 36
2 (fatigue tests)

smin
M.41 R Stress ratio (fatigue tests) 36
smax

da Fatigue crack
M.43 A1K2 m 44
dN propagation rate

ac


da
Nf
A1Ys 2pa2
a0
M.48 ac
Fatigue life 45

1 da

Ap 1s2 a0 Ymam/2
m/2 m

Goodman equation
sfs sfs0 a1 b
sm
M.49 fatigue strength for 47
TS nonzero mean stress

M.50 s alE T Thermal stress 50

D 0.140
c 1.60 a b d
dcGd Shear stress experienced
M.56 t 52
pD 2
d by a spring

List of Symbols
Symbol Meaning
A Material constant
a Length of a surface crack
D Spring coil center-to-center diameter
d Spring wire diameter
E Modulus of elasticity
G Shear modulus
N Number of stress cycles (fatigue)
TS Tensile strength
T Temperature difference or change
Y Dimensionless parameter or function
al Linear coefficient of thermal expansion
dc Amount of deflection per spring coil
gs Specific surface energy
r Density
rt Crack tip radius
s Applied stress
sfs0 Fatigue strength for zero mean stress
s0 Applied tensile stress
smax Maximum stress (cyclic)
smin Minimum stress (cyclic)
tf Shear strength
References 93

I M P O R TA N T T E R M S A N D C O N C E P T S
case hardening fracture mechanics plane strain
Charpy test fracture toughness fracture toughness
corrosion fatigue impact energy root cause (of
ductile-to-brittle transition Izod test failure)
fatigue materials selection chart stress intensity
fatigue life nondestructive testing factor
fatigue limit performance index stress raiser
fatigue strength plane strain thermal fatigue

REFERENCES
Principles of Fracture Mechanics Collins, J. A., Failure of Materials in Mechanical
Design, 2nd edition, Wiley, New York, 1993.
Griffith, A. A., The Phenomena of Rupture Flow
Courtney, T. H., Mechanical Behavior of Materials,
in Solids, Philos. Trans. R. Soc. London, 221A,
2nd edition, McGraw-Hill, Burr Ridge, IL, 2000.
163 (1920).This paper was republished with an-
Dieter, G. E., Mechanical Metallurgy, 3rd edition,
notations in Trans. ASM, 61, 871 (1968).
McGraw-Hill, New York, 1986.
Inglis, C. E., Stresses in a Plate Due to the Pres-
Esaklul, K. A., Handbook of Case Histories in Fail-
ence of Cracks and Sharp Corners, Trans. Inst.
ure Analysis, ASM International, Materials Park,
Nav. Archit., 55, 219 (1913).
OH, 1992 and 1993. In two volumes.
Irwin, G. R., Fracture, in Encyclopedia of Physics,
Hertzberg, R. W., Deformation and Fracture
S. Flgge (Editor),Vol.VI, Springer, Berlin, 551
Mechanics of Engineering Materials, 4th edition,
(1958).
Wiley, New York, 1996.
Nondestructive Testing (NDT) McEvily, A. J., Metal Failures: Mechanisms, Analysis,
Prevention, Wiley, Hoboken, NJ, 2002.
Cartz, L., Nondestructive Testing: Radiography,
Murakami, Y. (Editor), Stress Intensity Factors
Ultrasonics, Liquid Penetrant, Magnetic Parti-
Handbook, Elsevier Science, New York, 2001.
cle, Eddy Current, ASM International, Materi-
Wulpi, D. J., Understanding How Components Fail,
als Park, OH, 1995.
2nd edition, ASM International, Materials
Hellier, C. J., Handbook of Nondestructive Evalu-
Park, OH, 1999.
ation, McGraw-Hill, New York, 2001.
Mix, P. E., Introduction to Nondestructive Testing: Automobile Valve Spring
A Training Guide, 2nd edition, Wiley, Hobo-
ken, NJ, 2005. Juvinall, R. C., and K. M. Marshek, Fundamentals
Shull, P. J. (Editor), Nondestructive Evaluation: of Machine Component Design, 4th edition,
Theory, Techniques, and Applications, Marcel Chapter 12, Wiley, Hoboken, NJ, 2005.
Dekker, New York, 2002. Shigley, J., C. Mischke, and R. Budynas, Mechani-
cal Engineering Design, 7th edition, Chapter
General References on Failure and Fatigue 10, McGraw-Hill, New York, 2004.
ASM Handbook, Vol. 11, Failure Analysis and Pre-
Investigation of Engineering Failures
vention, ASM International, Materials Park,
OH, 2002. Dennies, D. P., How to Organize and Run a Failure
ASM Handbook, Vol. 12, Fractography, ASM Investigation, ASM International, Materials
International, Materials Park, OH, 1987. Park, OH, 2005.
Colangelo, V. J., and F. A. Heiser, Analysis of Met- Lewis, P. R., K. Reynolds, and C. Gagg, Forensic
allurgical Failures, 2nd edition, Wiley, New Materials Engineering Case Studies, CRC
York, 1987. Press, Boca Raton, FL, 2004.
94 Online Support Module: Mechanical Engineering

QUESTIONS AND PROBLEMS


Note: Those problems having an asterisk (*) by their (c) r  0.50 mm (2.0  102 in.), u  0

numbers are identical to (or refreshed versions of)


problems found in Introduction (8th edition) and (d) r  0.50 mm (2.0  102 in.), u  45

Fundamentals (3rd edition). A problem conversion M.6 The parameter K in Equations M.18a, M.18b,
guide is provided in the section that follows the and M.18c is defined in Problem M.5.
design problems. (a) For a surface crack of length 2.0 mm
(7.87  102 in.), determine the radial po-
Principles of Fracture Mechanics sition at an angle u of 30
at which the
normal stress sx is 100 MPa (14,500 psi)
*M.1 What is the magnitude of the maximum
when the magnitude of the nominal applied
stress that exists at the tip of an internal crack
stress is 150 MPa (21,750 psi).
having a radius of curvature of 2.5  104
mm (105 in.) and a crack length of 2.5  (b) Compute the normal stress sy at this
102 mm (103 in.) when a tensile stress of same position.
170 MPa (25,000 psi) is applied? M.7 Below is shown a portion of a tensile
specimen.
*M.2 Estimate the theoretical fracture strength of
s
a brittle material if it is known that fracture
occurs by the propagation of an elliptically
shaped surface crack of length 0.25 mm (0.01
40 mm
in.) and having a tip radius of curvature of
1.2  103 mm (4.7  105 in.) when a stress
of 1200 MPa (174,000 psi) is applied.
*M.3 If the specific surface energy for soda-lime
glass is 0.30 J/m2, using data contained in P
Table 12.5 of Introduction (Table 7.1 of
Fundamentals), compute the critical stress 4 mm
required for the propagation of a surface
crack of length 0.05 mm.
*M.4 A polystyrene component must not fail 6.5 mm
when a tensile stress of 1.25 MPa (180 psi)
is applied. Determine the maximum allow- 20 mm
able surface crack length if the surface en-
ergy of polystyrene is 0.50 J/m2 (2.86  103
in.-lbf/in.2). Assume a modulus of elasticity s

of 3.0 GPa (0.435  106 psi). (a) Compute the magnitude of the stress at
M.5 The parameter K in Equations M.18a, M.18b, point P when the externally applied stress
and M.18c is a function of the applied nom- is 140 MPa (20,000 psi).
inal stress s and crack length a as (b) How much will the radius of curvature
at point P have to be increased to reduce
K  s 1pa
this stress by 25%?
Compute the magnitudes of the normal M.8 A cylindrical hole 19.0 mm (0.75 in.) in di-
stresses sx and sy in front of a surface crack ameter passes entirely through the thick-
of length 2.0 mm (0.079 in.) (as depicted in ness of a steel plate 12.7 mm (0.5 in.) thick,
Figure M.7) in response to a nominal ten- 127 mm (5 in.) wide, and 254 mm (10 in.)
sile stress of 100 MPa (14,500 psi) at the fol- long (see Figure M.5a).
lowing positions:
(a) Calculate the stress at the edge of this
(a) r  0.10 mm (3.9  103 in.), u  0
hole when a tensile stress of 34.5 MPa (5000
(b) r  0.10 mm (3.9  103 in.), u  45
psi) is applied in a lengthwise direction.
Questions and Problems 95

(b) Calculate the stress at the hole edge *M.15 A structural component in the form of a
when the same stress in part (a) is applied wide plate is to be fabricated from a steel
in a widthwise direction. alloy that has a plane strain fracture
M.9 For each of the metal alloys listed in Table toughness of 77.0 MPa 1m 170.1 ksi1in.2
M.3, compute the minimum component and a yield strength of 1400 MPa (205,000
thickness for which the condition of plane psi). The flaw size resolution limit of the
strain is valid. flaw detection apparatus is 4.0 mm (0.16
in.). If the design stress is one-half of the
*M.10 A specimen of a 4340 steel alloy having a
yield strength and the value of Y is 1.0, de-
plane strain fracture toughness of 45 MPa 1m
141 ksi1in.2 is exposed to a stress of 1000
termine whether a critical flaw for this
plate is subject to detection.
MPa (145,000 psi). Will this specimen ex-
perience fracture if it is known that the M.16 A structural component in the shape of a
largest surface crack is 0.75 mm (0.03 in.) flat plate 25.4 mm (1.0 in.) thick is to be fab-
long? Why or why not? Assume that the ricated from a metal alloy for which the
parameter Y has a value of 1.0. yield strength and plane strain fracture
toughness values are 700 MPa (101,500 psi)
and 49.5 MPa 1m 145 ksi 1in.2, respectively;
*M.11 An aircraft component is fabricated from
an aluminum alloy that has a plane strain
fracture toughness of 35 MPa 1m 131.9
for this particular geometry, the value of Y
is 1.65. Assuming a design stress of one-half
ksi1in.2. It has been determined that frac-
of the yield strength, is it possible to com-
ture results at a stress of 250 MPa (36,250 psi)
pute the critical length of a surface flaw?
when the maximum (or critical) internal
If so, determine its length; if this computa-
crack length is 2.0 mm (0.08 in.). For this
tion is not possible from the given data, then
same component and alloy, will fracture oc-
explain why.
cur at a stress level of 325 MPa (47,125 psi)
when the maximum internal crack length is Impact Fracture Testing
1.0 mm (0.04 in.)? Why or why not?
*M.12 Suppose that a wing component on an air- *M.17 Following is tabulated data that were gath-
craft is fabricated from an aluminum alloy ered from a series of Charpy impact tests
that has a plane strain fracture toughness on a ductile cast iron:
of 40 MPa 1m 136.4 ksi1in.2 . It has been
determined that fracture results at a stress Temperature (C) Impact Energy (J)
of 365 MPa (53,000 psi) when the maxi- 25 124
mum internal crack length is 2.5 mm (0.10 50 123
in.). For this same component and alloy, 75 115
compute the stress level at which fracture 85 100
100 73
will occur for a critical internal crack
110 52
length of 4.0 mm (0.16 in.).
125 26
*M.13 A large plate is fabricated from a steel alloy 150 9
that has a plane strain fracture toughness of 175 6
55 MPa 1m 150 ksi 1in.2 . If, during service
use, the plate is exposed to a tensile stress of (a) Plot the data as impact energy versus tem-
200 MPa (29,000 psi), determine the mini- perature.
mum length of a surface crack that will lead (b) Determine a ductile-to-brittle transition
to fracture. Assume a value of 1.0 for Y. temperature as that temperature corre-
*M.14 Calculate the maximum internal crack sponding to the average of the maximum
length allowable for a 7075-T651 aluminum and minimum impact energies.
alloy (Table M.3) component that is loaded (c) Determine a ductile-to-brittle transition
to a stress one-half of its yield strength. As- temperature as that temperature at which
sume that the value of Y is 1.35. the impact energy is 80 J.
96 Online Support Module: Mechanical Engineering

*M.18 Following is tabulated data that were gath- Cyclic Stresses


ered from a series of Charpy impact tests The SN Curve
on a tempered 4140 steel alloy: *M.19 A fatigue test was conducted in which the
mean stress was 50 MPa (7250 psi) and the
Temperature (C) Impact Energy (J)
stress amplitude was 225 MPa (32,625 psi).
100 89.3
75 88.6 (a) Compute the maximum and minimum
50 87.6 stress levels.
25 85.4 (b) Compute the stress ratio.
0 82.9
25 78.9 (c) Compute the magnitude of the stress
50 73.1 range.
65 66.0 *M.20 A cylindrical 1045 steel bar (Figure M.75)
75 59.3 is subjected to repeated tension-compres-
85 47.9
sion stress cycling along its axis. If the load
100 34.3
125 29.3
amplitude is 22,000 N (4950 lbf), compute
150 27.1 the minimum allowable bar diameter to
175 25.0 ensure that fatigue failure will not occur.
Assume a factor of safety of 2.0.
(a) Plot the data as impact energy versus
*M.21 An 8.0-mm- (0.31-in.-) diameter cylindrical
temperature.
rod fabricated from a red brass alloy (Fig-
(b) Determine a ductile-to-brittle transi- ure M.75) is subjected to reversed tension-
tion temperature as that temperature cor- compression load cycling along its axis. If
responding to the average of the maximum the maximum tensile and compressive
and minimum impact energies. loads are 7500 N (1700 lbf) and 7500
(c) Determine a ductile-to-brittle transi- N (1700 lbf), respectively, determine its fa-
tion temperature as that temperature at tigue life. Assume that the stress plotted in
which the impact energy is 70 J. Figure M.75 is stress amplitude.

80 Figure M.75 Stress magnitude S versus the


logarithm of the number N of cycles to fatigue
500
70 failure for red brass, an aluminum alloy, and a
plain carbon steel. (Adapted from H. W. Hayden,
60 W. G. Moffatt, and J. Wulff, The Structure and
Stress amplitude (10 3 psi)
Stress amplitude, S (MPa)

400
Properties of Materials, Vol. III, Mechanical
1045 steel 50 Behavior, p. 15. Copyright 1965 by John Wiley
300 & Sons, New York. Reprinted by permission of
40
John Wiley & Sons, Inc. Also adapted from ASM
2014T6 aluminum alloy
Handbook, Vol. 2, Properties and Selection:
200 30
Nonferrous Alloys and Special-Purpose Materials,
1990. Reprinted by permission of ASM
20
International.)
100
Red brass 10

0 0
103 104 105 106 107 108 109 1010
Cycles to failure, N
Questions and Problems 97

*M.22 A 12.5-mm- (0.50-in.-) diameter cylindrical (a) Make an SN plot (stress amplitude ver-
rod fabricated from a 2014-T6 alloy (Figure sus logarithm cycles to failure) using
M.75) is subjected to a repeated tension-com- these data.
pression load cycling along its axis. Compute (b) What is the fatigue limit for this alloy?
the maximum and minimum loads that will
(c) Determine fatigue lifetimes at stress
be applied to yield a fatigue life of 1.0  107
amplitudes of 230 MPa (33,500 psi) and
cycles. Assume that the stress plotted on the
175 MPa (25,000 psi).
vertical axis is stress amplitude, and data were
taken for a mean stress of 50 MPa (7250 psi). (d) Estimate fatigue strengths at 2  105
and 6  106 cycles.
*M.23 The fatigue data for a brass alloy are given
as follows: *M.26 Suppose that the fatigue data for the cast
iron in Problem M.25 were taken for
Stress Amplitude bending-rotating tests, and that a rod of
(MPa) Cycles to Failure this alloy is to be used for an automobile
310 2  105 axle that rotates at an average rotational
223 1  106 velocity of 750 revolutions per minute.
191 3  106 Give maximum lifetimes of continuous
168 1  107 driving that are allowable for the follow-
153 3  107 ing stress levels: (a) 250 MPa (36,250 psi),
143 1  108
(b) 215 MPa (31,000 psi), (c) 200 MPa
134 3  108
(29,000 psi), and (d) 150 MPa (21,750 psi).
127 1  109
*M.27 Three identical fatigue specimens (denoted
(a) Make an SN plot (stress amplitude ver- A, B, and C) are fabricated from a nonfer-
sus logarithm cycles to failure) using rous alloy. Each is subjected to one of the
these data. maximum-minimum stress cycles listed in
the following table; the frequency is the
(b) Determine the fatigue strength at 5 
same for all three tests.
105 cycles.
(c) Determine the fatigue life for 200 MPa.
Specimen smax (MPa) smin (MPa)
*M.24 Suppose that the fatigue data for the brass
alloy in Problem M.23 were taken from tor- A 450 350
B 400 300
sional tests, and that a shaft of this alloy is to
C 340 340
be used for a coupling that is attached to an
electric motor operating at 1500 rpm. Give
the maximum torsional stress amplitude pos- (a) Rank the fatigue lifetimes of these three
sible for each of the following lifetimes of the specimens from the longest to the short-
coupling: (a) 1 year, (b) 1 month, (c) 1 day, est.
and (d) 2 hours. (b) Now justify this ranking using a
*M.25 The fatigue data for a ductile cast iron are schematic SN plot.
given as follows: *M.28 Cite five factors that may lead to scatter in
Stress Amplitude Cycles to fatigue life data.
[MPa (ksi )] Failure
Crack Initiation and Propagation
248 (36.0) 1  105 Factors That Affect Fatigue Life
236 (34.2) 3  105
*M.29 Briefly explain the difference between
224 (32.5) 1  106
213 (30.9) 3  106 fatigue striations and beachmarks in terms
201 (29.1) 1  107 of (a) size and (b) origin.
193 (28.0) 3  107 *M.30 List four measures that may be taken to in-
193 (28.0) 1  108 crease the resistance to fatigue of a metal
193 (28.0) 3  108 alloy.
98 Online Support Module: Mechanical Engineering

DESIGN PROBLEMS
Materials Selection Using Performance Indices accompanying figure. Derive strength and
stiffness performance index expressions
M.D1 (a) Using the procedure as outlined in
analogous to Equations M.9 and M.11 for
Section M.2, ascertain which of the metal
this beam. The stress imposed on the un-
alloys listed in Appendix B (of both
fixed end s is
Introduction and Fundamentals), have tor-
FLr
sional strength performance indices greater s (M.62)
than 10.0 (for tf and r in units of MPa and I
g/cm3, respectively), and, in addition, shear L, r, and I are, respectively, the length, ra-
strengths greater than 350 MPa. (b) Also dius, and moment of inertia of the beam.
using the cost database in Appendix C Furthermore, the beam-end deflection d is
(of both Introduction and Fundamentals), FL3
conduct a cost analysis in the same man- d (M.63)
3EI
ner as in Section M.2. For those materials
that satisfy the criteria noted in part (a), where E is the modulus of elasticity of the
and, on the basis of this cost analysis, which beam.
material would you select for a solid cylin- (b) From the properties database pre-
drical shaft? Why? sented in Appendix B (of both Introduc-
M.D2 In a manner similar to the treatment of tion and Fundamentals), select the metal
Section M.2, perform a stiffness-to-mass alloys with stiffness performance indices
performance analysis on a solid cylindrical greater than 3.0 (for E and r in units of
shaft that is subjected to a torsional stress. GPa and g/cm3, respectively).
Use the same engineering materials that (c) Also using the cost database in Ap-
are listed in Table M.1. In addition, con- pendix C (of both Introduction and Fun-
duct a material cost analysis. Rank these damentals), conduct a cost analysis in the
materials on the basis of both mass of ma- same manner as in Section M.2. Relative
terial required and material cost. For glass to this analysis and that in part (b), which
and carbon fiber-reinforced composites, alloy would you select on a stiffness-
assume that the shear moduli are 8.6 and per-mass basis?
9.2 GPa, respectively.
(d) Now select the metal alloys with
M.D3 (a) A cylindrical cantilever beam is sub- strength performance indices greater than
jected to a force F, as indicated in the 14.0 (for sy and r in units of MPa and
g/cm3, respectively), and rank them from
highest to lowest P.
(e) Using the cost database, rank the ma-
terials in part (d) from least to most costly.
Relative to this analysis and that in part
(d), which alloy would you select on a
strength-per-mass basis?
L
(f) Which material would you select if
both stiffness and strength are to be con-
sidered relative to this application? Justify
your choice.
r
M.D4 (a) Using the expression developed for
stiffness performance index in Problem
M.D3(a) and data contained in Appendix
F
B (of both Introduction and Fundamen-
tals), determine stiffness performance
Design Problems 99

indices for the following polymeric materi- (c) Also using the cost database in Ap-
als: high-density polyethylene, polypropy- pendix C (of both Introduction and Fun-
lene, poly(vinyl chloride), polystyrene, damentals), conduct a cost analysis in the
polycarbonate, poly(methyl methacry- same manner as in Section M.2. Relative
late), poly(ethylene terephthalate), polyte- to this analysis and that in part (b), which
trafluoroethylene, and nylon 6,6. How do alloy would you select on a stiffness-per-
these values compare with those of the mass basis?
metallic materials? (Note: In Appendix B, (d) Now select the metal alloys with
where ranges of values are given, use av- strength performance indices greater than
erage values.) 120 (for sy and r in units of MPa and g/cm3,
(b) Now using the cost database in Ap- respectively), and rank them from highest to
pendix C (of both Introduction and Fun- lowest P.
damentals), conduct a cost analysis in the (e) Using the cost database, rank the ma-
same manner as in Section M.2. Use cost terials in part (d) from least to most costly.
data for the raw forms of these polymers. Relative to this analysis and that in part
(c) Using the expression developed for the (d), which alloy would you select on a
strength performance index in Problem strength-per-mass basis?
M.D3(a) and data contained in Appendix B, (f) Which material would you select if
determine strength performance indices for both stiffness and strength are to be con-
these same polymeric materials. sidered relative to this application? Justify
M.D5 (a) A bar specimen having a square cross your choice.
section of edge length c is subjected to a uni- M.D6 Consider the plate shown in the accompa-
axial tensile force F, as shown in the accom- nying figure that is supported at its ends
panying figure. Derive strength and stiffness and subjected to a force F that is uniformly
performance index expressions analogous to distributed over the upper face as indi-
Equations M.9 and M.11 for this bar. cated. The deflection d at the L/2 position
F is given by the expression

5 FL3
d (M.64)
32 Ewt 3

w F

t
L


c
c L

(b) From the properties database pre-


sented in Appendix B (of both Introduc-
tion and Fundamentals), select the metal Furthermore, the tensile stress at the under-
alloys with stiffness performance indices side and also at the L2 location is equal to
greater than 26.0 (for E and r in units of 3 FL
s (M.65)
GPa and g/cm3, respectively). 4 wt 2
100 Online Support Module: Mechanical Engineering

(a) Derive stiffness and strength perform- toughness of the alloy is 50.0 MPa 1m
ance index expressions analogous to Equa- 145.5 ksi 1in.2 , the plate width is 60 mm
tions M.9 and M.11 for this plate. (Hint: (2.4 in.), and the design stress is 375 MPa
Solve for t in these two equations, and then (54,375 psi). The pa/W ratio is in radians.
substitute the resulting expressions into M.D9 Consider a steel plate having a through-
the mass equation, as expressed in terms thickness edge crack similar to that shown
of density and plate dimensions.) in Figure M.10a. If it is known that the min-
(b) From the properties database in imum crack length subject to detection is
Appendix B (of both Introduction and 3 mm (0.12 in.), determine the minimum
Fundamentals), select the metal alloys with allowable plate width assuming a plane
stiffness performance indices greater than strain fracture toughness of 65 MPa 1m
1.40 (for E and r in units of GPa and 159.2 ksi1in.2, a yield strength of 1000
g/cm3, respectively). MPa (145,000 psi), and that the plate is
(c) Also using the cost database in Ap- to be loaded to one-half of its yield
pendix C (of both Introduction and Fun- strength.
damentals), conduct a cost analysis in the M.D10 Consider a steel plate having a through-
same manner as in Section M.2. Relative thickness edge crack similar to that shown
to this analysis and that in part (b), which in Figure M.10a; the plate width (W) is 40
alloy would you select on a stiffness-per- mm (1.6 in.) and its thickness (B) is 6.0 mm
mass basis? (0.25 in.). Furthermore, plane strain frac-
(d) Now select the metal alloys with ture toughness and yield strength values
strength performance indices greater than for this material are 60 MPa 1m 154.6
5.0 (for sy and r in units of MPa and ksi1in.2 and 1400 MPa (200,000 psi), re-
g/cm3, respectively), and rank them from spectively. If the plate is to be loaded
highest to lowest P. to a stress of 200 MPa (29,000 psi), would
(e) Using the cost database, rank the ma- you expect failure to occur if the crack
terials in part (d) from least to most costly. length a is 16 mm (0.63 in.)? Why or
Relative to this analysis and that in part why not?
(d), which alloy would you select on a M.D11 A small and thin flat plate of a brittle ma-
strength-per-mass basis? terial having a through-thickness surface
(f) Which material would you select if crack is to be loaded in the manner of Fig-
ure M.10c; the KIc value for this material is
0.60 MPa1m 10.55 ksi1in.2 . For a crack
both stiffness and strength are to be con-
sidered relative to this application? Justify
your choice. length of 0.5 mm (0.02 in.), determine the
maximum load that may be applied without
Principles of Fracture Mechanics failure for B  1.5 mm (0.06 in.), S  10 mm
(0.39 in.), and W  2.5 mm (0.10 in.).
M.D7 Consider a flat plate of width 100 mm Assume that the crack is located at the S/2
(4.0 in.) that contains a centrally posi- position.
tioned, through-thickness crack (Figure *M.D12(a) For the thin-walled spherical tank
M.9) of length (i.e., 2a) 25 mm (1.0 in.). discussed in Design Example M.1, on the
Determine the minimum plane strain basis of critical crack size criterion [as
fracture toughness necessary to ensure addressed in part (a)], rank the following
that fracture will not occur for a design polymers from longest to shortest critical
stress of 415 MPa (60,000 psi). The pa/W crack length: nylon 6,6 (50% relative
ratio is in radians. humidity), polycarbonate, poly(ethylene
M.D8 A flat plate of some metal alloy contains terephthalate), and poly(methyl methacry-
a centrally positioned, through-thickness late). Comment on the magnitude range of
crack (Figure M.9). Determine the critical the computed values used in the ranking
crack length if the plane strain fracture relative to those tabulated for metal alloys
Design Problems 101

as provided in Table M.4. For these com- M.D16 (a) Calculate values for the A and m pa-
putations, use data contained in Tables B.4 rameters in Equation M.43 (in both SI and
and B.5 in Appendix B (of both Introduc- customary U.S. units) for the crack propa-
tion and Fundamentals). gation rate of the Ni-Mo-V steel for which
(b) Now rank these same four polymers the log da/dN versus log K plot is shown
relative to maximum allowable pressure in Figure M.35.
according to the leak-before-break crite- (b) Suppose that a metal component fabri-
rion, as described in the (b) portion of cated from this Ni-Mo-V steel alloy is ex-
Design Example M.1. Comment on these posed to cyclic tensile-compressive stresses.
values in relation to those for the metal al- If the fatigue lifetime must be a minimum
loys that are tabulated in Table M.5. of 3  105 cycles and it is known that the
critical surface crack length is 1.5 mm and
Crack Propagation Rate the maximum tensile stress is 30 MPa,
M.D13 Consider a flat plate of some metal alloy compute the maximum initial surface crack
that is to be exposed to repeated tensile- length. Assume that Y is independent of
compressive cycling in which the mean crack length and has a value of 1.25.
stress is 25 MPa. If the initial and critical sur- M.D17 Consider a thin metal plate 25 mm wide that
face crack lengths are 0.25 and 5.0 mm, re- contains a centrally positioned, through-
spectively, and the values of m and A are 4.0 thickness crack in the manner shown in Fig-
and 5  1015, respectively (for s in MPa ure M.9. This plate is to be exposed to re-
and a in m), estimate the maximum tensile versed tensile-compressive cycles of stress
stress to yield a fatigue life of 3.2  105 amplitude 120 MPa. If the initial and criti-
cycles. Assume the parameter Y has a value cal crack lengths (i.e., 2a0 and 2ac) are 0.10
of 2.0, which is independent of crack length. and 6.0 mm, respectively, and the values of
M.D14 Consider a large, flat plate of a metal al- m and A are 4 and 6 1012, respectively
loy that is to be exposed to reversed ten- (for s in MPa and a in m), estimate the
sile-compressive cycles of stress amplitude fatigue life of this plate.
150 MPa. If initially the length of the M.D18 For an edge crack in a plate of finite width
largest surface crack in this specimen is (Figure M.10a), Y is a function of the crack
0.75 mm and the plane strain fracture length-specimen width ratio as
toughness is 35 MPa 1m, whereas the
values of m and A are 2.5 and 2  1012,
b
0.2a
1.1a1 
respectively (for s in MPa and a in m), W
estimate the fatigue life of this plate. Y (M.66)
a 32
Assume that the parameter Y has a a1  b
W
value of 1.75 that is independent of crack
length. Now consider a 60-mm-wide plate that is
M.D15 Consider a metal component that is ex- exposed to cyclic tensile-compressive
posed to cyclic tensile-compressive stresses (reversed stress cycle) for which
stresses. If the fatigue lifetime must be a smin = 135 MPa. Estimate the fatigue life
minimum of 5  106 cycles and it is of this plate if the initial and critical crack
known that the maximum initial surface lengths are 5 mm and 12 mm, respectively.
crack length is 0.02 in. and the maximum Assume values of 3.5 and 1.5  1012 for
tensile stress is 25,000 psi, compute the the m and A parameters, respectively, for
critical surface crack length. Assume that s in units of megapascals and a in meters.
Y is independent of crack length and has M.D19 The spherical tank shown in Figure M.12 is
a value of 2.25, and that m and A have alternately pressurized and depressurized
values of 3.5 and 1.3  1023, respec- between atmospheric pressure and a pos-
tively, for s and a in units of psi and itive pressure p; thus, fatigue failure is a
in., respectively. possibility. Utilizing Equation M.48, derive
102 Online Support Module: Mechanical Engineering

an expression for the fatigue life Nf tively corrosion resistant, and, of course,
in terms of p, the tank radius r and thick- capable of being fabricated into wire form.
ness t, and other parameters subject to the Justify your decision.
following assumptions: Y is independent M.D22 A spring having seven coils and a coil-to-
of crack length, m 2, and the original coil diameter of 0.5 in. is to be made of
and critical crack lengths are variable cold-drawn steel wire. When a tensile load
parameters. of 15 lbf is applied, the spring is to deflect
no more than 0.60 in.The cold-drawing op-
Design and Materials Selection for Springs
eration will, of course, increase the shear
M.D20 A spring having a center-to-center diame- yield strength of the wire, and it has been
ter of 20 mm (0.8 in.) is to be constructed observed that ty (in ksi) depends on wire
of cold-drawn and annealed 316 stainless diameter d (in in.) according to
steel wire that is 2.5 mm (0.10 in.) in 63
diameter; this spring design calls for eight ty  (M.67)
d 0.2
coils.
(a) What is the maximum tensile load that If the shear modulus for this steel is 11.5
may be applied such that the total spring  106 psi, calculate the minimum wire di-
deflection will be no more than 6.5 mm ameter required such that the spring will
(0.26 in.)? not plastically deform when subjected to
the preceding load.
(b) What is the maximum tensile load that
may be applied without any permanent de- M.D23 A helical spring is to be constructed from
formation of the spring wire? Assume that a 4340 steel. The design calls for five coils,
the shear yield strength is 0.6sy, where sy a coil-to-coil diameter of 12 mm, and a wire
is the yield strength in tension. diameter of 3 mm. Furthermore, a maxi-
mum total deflection of 5.0 mm is possible
M.D21 You have been asked to select a material
without any plastic deformation. Specify a
for a spring that is to be stressed in ten-
heat treatment for this 4340 steel wire in
sion. It is to consist of ten coils, and the
order for the spring to meet the preceding
coil-to-coil diameter called for is 15 mm;
criteria.Assume a shear modulus of 80 GPa
furthermore, the diameter of the spring
for this steel alloy, and that ty 0.6sy.
wire must be 2.0 mm. Upon application of
[Note: heat treatment of the 4340 steel is
a tensile force of 35 N, the spring is to ex-
discussed in Section 10.8 of Introduction
perience a deflection of no more than 12
(Section 11.8 of Fundamentals).]
mm, and not plastically deform.
(a) From the materials included in the
database in Appendix B (of both Intro- Failure Analysis
duction and Fundamentals), make a list of *M.D24 Each student (or group of students) is to
candidate materials that meet the preced- obtain an object/structure/component that
ing criteria. Assume that the shear yield has failed. It may come from your home,
strength is 0.6sy, where sy is the yield an automobile repair shop, a machine
strength in tension, and that the shear shop, and so on. Conduct an investigation
modulus is equal to 0.4E, E being the mod- to determine the cause and type of failure
ulus of elasticity. (i.e., simple fracture, fatigue, creep). In ad-
(b) Now, from this list of candidate mate- dition, propose measures that can be taken
rials, select the one you would use for this to prevent future incidents of this type of
spring application. In addition to the pre- failure. Finally, submit a report that ad-
ceding criteria, the material must be rela- dresses these issues.
Problem Conversion Guide 103

PROBLEM CONVERSION GUIDE

Problem Number, Problem Number Problem Number,


Module Problem Introduction, 8th Fundamentals, 3rd Previous Version
Number Edition Edition of ME Module
M.1 8.1 9.1
M.2 8.2 W9.2
M.3 8.3 W9.3
M.4 8.4 9.3
M.5 M.1
M.6 M.2
M.7 M.3
M.8 M.4
M.9 M.5
M.10 8.5 W9.4
M.11 8.6 9.4
M.12 8.7 W9.5
M.13 8.8 9.5
M.14 8.9 W9.6
M.15 8.10 9.6
M.16 M.6
M.17 8.12 9.9
M.18 8.13 W9.10
M.19 8.14 9.10
M.20 8.15 W9.11
M.21 8.16 9.11
M.22 8.17 W9.12
M.23 8.18 W9.13
M.24 8.19 W9.14
M.25 8.20 9.12
M.26 8.21 9.13
M.27 8.22 9.14
M.28 8.23 W9.16
M.29 8.24 W9.17
M.30 8.25 9.16
M.D1 M.D13
M.D2 M.D14
M.D3 M.D15
M.D4 M.D16
M.D5 M.D17
M.D6 M.D18
M.D7 M.D1
M.D8 M.D2
M.D9 M.D3
M.D10 M.D4
M.D11 M.D5
M.D12 8.D2 9.D1
M.D13 M.D6
M.D14 M.D7
M.D15 M.D8
M.D16 M.D9
M.D17 M.D10
M.D18 M.D11
M.D19 M.D12
104 Online Support Module: Mechanical Engineering

Problem Number, Problem Number Problem Number,


Module Problem Introduction, 8th Fundamentals, 3rd Previous Version
Number Edition Edition of ME Module
M.D20 M.D19
M.D21 M.D20
M.D22 M.D21
M.D23 M.D22
M.D24 8.D1 W9.D1

You might also like