Understanding The Failure of Materials and Structures An Introduction
Understanding The Failure of Materials and Structures An Introduction
Understanding The Failure of Materials and Structures An Introduction
@seismicisolation
i
@seismicisolation
@seismicisolation
@seismicisolation
@seismicisolation
iii
Understanding the
Failure of Materials
and Structures
An Introduction
David Jesson
@seismicisolation
@seismicisolation
iv
@seismicisolation
@seismicisolation
v
For my parents, without whom I could not have started this book,
And for Ellie, without whom I could not have finished it.
@seismicisolation
@seismicisolation
@seismicisolation
@seismicisolation
vii
Contents
Preface....................................................................................................................... ix
Acknowledgements................................................................................................ xi
10. Strain Rate Dependence:Why It Matters How Fast You Test.............. 109
Index...................................................................................................................... 157
@seismicisolation
@seismicisolation
vii
@seismicisolation
@seismicisolation
ix
Preface
At the time of writing this preface, one of the very last parts of the book
to be written, several announcements have been made about new materials,
including “High fatigue resistance in a titanium alloy via near-void-free
3D printing”, which was published in Nature, on the 28th February, 2024.
As someone who has broken a wide variety of specimens whilst investi-
gating materials for automotive, aerospace, defence, and civil engineering
applications, I find the testing of new materials every bit as fascinating as the
new materials themselves.
As well as undertaking my own research, I have also managed a laboratory,
and have therefore had the exciting opportunity of supporting students as
they study new materials, or new applications. Whether undergraduate or
postgraduate, engineer, physicist, chemist, the vast majority had one thing
in common: this was their first experience of mechanical characterisation.
Some had a theoretical understanding of one aspect or another. Some arrived
gripping a standard. But no-one arrived knowing exactly what they needed
to do, or the implications of what their test required.
This book was born from a number of repeated conversations: engin-
eers usually understand the mechanics of the test, but very little about the
materials and what effect their fundamental nature has on the test, whilst
physicists and chemists usually understand the materials but need a helping
hand to get started with the testing.
I cannot claim to have captured every single aspect of mechanical character-
isation, but I have done my best to lay out the fundamentals for undertaking
the mechanical characterisation of materials and the impact of microstructure
on testing. I hope those coming to mechanical characterisation for the first
time will find the book of use as a primer, and those coming back to testing
will find something in these pages which will make their testing easier.
Happy researching!
David Jesson,
March 2024
@seismicisolation
@seismicisolation
ix
@seismicisolation
@seismicisolation
newgenprepdf
xi
Acknowledgements
@seismicisolation
@seismicisolation
xi
@seismicisolation
@seismicisolation
1
Learning from Failure
1.1 Introduction
On July 23rd, 1983, Air Canada Flight 143 was forced to make an emergency
landing. The flight became known as the Gimli Glider –not for the dwarfish
member of the Fellowship of the Ring, but because the pilot effected an emer-
gency glider landing at a former Royal Canadian Airforce base at Gimli,
Manitoba. The cause of the emergency is commonly stated to be a mistake in
fuelling caused by confusing metric and imperial units, leading to the plane
carrying less than 50% of the required fuel (e.g., Witkin, 1983). In fact, the
cause of the accident was much more complex than this simple error –the
fuel issue was just the last in a number of incidents, leading to what is termed
a ‘Swiss cheese failure’1 (Klein, 2011). The outcome of this event, happily,
resulted in no loss of life: the pilot was skilled and was able to land the plane
despite a lack of emergency protocols for the situation he found himself in.
Unusually, he had experience as a glider pilot, and this stood him in good
stead. The ‘Swiss cheese failure’ could have ended in tragedy if a few more
factors had lined up.
Fundamentally though, whilst a faulty gauge was at the heart of the
problem, the failure was one of human error. People made calculation errors,
because they used incorrect conversion factors. A technician was distracted
and forgot to turn an indicator off. The aircraft structure, however, was
sound: the failure in this instance was fundamentally procedural.
1
A ‘Swiss cheese failure’ is one where all the holes, which is to say the things that could go
wrong, line up, and give rise to an event that leads to failure. It might not matter that a draught
causes a door to slam, but if that door is your front door and you don’t have your keys, you
will be locked out.
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-1 1
2 Understanding the Failure of Materials and Structures
2
“Houston, we’ve had a problem.” –James Lovell
3
They also comment on the nature of the investigation, one of the first systematic inquiries of
its kind, and probably pioneering in its use of photography to log certain kinds of evidence.
@seismicisolation
@seismicisolation
Learning from Failure 3
associated with the Class I. Eventually it was discovered that the square-
framed windows were acting to concentrate the stress, leading to the devel-
opment of fatigue cracks.
Setting aside human error, failures are a result of the action of the environ-
ment on the structure of interest, where such action exceeds the material’s
ability to cope. By causing failure on our own terms, by examining failures
that occur unexpectedly, we can collect data that inform models and decision-
making processes. If we are careful, we can design better, safer, longer-lasting
structures –if that is our focus. Or we can design materials that will fail in a
manner of our choosing, for example biodegradable packaging, safety glass,
and props for stunts: in these instances, we need to ensure that the item has
sufficient strength for the job it is to be used for, but that it will fail in the right
way at the right time.
The trend is clear, but only #4 is appropriate for the purposes of this book, and
even this is too limited. From an engineering perspective, failure is as simple
@seismicisolation
@seismicisolation
4 Understanding the Failure of Materials and Structures
as the thing of interest being unable to do the job for which it is designed, but
failure can look like the sinking of a ship (‘unsinkable’ or otherwise), a cave-
in in a mine, faulty items coming off a production line, or a great many other
things, including seemingly trivial things like a liquid leaking from a failed
seal.4 The thing in question has ceased to function, it has failed: but simply
saying that it has failed is seldom enough. We need to understand why it has
failed.
Let us consider the timeline of a failure:
Inciting incident(s) → Event → Aftermath
The event itself is typically a moment of time which might be as short as
a few microseconds or potentially as long as a few moments.5 It is the point
at which the key characteristic of the item is overwhelmed by the situation it
finds itself in. A defective item can survive a surprisingly long time, provided
it is not subjected to a critical event. A wooden bench, for example, might be
rotten to the point of instability, but will only fail when someone sits on it.
The aftermath is the province initially of those who must deal with whatever
happened. This might be as simple as mopping the floor but might require
emergency services. Once all is made safe, investigators will determine what
happened. This might be as simple as someone saying “Well, I won’t pick up
a can with oily hands again” or it might require the painstaking collection of
every single piece of an aircraft and months of examinations.
During the aftermath, as well as looking at the physical evidence of the
broken item, the investigators will take a history, looking at the factors that
might have contributed to the failure. This history may be very straightfor-
ward or may take in multiple contributing factors; it may be of the order of
seconds, minutes, or days between an inciting incident and the failure itself,
or it may be years, decades, or centuries. The timescale will be dependent on
the mechanisms at work.
Failure does not have to be catastrophic of course. A worn item can fail and
stop doing its job. This might happen if a manufacturing process miscues, for
4
The consequences of such a leakage can seem trivial if the liquid in question is water, and more
obviously problematic if the liquid is oil or brake fluid, for example. However, it depends
on how widely we are thinking. On the one hand, given the over-consumption of water in
the West and the potential for water-stress, water wasted through leaks has some serious
socio-economic considerations. On the other hand, a leak underground, going unobserved
for many years can give rise to perfect conditions for accelerated corrosion, and/or a form
of in-soil turbulence which can cause pebbles to rotate and wear a deeper hole in the pipe,
leading to a catastrophic failure in due course. Hence, even relatively minor ‘failures’ should
be prevented or addressed in a timely fashion. A trunk main burst, for example, can cause the
release of millions of litres of water, untold water damage to local properties, disruption, and
the shutting off of the supply to local householders.
5
Where a moment was historically quantified as a fortieth of a solar hour, or about 90 seconds,
depending on the time of day and the time of the year. If this sounds imprecise, it is: it’s a
function of attempting to determine accuracy with inaccurate tools, at a time when 99% of the
population measured time by religious ritual, meal times, and whether there was enough nat-
ural light to work or not.
@seismicisolation
@seismicisolation
Learning from Failure 5
• Wear: two parts moving against each other, giving rise to damage, e.g.,
a rope in contact with a hard surface; gears in a mechanism; footfall.7
6
In some instances, such failures can become more valuable than the original item, such as in
the case of coins and stamps where striking and printing errors give rise to rare defects.
7
In this instance it is more likely to be the soles of the shoes that become worn, especially in the
short term, but consider stone steps in an old building and the distinctive dip that occurs from
generations of feet passing up and down, for example.
@seismicisolation
@seismicisolation
6 Understanding the Failure of Materials and Structures
8
Some chemical attacks can be self-limiting, leading to passivation and subsequent protective
coatings. Stainless steel, for example, is protected by the thick oxide layer which forms on the
surface and prevents further oxidation from occurring, unless the layer is damaged. Unless
treated, the layer on aluminium is porous and so attack can continue, although a small quan-
tity of aluminium added to silver leads to a passivating oxide layer which prevents tarnish
from occurring. Protective patinas of various copper-based minerals can be formed on copper.
@seismicisolation
@seismicisolation
Learning from Failure 7
9
Barring, of course, some unfortunate event such as a power-surge that killed all the bulbs
simultaneously.
@seismicisolation
@seismicisolation
Learning from Failure 9
The book finishes with a look at some of the challenges faced by some
extreme situations, and with some thoughts on ‘unbreakable’ materials and
the structures of the future.
Rather than gather all references at the end of the book, those for a specific
chapter follow on from a summary. Some, but not all, chapters also include
some recommended reading, which whilst not explicitly referenced in the
text has informed the book and is considered useful for anyone looking to
improve their understanding of materials, their mechanical characterisation,
their engineering uses, and ultimately their failure.
References
Court of Enquiry. Report upon the circumstances attending the fall of a portion of the
Tay Bridge, 1880. (See e.g., www.railwaysarchive.co.uk/documents/BoT_Tay
Inquiry1880.pdf, accessed 07/06/21)
Hemeda, S. and Sonbol, A., 2020. Sustainability problems of the Giza pyramids.
Heritage Science, 8(1), pp.8.
Jones, D.R.H., 1994. The Tay bridge disaster—Faulty materials or faulty design?.
Engineering Failure Analysis, 1(3), pp.243–253.
Klein, G.A., 2011. Streetlights and shadows: Searching for the keys to adaptive decision
making. MIT Press.
Lewis, P.M.R. and Reynolds, K., 2002. Forensic engineering: A reappraisal of the Tay
Bridge disaster. Interdisciplinary Science Reviews, 27(4), pp.287–298.
Martin, T. and Macleod, I.A., 1995, May. The Tay Rail Bridge Disaster-A Reappraisal
Based on Modern Analysis Methods. In Proceedings of the Institution of Civil
Engineers-Civil Engineering (Vol. 108, No. 2, pp. 77–83). Thomas Telford-ICE
Virtual Library.
United States. National Aeronautics and Space Administration and Cortright, E.M.,
1970. Report of Apollo 13 review board. National Aeronautics and Space
Administration.
Witkin, R., 1983. Jet’s fuel ran out after metric conversion errors. New York Times.
@seismicisolation
@seismicisolation
2
Why Break Things?
1
To date, the oldest known cache of tools, indicating the use of both stone and wood, were
found in Schöningen, Germany. Originally believed to have been up to 400,000 years old, a
recent assessment (Richter and Krbetschek, 2015) suggests that they are as recent as 300,000.
In either case they are associated with H. heidelbergensis rather than H. sapiens, i.e., an extinct
ancestor to the humans now inhabiting the Earth.
@seismicisolation
@seismicisolation
10 DOI: 10.1201/9780367822347-2
Why Break Things? 11
Today as you read this book, unless you have gone to some special effort
to remove yourself from the daily routine, you are surrounded by all sorts of
materials: fabrics for clothing and furnishings; timber, brick, concrete, and
other construction materials; metals for wiring, cookware, white ware; paper
for books, letters, packaging; plastics –everywhere!; and of course the dense
concentration of a plethora of different materials used in the range of elec-
tronic devices.2
All these materials serve a purpose; each represents a considered selection
of a set of properties. Materials may be chosen for a property other than their
strength, although this may well be a secondary consideration in such cases.
The suite of properties that might be important for a design include optical,
electrical, and corrosion resistance, to name but three.
But, the critical issue at this point, is that no one considers a material for
its own sake. We choose materials because we want to make things: bridges,
cars, aircraft, spacecraft, phones –the list is that of life, ranging from everyday
existence through to the challenges that we set ourselves as a species, and not
forgetting artistic expression. Every structure that we make is dependent on
design choices, and every design choice is dependent on an understanding of
materials and what they can do for us.
People have been observing the natural world since before we could really
be considered people. Our primitive ancestors would have behaved much
as the creatures we see around us today: they would have looked for tell-
tale signs of predators, for food, and for changes in the environment that
would have hinted at short-term weather patterns. Those that were good at
this survived and passed on their knowledge; those that were less skilled did
not. This is of course an over-generalisation, but since we are not considering
the evolution of species, we will leave the point there to focus on the evolu-
tion of scientists and engineers.
In Guns, Germs, and Steel, Diamond (1998) summarises the evolution of
societies, amongst other factors, to explain why the range of societies that
we have today exists in the way that it does. One factor that is considered in
detail is the rise of farming, at the expense of the hunter-gatherer lifestyle.
Whilst there are a surprising number of disadvantages to sticking in one spot,
one of the eventual advantages is that instead of the entire band (typically 10–
20 people comprising one or two extended families), several bands can come
together as a tribe (hundreds of people) or eventually a chiefdom (thousands),
and a proportion of the people can grow all the food for everybody. This, of
course, gives rise to all sorts of issues that are the domain of politicians, social
scientists, and anthropologists: a class structure, bureaucracy, specialists. It
2
There are around forty different elements used in a mobile phone, for example (UNEP, 2009;
OECD, 2010). The actual number of materials is harder to assess because there are several
different plastics (which are, in general, variations on the theme of carbon and hydrogen,
sometimes with oxygen, nitrogen, and/or one or other halogens). On the other hand, several
elements form unique alloys.
@seismicisolation
@seismicisolation
12 Understanding the Failure of Materials and Structures
is the last that we are most interested in: specialists do not have to spend
their days collecting food to live (unless they are the specialist farmers, of
course), and can therefore develop their skills, learning to improve on the
basics. Amongst other things, this evolution of society gives the specialists
more time to think, and to undertake experiments.
At this point, everyone was a scientist in some respect, but there were no spe-
cialist scientists. The farmers were undertaking genetic breeding experiments,
of a sort, the flint knappers were learning to discriminate between good-and
inferior-quality flints. Potters were learning to mould clay; metal workers
were discovering how to work copper, bronze, iron. Leaders were learning
how to make use of the workforce available to embark on civic works.
2.2 Specialist
In modern times, it might be argued that we are drifting into a state of over-
specialisation at too early a stage. Robert Heinlein once said that “specializa-
tion was for insects”: he felt that people ought to be able to take responsibility
for everything that touched their lives, ranging from building your own
house through to growing and cooking your own food, and beyond. The
problem with this is that there are so many things which work better with
more sophisticated tools and machinery than are available in the home envir-
onment. So, for example, it would be possible to make the paper for (the
physical edition of) this book at home, but it would not be as smooth as the
paper manufactured in a paper-mill.
In terms of specialisation, the role of the scientist as we would understand
it today is a recent construct when compared with the emerging specialists of
pre-history. It is accepted that it was not until the 19th century that the pro-
fessional scientist really came into being.3 This is not to say that scientists did
not exist before this, but they tended to be hobbyists, often working on their
experiments part time whilst employed as government officials or teachers,
for example. The earliest people we think of as scientists are usually referred
to as philosophers4 and were simultaneously engaged at several points along
a spectrum that takes in the various disciplines (including subjects now
discredited, such as astrology) and engineering. For example, Imhotep, the
3
Ironically, it was one of the last great generalists, William Whewell, who proposed the term
‘scientist’ in 1833 to overcome issues with descriptions. By this time, specialisation was
becoming the norm, with increasingly niche areas of focus becoming at once popular and
derided. At this point, a university education was still something of a rarity and one of the
greatest professional scientists of the time, Michael Faraday, had no formal education in this
regard, but rather served his time as a laboratory assistant and learned his craft in this manner.
4
Literally, lovers (philo) of knowledge (sofia).
@seismicisolation
@seismicisolation
Why Break Things? 13
builder of the Step Pyramid (built between approximately 2630 and 2611
bc), was an important government official and is credited with being the
architect of this early pyramid. Contemporaneous evidence is scarce, but he
was clearly an influential figure: after his death he was deified and credited
with the writing of various wisdom texts and with being a great physician.
Several thousand years later, and the pattern continues with people such as
Archimedes, who was clearly doing some deep thinking about statics and
hydrostatics, and equally clearly was making ‘engines’. Again, looking at a
body of work from a distance in time, it can be difficult to determine exactly
what he was responsible for inventing, what he improved upon, and what
has been corrupted down the centuries. It is interesting to note, however, that
he is supposed to have been most proud of his proof that both the volume
and the surface area of a sphere are two-thirds that of a cylinder (including
the ends), where the height of the cylinder is equal to the diameter of the
sphere.
By contrast, the concept of a professional engineer is much older. The word
engineer dates to the Middle Ages, and is associated particularly with the
building of fortifications, but a similar role can be found in the Roman mili-
tary. The concept of an engineer was also applied in civil applications, in
terms of the construction of large buildings, naval architecture, and in areas
that today we would consider to be the domain of mechanical engineering.
Fundamentally, the learning point that can be taken from this is that there
is a great overlap between the function of the engineer and that of the scien-
tist, particularly in certain areas of study. One of the things that we associate
more closely with engineering rather than science, however, is the process of
design.
@seismicisolation
@seismicisolation
14 Understanding the Failure of Materials and Structures
1970. Equally surprising is that less than 1600 planes based on the 747 design
have ever been produced. There are several variants, including the VC-25,
better known by its call-sign of Airforce One, the specially fitted version used
to transport the US president and entourage, and the E4 variant used as a
mobile command post by the US Government.
The design process is invariably complicated and fraught with last-
minute changes. An extreme example, presented anecdotally, was that of
a container to hold reaction mass5 for a satellite. Payloads for spacecraft
are limited by both volume (in terms of shape as well as total space avail-
able) and mass. Equally though, a launch is expensive, and it makes sense
to maximise the usage of the available space. But if the reason that you
are undertaking the launch in the first place is to put a communications
satellite weighing several tons into orbit, the left-over space can be oddly
shaped. Still, you want to make the most of the pocket available, and many
scientific experiments hefted into orbit do not need to be beautiful, or even
to last exceptionally long, but they do need to have specific features. In this
example, a ‘nano-sat’6 was taking advantage of some dead space that was
present in the cargo hold of a launch: therefore, there was a specific volume
available, both in terms of shape and size. The focus was on the mission-
critical features of the satellite –instrumentation and the like. The remaining
space for the fuel tank was, surprisingly, triangular in cross-section. This is
a less than ideal shape to work with, because under pressure, the triangular
corners will be weak points. Happily, it was not necessary to use the entire
volume for the reaction mass required for the mission, and this enabled
some creative design approaches to be considered. The final design was
for some piping to be coiled into a shape that might be described as a tri-
angular spring.
There are therefore several facets to the design round, which include:
I. The purpose of the overall structure or device. A car, for example, will
require distinct characteristics dependent on whether it is a sports car,
designed to go fast, or a family car, designed to carry everything and the
kitchen sink. Similarly, if we consider a laptop, an ultralight device is
unlikely to have much capacity in terms of integral memory or battery
5
Reaction mass is used to manoeuvre a satellite in space. The mass is not burned as in a com
bustion engine but is typically pressurised, and sometimes additionally heated by electricity
produced from solar panels on the satellite. The reaction mass is then released from various
nozzles placed around the satellite to move it. The usual reason for this is to return it to its
orbit when the orbit degrades over time and gravitational attraction of the Earth starts to draw
it home. In some instances, it might be necessary to move the satellite’s position relative to the
Earth, but in reaction mass terms this can be expensive. Some satellites can be ‘refuelled’, but
this can be dangerous and not entirely straightforward, so in many cases once the reaction
mass has been used up, this marks the end phase for the satellite. It may still have many useful
years ahead of it, but its functionality has been reduced because it can no longer be moved.
6
A satellite massing only a few kilograms.
@seismicisolation
@seismicisolation
Why Break Things? 15
capacity. On the other hand, there are laptops that might be described
as ‘luggable’ rather than portable and these tend to have higher per-
formance such as in the case of a gaming computer which will have
power-and memory-hungry dedicated graphics cards on board. Then
again, a laptop might be designed with robustness in mind, being
required for use in difficult field conditions. Such toughness might be
achieved by using a flight case, or similar hard-skinned, reinforced
box, but these can reduce the usability of the device. Further, whilst
the laptop is removable from the bulky protection, this is not always
desirable. Whilst obvious, having removed the protection, the device
is no longer protected. The resilience required might be achieved more
permanently using a more robust (but expensive) electronic architec-
ture and using a more rugged casing.
II. The interaction between various subsystems. Most things that are
made are formed of more than one component. Some are passive,
for example a protective coating. In other situations, there is a range
of systems that need to work together either directly or indirectly.
Returning to our automotive example, working from the prime func-
tionality of a device to move from A to B carrying passengers and
cargo, we can see that there is something that needs to move and some-
thing that needs to provide power to enable that movement to occur.
The wheels move, so they need to be attached to the chassis in such
a way that they can rotate, and there needs to be a linkage between
the engine and the wheels to transfer the power generated to where
it needs to be. This linkage may be simple or complicated, depending
on whether the car is four-wheel drive, rear-wheel drive, or front-
wheel drive. But, one has only to look at a modern car in comparison
to an early one to know that things have become more complex. For
example, to start cars one used to have to crank a starting handle to
initiate the combustion process by providing the initial compression
cycle –get this wrong and it was all too easy to end up with a broken
wrist or arm. This has evolved to the electric self-starter that we know
and love today –but this then requires an electrical system. On the
other hand, this electric system can also be used to replace candle or
oil lamps, and these days to power the entertainment systems that
have superseded simple radios, and GPS navigation.
III. Safety requirements, which may be dependent on location, or
overall function, and which may need to consider bystanders as
much as the user. Again, the automotive industry provides obvious
examples. Bonnet ornaments were required to be removed to protect
pedestrians in the event of a collision, and indeed there have also
been requirements to change the shape of cars so that in the event
of a collision a pedestrian is not flung up in such a way as to cause
greater injury. Further, whilst advances in materials have led to cars
@seismicisolation
@seismicisolation
16 Understanding the Failure of Materials and Structures
7
Although it is unlikely in the extreme that there was sufficient exchange of ideas for these
structures to influence the others.
8
Jerry Lord of Boeing has suggested that there should be a second pyramid that should be
considered at the same time. Whilst not entirely relevant to the current context, it is worth
noting for completeness. This second pyramid considers the validation of the manufacturing
process.
@seismicisolation
@seismicisolation
Why Break Things? 17
FIGURE 2.1
The Pyramid of Testing: An understanding of the behaviour of a structure or complex assemblage,
such as an airplane, is built on a fundamental understanding of the behaviour of materials. This
understanding helps with the design of coupons, and once these are tested, the information
gathered informs the building of structural elements and more complex components.
the fibre architecture.9 Rouchon (1990) moves on from these simple coupons
to ‘elements’ (shapes produced with specific fibre architectures), through to
‘details’ (i.e., these shapes, but now with holes, changes in fibre architecture,
or changes in thickness, the focus on this last being the effect of ply-drops).
From details we move to sub-components (a box section or another part with
structural and functional features) and eventually to components, e.g., an
entire wing.
Rouchon’s (1990) pyramid is exactly what is required in the context in which
it is used, but understanding how a manufactured component will behave
is dependent on a thorough understanding of the constituent materials.
Composites have properties that change as a result of being combined –and
so the first level of the pyramid should consider the constituent resins and
fibres (although information on these is usually provided by the manufac-
turer) and leading to their combination. Different processing routes should
also be considered at this stage, and the properties achievable that are
dependent on how the material is manufactured. From coupons we move
9
I.e., the directions of fibres with reference to principal stresses, the nature of the weave, if
woven, or other method of keeping the fibres together so that they can be handled, and a com-
ponent manufactured.
@seismicisolation
@seismicisolation
18 Understanding the Failure of Materials and Structures
10
Such us wind loading, as in the case of the Tay Bridge discussed in Chapter 1, traffic, the flow
of fluids of granulated goods through piping, expansion, and contraction due to heating,
and so on.
@seismicisolation
@seismicisolation
Why Break Things? 19
References
Diamond, J.M., 1998. Guns, germs, and steel: A short history of everybody for the last
13,000 years. Random House.
MacKenzie, D., Zoepf, S. and Heywood, J., 2014. Determinants of US passenger car
weight. International Journal of Vehicle Design, 65(1), pp.73–93.
OECD, 2010. Materials case study 1: Critical metals and mobile devices. OECD
Environmental Directorate.
Richter, D. and Krbetschek. M., 2015. The age of the Lower Paleolithic occupation at
Schöningen. Journal of Human Evolution, 89, pp.46–56.
Rouchon, J., 1990. Certification of large airplane composite structures. In ICAS Congress
Proceedings (Vol. 2, pp. 1439–1447). International Council of the Aeronautical
Sciences.
UNEP, 2009. Sustainable innovation and technology transfer industrial sector studies:
Recycling —From E-Waste to resources.
@seismicisolation
@seismicisolation
3
Materials Science 101
3.1 Introduction
There are innumerable textbooks on the subjects of Materials Science and
Engineering, as well as on specific aspects both in respect of materials and
on the techniques used by the practitioner to analyse them in various ways.
Some of these books are incredibly niche, whilst others take a broader view of
a particular topic. The purpose of this book is not to contribute to this myriad
with yet another introduction to the intricacies of Materials Science, but it is
important to understand some basic principles and how these affect the per-
formance of particular materials, and consequently the performance of an
engineering structure.
Take a group of light bulbs for instance, or a collection of water pipes, or
pretty much anything that is manufactured in a batch. The group is nomin-
ally identical, but some will fail very early on, the majority will perform well
for a certain amount of time, and a handful will continue to function for a
surprisingly long time after the rest of the cohort have failed. Some will last
so long, in fact, that they are taken out of service before failure, because it is
cheaper to replace a cohort in one go, rather than piece-meal. Why is it that
one item fails early, others last seemingly forever, and others fail at a point
in-between these two extremes? Can we finesse things to move the failure
curve? What causes these failures to occur in the first place?
Whilst it is possible to undertake a campaign of mechanical character-
isation without understanding anything about the materials that form the
structures being tested, it helps a lot to know something about the way that
atoms work together to make the materials we use.
At its simplest, Materials Science can be described as the study of the
behaviour of materials and how this behaviour is affected by the processing
route. Consider Figure 3.1, this is a visual representation of this definition,
@seismicisolation
@seismicisolation
20 DOI: 10.1201/9780367822347-3
Materials Science 101 21
FIGURE 3.1
The Materials Tetrahedron, demonstrating the relationship between microstructure, properties,
performance, processing, and the characterisation of materials.
1
The impact of the COVID-19 pandemic is still to be completely accounted for, but history
suggests that any downturn in production will be a blip in the long term. In 2024, it was
announced that a major UK steelworks would shut down production of primary steel (i.e.
using a blast furnace to reduce mined ore and produce steel and other byproducts), causing
significant national angst: what price a country that cannot produce its own steel? More
broadly, and looking to the longer term, the production of steel is a major producer of CO2,
@seismicisolation
@seismicisolation
22 Understanding the Failure of Materials and Structures
its simplest we can consider the carbon content: steels are, fundamentally, an
alloy of iron and carbon. The carbon content can be as low as 0.002 wt.% and
as high as 2.14 wt.% –more than this and we move out of the range of steels
and into the realm of cast irons. By varying alloying additions, processing
routes, and heat treatments, it is possible to produce mild steels, stainless
steels, and maraging steels, to name but three families. Most people would
recognise the very different performance in terms of corrosion that would be
exhibited by mild and stainless steels. What might be harder is determining
which of the fifty-seven types of stainless steel is best for a particular appli-
cation, not to mention ensuring that the microstructure is not compromised
when the material is formed into the structure or component.
In the following sections, we will consider the nature of materials, begin-
ning with a brief overview of the structure of the atom, how atoms bond
together and how, as we increase the numbers of atoms collected, we see the
creation of defects that will impact on performance.
and this will need to be addressed if Net Zero targets are to be met. Do we need so much pri-
mary steel? Some will say yes, of course we do if we are to grow the economy, but the answer
must be more nuanced if we wish to grow the economy sustainably. Better recycling practices,
changes in the way that we use resources, and other factors could reduce the demand for pri-
mary steel, although the full discussion of this would take a book.
2
The usual four are earth, air, fire, and water. Aether is sometimes added to the mix, and some
cultures replace one or more items on the list, or add to it, with wood and/or metal.
3
This is sometimes thought of in terms of the things such as electrons, protons, and neutrons,
and sometimes in terms of the quarks and other fundamental building blocks of matter.
@seismicisolation
@seismicisolation
Materials Science 101 23
FIGURE 3.2
The evolution of human understanding of the atom. (a) Dalton’s ball-bearing model (the atom
is a discrete body with no internal characteristics); (b) JJ Thompson’s plum pudding model
(the atom is a discrete body with no overall charge, but regions within the atom carry negative
charge, whilst the rest of the atom carries a counter-balancing positive charge); (c) Rutherford’s
model (negatively charged electrons orbit a positively charged nucleus; (d) Bohr’s model,
demonstrating that electrons can occupy different energy states; and (e) Schrodinger’s model
which shows that the ‘orbit’ of an electron is actually a volume in which there is a probability
of finding it.
The modern story of the atom begins around 1803 with John Dalton’s work
on elements, and he proposed what is, today, commonly called the ball-bearing
model: an atom is like a ball-bearing, indivisible, and atoms of different elem-
ents are ball-bearings of different sizes, densities and so on. We can mix them
all up and separate them out again and they do not really change. This model
is perfectly adequate, and gives us something with which we can work. In
some circumstances though, it is completely wrong and unhelpful –it does
not consider electrons, protons, and neutrons, for example. It took slightly
over one hundred years from Dalton’s work for these building blocks to be
recognised and understood, and then things changed rapidly. Between 1904
and 1926, JJ Thompson gave us the plum-pudding model (the atom is still
solid, but composed of volumes with positive and negative charges, like fruit
dispersed in a cake), Rutherford showed that the electrons were separate
and orbited a dense core, Bohr showed that the electrons had different ener-
gies, and that electrons could be shielded from the positive core, and finally
Schrödinger showed that the electrons might be anywhere, although it is pos-
sible to define a volume where they are likely to be.4 This last model is still
considered to be the most accurate but, in many circumstances, it is preferable
to stick with Dalton’s or Bohr’s models (Figure 3.2).
A great deal of research is still conducted on the fundamental behaviour of
protons, neutrons, and electrons, especially the interactions between protons
and neutrons, and the implications of the wave-particle duality of electrons,
including the presence of electrons in the nucleus of an atom.
Happily, for the current purpose we can proceed with a relatively simplistic
structure of the atom: the nucleus, the cluster of positively charged protons
4
It should be noted though, that the neutron was not discovered until 1932 (Chadwick, 1932).
@seismicisolation
@seismicisolation
24 Understanding the Failure of Materials and Structures
and neutral neutrons at the centre of the atom pulls electrons towards it but
because electrons can only come so close to each other, electrons are only able
to fill certain stable energy shells, called orbitals. The closest orbital to the
nucleus is the 1s orbital and can hold up to two electrons.
The simplest atom is therefore that of hydrogen, which is a single elec-
tron orbiting a single proton;5 as a consequence of its structure it is extremely
reactive, seeking to bond with other atoms and so become more stable.
Helium, the first of the ‘noble’ gases, is two protons, two neutrons, and two
electrons; therefore, its outer shell is complete, and atoms of helium are
almost completely unreactive because they are stable. As more protons and
electrons become involved, the orbitals encompass greater volumes, and
more electrons can be contained in each orbital. Energetic events can lead to
electrons being ejected from the atom. In terms of particle physics, this is as
far as we need to go, although it is worth noting that the energy required to
remove an electron can be measured and the element present inferred from
this characteristic. Further, the energy required to remove a particular elec-
tron is affected by its relationship to the atoms around it, and this can be used
as an investigative tool, see e.g., Watts and Wolstenholme (2019). Beyond this,
the nuances of electron orbitals can be ignored, except where they relate to
atomic bonding.
5
Deuterium and tritium are rare but naturally occurring forms of hydrogen which include one
and two neutrons, respectively. Tritium is radioactive. This becomes important when we start
to consider the impact of the operational environment on future performance, especially where
such atoms are present in large quantities over time. For example, whilst most of the difficul-
ties with electricity generated from nuclear fusion are associated with the physical reactions
required, generating the initial starting conditions, and preventing breakaway reactions, con-
siderable effort is also required in developing the materials that contain and control these
reactions, and extract energy from them. Hydrogen embrittlement is a well- understood
problem in many sectors, but there is still some way to go to be sure that there will not be
unexpected issues with deuterium-and tritium-doped hydrogen causing the embrittlement.
6
Which can be visualised as ‘hand-grasps’: if covalent bond are a firm hand-clasp, metallic
bonds could be considered as a grip further up the arm, whilst ionic bonding is, in a sense,
more like a ‘monkey-grip’ with only the fingers curled around, strong enough under some
circumstances but easier to break in others. The analogy breaks down of course when we think
of ionic solutions, where the ionically bonded atoms dissolve and are dispersed, albeit that
they have claimed the electron or the hole, as preferred.
@seismicisolation
@seismicisolation
Materials Science 101 25
FIGURE 3.3
Primary and secondary bonding between atoms. (a) Metallic bonding, with electrons shared
(and free to move) within the lattice formed; (b) covalent bonding, where two atoms complete
an electron pair by sharing electrons; (c) ionic bonding, where one atom gives up its outermost
electron and donates it to an atom with a hole in its orbital; and (d) an example of secondary
bonding based on water, where a slight positive charge (δ+) is put on the hydrogen atoms due
to the covalent bond with oxygen. This δ+ charge forms a weak bond with the electron pair of
another oxygen atom.7
7
More of a ‘pinkie-finger promise’.
@seismicisolation
@seismicisolation
26 Understanding the Failure of Materials and Structures
Ionic bonding: atoms with loosely bound outer electrons can enter a
state that is more stable by allowing that electron to become disassociated.
Similarly, some atoms have outer shells that are nearly, but not quite, full and
so are willing to accept electrons and therefore become more stable. Because
one loses a negatively charged electron, the resultant ion is positively charged,
and conversely the one that accepts the electron becomes negatively charged.
In a low energy state, these charged atoms, called ions, will form them-
selves up in nice, neat rows like metals, leading to the formation of crystals.
The most obvious example of this behaviour is table salt, sodium chloride,
NaCl. This behaviour is not dependent on equal numbers of atoms, so long
as charge parity is observed. So, for example, iron can form different oxides
because it can lose two (Fe II/Fe2+) or three (Fe III/Fe3+) electrons. Oxides
include FeO, FeO2, Fe2O3, and there are also forms that mix Fe II and Fe
III ions.
Covalent bonding: covalent bonding sees a sharing of electrons between
two atoms. The difference compared with ionic bonding is that the orbitals
overlap so that electrons are shared equitably. The usual example of this
is carbon, which happens to have an outer shell which could hold eight
electrons, but actually has four, and hence four spaces. It could give up those
four electrons, or it could accept four to fill the gaps, but there is no strong
driver to do either. Instead, an electron pair is shared between two atoms.
Two carbon atoms could, therefore, form up to four covalent bonds. However,
the greater the number of bonds, the greater the bond strain becomes, as the
electrons are all pulled around to one side. Imagine that you are standing
side-by-side with a friend, holding their hand. Now, without moving your
feet or body, bring the other arm around to hold their other hand. It is pos-
sible, but places significant strain on your back and the arm reaching across
your body.8
Secondary bonding, also called dipole bonding, arises from the movement
of electrons. In an incomplete shell, a dipole can be created when an electron
is shielded by the nucleus. Consider the Moon orbiting the Earth. The Moon
cannot be seen from all points at once, and if we were orbiting the Earth at
a point diametrically opposed to the Moon, we could not see it because the
8
In the case of water, this secondary bonding has an important part to play in the states that
water adopts, dependent on the overall energy in the system. Between 0oC and 100oC, the
secondary bond between water molecules undergoes a process of formation and reformation
which allows the molecules in a volume to move around relative to each other, but also gives
the volume some coherence. Above 100oC and the energy in the system becomes such that sec-
ondary bonds are too weak to keep water molecules in contact with each other and the water
becomes gaseous. Below 0oC and the energy in the system is so low that the water becomes
solid. During the formation of ice, the secondary bonds take on a more permanent character-
istic: the bond becomes more uniform throughout, so that the water molecules are locked in
place in a regular pattern. It is a direct consequence of the secondary bonding between water
molecules that leads to ice floating on water, rather than sinking, as would happen if the
solidified water were more densely packed.
@seismicisolation
@seismicisolation
Materials Science 101 27
Earth would be in the way. In the case of an atom, with its central positive
and orbiting negative charges, this means that there are times when there
is a greater concentration of negative charge, and of positive charge, which
can be ‘seen’ by other charged objects. In a similar manner to a magnet,
opposites attract. In a covalently bonded molecule, there are regions where
there is a permanent dipole, because the sharing of an electron pair means
that electrons are held in position.9 The usual example at this point is to
consider water, H2O. Oxygen can form two covalent bonds, hydrogen, one.
A molecule of water is therefore made up three atoms forming two covalent
bonds, giving rise to four permanent dipoles (Figure 3.3a), two negative
(on the other side of the oxygen atom from the bonds, arising from already
complete electron pairs), and two positive (one on each hydrogen atom).
Hence, the three states of water (solid ice, liquid water, and gaseous steam)
arise from the secondary bonding present, controlled by the level of energy
in the system. When there is sufficient energy, the molecules vibrate so rap-
idly that there is little opportunity for dipole–dipole interactions to occur,
and the water molecules move as far away as possible from each other, to
fill the entire volume available to them. By contrast, when there is very
little energy in the system, the molecules are unable to vibrate and move
about, and the water molecules organise so that the negative dipoles on
the side of the oxygen away from the hydrogen atoms attract the positive
dipoles formed by the unbalanced hydrogen atoms of another water mol-
ecule (Figure 3.3b)
9
The careful reader will have noted that this does not necessarily match with the previous
comments about the probabilities associated with where an electron is located. The even more
careful reader will realise that the level of complexity that this gives rise to is unnecessary in
the current context. Whilst the briefest of summaries, the arguments laid out above are the
convention for most basic chemistry and physics teaching and all that is required for this book.
@seismicisolation
@seismicisolation
28 Understanding the Failure of Materials and Structures
FIGURE 3.4
Examples of unit cells: (a) a unit cell is usually defined in an xyz cartographic system, where
each corner is a node which is a junction between eight unit cells; (b) the simplest arrangement
is the primitive cubic cell where there is only an atom at each node, which is shared between
eight unit cells –the overall number of atoms associated with each cell is 1; (c) the body-centred
cubic cell adds an atom in the centre of the cell –with the one atom associated with the eight
nodes, the atom in the centre means that there is a total of two atoms associated with each unit
cell; (d) rather than an atom in the centre of the cell, the face-centred cubic structure has an atom
in each face of the cell, i.e. each atom (not at a node) is shared between two cells, and hence there
are four atoms associated with each cell; and (e) the one exception to the cubic structure is the
hexagonal close-packed cell, where there are six atoms associated with the cell, with the three
in the central layer being completely within the cell, the two in the centre of the top and bottom
faces being shared between two cells, and the atoms at the corners being shared with six unit
cells, hence there are six atoms per unit cell.
the centre of the unit cell, in addition to the eighths of an atom present at each
corner. So a body-centred cubic cell has two atoms per cell. Instead of placing a
single cell in the centre of the cell we can place half an atom in the centre of each
face, giving a face-centred cubic cell, with four atoms associated with each cell.
There is one further kind of until cell, which is hexagonal rather than
cubic: hexagonal close-packed structures.
The full complexities of crystal structure and the terminology associated
with describing lattices, which occupies entire textbooks and is the subject
of entire courses within Materials Science degrees, cannot be summed up in
a few paragraphs. However, the basics presented here are sufficient to allow
us to understand that:
10
One of the complexities that we have not explored here is that these lattice structures are also
relevant to alloys, i.e. mixtures of different kinds of atoms, and molecules.
@seismicisolation
@seismicisolation
Materials Science 101 29
11
The tendency is to visualise basic crystal structures in two dimensions, as this makes it easier
to see the point of interest, but of course bulk structures will have multiple layers within the
crystal, which we begin to see in Figure 3.3.
12
And in some contexts, magnetic fields.
@seismicisolation
@seismicisolation
30 Understanding the Failure of Materials and Structures
FIGURE 3.5
Examples of defects observed in crystal structures: (a) a vacancy in the lattice; (b) an interstitial
insertion in the lattice; (c) a substitution in the lattice by an impurity; and (d) an interstitial
insertion of an impurity in the lattice.
13
Some definitions of microscale, or microstructure, suggest that features within the micro
structure require a minimum of 25× magnification to be observed. This is to say that if we
were to observe a human hair, of the order of 100 μm in diameter, under a microscope, then
anything that is observable only when the hair is magnified to 2500 μm (or 2.5 mm) in diam-
eter counts as microstructure.
@seismicisolation
@seismicisolation
Materials Science 101 31
However, the gross defects that we can observe at the macroscale can lead to
microstructural defects: large cracks can become atomically sharp and set the
stage for catastrophic fracture.
A familiar illustration of this hierarchy of structure is the tree, and the
useful timber that can be extracted. The whole tree is the macrostructure,
and we can extract parts as necessary for industrial use. We can observe the
grain, and there will be variation based on the type of tree felled, its age, loca-
tion, and the history of the environment during the time it was growing. At
the microstructural scale we can look at individual plant cells and the way
that molecules14 interact to form different parts of the cell wall. However, in
between there is a rich world of detail, incorporating the formation of annual
growth rings, the living sap wood, the heart wood, which is essentially a
repository of unwanted chemicals, radial cells, and so on. Depending on
what you are looking for, the mesostructure can be the most interesting part
to look at.
3.7 Summary
Materials and structures fail for a variety of reasons, and in the event of a dis-
aster, there are likely to be a number of events that coincide: poor materials
selection, poor condition monitoring, and human error can turn a mistake
in production into a significant loss of life. However, by understanding
materials behaviour, by making good choices, by testing materials beyond
their normal performance, we can mitigate, to some extent, human error.
Beyond the incidence of catastrophic failures, by understanding the nature of
materials, we can make an assessment as to the feasibility of extending oper-
ational lifetimes. By looking at the environmental conditions seen during the
service lifetime, by reviewing the loading we might expect to see we can pre-
dict the future performance of the asset when it approaches the end of its
design life, or, as is increasingly common, the asset is exposed to forces that
were predicted to occur only once in a hundred or a thousand years. Key to
this is understanding how materials behave.
This chapter has not considered the nuances of electrochemistry, for
example, and the impact that galvanic corrosion can have on the lifetime
of a component when poor material choices are made. This chapter has not
considered the impact of magnetic or electrical fields. When getting into the
detail of operational life these are certainly factors that need to be considered,
14
Technically macromolecules, because as molecules go, they are very long. This is one of those
situations in science where a linguistic toolkit based on certain well-understood principles
can become confusing when different contexts at different length scales begin to overlap. If
we are operating at the macroscale, we cannot observe macromolecules.
@seismicisolation
@seismicisolation
32 Understanding the Failure of Materials and Structures
References
Chadwick, J., 1932. Possible existence of a neutron. Nature, 129(3252), pp.312–312.
Cottrell, A., 1997. An introduction to metallurgy, Second Edition. Edward Arnold.
Hayward, R.C., Alberius-Henning, P., Chmelka, B.F. and Stucky, G.D., 2001. The
current role of mesostructures in composite materials and device fabrication.
Microporous and Mesoporous Materials, 44, pp.619–624.
Hull, D. and Bacon, D.J., 2011. Introduction to dislocations, Fifth Edition. Elsevier.
Jesson, D.A. and Watts, J.F., 2012. The interface and interphase in polymer matrix
composites: effect on mechanical properties and methods for identification.
Polymer Reviews, 52(3), pp.321–354.
Watts, J.F. and Wolstenholme, J., 2019. An introduction to surface analysis by XPS and
AES, Second Edition, John Wiley & Sons.
Xu, F.J., Neoh, K.G. and Kang, E.T., 2009. Bioactive surfaces and biomaterials via atom
transfer radical polymerization. Progress in Polymer Science, 34(8), pp.719–761.
Yurgartis, S.W., 1995. Techniques for the quantification of composite mesostructure.
Composites Science and Technology, 53(2), pp.145–154.
Recommended Reading
Now in its fourth edition, Ashby and Jones is indispensable for anyone looking to
ground themselves in the basics of materials and materials behaviour. Part 1
address properties applications and design, whilst part 2 considers microstruc-
ture and processing.
Ashby, M.F. and Jones, D.R., 2012. Engineering materials 1: An introduction to properties,
applications and design, 4th Ed., Butterworth-Heinemann (Elsevier).
Jones, D.R. and Ashby, M.F., 2012. Engineering materials 2: An introduction to
microstructures and processing, 4th Ed., Butterworth-Heinemann (Elsevier).
Another excellent undergraduate textbook on Materials Science and Engineering
is Callister; modern editions come with useful digital tools, but even an older
edition will stand the reader in good stead.
Callister Jr, W.D. and Rethwisch, D.G., 2020. Callister’s materials science and engineering.
10th Ed. John Wiley & Sons.
For those who feel that they need to expand their knowledge of physical chemistry
in support of the basics presented in this chapter, Atkins’ Physical Chemistry is
a standard textbook on many courses, although the lighter Elements of Physical
Chemistry may be preferable.
@seismicisolation
@seismicisolation
Materials Science 101 33
Atkins, P. and De Paula, J., 2013. Elements of physical chemistry. Oxford University
Press, USA.
Atkins, P.W., De Paula, J. and Keeler, J., 2023. Atkins’ physical chemistry. Oxford
University Press.
In a footnote, I raised the issues of the continuing trajectory of the growth of the steel
industry. If there is a desire to pursue this line of thinking further, then two
texts that are worth looking at are Julian Allwood’s Sustainable Materials and Tim
Jackson’s Prosperity without growth.
Allwood, J.M., Cullen, J.M., Carruth, M.A., Cooper, D.R., McBrien, M., Milford, R.L.,
Moynihan, M.C. and Patel, A.C., 2012. Sustainable materials: With both eyes open.
Cambridge, UK: UIT Cambridge Limited.15
Jackson, T., 2016. Prosperity without growth: Foundations for the economy of tomorrow.
Taylor & Francis.16
15
Republished in 2015 by Bloomsbury as Sustainable materials without the hot air: making
buildings, vehicles and products efficiently and with less new material.
16
This is the second edition, revised and updated from Prosperity without growth: Economics for
a finite planet published by Routledge in 2009.
@seismicisolation
@seismicisolation
4
Defects and Degradation
4.1 Introduction
In Chapter 3, the concept of defects within crystal structures was introduced.
These are defects writ small, at the level of the misplacement of atoms, but
which in quantity and repetition can have a significant effect on the behav-
iour of the material and hence the structures into which they are made.
The concept of different levels of structure was also introduced, with this
interest in the placement of atoms being at the microstructural (or occasion-
ally the nano-structural level). In this chapter we will explore the defects that
are found at the macrostructural level. These defects can be split into three
categories:
But what then is degradation, and what is the difference between this and a
defect? In the context of the last point on the list above, we could use deg-
radation as a special case, although defects that arise throughout service can
be temporally discrete (such as an impact leading to plastic deformation or
cracking), or temporally continuous, such as the cracking engendered through
the freeze–thaw cycle of water entering a non-critical defect, becoming ice
and causing the defect to grow: this process may occur hundreds or even
thousands of times over the course of years, decades, or centuries before the
defect becomes a critical one.
@seismicisolation
@seismicisolation
34 DOI: 10.1201/9780367822347-4
Defects and Degradation 35
The issue is exacerbated by the fact that whilst ‘defect’ can only be a noun,
degradation can be both a noun and a verb. In the context of this book, i.e.,
the failure of materials and structures we can use the following definitions:
Defect, noun, a feature of the structure of a material that is different from the
general configuration, and which will give rise to localised mechanical prop-
erties that are different from the bulk. Defects may be critical or sub-critical,
where a critical defect is defined as one which will cause failure if the loading
on the part or structure reaches the yield or ultimate strength of the material
in the location of the defect. A sub-critical defect is one which is small enough
that it will not cause failure during normal operation, but which has the cap-
acity to grow over time, e.g., because of fatigue, and become critical at some
point in the future.
Degradation, verb, any of a number of processes including but not limited to
corrosion and wear, which will eventually lead to the failure1 of the material,
component, or structure. The effects of degradation can be observed, in some
circumstances, before failure occurs, although it is not usually possible to
quantify these effects by eye.2 Specialist equipment3 will need to be used to,
e.g., measure the depth of degradation, or its impact on electrical resistance
or conductivity, for example.
Degradation, noun, the impact of a process of degradation on a component
or structure.
Or, to put it another way, not all defects are the result of degradation, and
not all degradation leads to a defect that we would see as a problem in the
context of, e.g., fracture mechanics (see Chapter 6). There is, however, cer
tainly some overlap (Figure 4.1).
The first step to understanding the impact of a defect is to understand how
it comes about. This has been explored in Chapter 3, at least in as much as it
relates to the microstructure. The next section considers macro-scale defects
and their origin, and afterwards some examples are presented of how to
eliminate defects, or how to live with them.
1
Recalling that failure need not be catastrophic, but simply that a part is no longer able to per
form its function, or even that it no longer meets an operational standard. For example, if the
standard calls for an unbroken layer of paint on the surface, and the paint becomes scratched
or worn through, it has failed, even if the component it is on is still able to function normally.
2
Some may argue that you can look at a surface and determine that it is worn. This may be
true, to an extent, but comparing two items from the same batch and determining which is the
more worn is trickier and determining which (either? both? neither?) is critically worn when
compared with a standard is trickiest of all. Even experienced practitioners struggle to rank
defects across a surface. And this is just dealing with surface defects and degradation –what
about those on hidden surfaces, or those within the bulk volume of the component of interest?
3
Occasionally, specialist equipment can actually be very basic, although it tends to be more
destructive in the long run: if you can break a piece off something with either your bare hands,
or brute force applied through a chisel or even a broom handle, you can get a feel for how
badly degraded something is. If you are at that point though, it is probably too late.
@seismicisolation
@seismicisolation
36 Understanding the Failure of Materials and Structures
FIGURE 4.1
Venn diagram demonstrating overlap of defects and degradation: the terms are connected but
are not synonyms.
@seismicisolation
@seismicisolation
Defects and Degradation 37
placed vertically and a sacrificial portion which contained all the potential
defects could be removed.
Prior to the introduction of materials standards, and indeed prior to a full
understanding of the nature of iron, cast iron, and steel, and the introduction
of mass production, it was not unusual for small foundries to pop up where
they were needed, move on when they were not, and for production methods
to be controlled by instructions passed on by word of mouth. Hence, control
of temperature throughout the pour and production of the components could
be variable, and the additions to the melt to control various aspects would be
done by eye, or judgement.4
The properties of cast iron are controlled by the graphite flake struc-
ture (Angus, 1976), and so it is not surprising that variable properties were
observed even after taking the variable defect population into account.
Manufacturing defects are sometimes referred to as ab initio defects, liter-
ally ‘from the beginning’. Hence, ab initio defects are those ‘baked in’ during
manufacture; these include, but are not limited to:
• microstructural defects;5
• porosity;
• slag inclusions (from the source material);
• impurities and contamination from the manufacturing site and improp-
erly maintained equipment;
• cold shuts (where two fronts of molten iron fail to fuse properly due to
premature cooling); and
• geometrical issues arising from improper alignment of moulds,
Whilst we have been using cast iron as an example, it can be understood that
these issues can arise in any poorly conducted manufacturing endeavour,
with any material which is not treated appropriately. In practice, these sorts
of defects are much rarer, and the quality assurance processes are more
robust, ensuring that items with such defects are now much more seldom
encountered. However, where small items are produced in large quantities,
such as screws and other fastenings, it can be interesting to look at items that
fail when they are first used in an attempt to identify a defect that has led to
failure.
4
Recycling was surprisingly common too, and at one time it was not unusual to find unmelted
bits and pieces of prior incarnations embedded in a casting, including unburned lumps
of wood.
5
Cast iron is probably the poster child for this issue. Carbon was added to iron to help reduce
the temperature at which iron melts and hence make casting easier. However, the flake struc-
ture forms when the iron cools and the carbon is rejected from the melt. The flakes act as crack
initiators, significantly affecting the tensile strength of cast iron.
@seismicisolation
@seismicisolation
38 Understanding the Failure of Materials and Structures
1. The cracks would need to be a very precise length not to cause imme-
diate failure, and to grow under fatigue cycling over the course of the
time that we know such pipes have spent in service.
2. Cracks in the bell, and to some extent the spigot, are less of an issue
than other defects.
3. The loads required to grow such cracks would have to be very pre-
cise –any less and nothing would happen, any more and other defects
would be more of an issue.
Other defects arise from the use of chains for pipe handling rather than
canvas slings: this is known to have damaged the protective outer coal tar or
bitumen coating, leaving areas of bare metal, which would then be prone to
localised corrosion.
Once the pipes were in situ, the joint would be packed with hemp to form
a seal, filled with molten lead and finally caulked by hammering the solidi-
fied lead into the joint (Rajani and Abdel-Akher, 2013). If the caulking were
carried out incorrectly, the lead would be hammered too far into the joint,
leading to an increase in stress around the joint, potentially causing a crack
to initiate. A worst-case scenario would see poor caulking practice expand a
crack created during the siting of the pipe. This could result in a structurally
flawed pipe with a reduced service life.
Generalising, we can see that there are a range of ways in which defects
can be introduced after manufacture but before use. Many components
are left stockpiled for a time, at the point of production, at storage facilities
6
In much the same way as a bell. A similar method was used for checking the integrity of the
wheels of steam locomotives in service, a fact that provided a vein of humour for observers
and comedians in late 19th and early 20th centuries.
@seismicisolation
@seismicisolation
Defects and Degradation 39
7
An example from personal experience would be the carpenter fitting the handrail on some
new stairs and drilling through a water pipe. A very visible and obvious failure, but it can
be imagined that a glancing blow or damage from a drill that fails to penetrate the wall of
the pipe could sit quietly for some time until an inciting incident caused a more problematic
failure. Water escaping from such a pipe could easily cause more damage and degradation,
leading to bigger problems with the building.
@seismicisolation
@seismicisolation
40 Understanding the Failure of Materials and Structures
applied to the part, and the final design will encompass not only the general
operational requirements, but the loads that may be placed on the compo-
nent in a predictable emergency,8 and some level of safety factor. However,
the capacity for a structure to carry load can be degraded by biological,
chemical, or mechanical action. In some instances, the loss of capacity can
be calculated simply by determining a ratio of the thickness of the material
affected with the original thickness, as a so-called loss-of-section approach.
However, where the degradation leads to cracking or other penetrating
damage such as corrosion pits, even very little damage can lead to cata-
strophic failure, as the damage leads to stress concentration. Jesson et al.
(2013) demonstrate the different effects that the same basic corrosion mech
anism can have on a material when the corrosion is generalised (loss-of-
section) or more pit-like (fracture mechanics). As an aside, but in keeping
with the concept that understanding failure can lead to better materials
modelling, and therefore more focused mechanical test programs, Fahimi
et al. (2016) and Ugoh et al. (2019) were able to use this understanding to
provide predictive models.
Of course, life is never entirely straightforward. Having developed a nice,
neat classification for the description of different kinds of defects, we imme-
diately come to those that don’t fit nicely into this system. For example, if
we look at barely visible impact damage (BVID9), this is something that is
usually associated with the operational phase. In some respects, it looks a
lot like some of the issues discussed in the section on installation defects.
BVID is usually characterised as being low energy, of the order of 15 J
and below, and is often illustrated as being the situation where someone
effecting repairs on a structure drops a tool which hits the surface, pos-
sibly, but not necessarily, leaving a mark or very small dent, but otherwise
apparently no damage. The important damage lies below the surface. In the
context of composite materials, the key issue is the cracking that can occur
in the matrix. Whilst it begins by being sub-critical, this damage can grow
by fatigue once the component returns to service. It is not really a service-
induced defect as described at the beginning of this section, but neither is it
an installation defect.
8
Predictable in the sense that it is possible that it will occur, not necessarily that it is probable.
For example, the chances of a car driving into a static object such as a telephone pole, or a
building, are on the whole unlikely, but not, as we know from the news, impossible. Under
normal circumstances, if the object has been built or sited correctly, the event should not occur,
but factors such as the driver speeding, or a previous spill on the road, or an animal leaping
out at an inopportune moment can lead to a crash.
9
BVID is usually considered in the context of composite materials, but there is potential for it to
occur in other contexts as well.
@seismicisolation
@seismicisolation
Defects and Degradation 41
@seismicisolation
@seismicisolation
42 Understanding the Failure of Materials and Structures
FIGURE 4.2
Idealised grain structures within a crystal: (a) the crystal is made up of a number of grains with
various orientations arising from independent initiation of grains within the crystal; (b) grains
are aligned in a specified direction although multiple grains with different orientations are still
present, each grain runs the length of the component; and (c) the crystal is formed from a single
grain derived by control of the initiation and/or elimination of competing orientations.
FIGURE 4.3
Illustrative example of the creation of a component formed from a single crystal.
on whether the crystal is grown from solid, melt, solution, or vapor, with
different ‘flavours’ arising from processing temperatures and other variations
in the overall process. Synthetic production of various mineral structures has
been of interest since around 2500 bce, beginning with the purification of
salt,10 continuing through the early understanding of materials and attempts
to recreate what was found in nature and arriving, eventually, at the need
for high-grade electronics and optical components. Today, the largest producer
and consumer of high- grade, defect-free, large-
scale single crystals is
the semiconductor market.
4.3.3 Fibreisation
A fundamental aspect of the production of composite materials, at least those
based on fibres, is that by turning a three-dimensional volume into a 2.5-
dimensional one, critical defects can be eliminated. Consider an expanse
of glass (Figure 4.4a): for the purposes of this example, we can ignore the
thickness. The glass is placed into omni- directional stress, i.e., an equal
amount of the tensile stress in every direction (Figure 4.4b). Based on our
calculations, this should be sub-critical, yet the glass shatters. Perhaps an
issue of the stress state? This being a thought experiment, we can use the
same sheet of glass again. Let us make the stress-state simpler, still tensile,
still apparently sub-critical, but only applied in a single direction. The glass
breaks. We have nothing but time, and our thoughts, so we repeat this situ-
ation, testing the plate by applying tensile stress in different directions. Every
time, the sheet of glass breaks at an unexpectedly low load (Figure 4.4c).
The problem is that the glass is full of defects that act as stress concentrators.
These defects are randomly dispersed within the sheet of glass, vary in size,
10
Recalling of course that salt was an essential preservative and flavour enhancer, and of course
that the word salary derives from the daily salt ration that was given to Roman soldiers.
@seismicisolation
@seismicisolation
44 Understanding the Failure of Materials and Structures
FIGURE 4.4
The mechanical properties of a piece of glass –why does it keep failing? (a) A typical piece of
glass, (b) load the glass equally in all directions, (c) simplify the load case to a single tensile
load, (d) we become aware of a population of microstructural defects preventing the glass from
achieving its full potential, and (e) we reduce the impact of the defects by cutting the glass
into smaller pieces, eliminating some defects completely and reducing the size below critical of
others.
shape, orientation (Figure 4.4d), and can arise from various causes including
dust, porosity, uneven cooling, and the like. However, if we can eliminate or
at least mitigate the presence of the defects, we will have a stronger piece of
glass. If we cut the sheet into smaller pieces, we can play a sort of Russian rou-
lette, by reducing the chance of a critical defect –although we may introduce
more defects in the cutting process (Figure 4.4e). Make the pieces smaller still
and we stand a good chance of eliminating or isolating more critical defects.
However, there is then the danger that we make the pieces so small that they
are no longer useful.
By casting fibres from the melt, we create long, thin pieces of glass, which
are too small to contain defects that are problematic should the glass fibre
be placed in tension. By bundling these fibres together, we can create some-
thing that approaches the maximum strength of glass. For those interested
in composite materials, we then need to consider the manner in which we
align the fibres in order to deal with complex stress states –if our fibres are all
aligned in a single direction, then the overall material will be strong parallel
to the fibres but incredibly weak perpendicular, having only the strength of
the resin (adhesive) that is used to bond the fibres together –the matrix in a
fibre-reinforced polymer matrix composite.
So far, we have considered only glass fibres, and by extension some of the
other mineral fibres derived in a similar way, such as basaltic fibres; how-
ever, a similar approach can be taken with carbon fibres, and those formed
of polymers and other macromolecules. For example, spider silk is known
to be incredibly strong, and we can therefore intuit that at least part of the
performance of the silk comes down to the size of the fibre, typically 2.5–4
μm in diameter. Currently, commercial attempts at wet-spinning various syn-
thetically produced proteins into silks can ‘only’ achieve diameters of around
10–60 μm. These are giving rise to a range of properties which are comparable
with naturally produced spider silks, but it can be understood from the dis-
cussion above that any attempt to produce a sheet of material will give rise to
the generation of defects that will impact on the performance of the material.
@seismicisolation
@seismicisolation
Defects and Degradation 45
4.4 Summary
Defects can be problematic, but they can also have their advantages.
Fundamentally though, by understanding that defects:
1. exist,
2. can be accounted for,
@seismicisolation
@seismicisolation
46 Understanding the Failure of Materials and Structures
References
Angus, H.T., 1976. Cast iron: Physical and engineering properties. Butterworth
London, UK.
ASTM A247-19, 2019. Standard test method for evaluating the microstructure of graphite in
iron castings. ASTM.
Fahimi, A., Evans, T.S., Farrow, J., Jesson, D.A., Mulheron, M.J. and Smith, P.A., 2016.
On the residual strength of aging cast iron trunk mains: Physically-based models
for asset failure. Materials Science and Engineering: A, 663, pp.204–212.
Jesson, D.A., Mohebbi, H., Farrow, J., Mulheron, M.J. and Smith, P.A., 2013. On the
condition assessment of cast iron trunk main: The effect of microstructure and
in-service graphitization on mechanical properties in flexure. Materials Science
and Engineering: A, 576, pp.192–201.
Logan, R., Mulheron, M.J., Jesson, D.A., Smith, P.A., Evans, T.S., Clay-Michael, N. and
Whiter, J.T., 2014a. Graphitic corrosion of a cast iron trunk main: implications for
asset management. WIT Transactions on The Built Environment, 139.
Logan, R., Mulheron, M.J. and Jesson, D.A., 2014b. Observations on the graphitic corrosion
of cast iron trunk main: Mechanisms and implications, Eurocorr 2014: European
Corrosion Congress Pisa, Italy.
Rajani, B. and Abdel-Akher, A., 2013. Performance of cast-iron-pipe bell-spigot joints
subjected to overburden pressure and ground movement. Journal of Pipeline
Systems Engineering and Practice, 4(2), pp.98–114.
Rajani, B. and Kleiner, Y., 2010. Fatigue failure of large diameter cast iron mains. In
Water Distribution Systems Analysis, pp.1146–1159. American Society of Civil
Engineers.
Ugoh, G., Cunningham, R., Farrow, J., Mulheron, M.J. and Jesson, D.A., 2019. On
the residual strength of ageing cast iron wastewater assets: Models for failure.
Materials Science and Engineering: A, 768, pp.138221.
@seismicisolation
@seismicisolation
5
Mechanical Properties I:
Strain, Stress, Stiffness
5.1 Introduction
To begin, it is important to understand that whatever the material, there are
two types of deformation that can occur. The first is elastic deformation and
the other is plastic1 deformation. Elastic behaviour is recoverable, plastic
behaviour is not. At this point it is useful to note two classes of material
which do not conform to the normal rules that will be explored in this
chapter. The first are auxetic materials, which will be discussed more fully
when we explore the concept of Poisson’s contractions, and the second is
shape-memory materials, which are able to recover from plastic deformation,
typically on the application of heat, without the need to melt or otherwise
turn the material back to its constituents and begin again.
Fundamental to an understanding of deformation, and ultimately the
failure of materials and structures, is the trifecta of strain, ɛ, stress, σ, and
Young’s modulus, E. In the case of a material which deforms elastically, the
three are linked by equation (5.1):
σ =Eɛ (5.1)
Consider a rubber band. These are notable for their ability to stretch and then
return to their original shape and size –elastic behaviour. However, it is pos-
sible to over-stretch them, without rupturing them completely. They become
noticeably flaccid and are unable to return to their original shape and size –
they have been plastically deformed. How much force was required to cause
1
On that basis, it is preferable to avoid talking about plastic materials when one means
polymeric.
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-5 47
48 Understanding the Failure of Materials and Structures
5.2 Strain
In a non-engineering setting, strain and stress are often used interchangeably
to denote a mental state.2 In an engineering context, the two are linked, but
have distinct meanings.
Strain is a measure of how much a material or structure (say the rubber
band from Section 5.1) will stretch when placed under load. More specific
ally, strain is the ratio of how much the band stretched, divided by its ori-
ginal length. Typically, the change in length is labelled Δl.3 By dividing the
change in length by the original length (Δl/l) we get the strain, which is usu-
ally given the symbol epsilon (ε):
ε =Δl/l (5.2)
So, if the rubber band starts with a length of 10 cm and, as we apply load, it
stretches by 20 cm then the strain is 20 cm/10 cm, or 2 ε, or 200%ε. From this,
it can be seen that strain is dimensionless. Now compare that to a 10 cm long
bit of string, it might stretch by just 1 mm for the same applied load, so its
strain is just 0.1 cm/10 cm, or 0.01 ε or 1%ε. Table 5.1 provides some examples
of equivalence between strain, microstrain, and percent strain.
It is worth noting that there are other measures or definitions of strain.
The strain defined above is also referred to as ‘engineering strain’, or some-
times the Cauchy strain. An alternative to this is to define what is called ‘true
strain’. Where the engineering strain considers only the total extension at a
given point with reference to the original length of the specimen tested, in the
case of true strain, the measurement takes into account the change in length
occurring as the material is loaded, or, to put it another way, the sum of incre-
mental strains (δεt), which represent the strain at a given moment, where that
incremental strain is defined as the incremental change in length (δl) divided
by the length prior to the application of that increment:
2
There is perhaps a nuance that stress sees someone under emotional pressure, whilst strain
suggests that emotions are being pulled taught, that a person is being stretched by their
commitments.
3
In physics, the Greek letter delta (Δ) is generally used to denote the change in something. C.f.
Δv, the change in velocity due to an accelerating force.
@seismicisolation
@seismicisolation
Mechanical Properties I: Strain, Stress, Stiffness 49
TABLE 5.1
Equivalency in comparative measurements of strain
ε με %ε
1 1 × 106 100
0.1 1 × 105 10
0.01 1 × 104 1
0.001 1 × 103 0.1
0.0001 1 × 102 0.001
0.00001 1 × 101 0.0001
δl
εt = ∫ (5.3)
L
5.3 Stress
At its simplest, stress is a measure of the overall force, F (or sometimes, P, for
pressure) applied to a material, normalised by the area, A, over which it is
applied:
σ =F/A (5.4)
4
It is worth noting that 1 MPa is equal to 1 N.mm–2.
@seismicisolation
@seismicisolation
50 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
Mechanical Properties I: Strain, Stress, Stiffness 51
FIGURE 5.1
Examples of stress–strain behaviour: (a) a strong material with a linear relationship between
load and extension but which resists deformation, and with a brittle failure mechanism, such
as a ceramic; (b) a strong material, again with a linear relationship between load and extension,
but which shows some elongation with load, and which yields, with further deformation
until failure, such as a metal; (c) a material with a non-linear component to the load–extension
relationship, which again shows some yielding behaviour, typical of some polymers; and (d) a
material which yields early, but continues to deform in a uniform manner for some time before
undergoing further yielding and failure, typical of a foam in compression.
• There will be some level of noise in the output. This will be exacerbated
when the test is in the very early stages, and it is important to ensure
that a load cell of the correct capacity is selected for the test. Collection of
strain data: material type, test geometry, and environmental conditions
all need to be considered when selecting strain gauges.
• There are a number of microstructural features that impact on the
overall behaviour. For example, Figure 5.1 does not illustrate the kind
of stick–slip behaviour that is seen in some steels, whereby immedi-
ately post-yield a saw-tooth effect is seen with a small increase in stress
followed by a small decrease in stress. Sometimes referred to as dis-
continuous yielding or plastic flow, this is a function of internal crystal
restructuring occurring at a nominally constant level of loading, usually
slightly lower than the yield stress, but with some elongation occurring,
until the restructuring has concluded, and the material enters the more
familiar plateau region.
• Non-linear elasticity is a frequent source of surprise when it is seen for
the first time.
@seismicisolation
@seismicisolation
52 Understanding the Failure of Materials and Structures
1. When removing the specimen, look for witness marks from the test
protocol. These come in different forms, depending on the test geom-
etry, the material being tested, and the method of holding it secure;
however, if the sample has moved during testing, this will be evident
in the witness marks, which will become elongated during the test, if
the specimen moves.
2. When securing the specimen, look at the fracture surfaces and see if
they look as you might expect for the kind of failure you are expecting
(ductile or brittle), whether there are any obvious features that show
crack growth, or if there are any gross defects.
3. When storing the specimens, whether or not you are planning to
study the sample further at this point, act as if you were and ensure
that the fracture surface is protected. There is a tendency to want to
put the fracture surface back together –do not do this, as you will
damage the fracture surface further such that it can provide no further
information. If it is likely that you will want to undertake chemical
analysis of the material after testing, especially of the fracture sur-
face, then wrap the sample in baking foil. Plastic specimen bags will
tend to be contaminated with PDMS (polydimethylsiloxane, a release
agent). PDMS can cause problems with chemical analysis, as its preva-
lence and strong chemical signature can result in even small quantities
masking chemical features of interest.
If there is any concern with the specimen, specifically the manner of its
testing and failure, then this can be addressed before moving onto the next
specimen. At this point, before abandoning a test schedule a mental review
of the actions so far can go a long way to understanding the results. Have
you in fact tested the right specimen? Was there a specific orientation that
needed to be tested? Were there any safety or control features that need to be
addressed?
It is also worth consulting other practitioners to see if they have any
insight. But a fundamental rule, especially when it comes to research, is that
just because it does not look like a textbook response, it doesn’t mean it’s
wrong: most textbooks deal in the abstract, in the general and hence examples
will tend to be representative, not specific.
@seismicisolation
@seismicisolation
Mechanical Properties I: Strain, Stress, Stiffness 53
5.4 Stiffness
In the 17th century, Robert Hooke stated “ut tensio, sic vis”, which means
“as the extension, so the force”. What he observed was that elastic materials
followed a law by which the extension was proportional to the force applied.
Different materials extended by different amounts for a given applied force,
but whatever the extension of an elastic material, it will double if you double
the applied force, triple if you apply three units of force and so on. Of course,
there is a limit, where things go PING, which we will come onto in a moment.
Hooke was a renowned polymath with a range of interests, including the
microscopic world and his architectural ambitions for a London being rebuilt
after the Great Fire of London, not to mention his long-running and ultim-
ately career-ruining feud with Sir Isaac Newton. This perhaps explains why
he didn’t take this to its logical conclusion, i.e., that there is a relationship
between stress and strain, and that when an elastic material is extended in
this way, the resistance to extension, as defined by this relationship between
stress and strain can be defined as the stiffness. This relationship was first
defined by Leonard Euler, in the early 18th century, but, for reasons which
do not concern the present text, we talk about the Young’s modulus,5 E, of a
material: E =σ/ε. The Pyramid of Tests sets out a hierarchy of testing which
starts with materials at the bottom and finishes with structures at the top.6
Humans have been building things in a serious way for at least 5000 years –if
you start with megalithic stone structures. In a mere six hundred or so years
we go from ‘simple’ structures, with one stone lintel placed on top of two
uprights, to the pyramids of Egypt, and on to the large-scale structures of
ancient China and Mesoamerica. These were built with the materials avail-
able and were designed on the basis of tried and tested methods: empirical
design. Archaeologists are still trying to determine how some effects were
achieved with what is known of the tools available at the time. Empirical
methods were still employed during Europe’s Medieval period when many
of the great cathedrals and castles were built, sometimes over the course of
hundreds of years.
But it is only within the last hundred or so years that we have moved
towards predictive design, based on models, and it is only in the last fifty or
5
As in Thomas Young, another polymath, who in addition to many other feats helped to trans
late Egyptian Hieroglyphs using the Rosetta Stone. His paper on the subject of E dates to the
very early 19th century.
6
Jerry Lord of Boeing has suggested that there should be a second pyramid that should be
considered at the same time. Whilst not entirely relevant to the current context, it is worth
noting for completeness. This second pyramid considers the validation of the manufacturing
process.
@seismicisolation
@seismicisolation
54 Understanding the Failure of Materials and Structures
so years that this design process has been dependent on the materials avail-
able, especially those of a synthetic nature.
ε transverse
ν =− (5.6)
ε axial
FIGURE 5.2
Poisson’s ‘reactions’: (a) in tension, we would normally expect to see the sample elongate in the
direction of load, whilst the transverse direction contracts; (b) in compression we observe the
opposite effect in that the specimen reduces in length in the direction of the applied load and
the transverse width increases; and (c) in an auxetic material in tension, all dimensions increase.
@seismicisolation
@seismicisolation
Mechanical Properties I: Strain, Stress, Stiffness 55
The reverse is also true: if we place the material under compression, the
material will expand in the perpendicular direction(s).7 When it comes to a
physical test, in the case of a compressive test, it is important to consider the
level of friction between the sample and the platens constraining the spe-
cimen. If the friction is too great, the material will bulge, in a form called
barrelling (Figure 5.3). If the friction is too small, then there is a danger that
the sample will slip out from between the platens.
Typically, the Poisson’s ratio of a material lies somewhere between 0.3 and
0.5, but some materials, usually engineering foams and the like, can have
a ratio approaching 0, which is to say that there is very little in the way of
Poisson’s reaction.8 In terms of extreme behaviours, this brings us to the
case of auxetic materials. These unusual materials actually increase in all
directions when placed in tension (Figure 5.3c), or contract in all directions
when compressed in one, because of the internal structure. Here again is
an example of a mesostructure, where internal structuring of the material
at a level between the micro-and macrostructures is such that instead of
the normal process of deformation, the material opens (or closes) with the
tensile (or compressive) load.9 For a more in-depth consideration of auxetic
materials, see Lim (2015).
5.6 Summary
The relationship between stress, strain, and Young’s modulus is a funda-
mental one. It is important to remember that at different times, a given design
may be stress limited or it may be strain limited. That is to say, the design may
be required to carry a particular load, which will require the design to have a
certain cross-section. If the cross-section is greater than can be accommodated
7
To avoid confusion between orthogonal directions, here the terms axial, i.e., parallel, and trans
verse, i.e., perpendicular, are used, with reference to the loading direction. (Some conventions
would align the x direction with the direction of loading, others the z direction.) The default
convention is also to consider Poisson’s contractions as a two-dimensional problem, with
the through thickness direction effectively ignored. This is not generally a problem. Where
a flat specimen is considered, the relative through thickness contraction will be compara-
tively small, and where the material is isotropic, at least in the directions perpendicular to the
applied load, then the Poisson’s ratio will be the same in both transverse directions. Things
become more complex when there is a question of anisotropy to account for, i.e., there will be
a different Poisson’s ratio for each transverse direction.
8
This is particularly the case with foams placed in compression, where the cellular structure
tends to collapse in on itself rather than distorting in other ways. See also Section 4.6.
9
Whilst there are some materials that display auxetic behaviour, the majority of examples are
not really materials, in a sense, but are structures, albeit at a small scale. In section, they are
much like a foam, with volumes of open space. The space is bounded by material, which is not
necessarily special of itself, but rather by the way it is shaped.
@seismicisolation
@seismicisolation
56 Understanding the Failure of Materials and Structures
FIGURE 5.3
The issue of barrelling. When a sample is placed in compression (a), instead of compressing
uniformly throughout the volume, the sides bulge (b). This arises due to friction between the
platen and the sample, which prevents the sample from accommodating the change in shape
arising from the applied load.
References
Goodno, B.J. and Gere, J.M., 2020. Mechanics of materials, 9th Edition. Cengage
Learning.
Lim, T.C., 2015. Auxetic materials and structures (pp. 55–56). Springer.
Timoshenko, S. and Gere, J.M., 1972. Mechanics of materials, 1st Edition, D. Van
Nostrand Company.
10
This presents the bias of a materials scientist. A designer might default to redesigning the
component or structure, especially if there is a requirement to use a particular material. It
will of course depend on the overall constraints of the project as a whole, including budget,
the time available to develop the design, environmental concerns, as well as the footprint/
volume available.
@seismicisolation
@seismicisolation
Mechanical Properties I: Strain, Stress, Stiffness 57
Recommended Reading
In addition to the textbook that is required reading for Engineering students around
the world, Timoshenko wrote up lecture notes for his course on the history
behind the development of fundamental engineering theory:
Timoshenko, S., 1983. History of strength of materials: With a brief account of the history of
theory of elasticity and theory of structures. Courier Corporation.
@seismicisolation
@seismicisolation
6
Mechanical Properties II: Toughness
and Fracture Toughness
6.1 Introduction
It can be tempting to think that so long as we know the strength of a material,
its strain to failure, and its Young’s modulus, then we know everything we
need in order to design structures using that material. In some senses that is
true, but it only deals with one facet of the design process. Two further mech-
anical properties that need to be considered are the toughness and the fracture
toughness. Together, these properties form the core of fracture mechanics.
Fracture mechanics is the study of cracks, specifically how big does a crack
need to be in order to be a problem. There are two aspects to consider here,
the amount of energy required to create two new surfaces (the toughness),
and the resistance to crack growth (the fracture toughness). The latter in par-
ticular is a function of the material, the sharpness of the crack and its orien-
tation with respect to applied loads, the load applied, the geometry of the
component and the position of the crack in it, and various environmental
conditions. The former is more related to the nature of the material itself, and
its ability to deform.
Fracture mechanics is divided into various sub-groupings, the two most
important being linear elastic fracture mechanics (LEFM) and elastic–plastic
fracture mechanics (EPFM).
@seismicisolation
@seismicisolation
58 DOI: 10.1201/9780367822347-6
Mechanical Properties II: Toughness and Fracture Toughness 59
Andrews (1968) also specifies that this definition means that whilst some
words (e.g. ‘rupture’) could be used interchangeably with fracture, ‘failure’
is a word that simply indicates that a component or test piece can no longer
be used for its original purpose, with additional information being required
to specify how failure occurred. Equally, ‘cracking’ is not interchangeable
with fracture since whilst it clearly involves fracture, there are other forms
of fracture which have nothing to do with cracking. In fact, Andrews (1968)
proposes six separate forms of fracture:
1
Recall the point made in the previous chapter regarding the difference between stress (a gen
eral term for the effect of an applied load on a given surface area) and strength (a specific
stress for a given material at which some significant behaviour occurs, such as plastic yielding,
the yield strength, or complete failure, the ultimate tensile strength). For some materials, the
presence of some kind of degradation will just lead to a reduction in the overall area that can
carry the load. For others, the presence of a crack will lead to a stress state at the crack tip
which will exceed the strength of the material at much lower loads than might be expected.
@seismicisolation
@seismicisolation
60 Understanding the Failure of Materials and Structures
us to the question of cause and effect. We can observe a failed specimen and
attribute failure to a particular mechanism, but the failure observed may not
necessarily be a fracture: the generation of a crack and its evolution to a crit-
ical crack can be attributed to a specific mechanism, but the mechanism in
question can also cause other forms of failure. For example, abrasion may
lead to the formation of a crack, or it may lead to a more stable thinning of
the component in a specific location which means that the component fails,
in that it is no longer performing its proper function, or which may lead more
extreme issues in due course.2
“According to these hypotheses rupture may be expected if (a) the maximum ten-
sile stress, (b) the maximum extension, exceeds a certain critical value. Moreover,
as the behaviour of the materials under consideration, within the safe range of
alternating stress, shows very little departure from Hooke’s Law, it was thought
that the necessary stress and strain calculations could be performed by means of
the mathematical theory of elasticity…
“…Thus, on the maximum tension hypothesis, the weakening of, say, a shaft
1 inch in diameter, due to a scratch one ten-thousandth of an inch deep, should be
almost exactly the same as that due to a groove of the same shape one hundredth
of an inch deep…
“…These conclusions are, of course in direct conflict with the results of alter-
nating stress tests…
“…To explain these discrepancies, but one alternative seemed open. Either the
ordinary hypothesis of rupture could be at fault to the extent of 200 or 300 per
cent., or the methods used to compute the stresses in the scratches were defective
in a like degree.”
2
For example, leakage at the joint of a buried water main can lead to a hard object, such as
a pebble, in a soft object, such as soil or backfill, rotating, causing the pipe to wear and a
hole to form. Technically the pipe has failed twice over at this point, as water is escaping
from the joint, as well as through the hole created by the wear caused by the rotating pebble.
However, the pipe is still able to perform its primary function with a greater or lesser loss of
efficiency. However, the greater volume of water now escaping can lead to the pipe becoming
undermined as the pipe’s support is washed away. In due cause, a more significant failure will
occur if the problem remains unidentified for any length of time.
@seismicisolation
@seismicisolation
Mechanical Properties II: Toughness and Fracture Toughness 61
Hence, the common theories, prior to 1920, regarding the rupture of materials
had started to come into disrepute, due to non-reproducibility of results and
variations in results developing from variations in test conditions.
The ‘Molecular Theory of Strength’ (based on the energy required to sep-
arate two atoms so that they no longer interact) suggests that the stress at
fracture (or fracture strength), σmax, is given by:
Eγ
σ max = (6.1)
b0
where E is the Young’s modulus, γ is the surface tension, and b0 is the inter-
atomic spacing. This typically leads to predicted strengths of ~E/10. In prac-
tice it was found that the stress at fracture was much lower, at ~E/1000. This
is approximately two orders of magnitude lower than predicted. Griffith
(1920) was the first to propose a theory to explain this discrepancy.
Griffith built on early work by Inglis (1913) who provided an analysis
which showed the effect of stress concentration by elliptical cut outs3 in uni
formly stressed plates. Whilst the mathematical treatment provided by Inglis
(1913) was significant, it begged the question of why large cracks were more
detrimental than smaller cracks despite the apparent size-independence of
the stress concentration factor.
3
By making one axis much smaller than the other, the ellipse may be assumed to be a mathem
atical model of a crack.
@seismicisolation
@seismicisolation
62 Understanding the Failure of Materials and Structures
result of an applied stress, σ, equation [6.2]) and the surface energy of a crack
of length 2a, US (where γ is the surface energy per unit area, equation [6.3]).
σ 2 πa 2
UE = (6.2)
E
U S = 4γ a (6.3)
Under a given stress the total energy reaches a maximum at a particular value
of crack length. Therefore, crack growth will occur when the strain energy
release rate is equal to or greater than the rate of change of surface energy
with crack length:
dU E dU S
≥ (6.4)
da da
By considering the change in the energy of the system with crack length it is
possible to show that for crack growth to occur:
dU dU S dU E
= − =0 (6.5)
da da da
Substituting equations (6.2) and (6.3) into equation (6.5), and differentiating,
it may be shown that:
σ πa = 2 γ E (6.6)
and since the critical strain energy release rate, GC, is equal to twice the sur-
face energy, γ:
GC = 2 γ (6.7)
and introducing the critical stress intensity factor, KC, which is equal to σ√πa,
equation (6.6) may be written in the form:
KC = GC E (6.8)
@seismicisolation
@seismicisolation
Mechanical Properties II: Toughness and Fracture Toughness 63
then the crack formed by the previously applied stress should ‘heal’ or
close up. The healing of cracks has been reported, and there are very good
reasons why it can potentially occur. However, it is important to maintain a
healthy scepticism when considering a report of self-healing: there must be
a compelling reason for the cracking process to be undone and for bonds to
reform. Simply juggling energy flows is not sufficient. In this context, it is also
important to acknowledge that the term self-healing is also used in materials
science to denote a material that has been designed to incorporate a repair
mechanism.4
Also, the Griffith construct is based on energy changes between an initial
and final state and so represents only a necessary condition for fracture. As it
does not deal with the actual mechanism of the fracture concerned there may
be other criteria which must also be fulfilled for fracture to occur.
(
Gc = 2 γ s + γ p ) (6.9)
Irwin (1958) presented a criterion for the minimum specimen thickness for
the specimen to be in plane strain. This has subsequently been revised, and
here it is presented in the more general form, as it is found, for example, in
ASTM D5045-14R22 (2022):
4
Biological organisms such as the human body are excellent examples of self- healing
mechanisms, and there are certain cements (and derived concretes) and composite materials
which are designed to repair cracks that occur during operation. However, such repair
mechanisms are not self-healing within the context of the damage of a crack being undone
simply by unloading the component.
5
The concept of plane strain, and indeed plane stress, is examined more fully in Chapter 8.
@seismicisolation
@seismicisolation
64 Understanding the Failure of Materials and Structures
2
K
B > 2.5 (6.10)
σy
K =σ√πa (6.11)
where K is the stress intensity factor generated for a given applied stress,
σ, and a is half the length of the crack under consideration. When K is
Kc, the critical stress intensity factor, then this is a property of a given
material.6 However, equation (6.11) is something of special case –it
considers a crack in a semi-infinite plate, which is to say that there are
various factors of constraint that we don’t have to think about. A generic
form of equation (6.11), which begins to take into account various geo-
metric factors is:
K = Yσ √a (6.12)
The factor of π is subsumed into the geometric term, Y, where Y can be modi-
fied to take into account, for example, the ratio of the length of the crack with
respect to the width of the plate (a/w) (see e.g., Hertzberg et al., 2020), the
angle of the crack with respect to the applied load, and the overall shape of
the component (see e.g., Boresi and Schmidt, 2002). A seminal example of the
development of a geometrical factor for a specific application is Newman and
Raju’s (1980), calculated for a crack with a specific profile and location in a
thin-walled pressure vessel.
6
As with stress and strength, K represents a state that exists when a load is applied to a sample
that contains a crack. KC is the material property, which means that if we know KC and either
the applied stress or the crack length, then we can determine the other. If we know that there
will be a particular load applied to a structure, then we can determine the length of the crit-
ical crack. Conversely, if we have identified a crack, perhaps through NDE, or even simple
observation, and we can determine the length of the crack, then if we also know the KC for the
material, then we can calculate the maximum stress that can be applied before failure, and act
accordingly.
@seismicisolation
@seismicisolation
Mechanical Properties II: Toughness and Fracture Toughness 65
∫ δu
J = ∫ We dy − T. ds (6.13)
Γ δx
FIGURE 6.1
Simplified illustration of the application of the J-integral.
Source: Adapted from Rice (1968).
7
y being perpendicular to the crack surface.
@seismicisolation
@seismicisolation
66 Understanding the Failure of Materials and Structures
FIGURE 6.2
Determination of -ΔU. After Begley and Landes (1972).
−∆U
J= (6.14)
∆a
Begley and Landes (1972) presented a method for determining –ΔU from
experimental results, as presented in Figure 6.2. In essence, the difference in
the load–displacement curve for identical material with cracks of different
lengths can be used to calculate the difference in energy absorbed by the
material until the point at which the crack grows.8
In fact, many samples present a sequence such as that demonstrated in
Figure 6.3. Here it can be seen that with increasing crack length, both the max
imum load and maximum displacement that a sample can sustain decreases.
Agarwal et al. (1984) and Singh and Parihar (1986) showed that this type of
behaviour was due to significant yielding away from the crack tip for samples
with short initial cracks. For different materials they were able to show that
yielding away from the crack tip was linked to a critical crack length so that
for deeply notched specimens only yielding at the crack tip was observed (a/
W =0.35 for Agarwal et al., 1984, and a/W =0.34 for Singh and Parihar, 1986).
Kim and Kim (1989) reported that they could not find such a critical crack
8
ΔU may also be calculated from a pair of samples with the same maximum load but where
–
the displacement due to a+Da is greater than that due to a.
@seismicisolation
@seismicisolation
Mechanical Properties II: Toughness and Fracture Toughness 67
FIGURE 6.3
Evaluation of –ΔU for the case where crack propagation occurs at decreasing load and decreasing
displacement for samples with increasing crack length. After Kim and Joe (1988).
length but stated that at a value of a/W =0.4 the relationship between energy
dissipation away from the crack tip and the initial crack length became linear.
Kim and Joe (1987, 1988) presented a ‘locus’ method which seeks to separate
the contributions to the total deformation energy required for the propaga-
tion of the initial crack for a sequence of samples with increasing initial crack
length. The locus method seeks to provide a ‘partition’ between the energy
absorbed at the crack tip and that absorbed away from the crack tip.
Kim and Kim (1989) have used this method to assess a ‘short glass fibre
reinforced thermoplastic polyester’ and found that JC and GC were in
good agreement since the material showed only a minor plastic compo-
nent to its behaviour. Their work shows some unexpected results, however.
For example, in Figure 6.3, it is clear that the displacement due to a sample
decreases with increasing crack length. Kim and Kim (1989) present samples
where a similar relationship of four crack lengths has a4 with a displacement
greater than that of a2 and a3.
Rice et al. (1973) showed that J could be determined using a single spe
cimen. In this method, the displacement is partitioned into two components,
the elastic, δel, and the plastic, δpl, and hence J is made up of two components,
the elastic, Jel, and the plastic, Jpl, contributions to fracture:
@seismicisolation
@seismicisolation
68 Understanding the Failure of Materials and Structures
Jel is essentially the LEFM critical strain energy release rate or toughness, G,
and may be determined by:
P 2 dC
J el = G = (6.16)
2 B da
J pl =
1
B (W − a ) (δ pl
2∫ Pd δ pl − Pδ pl
0 ) (6.17)
2 A*
J pl = (6.18)
B (W − a )
where A* is the area described by the relation between stress and strain and
a straight line between 0 and the maximum load, as illustrated in Figure 6.4.
As may be imagined, this analysis is dependent upon specimen geometry. In
equation (6.18), a single deep notch is considered. Equation (6.18) is suitable
FIGURE 6.4
Demonstrating the derivation of A*.
@seismicisolation
@seismicisolation
Mechanical Properties II: Toughness and Fracture Toughness 69
6.8 Summary
This chapter has considered material properties of toughness and fracture
toughness, and the impact of defects on mechanical performance, where
those defects act to cause stress concentration. When undertaking testing of
materials to determine toughness and fracture toughness, it is important to
ensure that the correct standard is selected for the materials being tested,
and that factors such as whether the material is in plane strain or plane stress
have been taken into account. It is also important to consider the structure
of the material to be tested and whether there are other factors that may
impact on the specimen to be tested. For example, if testing material that
contains particulate reinforcement, the minimum depth of the specimen will
need to be several times the diameter of the reinforcement to ensure that the
reinforcement does not sit proud of the surface and does not have an undue
effect on the matrix, for example, constraining its ability to deform. Hence,
whilst there are fundamental principles to be observed in every test, there are
caveats and modifiers for every material to be tested.
References
Agarwal, B.D., Patro, B.S. and Kumar, P., 1984. “J integral as fracture criterion for short
fibre composites: An experimental approach”. Engineering Fracture Mechanics,
19, pp.675–684.
Andrews, E.H., 1968. Fracture in polymers. Oliver and Boyd.
ASTM D 5045 -96, 1996. “Standard test methods for plane-strain fracture toughness
and strain energy release rate of plastic materials”, Annual Book of ASTM
Standards (pp.325–333). ASTM, Philadelphia.
Begley, J.A. and Landes J.D., 1972. “The J Integral as a fracture criterion”, Fracture
toughness, Proceedings of the 1971 National symposium on fracture mechanics, Part II,
ASTM STP 514 (pp.1–20). American Society for Testing and Materials.
Boresi, A.P. and Schmidt, R.J., 2002. Advanced mechanics of materials, Sixth Edition. John
Wiley & Sons.
Griffith, A.A., 1920. “The phenomena of rupture and flow in solids”. Philosophical
Transactions of the Royal Society, A221, pp.163–198.
Hertzberg, R.W., Vinci, R.P. and Hertzberg, J.L., 2020. Deformation and fracture mechanics
of engineering materials, Sixth Edition. John Wiley & Sons.
@seismicisolation
@seismicisolation
70 Understanding the Failure of Materials and Structures
Inglis, C.E., 1913. “Stresses in a plate due to the presence of cracks and sharp corners”.
Transactions of the Institute of Naval Architects, 55, pp.219–241.
Irwin, G.R., 1958. “Fracture”, in Encyclopaedia of Physics (Vol. VI, pp.551–590), ed. S.
Flügge, Springer-Verlag OHG.
Kim, B.H. and Joe, C.R., 1987. “A method to determine the critical J-Integral value
independent of initial crack sizes and specimen lengths”. International Journal of
Fracture, 34, pp.R57-R60.
Kim, B.H. and Joe, C.R., 1988. “Comparison of the locus and the extrapolation
methods that determine the critical J-Integral in the presence of remote energy
dissipation”. Engineering Fracture Mechanics, 30, pp.493–503.
Kim, B.H. and Kim, H.S., 1989. “Fracture characterization of short glass fibre
reinforced thermoplastic polyester by the J-Integral”. Journal of Materials Science,
24, pp.921–925.
Knott, J.F., 1973. “Fundamentals of Fracture Mechanics”, Butterworths.
Lawn, B.R., 1993. Fracture of Brittle Solids, Second Edition. Cambridge University Press.
Newman Jr, J.C. and Raju, I.S., 1980. Stress-intensity factors for internal surface cracks
in cylindrical pressure vessels. Journal of Pressure Vessel Technology, 102, pp.343.
Orowan, E., 1955. “Energy criteria of fracture”. Welding Research Supplement, 34,
pp.157–160.
Rice, J.R., 1968. “A path independent integral and the approximate analysis of
strain concentration by notches and cracks”. Journal of Applied Mechanics, 35,
pp.379–386.
Rice, J.R., Paris, P.C. and Merkle, J.G., 1973. “Some further of J-Integral analysis and
estimates” in Progress in Flaw Growth and Fracture Toughness Testing (ASTM STP
536) (pp. 231–245). American Society for the Testing of Materials.
Singh, R.K. and Parihar, K.S., 1986. “The J- integral as a Fracture Criterion for
Polycarbonate Thermoplastic”. Journal of Materials Science, 21, pp.3921–3926.
@seismicisolation
@seismicisolation
7
Fatigue, Wear, Creep, and Note on Corrosion
7.1 Introduction
Where strength, Young’s modulus, toughness, and fracture toughness are
material properties, fatigue, wear, and creep are a material’s response to
sub-critical loads. Corrosion is something of a special case, a form of deg-
radation that reduces the carrying capacity of a structure. The thing all four
have in common is that failure does not occur because the structure has been
bought to the ultimate tensile strength of the material in an instant, but rather
because, through continual use and exposure to the environment, the amount
of material that is capable of carrying load is reduced.
7.2 Fatigue
If we have had a long day, we might say that we are tired, or perhaps fatigued.
There are tasks that we perform easily, but at the end of such a day, or if we
repeat them a number of times, we might find ourselves making mistakes, or
perhaps suffering physically more than we do if we complete the tasks when
fresh. The greater the number of repetitions, the greater the impact.
Materials and structures, in some senses, are no different. Recalling our
determination of the strength of a material, this property relates to the value
at which an applied load causes certain changes to occur in the microstructure
of the material (yield strength) or which causes the destruction of the sample
tested (ultimate strength). But what happens if we load the material to some
point short of an upper limit? What happens if we repeat a test a number of
times? Like us repeating an exercise a number of times, the material becomes
fatigued, and damage that was non-critical when the sample was tested short
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-7 71
72 Understanding the Failure of Materials and Structures
of complete failure for one test becomes a problem when we repeat the test.
How much of a problem will depend on the number of times that the spe-
cimen goes through a loading cycle.
Some things, therefore, to keep in mind: every time a vehicle goes over a
bridge, every time an aeroplane takes off and lands, every time a car goes
over a speed bump, part of the structure undergoes a loading cycle. Have a
think about things that you do repeatedly in your daily life: putting on the
kettle, switching on a light, opening a door…
The key issue with fatigue is determining the number of loading cycles that
a material can survive at a given applied stress. Clearly, a material cannot
undergo more than one cycle at a critical stress before yielding or catastrophic
failure prevent the structure from being used again. It is necessary to develop
an understanding of the behaviour of the material for different applied loads,
and determine a model that can be used to predict behaviour, for a given load
cycle (Figure 7.1).
Whilst a single curve can be defined, a distinction is usually drawn between
low-cycle and high-cycle fatigue, with different failure mechanisms being
observed, depending on the balance of applied load and the number of cycles
to failure. For example, Murakami and Miller (2005) describe the impact of
high-stress, low-cycle fatigue on the fracture ductility of materials. Samples
with pre-drilled holes became more susceptible to crack propagation after
low-cycle fatigue, no matter how small the defects already present.
Commonly, the Coffin- Manson equation is used to describe low- cycle
fatigue. As it is easy to misinterpret this equation, it is often helpful to con-
sider it as one equation formed of two parts:
FIGURE 7.1
Indicative stress–number (S–N) curve demonstrating the relationship between applied stress
and the number of cycles that a structure can go through at a given load before failure.
@seismicisolation
@seismicisolation
Fatigue, Wear, Creep, and Note on Corrosion 73
∆εT ∆ε e ∆ε p
= + (7.1)
2 2 2
∆εT
i.e., that the total strain amplitude, , is comprised of an elastic strain
2
∆ε e ∆ε p
amplitude, , and a plastic strain amplitude, . The elastic part of the
2 2
equation is defined as:
∆ε e E
( )
b
= σ a = σ ′f 2 N f (7.2)
2
∆ε p
( )
c
= ε ′f 2 N f (7.3)
2
∆ε p
where is the plastic strain amplitude, ε ′f is the fatigue ductility coeffi-
2
cient, an empirical constant defined by the strain intercept at 2N =1, and c is
the fatigue ductility exponent. This latter is a material property, but will gen-
erally lie between –0.5 and –0.7.2
Hence:
∆εT σ ′f
(2 N ) ( )
b c
= f
+ ε ′f 2 N f (7.4)
2 E
σ ′f
It is tempting to try to simplify the equation by assuming that =ε ′f , but as
E
the two are defined on the basis of separate graphs, this is not an appropriate
step to take.
Whilst it is normal to think in terms of the fatigue cycle as being a sinus-
oidal wave function (Figure 7.2), other wave shapes are used in certain
circumstances, and it is important to check what shape is required for a
1
That is, twice the number of cycles, Nf, to failure, because there are two strain reversals
per cycle.
2
There is no correlation between b and c, although b is always smaller, and the exact ratio varies
between approximately 1/10 and 1/4.
@seismicisolation
@seismicisolation
74 Understanding the Failure of Materials and Structures
As a result of any of the factors above, or a combination, there are three key
issues to watch out for:
• Failure to reach the maximum or minimum loads before the next cycle
starts;
• ‘Smearing’ of the peak, where there is some flattening of the wave-
peak; and
• Asymmetry of the wave-form where one side of the wave takes longer
and the other is slower.
Figure 7.2 represents the typical features of a fatigue curve, but one where the
peak tensile stress and the peak compressive stress are equal. In practice, such
a cycle, where the specimen undergoes full reversal, is at the more difficult
end of testing, and is not the typical fatigue loading cycle to be considered.
More normally, fatigue will be carried out in a fully tensile or fully compres-
sive manner, which is more representative of how most structural members
will behave in operation. Hence, the cycle is usually defined by a ratio of
peak and minimum applied stress, R, which will usually equal 0.1, although
other values of R may be used on occasion:
σ min
R= (7.5)
σ max
Considering, Figure 7.2 again, another feature that we can see that has been
idealised is that the fatigue cycle is uniform over the whole number of cycles
considered. In terms of testing, this is generally the case; in practice, over the
3
What constitutes high in this case will depend on the capacity of the test frame and its mode
of operation. An electrodynamic frame can respond more rapidly than a hydraulic frame, for
example.
4
A particular issue with hydraulic test frames is the potential for the oil in the power-pack to
become overheated.
@seismicisolation
@seismicisolation
Fatigue, Wear, Creep, and Note on Corrosion 75
FIGURE 7.2
Key features of fatigue. A fatigue cycle can be defined from any starting point, and this point will
be seen at three points in any cycle, at the beginning of the cycle, at the end of the cycle, and at
some point in the middle, depending on exactly which starting point is chosen. The exception to
this is if the maximum or minimum point of the cycle is chosen as the starting point.
da
= C ( ∆K )
m
(7.6)
dN
where da/dN is the change in crack length in a single fatigue cycle, ΔK is the
difference between the maximum and minimum stress intensity factors:
@seismicisolation
@seismicisolation
76 Understanding the Failure of Materials and Structures
FIGURE 7.3
Illustration of the crack growth rate vs. stress intensity range. At very low (Region A) or very
high (Region C) ΔK, crack growth is unstable. The Paris Law only applies to stable crack growth
(Region B).
should also be noted that the Paris Law only applies to stable crack growth
(see Figure 7.3).
7.3 Wear
The focus of this chapter is very much on the aspects of failure that relate to
the mechanical properties of materials, but it would be remiss not to touch
on other aspects. Returning to Andrews’ (1968) list, there is one glaring inclu
sion, which deals with an application of load in a manner that does not dir-
ectly relate to the mechanical properties described previously: wear.
There are two obvious examples of wear that can be described to help us
define what we are talking about. Firstly, clothing is particularly prone to
wear, even if it has not been fashionably distressed. Once again popular, the
concept of putting patches on elbows was considered laughable at one point,
but came about because otherwise perfectly serviceable jackets were worn
out at the elbows, when the arms inside were in contact with, typically tables.
This kind of wear is also seen on trousers at the knee. Again, this is the most
likely part of the piece of clothing to become worn, as the clothing becomes
stretched over this part of the limb, and hence is under tension when it
comes into contact with another surface, such as furniture or the ground. The
second example to consider is that of the car tyre. There is, of course, a legal
@seismicisolation
@seismicisolation
Fatigue, Wear, Creep, and Note on Corrosion 77
requirement in many countries for the tread cut into the tyre to be a certain
depth, and this is part of the annual health check of a car, as well as the respon-
sibility of the motorist to carry out spot checks. The road-handling perform-
ance of the car is dependent on the condition of the tyre. Tyres become worn
through normal driving, although it is worth noting that power-steering can
lead to greater wear in certain locations: it is good practice to allow the car to
move slightly whilst steering to limit this kind of damage.
7.4 Creep
Creep can be considered as a form of fatigue, in as much as long-term applied
load gives rise to changes in the microstructure that lead to eventual failure,
which occurs at a load significantly lower than the yield or ultimate tensile
strength of the material. However, the applied load in this instance is static,
rather than dynamic, and frequently the load in question is self-load. The
classic example in this context would be that of the lead used to protect church
roofs. In the past the lead would be removed approximately every hundred
years, melted, recast, and reapplied for another century. Over this time, the
weight of the lead would cause the sheet to stretch, eventually leading to
cracking of the sheet, and the potential for water ingress. Mathematically,
creep is defined as follows:
d ε C σ m −kTQ
= b e (7.8)
dt d
@seismicisolation
@seismicisolation
78 Understanding the Failure of Materials and Structures
‘rot’ observed in, e.g., wood, because rot is essentially a biological function
where living organisms eat biological materials (the macromolecules that
form the wood). Corrosion, meanwhile is non-biological.
Corrosion can be prevented by the careful selection of materials, and by
specifying protective coatings, environmental conditions, and the like. From
the perspective of an interested observer, it is interesting to note that arch-
aeological relics can survive for significant periods of time in what might
be considered unpromising environments, only to fall apart or corrode rap-
idly when removed to room temperature, pressure, and humidity. From the
perspective of a curator such phenomena can be perplexing and worrying;
from that of an observant engineer, it should be terrifying. Take for example
the case of artefacts recovered with the excavation of the Mary Rose. The
Mary Rose sank in 1545 and spent the next four centuries on the seabed
of the Solent. Significant portions of the wreck rotted away, but a mean-
ingful amount was protected by deposits of sediment over time. Personal
possessions were recovered that provide insight not only to life on board a
Tudor warship, but also to Tudor life more generally.
The Mary Rose Museum in Portsmouth, UK, is well worth a visit: the
efforts that have been undertaken to present the ship and the lives of sailors
on board at the time of its sinking, not to mention its 33 years of service up to
this point, is fascinating, and an exemplar of what such a museum should be.
The curators and conservators have gone to great pains to protect the
remains of the vessel: originally raised in 1982, it has taken much of the inter-
vening four decades to preserve the timbers, drying them out by displacing
water with propylene glycol, which prevents the wood from becoming too
dry and brittle. Today, the remains can be seen close too, and the ship provides
an intimate setting for history that can otherwise be somewhat dry and dusty.
But what of understanding failure? The conservators have not only had to
deal with preventing wood from degrading but also the unexpected spalling
of corroded material from the surface of recovered cannonballs. Advanced
characterisation techniques utilising the Dimond Light Source were able to
determine which corrosion phases formed during conservation and hence
the best route to preventing the situation occurring again (Simon et al., 2018).
Corrosion is also awkward because the same fundamental corrosion
processes can lead to different results. For example, Jesson et al. (2013) were
able to show that graphitisation, a form of corrosion unique to cast irons, in
which the graphite flake within the structure is broken down and back-filled
with corrosion products in a manner akin to fossilisation, can occur over a
large area leading to a gradual loss of section, or can penetrate into the bulk,
effectively creating cracks that act as stress concentrations.
A final thought on corrosion, is that it is not all bad. Corrosion can, surpris-
ingly, be used to protect. When controlled, corrosion processes can be used to
create protective coatings. In a sense, this is exactly what occurs with stainless
steel, where protective oxides are formed thanks to alloying additions to the
@seismicisolation
@seismicisolation
Fatigue, Wear, Creep, and Note on Corrosion 79
steel. In the case of copper, a protective patina can be created through a mix
of various copper salts being allowed to form at the surface (see, e.g., Rosales
et al., 1999).
7.6 Summary
In previous chapters we have considered the mechanical behaviour of
materials and structures on the basis of the microstructure engendered
through the manufacture of the component of interest, and on the defects that
arise during manufacture. Here, the question of the accumulation of a life-
time of relatively minor degradation processes has been considered, and the
impact of these on the ability of the material or structure to withstand oper-
ational stresses. These are the kind of issues that lead to structures that have
been fine up until now, ‘suddenly’ failing. The failure wasn’t sudden: it may
have been approaching instantaneous when the catastrophic event occurred,
but the failure has been a lifetime in the making. Understanding the specifics
of degradation processes that apply to specific materials and situations can
help prevent or at least delay the catastrophic failure, by targeting appropriate
maintenance, and predicting when replacement should occur. Monitoring of
structures, particularly expensive structures, can also help, especially with
extending the life of the structure not through remediation per se, but by
understanding whether the rate of degradation has occurred as expected, or
more slowly. It can also help in preventing catastrophe by determining that
the degradation has occurred in an unexpected way.
References
Andrews, E.H., 1968. Fracture in Polymers, Oliver and Boyd.
Jesson, D.A., Mohebbi, H., Farrow, J., Mulheron, M.J. and Smith, P.A., 2013. On the
condition assessment of cast iron trunk main: The effect of microstructure and
in-service graphitisation on mechanical properties in flexure. Materials Science
and Engineering: A, 576, pp.192–201.
Murakami, Y. and Miller, K.J., 2005. What is fatigue damage? A view point from the
observation of low cycle fatigue process. International Journal of Fatigue, 27(8),
pp.991–1005.
Rosales, B., Vera, R. and Moriena, G., 1999. Evaluation of the protective properties of
natural and artificial patinas on copper. Part I. Patinas formed by immersion.
Corrosion Science, 41(4), pp.625–651
@seismicisolation
@seismicisolation
80 Understanding the Failure of Materials and Structures
Simon, H., Cibin, G., Robbins, P., Day, S., Tang, C., Freestone, I. and Schofield, E., 2018.
A Synchrotron-based study of the Mary Rose iron cannonballs. Angewandte
Chemie International Edition, 57(25), pp.7390–7395.
Sonsino, C.M., 2007. “Fatigue testing under variable amplitude loading.” International
Journal of Fatigue, 29(6), pp.1080–1089.
@seismicisolation
@seismicisolation
8
Testing Modes and Application of Load
8.1 Introduction
When designing a structure that must exist in the real world, it is necessary
to think about the kinds of loads that will be applied and how these will
be accommodated. For example, Gordon (1978) illustrates the importance
of pre-tension with the nature of trees to grow wood that is naturally pre-
stressed, so that when the tree is placed in bending by a strong wind it is
able to bend further than if it were not prestressed. In part this comes from
the nature of heartwood and sapwood, with the heartwood being stronger
in compression and the sapwood in compression, which is why traditional
long-bows are made from a single stave of wood which has both heart and
sapwood, so that when the bow is bent, the heartwood is in compression and
the sapwood is in tension. It is also interesting to see the effect of the wind on
tree growth. Where a tree is subjected to a single predominant wind, then it
will tend to grow in a lopsided manner to compensate for this loading, which
is effectively only acting in one direction. Where a tree is subjected to winds
from many directions, then it will grow in a more uniform manner.
Without getting too focused on specifics, there are a range of loading
conditions that may need to be considered:
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-8 81
82 Understanding the Failure of Materials and Structures
FIGURE 8.1
Consider a cube of material: (a) three orthogonal directions are defined and load may be applied
in (b) one, (c) two, or (d) three directions at once.
@seismicisolation
@seismicisolation
Testing Modes and Application of Load 83
forms of loading, although sometimes more than one will be used at once;
tension-
torsion is probably the most common form of compound test.
Arguably, shear is a subset based on a specific geometry type combined with
a form of loading.
Uniaxial Loading
In some respects, uniaxial loading is the simplest form of testing that is
available for mechanical characterisation of materials and structures and
it is the basis of much that is used to define mechanical properties and
performance. Uniaxial loading encompasses both tension and compression
however, and the failures observed through each of these behaviours are
different. The failure mode observed is also dependent on the structure of
the material. Uniaxial testing forms the basis of most material properties
characterisation.
Bending
The simplest form of non-uniaxial loading is bending. This will occur, to some
extent, due to the self-weight of the structure, although this will typically
be small unless the structure is very large, and/or slender. So, for example,
a long hair placed on two supports will tend to sag under its own weight,
as will a filament of wire of the same thickness. If we reduce the distance
between the supports, the hair will sag less and less perceptibly. However,
when it comes to the design process, it is necessary to specify a maximum
amount of displacement that can be allowed, and this, together with the
overall load to be carried, will define the amount of material required for
a beam or similar to cross a known width without exceeding the maximum
permitted displacement.
The load to be carried may be localised or distributed. For the latter, think
of a lintel above a door or window. Bridges can be more complicated still,
depending on the total traffic that they can carry and how much this varies
over time. In determining mechanical properties, a sample is usually placed
into three-or four- point bending (Figure 8.1). In bending, a material is
subjected to both tensile and compressive forces (Figure 8.1b). The maximum
strain that the beam sees occurs at the outermost elements, hence I-beams,
box sections, and sandwich panels are designed to place the majority of
material at the place where the greatest strain is observed. As there is less of
a need for materials in the centre of the section, this material can be reduced
without compromising the integrity of the structure, producing a beam or
panel which is relatively light, thus reducing the potential for deflection due
to self-weight. To put it another way, we have increased the specific stiffness
(Figure 8.2).
@seismicisolation
@seismicisolation
84 Understanding the Failure of Materials and Structures
FIGURE 8.2
Illustration of the tensile and compressive forces placed on a structural member in bending.
(a) A sample placed in three-point bending, (b) the specimen in (a) sees compressive force on
the top face and tensile force at the bottom, because the curvature engendered in the specimen
leads to the bottom face becoming stretched and the top face compressed. (c) A sample placed
in four-point bending. (d) A sample in bending demonstrates a neutral axis, nominally through
the centre of the specimen, where the material is neither in compression or tension. The neutral
axis will move towards the stiffer face, which may occur if, for example, the material is stiffer in
either tension or compression, or if the material is in any way damaged.
For example, simple bending theory, which can be applied to both three-
and four-point flexural tests, can be reduced to:
M σ E
= = (8.1)
I y R
@seismicisolation
@seismicisolation
Testing Modes and Application of Load 85
In all affairs it’s a healthy thing now and then to hang a question mark
on the things you have long taken for granted.
External advice was sought, and the sector was introduced to fracture
mechanics: cracks lead to stress concentrations at the crack tip and in this
1
A moment, also called a moment of force, acts to make an object rotate around a fixed point.
The size of the moment is a function of the distance from the pivot point, and the force applied.
A perfect example of this is the use of an allen key. When undoing a hex-bolt, you get a larger
moment if you use the longer arm for leverage. When the bolt is looser, it is quicker to undo it
when the shorter arm is used to turn the key.
2
The second moment of area, sometimes called the moment of inertia, is a geometric function
which addresses the impact of the cross-section on various other functions. Perhaps the sim-
bd 3
plest is a rectangular cross section, where I = .
12
@seismicisolation
@seismicisolation
86 Understanding the Failure of Materials and Structures
FIGURE 8.3
Illustrating the action of shear: (a) deformation due to a shear load and (b) a sample failing
in shear.
instance some forms of corrosion can act in a crack-like manner. These were
shown to be the cause of the unexplained catastrophic failures, but it required
a re-evaluation of the material and its behaviour when degraded, to provide
an answer when the conventional wisdom of the industry had been shown to
lack an answer. This was considered in greater depth in Chapter 5.
The point is that not all materials behave in the same way. In the context of
materials performance there are a number of factors which affect the mechan-
ical properties of a material, which is why metals and ceramics, for example,
(tend to) behave in such different ways. But even if we focus on only one
material, various factors can affect performance. This will be explored further
in the context of strain-rate dependence in Chapter 10.
Shear
Shear occurs when a specimen is loaded normal to the face that it is acting
on, rather than parallel, with one face perpendicular to the applied load fixed
in place (Figure 8.3a). In uniaxial tension or compression we take the sample
and pull (or push) along the main axis, and cause the specimen to elongate
(or compress). With shear, we take the same specimen, hold it in the same
way at the bottom, but then push at the top, leading to deformation, with
final failure occurring as seen in Figure 8.3b.
Torsion
Like shear, torsion occurs when force is applied in a non- axial manner
(Figure 8.4), such that deformation of the material occurs. Torsion is a fun
damental aspect of many engineering processes, including, for example,
the transmission of force from an engine to a wheel-set using a drive shaft.
@seismicisolation
@seismicisolation
Testing Modes and Application of Load 87
Another example familiar to many is that of a screw being driven into a bulk
material such as a masonry wall or a wooden fence. The screw is rotated,
and due to the force placed behind and the pitch of the thread, the screw
is driven in, thereafter to hold something in place. However, if the screw is
unable to move and we continue to apply force, the screw will fail through
torsional shear.
JT JT
T= τ= Gϕ (8.2)
r l
Plane Stress
Recalling the sample of material in Figure 8.1, and identifying the orthogonal
directions (Figure 8.5a), we can undertake mathematical operations to calcu
late the various stresses we are interested in. However, we can simplify these
operations by considering a situation of plane stress, whereby if we define
a plane xy, the through-thickness direction, z, is infinitely thin (Figure 8.5b).
Hence, there will be no stress, other than that in the x and y directions. In
practice, the materials that we consider are not going to be infinitely thin, but
@seismicisolation
@seismicisolation
88 Understanding the Failure of Materials and Structures
FIGURE 8.4
Effect of torsion on a cylindrical sample.
FIGURE 8.5
Recalling Figure 8.1, we can (a) identify specific orthogonal directions. In the case of an infinitely
thin xy plane (b), there can be no through-thickness stress in the z direction. For a real material,
there is a finite thickness (c) for which this assumption can be considered to be true.
r
< 10 (8.3a)
t
@seismicisolation
@seismicisolation
Testing Modes and Application of Load 89
or
r
t< (8.3b)
10
In other words, if the radius of a tube were 10 mm, then the wall thickness
would need to be less than 1 mm for this assumption to be valid. Happily, this
is frequently easy to achieve, although it should be noted that the microstruc-
ture of the material may become important, as indeed will the processing
route. Some microstructural features may be suppressed or take unusual
forms if the thickness of the material is too small, and this may affect the
properties of the material. As we increase the size of the component, different
microstructures might be observed, and hence a different strength will be
seen in a smaller component compared with a bigger one, unless the pro-
cessing route is controlled carefully.
Plane Strain
By contrast, in the case of plane strain, strain is assumed to be zero in the
through-thickness direction, when the specimen is sufficiently thick that it
constrains the material from moving when the sample is loaded. This sort of
condition will occur in long billets of material, and in structures with thick
walls, a prime example being hydraulic dams. A stress is generated in the
through-thickness direction (Figure 8.6) which is defined as:
(
σz ≈ υ σ x + σ y ) (8.4)
FIGURE 8.6
In plane strain, a stress is generated in the through-thickness direction (as indicated by the
dashed arrows) due to constraint by the bulk when the material is loaded in other directions.
@seismicisolation
@seismicisolation
90 Understanding the Failure of Materials and Structures
The key thing to make note of, however, is that around crack tips, two very
different behaviours are observed. In the case of plane stress, there is sig-
nificant plastic yielding and the maximum fracture toughness is achieved.
Where a specimen is in plane strain, a triaxial stress state is found at the crack
tip such that plastic yielding is constrained, and hence the minimum fracture
toughness is found. There is no point of inflexion between the two conditions,
but rather a region of mixed-mode behaviour between the two extremes.
Recalling Figure 8.4b and the associated discussion, it is also worth noting
that, of course, an infinitely thin plate will not have any volume to plastically
deform and so there will be a drop off in fracture toughness beyond a certain
thickness.
8.6 Summary
Testing specimens to determine material properties can be very straightfor-
ward. It can also be very complicated, with the difference frequently arising
from the nature of the material being tested and the requirements to constrain
the specimen in particular ways to prevent certain behaviours occurring. For
example, the single lap shear test is fairly easy to prepare for, although care
must be taken to ensure that the individual specimens are properly aligned.
However, there is an inherent danger with this specimen geometry that could
lead to out-of-plane bending of the specimen. The double lap joint is more
time consuming to prepare, but more stable to test.
@seismicisolation
@seismicisolation
Testing Modes and Application of Load 91
FIGURE 8.7
Modes of testing: (a) I –opening, or tensile, (b) II –sliding, or in-plane shear, and (c) III –tearing,
or antiplane shear.
@seismicisolation
@seismicisolation
92 Understanding the Failure of Materials and Structures
References
Fahimi, A., Evans, T.S., Farrow, J., Jesson, D.A., Mulheron, M.J. and Smith, P.A., 2016.
On the residual strength of aging cast iron trunk mains: Physically-based models
for asset failure. Materials Science and Engineering: A, 663, pp.204–212.
Gordon, J.E. 1978. Structures or Why things don’t fall down, pp.149–167, Penguin.
Rajani, B. and Makar, J., 2000. A methodology to estimate remaining service life of grey
cast iron water mains. Canadian Journal of Civil Engineering, 27(6), pp.1259–1272.
@seismicisolation
@seismicisolation
9
How to Measure Strain
I often say that when you can measure what you are speaking about, and express
it in numbers, you know something about it; but when you cannot measure
it, when you cannot express it in numbers, your knowledge is of a meagre and
unsatisfactory kind; it may be the beginning of knowledge, but you have scarcely,
in your thoughts, advanced to the stage of science, whatever the matter may be.
–William Thomson, Lord Kelvin
“Electrical Units of Measurement” in
Popular Lectures (1883) Vol. I, p. 73
9.1 Introduction
Strain is one of the fundamental measurements that can be made when under-
taking mechanical characterisation. On applying a load, we begin to change
the specimen that we are testing. Initially this change will be all but impos-
sible to identify, but whether we are pulling, pushing, bending, or twisting,
we are making the atoms move relative to each other. This was discussed in
detail in Chapter 4. From that discussion, it was shown that we can identify
a property that represents the change in the specimen, which we call strain.
As will be recalled, strain is the change in original length, divided by the ori-
ginal length. The original length is relatively easy to determine, and to deter-
mine accurately –although it is important to beware of spurious precision
in quoting the dimensions of a component. For example, if you have a rule
marked off in millimetres, and make multiple measurements, you cannot
provide an average measurement to the nearest 100 microns.
When the test is underway though, how are we to determine the change in
length? In a very basic test, we might use the same millimetre marked rule
used to determine the original length, but that is rather unsatisfactory, for a
number of reasons, as a moment’s thought will show. At the very least, there
is the question of keeping the rule steady against the specimen, which might
affect the specimen as it is being tested, not to mention the potential safety
problems, especially if you are taking your test to failure. A common meth-
odology in civil engineering is to use a DEMEC (demountable mechanical)
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-9 93
94 Understanding the Failure of Materials and Structures
strain gauge, developed by Morice and Base (1953). Today there are numerous
suppliers of the equipment for this technique, but the basic principle is that
small discs of, typically, mild steel1 which have a central dimple, are bonded
to the surface of interest. A bar, usually of invar2 (although sometimes this is
simply used for calibration purposes), with two conically tipped pointers and
a dial gauge is placed with the tips of the pointers fitting into the dimples.
One pointer is fixed, and the other is on a knife-edge pivot, causing the dial
gauge to register a change in position, which is typically expressed as a
strain.3 The original paper suggests that a skilled operative can conduct some
200 measurements an hour, with an accuracy of ~ ± 5 × 10–6. The accuracy
can be improved upon under ideal laboratory conditions. Like many simple
concepts, it is easy to understand, easy to learn the basics, but does require
some practice to do it well and develop the skill and precision to achieve the
rate of testing and accuracy stated in the paper.
This is just one way of determining strain: it has its advantages and
disadvantages. This chapter will explore some of the most commonly used
methods to determine strain in and out of the laboratory environment,
ranging from simple measurements to the latest computer- vision-
based
approaches.
1
Sometimes stainless steel or other materials with good weathering characteristics are used if
the installation is to be used for a longer-term programme of strain monitoring.
2
An alloy of iron and nickel, invar is frequently used in scientific instruments and mechanisms
thanks to a co-efficient of thermal expansion that is virtually zero.
3
With respect to the point made about health and safety, there are two modes in which this
kind of measurement is made. Firstly, the system is under natural load and the structure
being assessed is undergoing routine and long-term asset management in order to ensure that
all is well. Secondly, whilst the measurements may be made on a sample undergoing active
loading, this will not be a quasi-static test, but rather increases in load will not continue until
the measurement is made.
4
Frequently misspelled as ‘Iosepescu’; the methodology for this test is found in ASTM D5379,
originally published in 1992 and last revised in 2012.
@seismicisolation
@seismicisolation
How to Measure Strain 95
there has to be some kind of frame to hold the test, there must be some means
of measuring the load applied to the specimen to be tested, there must be
some way of holding the specimen, and there must be a moving part in order
that the specimen can be loaded.5 In the case of the universal test frame, it is
the crosshead that moves. Servo-hydraulic and linear motor frames have a
slightly different design, whereby a piston is driven up and down either by
pumped fluid, usually oil, or by an electric motor. One set of jaws, or some
other attachment which connects with the specimen, are directly attached to
the piston. Even so, similar principles are in play, and so we will consider the
case of crosshead displacement and treat all three frames the same knowing
that there are some fundamental differences in operation but that the mech-
anical problem that we are currently interested in is the same whatever test
frame we actually use for the test.
Crosshead displacement is perhaps the simplest method available for
determining strain, especially as most commercial test frames will record this
information automatically when a test is underway. When this happens, the
specimen will resist the deformation that is caused by the movement of
the crosshead (or the piston), and it is this resistance that is measured by
the loadcell. The problem, however, is that no test frame is infinitely stiff. The
consequence may be expressed in the same terms as Newton’s Third Law
of motion: for every action, there is an equal and opposite reaction. To put
it another way, if testing an infinitely stiff specimen, the test frame would
fail in the attempt of trying to test it (which, incidentally, is why it should be
standard operating procedure to ensure that physical trip limits are set on the
test frame as well as virtual ones on the control software). Every component
of the test, as detailed in Figure 9.1, contributes something to the stiffness
observed in the load–displacement output of the test. Once plastic deform-
ation, or some other indication of ongoing failure,6 is underway, the effect
of the test frame can be ignored. However, whilst mechanisms of failure are
important, it is the behaviour of the material prior to this permanent change
that is typically most important to designers.
Compliance, C, can, in some respects, be considered as the inverse of
stiffness; it is the ratio of displacement, δ, and load, P, i.e.:
5
In some circumstances the specimen changes due to some external stimulus and hence causes
the linkage to the load cell to move, and the load measured is a consequence of this movement.
Jesson et al. (2010) describe a test on exhumed cast iron water pipes, which included tests with
the specimen –a 4”/100 mm internal diameter pipe from a distribution main –fitted with spe-
cial end caps which allowed the specimen to be held in place, and which also allowed the pipe
to be filled with water whilst under test, and for the water to circulate. The water was passed
through a chiller and the temperature of the water was varied. As a consequence of thermal
contraction, the load induced as the pipe cooled could be measured.
6
See, for example, the behaviour of fibre-reinforced cements and concretes, where the slip of
fibres is a non-reversible form of damage but does not conform to the definition of plastic
failure given in Chapter 4.
@seismicisolation
@seismicisolation
96 Understanding the Failure of Materials and Structures
FIGURE 9.1
Schematic of the principal components of a test frame: (a) a universal test frame, where the entire
crosshead is moved and (b) a test frame where only the actuator moves, driven either by a servo-
hydraulic or electro-mechanical system.
@seismicisolation
@seismicisolation
How to Measure Strain 97
C =δ/P (9.1)
CT =Ʃ(CS+C1+C2+C3…) (9.2)
where CM is the machine compliance (if we take a broad view and include
the full load train with this), which is to say that in some respects the actual
test geometry is of relatively little significance. It is possible to determine
the compliance of the test frame by taking a very stiff sample, typically a
large piece of metal, and applying a load to it. The applied load should be no
more than a few percent of the load cell capacity.8 Under such circumstances
the specimen will be loaded but will not be deformed, and any crosshead
displacement registered will therefore be due to the machine compliance.
Therefore, when CS → 0:
CT =CM (9.4)
When undertaking actual tests using this setup, the load–displacement data
can then be corrected, firstly by subtracting the machine compliance:
7
And indeed, even tensile specimens of different thicknesses might have an effect, because
different jaw faces will need to be used.
8
In some instances this will be the same as the frame capacity, but some frames are designed
to be used with a range of load cells, and some of these can be considerably smaller than the
frame. Applying 1 kN to a 100-kN load frame with a full-capacity load cell is perfectly reason-
able, but when using a 10-N load cell, it is not.
@seismicisolation
@seismicisolation
98 Understanding the Failure of Materials and Structures
But the load that is applied to achieve the total displacement is the same
load seen by any part of the load train, including the specimen that we are
interested in. Therefore:
δS =δT – δM (9.9)
The theory is impeccable, and the practitioner should always attempt the
process at least once to convince themselves that it does indeed work, whilst
at the same time remaining the least accurate method of ultimately deter-
mining strain. The strain output from crosshead measurements is inherently
flawed due to machine compliance issues, and whilst some corrections can
be applied, corrections that will be unique to a particular test frame, the final
result is still likely to be overly compliant and give a spurious measure of the
stiffness of the specimen.
@seismicisolation
@seismicisolation
How to Measure Strain 99
@seismicisolation
@seismicisolation
100 Understanding the Failure of Materials and Structures
Recalling from Section 9.1 that a skilled operator can undertake of the
order of 200 measurements an hour using the DEMEC system, a well set
up datalogger, collecting information from a well installed strain gauge can
collect several times this number of measurements every second, and there-
fore the question becomes one of determining a meaningful sampling rate in
order that the glut of data does not become overwhelming.
The average practitioner will probably only need to keep a handful of types
of strain gauge in mind for the tests that they will typically undertake. It must
be remembered though that the key thing that sets human beings apart from
most other animals is imagination. Hence, if you can imagine a test, there
is probably a strain gauge already available for it and, if there isn’t, with
modern technology it would not be too hard to make a custom gauge for a
particular experiment.
However, a careful read of the catalogue9 is also a sensible idea, and a dis
cussion with your regular supplier can pay dividends. There are gauges
which are tuned to particular materials; there are gauges that are more suited
to fatigue tests; there are all manner of special gauges for determining strains
under odd loading conditions and in strange places.
One final note, born out of experiences with gauges to measure crack
growth, in the majority of situations, all things will be well. However, very
occasionally they won’t: it is entirely possible for spurious readings to be
recorded during a test, and for these to cause great consternation. In the case
of a crack-measuring gauge, the gauge is designed to determine the growth
of a crack during cyclic fatigue. It does this in a stepwise fashion: the gauge is
glued to the specimen where the crack is due to grow. Strictly speaking this is
not a strain gauge in the sense that we have been discussing, but it is used for
mechanical characterisation, and does operate on the same basic principles
as other wire-based gauges. In this instance though, the gauge forms a circuit
of wires in parallel and in pristine condition all the wires carry the current
that is passed through the gauge. As the crack grows, each wire is broken
when the crack tip crosses it. However, the broken wires of the gauge can end
up touching and hence conducting current. Careful observation is required
during the test, and some experience is helpful when reviewing the test data.
This is perhaps the most extreme thing that can occur, but it is entirely pos-
sible to install the gauge upside down (i.e. with the metal contacts against the
specimen), use the wrong gauge for the experimental set up, or even for it to
be not quite aligned properly with the loading direction. It is important to
review the information collected: the practitioner must assume that anyone
reviewing their data will be judging it, and that it will need to be defended.10
9
As a certain generation of lecturers are keen to point out, a month in the laboratory can save
an hour in the library.
10
Although it is often said that the modeller always believes their model, and no one else
does, whilst a physical researcher never believes their data, when everyone else takes it at
face value.
@seismicisolation
@seismicisolation
How to Measure Strain 101
This is good practice in any branch of research, and a good defence is one that
is grounded in a thorough knowledge of the material tested and its normal
behaviour, the operation of the test, and an understanding of the typical
sources of error.
11
It is worth seeking the paper out, not just for its novelty value. It provides a useful exercise
for the researcher in considering the fundamental issues to be presented when discussing
the results of an experiment, whilst at the same time emphasising the importance of a good
literature review.
@seismicisolation
@seismicisolation
102 Understanding the Failure of Materials and Structures
FIGURE 9.2
Schematic of an extemsometer attached to a specimen.
@seismicisolation
@seismicisolation
How to Measure Strain 103
as an upgrade. Most are also compatible with other systems or can be used in
conjunction with custom-built self-reacting frames.
A variant of this is laser extensometry: in some ways this is even simpler
than video extensometry, in that you do not need to apply markers to the spe-
cimen for the tracking algorithm to follow. Instead, a laser beam is directed
onto the specimen and the reflections are captured by a CCD camera and
analysed.
Virtual extensometry is like other forms of extensometry in that you can get
continuous and live strain data output whilst you are carrying out the test.
In some respects, it is more complicated because now there is the additional
issue of lighting to consider: even in a well-lit laboratory, additional lighting
is usually necessary to ensure that there are no opportunities for visual
artefacts to be created, such as shadows or point reflections. That said, the
engineering principles behind the technique are straightforward: instead of
measuring the change in resistivity of a wire as it becomes stretched, a com-
puter program measures the distance between two markers. If continuous
data are not required, those so inclined can achieve a similar effect using
an SLR camera and digital image management software, such as ImageJ.12
Originally developed for the analysis of images of biological origin, such as
cell tissue, the program has a great deal of functionality which can be applied
in a number of other contexts. One of the main features of the program is the
ability to batch process images, and to track features. For example, Boughanem
et al. (2015) assessed the orientation of fibres in a reinforced cement sample,
analysing hundreds of images per sample, whilst Matthews et al. (2020) used
ImageJ to (1) produce scale bars on images; (2) undertake greyscale profile
plotting for an analysis line and as a histogram of an image; and (3) stitch
together multiple images. The program is open-source software13 and there
are a number of plugins available, some of which are bundled with the main
program in ‘flavours’ available from other sources. As the program is open
source, it is relatively easy for the user to develop their own plugins, and
there is a wide array of user forums for help with coding issues.
Digital image correlation may be thought of as a form of virtual extensometry
in as much that the sample is photographed as the test proceeds. The point
made above about lighting remains, if anything becoming more critical.
However, unlike other forms of virtual extensometry, significantly more
information is captured, although the consequence is that it is rarely feas-
ible to record strain information for the entire duration of the test. Instead,
it is more usual to capture data every few seconds, or at significant points in
the test. As a consequence, DIC should not be considered as a default alter-
native to the humble strain gauge or extensometer: there is a trade-off to be
considered in terms of the set-up time vs. the (usefulness of) the information
collected. Fundamentally, do you need full field strain data?
12
ImageJ is a java-based program developed by the US National Institutes of Health.
13
It can be downloaded from https://imagej.nih.gov/ij.
@seismicisolation
@seismicisolation
104 Understanding the Failure of Materials and Structures
14
The picture on the cover of this book is an example of the output from DIC. The specimen is
in four-point bending, and the resulting rainbow effect is a result of the transition from com-
pressive strain at the top, through the neutral axis in the centre, to tensile strain at the bottom.
Concentrations of strain can be observed around the diamond cut out from the centre of
the specimen. The distribution of strain is, of course, more smoothly linear in reality, but
the output is ‘binned’. The number, and indeed the size, of bins can be varied as required,
depending on the software used. Again, different software packages provide the output in
different ways, but generally the scale will automatically update as you progress through
the specimen. This can be useful to show the evolution of strain concentration but can give a
spurious confidence in the overall magnitude of the strain: early in a test it may appear that
there is a significant strain concentration in a particular location, when in fact the bins are
separated by only a few hundred microstrain. However, it is usually possible to determine the
final maximum tensile and compressive strains and to set the scale for every image to these.
It is then possible to track backwards and forwards when strain concentrations emerge, espe-
cially where these lead to failure, and perhaps just as importantly, when they don’t. An early
lesson for the practitioner is in the interpretation of the colour patterns produced by DIC and
learning to match these to the physical measurement. There are a lot of pretty patterns that
are produced which are essentially meaningless because the whole scale is still in the noise.
@seismicisolation
@seismicisolation
How to Measure Strain 105
FIGURE 9.3
Understanding the speckle pattern for digital image correlation. A subset, i.e., a unit area being
assessed, needs to have a certain number of features. For the purpose of illustration, we begin
with five (a), arranged as the five-spot on a die. Under load (b), the relationship between the
spots changes, and it is the measuring of these changes that gives rise to a determination of
strain in the specimen. If we idealise, this relationship, we give rise to something like (c), but if
we consider (d) we will see that the ideal falls short, because we need the pattern to be unique in
each subset. Therefore, we want the speckles to be randomly positioned within the subset, (e). In
reality, we would like more speckles in the subset, ideally around nine (f).
@seismicisolation
@seismicisolation
106 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
How to Measure Strain 107
The methods for applying speckles are numerous and limited only by
the ingenuity (or overall knowledge of the literature) of the operator. For
some time, the default method for applying speckles was by spray painting
the specimen, usually by applying a base coat of white, and then misting
black paint over the surface. However, this method has fallen out of favour
for two key reasons. Firstly, the technique gives rise to speckles that are
usually quite fine, which is to say that the density of speckles is high, but
that they are very small and lacking in detail, and with limited contrast
to the basecoat. This means that the output is prone to a problem called
aliasing, whereby the computer algorithms that track the speckle-features
get confused, which has an impact on the accuracy and resolution of the
output. The second issue is that it becomes difficult to ensure consistency,
even when the speckle pattern is applied by the same operator. Matthews
(2020) discusses these points in more detail and has proposed a new
approach to optimise the speckle for the application and a new method-
ology for printing a speckle pattern onto a kind of transfer paper. Whilst
not applicable in every situation, the approach is a significant improvement
over previous attempts to print speckles onto a vinyl sticker,15 because the
paper has no measurable impact on the performance of the test. The appli-
cation of the methodology in several other extreme situations gives confi-
dence in its wider use (Matthews, 2020).
There is some debate over the ‘perfect’ speckle pattern. Jones and Iadicola
(2018) have brought together the collected knowledge and experience of the
DIC user community, to produce A Good Practices Guide. This covers all of the
key areas to be considered when undertaking DIC and is essential reading.
With respect to speckles however, there is so much that is context specific,
and such a variety of advice in the literature, that the Guide does not give any
specific advice. Conventional wisdom suggests that the pattern produced
should be ~50:50 black and white, but Matthews (2020) includes an in-depth
study which suggests that in many contexts a ~70:30 balance in favour of
the speckle may be preferable. Perhaps more importantly, Matthews (2020)
provides a methodology for determining the optimum speckle in a given
context in a systematic manner.
A variant of this form of extensometry is called the grid method. Here,
instead of using a randomised pattern of speckles, a uniform array of identical
points is tracked (see, e.g., Pannier et al., 2006). This technique is considered
to be particularly useful when carried out in connection with the virtual
fields method, and for providing information for complex stress states. The
technique is usefully reviewed by Grédiac et al. (2016). It is limited in that,
like 2D-DIC, it does not provide any out-of-plane information.
15
Such as those used for applying company branding to vehicle coachwork.
@seismicisolation
@seismicisolation
108 Understanding the Failure of Materials and Structures
9.6 Summary
This chapter has introduced several methods for the determination of strain.
When dealing with very simple tests, determining strain is relatively easy.
It must be remembered that most methods of determining strain deal either
with a localised measurement, or one that is an average over a smaller or
greater distance. Determining the strain over a large area requires some
thought: simply plastering the specimen with strain gauges leads to a
complicated test that requires the careful handling of a lot of wires and data
capture equipment. However, full field capture is possible and can help when
dealing with anisotropic materials and structures. When choosing the method
for measuring strain, it is important to consider the accuracy required. For
example, some validation procedures require strain to be measured to greater
than 50 με accuracy, which is not generally feasible with DIC.
References
ASTM Standard D5379- 12, 2012. Standard test method for shear properties of com-
posite materials by the V-Notched beam method. American Society for Testing and
Materials.
Boughanem, S., Jesson, D.A., Mulheron, M.J., Smith, P.A., Eddie, C., Psomas, S. and
Rimes, M., 2015. Tensile characterisation of thick sections of Engineered Cement
Composite (ECC) materials. Journal of Materials Science, 50(2), pp.882–897.
Grédiac, M., Sur, F. and Blaysat, B., 2016. The grid method for in-plane displacement
and strain measurement: a review and analysis. Strain, 52(3), pp. 205–243.
Huston, C., 1879. The effect of continued and progressively increasing strain upon
iron. Journal of the Franklin Institute, 107(1), pp.41–44.
Iosipescu, N., 1967. New accurate procedure for single shear testing of metals. Journal
of Materials, 2(3), pp.537–566.
Jesson, D.A., Le Page, B.H., Mulheron, M.J., Smith, P.A., Wallen, A., Cocks, R.,
Farrow, J. and Whiter, J.T., 2010. Thermally induced strains and stresses in cast
iron water distribution pipes: An experimental investigation. Journal of Water
Supply: Research and Technology-Aqua, 59(4), pp.221–229.
Jones, E.M.C. and Iadicola, M.A. (eds.), 2018, A good practices guide for digital image cor-
relation. International Digital Image Correlation Society.
Matthews, S.J., 2020. Speckle pattern control in the application of digital image correlation
for detecting damage in helmets (Doctoral dissertation, University of Surrey).
Morice, P.B. and Base, G.D., 1953. The design and use of a demountable mechanical
strain gauge for concrete structures. Magazine of Concrete Research, 5(13), pp.37–42.
Pannier, Y., Avril, S., Rotinat, R. and Pierron, F., 2006. Identification of elasto-plastic
constitutive parameters from statically undetermined tests using the virtual
fields method. Experimental Mechanics, 46, pp.735–755.
Thomson, W., 1883. Electrical units of measurement. Popular Lectures and Addresses,
1(73), pp.73–136. @seismicisolation
@seismicisolation
10
Strain Rate Dependence: Why It
Matters How Fast You Test
10.1 Introduction
Imagine an egg. Assuming that it is fertilised, leave it under the hen that laid
it (a temperature of 35–40.5°C) for approximately 21 days and it will hatch,
and a newborn chick will emerge. Now take a box of similar, albeit unfer-
tilised eggs. Place one in a pan of water boiling at 100°C for 3 minutes and
you will have a delicious snack with the white almost fully cooked and the
centre runny. Leave it a little longer and it will be hard-boiled. Take another
egg. Place it in vinegar for one week (replacing the vinegar after the first 24
hours) and you will remove the eggshell and make the egg become some-
what rubbery. Leave one on the kitchen counter at room temperature for a
couple of months, and you will have something smelly and inedible.
This, in short, presents the fundamental problem with accelerated ageing
to predict long-term performance. This is not to say that accelerated ageing
tests are impossible, or that they should not be performed, but rather that
they need to be considered with care: chemistry and physics conspire to
change the rules when we step outside of ambient conditions.
But this book is about breaking things –why should we care about ageing
samples? On the one hand, one of the reasons that mechanical character-
isation is undertaken is in order to ensure that components have behaved
themselves, are behaving themselves, and will behave themselves. Such aged
specimens are often tested to see what the residual capacity of the specimen
is in order to put limits on the lifetime of the component as the materials
degrade. On the other, the problems associated with accelerated ageing are
also to be found when considering strain rate, and that is firmly in the court
of mechanical characterisation.
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-10 109
110 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
Strain Rate Dependence: Why It Matters How Fast You Test 111
1
Noting that there need to be slip planes free for them to move, c.f. work hardening which
prevents dislocations from moving.
2
Recalling that whether alloyed or simply present as contamination, there may be atoms of
other elements present in the lattice structure that are of a different size to the parent, so that
they disrupt the lattice, or sit in interstitial spaces. These will move under load.
@seismicisolation
@seismicisolation
112 Understanding the Failure of Materials and Structures
10.6 Summary
Materials will show a strain rate dependence of some sort or another. This
will be a function of the microstructure of the materials and their ability to
deform under certain conditions. It is important to factor this in when pre-
paring to test materials, but be careful when looking at the literature to deter-
mine what sort of strain rate you should be applying to your specimen. The
reference may be very clear as to what strain rate was applied in a specific
situation, but it’s important to remember that not all materials will behave in
the same way.
@seismicisolation
@seismicisolation
11
The Use of Statistics
There are three kinds of lies: lies, damned lies, and statistics.
– Unknown
11.1 Introduction
The first quote at the beginning of this chapter has some mystery about it,
in that it is not certain who first coined the phrase. Mark Twain attributed
it to Benjamin Disraeli; some think it was a favoured aphorism of Arthur
Wellesley, better known as the Duke of Wellington. Other famous politicians
and learned people have also been considered as the original source.1
Whenever it was first stated, today it is misused: it is frequently used to jus-
tify fake news, in the modern idiom, or to denigrate sound science. Statistics,
like any tool, can be misused either through a lack of understanding, or to
deliberately obfuscate a point, but when used properly can uncover meaning
within a mess of data.
The term statistics is much changed from its original meaning. As with so
much to do with mathematics, the earliest example of what we would take to
be statistics is to be found in Arabic writings, in this instance of the 9th century.
1
In fact, the phrase may be a recasting of a comment made in a legal context about witnesses.
It is usually attributed to the judge Sir William Bramwell (later Baron Bramwell of Hever).
“There are three kinds of witness. Simple liars, damned liars, and experts.” He’s also supposed
to have finished this statement by saying “And then there’s brother Fred”, speaking of his
brother Sir Frederick Bramwell, an eminent consulting engineer. The phrase was quoted in the
26th of November 1885 edition of Nature. The article explains “He did not mean that the expert
uttered things which he knew to be untrue, but that by the emphasis which he laid on certain
statements, and by what has been defined as a highly cultivated faculty of evasion, the effect
was actually worse than if he had” (Nature, 1885).
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-11 113
114 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
The Use of Statistics 115
2
Not, of course, to be confused with the sigma used for strength.
@seismicisolation
@seismicisolation
116 Understanding the Failure of Materials and Structures
of the population, squared, and this sum total divided by the size of the
population, N, i.e.:
FIGURE 11.1
Possible outcomes from flipping a coin ten times in a row, presented as a normal distribution.
(a) Total number of combinations for a given number of heads or tail; (b) the distribution,
indicating standard deviations from the mean; (c) the distribution presented as a cumulative
probability; and (d) the distribution presented as a probability density.
@seismicisolation
@seismicisolation
The Use of Statistics 117
lie within ± 1 s.d., 95% within ± 2 s.d., and 97.5% within ± 3 s.d.3 Hence,
there are two things that we can take from this. Firstly, that the more that we
move from the norm, the more of the population we capture. We can turn this
around and say that the outliers will always be difficult to assess, because
3
As an aside, this is where the six-sigma management methodology is derived from. This meth
odology has two guiding principles. Firstly, that any operation that gives rise to outcomes that
are more than three standard deviations away from the mean need investigating. Secondly,
that a company’s output should come down to less than four failures in one million operations,
i.e. six sigma. This could be a product, be it a loaf of bread or a car, or it could be money trans-
action on a website, or a phone call with someone at a support desk.
@seismicisolation
@seismicisolation
118 Understanding the Failure of Materials and Structures
they don’t turn up very often. The larger the population, the more likely it
is that we will see an extreme, but we need to be careful not to give undue
weight to this extreme, given the increasing number of satisfactory outcomes
that were required to find this extreme outlier.
Secondly, we do need to be a little careful when applying a continuous
distribution to discrete data: in the example that we are looking at the distri-
bution implies that we could look for examples of 3.5 or 4.75 out of 10 for a
series of heads or tails. This is of course impossible.
The normal distribution is important but implies that the population of
data to be considered will be well behaved and give rise to something that
is perfectly symmetric. This may or may not be the case in reality, and it is
possible that a population of data will not follow the normal distribution.
However, there are other tools.
4
This list as presented in the original paper.
5
These last two presented only in an appendix in the original paper.
@seismicisolation
@seismicisolation
The Use of Statistics 119
σ − σu m
( )
(− )
σo
Ps V0 = e
(11.2)
σ
( − )m
σo
Ps = e (11.3)
1 σ
ln( ) = ( )m (11.4)
Ps σo
1
ln ( ) = m ln σ − m ln σ o (11.5)
Ps
@seismicisolation
@seismicisolation
120 Understanding the Failure of Materials and Structures
FIGURE 11.2
Illustration of a Weibull plot, demonstrating the determination of the Weibull modulus and the
characteristic value.
A line of best fit through the data should yield a straight line with an
R2 correlation better than 0.9.6 There are two things that can affect both
the fit and the ultimate calculation of m and σ0. The first is that a data set
may have a slight tail to it, in terms of a small cluster of similar valued
specimens. There can also be a corresponding ‘neck’, which is to say a
small cluster of similarly high values. Under these circumstances, when
presenting information, it can be acceptable to de- select some of the
values in the cluster, when applying the fit, so long as this action is stated
clearly. This should, however, be treated as something of a last resort and
there must be a sufficiently compelling reason to do so. A second reason
that the data might not yield a good fit, is that two, or potentially more,
mechanisms may be in competition. For example, Smith, Mulheron, and
co-workers have applied Weibull statistics to the performance of cast iron
water main (Belmonte et al., 2007, 2008; Jesson et al., 2013). With respect
to smaller-diameter pipe (typically of the order of 100 mm in diameter,
but with examples up to 200 mm in diameter), it is possible to observe
the effect of the competition between ab initio defects7 and the impact of
corrosion. This example also yields a useful philosophical point –just
because a material has a high value for m, it does not mean that a material
6
This is rather arbitrary, and based on personal experience, however this suggests that a correl
ation of 0.999 or even 0.9999 is possible with a well-behaved set of samples. Anything below
0.9 should be treated with suspicion, and everything from the samples to the experimental
equipment double checked.
7
See Chapter 4.
@seismicisolation
@seismicisolation
The Use of Statistics 121
8
www.keramverband.de/brevier_engl/5/3/3/5_3_3_4.htm, accessed 02/04/2020.
@seismicisolation
@seismicisolation
122 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
The Use of Statistics 123
9
Précised: An economist, a logician, and a mathematician are on a train. They cross a border and they
see a brown cow standing in a field from the window of the train; the cow is standing parallel to the
train. The economist says, ‘Look, the cows here are brown.’ The logician says, ‘No. There are cows here
of which at least one is brown.’ And the mathematician says, ‘No. There is at least one cow in Scotland,
of which one side appears to be brown.’
@seismicisolation
@seismicisolation
124 Understanding the Failure of Materials and Structures
kind and, if we’re lucky, the spare one will be our favourite flavour. But if
we were to open 12 boxes, would we see one box having the ‘spare’ for each
flavour? No –because there were not an equal number to begin with. To
start, the sweets are not perfectly distributed between boxes, and whilst some
quality control is conducted to provide some sort of parity, not all flavours
are equally represented in the mix to begin with. Some flavours are more
expensive/difficult to make, and so fewer are produced: there is already an
inherent bias in the distribution. Depending on how the sweets are brought
together, there may also be mixing issues, in that some sweets may slip to the
bottom or rise to the top of a mix and so there are more or less of a particular
kind depending on where the box falls in the packing order.
If we wanted to investigate this more thoroughly, we would need to for-
malise a hypothesis, perhaps suggesting that there is an absolute upper limit
to the number of particular sweets in the tin, or perhaps suggesting that there
are only ever a certain number, or, taking the wish above, that on average
there will be seven of each type of sweet. How many tins would we need to
open to prove or disprove the hypothesis?
First, we need to develop our hypothesis, and then determine the null
hypothesis (H0) that we wish to test.
If we assume that all the boxes have, on average, seven of a particular kind
of sweet, this would mean that some have more, and some less, but we are
assuming that the mean =the mode =the median, i.e., that most boxes contain
seven, that the distribution is a ‘normal’ one, so that seven represents the centre
of the distribution, and hence the largest number of tins will have seven. We can
only have integer values, and we’re expecting a reasonably tight distribution, so
we’ll say that the standard deviation is 1. On that basis, if we opened 1000 tins
of sweets, we’d expect to get 400 tins with exactly seven sweets, 240 tins with
six sweets, and another 240 with eight. There will be 55 each with five and nine,
and another five each with four and ten. In theory, that is; in practice, we might
need to open more to get these proportions to come out to this level of accuracy
and there is a certain amount of rounding going on here. Properly there is a non-
zero probability of getting a box containing any other number between zero and
fourteen, but we can perhaps assume that the tails, representing chances of 1 in
10,000 or greater, are eliminated by the quality control process.
But what if we made all the same assumptions above, but we believe that
the mean is actually four? The same spread, so our lowest possibility is one,
and the highest is now seven. Given the numbers of tins that are sold every
year, if we opened any one tin, we wouldn’t be able to tell what kind of dis-
tribution it came from. So how many tins would we need to open to make
a determination? This is where ‘power’ comes in. Variants turn up in many
places, so you may have come across this elsewhere already, in another guise.
Let us break this down into the steps to be taken.
Firstly, there are two opposing theories: the probability that there are seven
of a particular kind of sweet in the tin is either 40% or 0.5%. The difference
here is approximately two orders of magnitude, so even if we were a few
@seismicisolation
@seismicisolation
The Use of Statistics 125
defined as the value such that 95% of the values assessed lie above this value,
with the 5% below being considered defectives (British Standards Institute,
2023; CEN CR 13901, 2000). This is a design parameter. The term ‘margin’
is used to define the difference between this characteristic strength and the
mean. The characteristic strength in this case sits at 1.64 standard deviations
below the mean.
But there is also the question of how much material is available to test.
As intimated in Chapter 2, we cannot test more than one of a large and/or
unique structure, and indeed we may not be able to test that structure in a
meaningful way. Instead, the approach outlined in Figure 2.1 is the Rouchon
pyramid. We ensure that a material is sufficiently well understood so that
predictive models are based on extremely accurate data. Of course, this is
all very well when you have a material in mass production, or one that is
reasonably well understood already. But what about the case where the
material is novel and is not available in significant quantities? Jesson (2005)
is an example of such a material: four different modified silicas, produced as
nanoscale powders to be incorporated into a resin system. The nanosilicas
were produced in a university laboratory and were available in quantities of
the order of 10s of grams. This precludes a completely robust testing regime
requiring kilograms of material, not to mention overcoming the issues with
dispersing the particles in the resin system in the first place.
At some point it is necessary to settle on a number of specimens to test, and
on a justification for that number: 6, 10, or more? If you can only test a single
specimen, what can you do to maximise the value of that test?10
11.6 Summary
Statistics is as valid a part of engineering sciences as any other form of
mathematical analysis, and indeed is a fundamental part of any aspect that
requires a probabilistic answer. In the context of failure, there are several
reasons to turn to statistics. There are many materials that are not especially
homogeneous, and so failure is dependent on the probability of a particular
flaw being present in a particular location. In other contexts, statistics can
be used to provide the boundary conditions for the design of a structure: in
some circumstances, a large, expensive structure will be built to withstand
a 1 in 25, a 1 in 50, or a 1 in 100 event, which is to say an event that happens
once in 25, once in 50, or once in 100 years. Such an event could be an earth-
quake, a storm, or a confluence of tides, for example. Whatever the event, the
expectation is that there will be variations in the intensity of the event, but an
extreme example of the event will happen infrequently. Using statistics, it is
10
Actually, that’s always a good question to ask, no matter how much material you have avail
able for testing –there is always a cost to manufacturing a sample, whether overt or implied.
@seismicisolation
@seismicisolation
The Use of Statistics 127
possible to provide some indication of how frequently the critical level event
will occur and to design accordingly. The caveat, of course, is that such crit-
ical events do not occur to an exact schedule.
This chapter has considered three key topics, which can be applied in most
circumstances. The meaningfulness of a statistical investigation, or the use-
fulness of going to the effort of conducting a Monte Carlo simulation, for
example, will be dependent on the study being undertaken.
In certain circumstances, it may be necessary to look to a particular model
that has been developed to deal with those circumstances. For example, in
ballistics, V50 is typically used, which is to say the strike velocity at which
a given threat projectile has a 50% chance of penetrating a particular target
material (NATO Standardization Agency, 2015). Other verification processes
by statistics are also used, and the practitioner will need to study the avail-
able tools and apply the most appropriate one to the specific setting. For
example, Gumbel statistics, sometimes called the log-Weibull distribution, is
popular in civil engineering, where it is used for investigating extreme value
distributions (see, e.g., Zhang et al., 2014).
It is also worth noting that it may be difficult to collect enough samples to
conduct an analysis to the level required in other settings. A novel material,
for example, might be available in gram quantities and testing is required
to justify scale-up. Under such circumstances, all one can do is conduct the
tests, provide the best assessment possible, and make it clear that there are
boundary conditions, and that the results observed may be overly optimistic
or pessimistic.
This chapter has given only the briefest overview of statistical methods.
As noted at the beginning, it is all too possible to draw unreasonable signifi-
cance from data that have been overly analysed –not necessarily deliberately
manipulated through the inclusion or exclusion of particular records, but just
organised and reorganised and put through various operations. Raw data are
seldom helpful or meaningful and need some level of tidying –much like an
animal being readied for a competition: usually it will look a little scruffy, but
when being judged it will look immaculate. We want people to see the sig-
nificance we see: there is a difference though between, say giving a dog’s tail
a good brush and surgically replacing it with one that is more ‘waggy’. One
way of giving it ‘a good brush’ would be to look for some statistical signifi-
cance, which is very different to adding to, or replacing some of, the data set.
One final note to conclude this chapter on statistical analysis: there are
many tools available, and the choice of tool may come down to personal
experience, specific training or the direction of a mentor, or exposure to the
tool through books such as this one. Whilst designed for a different audi-
ence, the ‘Statistical Package for Social Sciences’, SPSS, is also of use and can
usually be accessed for free, although do be careful to check the provenance
before downloading.
@seismicisolation
@seismicisolation
128 Understanding the Failure of Materials and Structures
References
Belmonte, H.M.S., Mulheron, M. and Smith, P.A., 2007. Weibull analysis, extrapolations
and implications for condition assessment of cast iron water mains. Fatigue &
Fracture of Engineering Materials & Structures, 30(10), pp.964–990.
Belmonte, H.M.S., Mulheron, M., Smith, P.A., Ham, A., Wescombe, K. and Whiter, J.,
2008. Weibull-based methodology for condition assessment of cast iron water
mains and its application. Fatigue & Fracture of Engineering Materials & Structures,
31(5), pp.370–385.
British Standards Institution, 2023. Concrete. Complementary British Standard to BS
EN 206: Method of Specifying and Guidance for the Specifier. British Standards
Institution.
CEN CR 13901, 2000. The use of the concept of concrete families for the production and con-
formity control of concrete; CEN Technical Report; National Standards Authority
of Ireland: Dublin, Ireland.
Garrett, S., 2021. Aspects of the Mechanical Properties of Wood Fibres for Engineered Wood
Products (Doctoral dissertation, University of Surrey).
Goldacre, B., 2014. I think you’ll find it’s a bit more complicated than that. Fourth Estate,
London.
Haddon, M., 2003. The curious incident of the dog in the Night- time. Vintage
Contemporaries, New York.
Hinde, N, 2020. This Man’s audit of quality street chocolates has the internet shook.
Huffington Post.11
Huff, D., 1993. How to lie with statistics. WW Norton & Company.
Jayatilaka, A.D.S. and Trustrum, K., 1977. Statistical approach to brittle fracture.
Journal of Materials Science, 12(7), pp.1426–1430.
Jesson, D.A., 2005. The interaction of nano-composite particles with a polyester resin and the
effect on mechanical properties. University of Surrey (United Kingdom).
Jesson, D.A., Mohebbi, H., Farrow, J., Mulheron, M.J. and Smith, P.A., 2013. On the
condition assessment of cast iron trunk main: The effect of microstructure and
in-service graphitization on mechanical properties in flexure. Materials Science
and Engineering: A, 576, pp.192–201
NATO Standarization Agency, 2015. AEP-2920 Procedures for the evaluation and
classification of personal armour, bullet and fragmentation threats. Allied
Engineering Publication. Brussels: NATO.
Nature, 1885. The whole duty of a chemist. Nature, 33, pp.73–77.
Spanos, P.D. and Kontsos, A., 2008. A multiscale Monte Carlo finite element method
for determining mechanical properties of polymer nanocomposites. Probabilistic
Engineering Mechanics, 23(4), pp.456–470.
Steele, J.M., 2005. Darrell Huff and fifty years of how to lie with statistics. Statistical
Science, 20(3), 205–209.
Vigen, T., 2015. Spurious correlations. Hachette books.
Wei, X., Ford, M., Soler-Crespo, R.A. and Espinosa, H.D., 2015. A new Monte Carlo
model for predicting the mechanical properties of fiber yarns. Journal of the
Mechanics and Physics of Solids, 84, pp.325–335.
11
Short form link: https://tinyurl.com/33p5vjcf; accessed 12/02/2021.
@seismicisolation
@seismicisolation
The Use of Statistics 129
@seismicisolation
@seismicisolation
12
Models vs. Reality
12.1 Introduction
As has been noted in several places in this book, physical testing of any kind
is expensive. One must have equipment to do the test, material to be tested,
and frequently the material must be shaped or in some other way made fit for
testing. Then, as noted in the previous chapter, it is not sufficient to test just
one specimen, there is the need to test a reasonable number and calculate an
average and the likelihood that this average is meaningful. Further, the more
complex a condition to be tested, the more complex the testing set-up needs
to be; see, for example, Hafiz et al.’s (2010) work on mixed-mode testing of
adhesives, and Yussof et al.’s (2017) studies on glass facades.
Models can therefore be incredibly useful. They allow us to explore situ-
ations where we cannot otherwise get information, such as in very high-
speed tests where the event is so transient that it cannot be recorded by
conventional means. They allow us to predict what will happen under certain
conditions, so that in the case of an expensive test, such as the testing of the
containment of an aero-engine in the event of the failure of a fan-blade, only
one iteration is required. They allow us to repeat experiments and change
the variables from the comfort of our desks, iterating to a point where we
can make a single test, that will hopefully vindicate the model that has been
produced. The danger is when we become too comfortable at our desks and
fail to set foot in the lab. It is all too easy for students, who do not usually
spend a lot of time undertaking physical testing, to believe that the results of
@seismicisolation
@seismicisolation
130 DOI: 10.1201/9780367822347-12
Models vs. Reality 131
the physical test are incorrect because they don’t match the theory. It is said
that a modeller always believes their models when no one else does, whilst
an experimentalist never believes their results but everyone else does. A lot
rides on a physical model, and an experimentalist will ask all sorts of (awk-
ward) questions about the assumptions that underpin the model. Life is very
much better when experimentalists and modellers work together and learn
from each other. Both are invaluable and have their place: the model can have
no validity without experimental data to inform it, and experimental testing
to validate it, but the experimentalist can save a lot of time and expense by
targeting their experiments to the gaps highlighted by meaningful modelling.
1
Emphasis here on ‘difference’: each material will have its positive and negative aspects, but
overall it is important to understand that the materials are not directly interchangeable. It is
necessary to understand all facets of the material, including those that might not be important
until later in the life-cycle.
@seismicisolation
@seismicisolation
132 Understanding the Failure of Materials and Structures
at the same time structures made from composites fail unexpectedly and the
material is blamed when it was a design error that caused the problem right
back at the beginning of things.
The Worldwide Failure Exercise has attempted to assess the viability of
a range of composites failure models (Hinton, Kaddour, and co-workers,
2004, 2012, 2013a,b). It has been through three rounds, with each successive
round looking at more complicated loading conditions and more involved
composite layups. In all three cases, physical test data were collected and
used for comparison with the model data produced using the specifications
provided. The disturbing conclusion is that there are no models that perfectly
predict the failure of composite materials, even in relatively simple cases,
although there are some that are better than others. It is therefore necessary
to be conservative when working with composite materials, more so than in
other sectors –however, the point remains that it is possible to do more with
less, if you understand the physical properties that you are working with. In
this instance, a model can give insight, potentially, but cannot give a perfect
solution.
@seismicisolation
@seismicisolation
Models vs. Reality 133
@seismicisolation
@seismicisolation
134 Understanding the Failure of Materials and Structures
12.5 Summary
Whilst the focus of this book is very much on the issues which face the experi-
mentalist and the steps that need to be taken to ensure that a mechanical test
is carried out in an appropriate way, and that the maximum value is gleaned
from each physical test, the ethos is that of ensuring good relations between
models and reality. Physical tests are expensive, and computational power is
becoming increasingly cheap. With a good understanding of the material’s
behaviour, robust models can be developed so that one can minimize the
need for the more difficult tests, and can inform the experimentalist of the
critical tests to be undertaken. Going the other way, fundamental mechanical
testing can provide good-quality data with which to inform models, ensuring
that the models more closely relate to reality.
References
Courant, R., 1943. “Variational methods for the solution of problems of equilibrium
and vibrations”. Bulletin of the American Mathematical Society, 49, pp.1–23.
Hafiz, T.A., Wahab, M.A., Crocombe, A.D. and Smith, P.A., 2010. “Mixed-mode frac-
ture of adhesively bonded metallic joints under quasi-static loading”. Engineering
Fracture Mechanics, 77, pp.3434–3445.
Hinton, M.J., Soden, P.D. and Kaddour, A.S. eds., 2004. Failure criteria in fibre reinforced
polymer composites: The world-wide failure exercise. Elsevier.
Hinton, M.J. and Kaddour, A.S., 2012. “The background to the second world-wide
failure exercise. Journal of Composite Materials, 46, pp.2283–2294.
Hinton, M.J. and Kaddour, A.S., 2013a. The background to part b of the second
world-wide failure exercise: Evaluation of theories for predicting failure in
polymer composite laminates under three-dimensional states of stress. Journal
of Composite Materials, 47, pp.643–652.
Hinton, M.J. and Kaddour, A.S., 2013b. “Triaxial test results for fibre- reinforced
composites: The second world-wide failure exercise benchmark data”. Journal of
Composite Materials, 47, pp.653–678.
Hrennikoff, A., 1941. “Solution of problems of elasticity by the framework method”,
Journal of Applied Mechanics, 8, pp.169–175.
Kaddour, A.S., Hinton, M.J., Smith, P.A. and Li, S., 2013a. “The background to the third
world-wide failure exercise”. Journal of Composite Materials, 47, pp.2417–2426.
Kaddour, A.S., Hinton, M.J., Smith, P.A. and Li, S., 2013b. “A comparison between the
predictive capability of matrix cracking, damage and failure criteria for fibre
reinforced composite laminates: Part A of the third world-wide failure exercise”.
Journal of Composite Materials, 47, pp.2749–2779.
@seismicisolation
@seismicisolation
Models vs. Reality 135
Soden, P.D., Kaddour, A.S. and Hinton, M.J., 2004. “Recommendations for designers
and researchers resulting from the world- wide failure exercise”, In Failure
Criteria in Fibre-Reinforced-Polymer Composites (pp. 1223–1251). Elsevier.
Spendley, P.R., 2012. Design allowables for composite aerospace structures, Doctoral disser-
tation, University of Surrey.
Yussof, M.M., Parke, G. and Kamarudin, M.K., 2017. “Glass stiffness contribution of
flat and curved cable-net supported glass Façade systems”. IPTEK Journal of
Proceedings Series, 3, 531–536.
@seismicisolation
@seismicisolation
13
Biomechanics
13.1 Introduction
The $6 Million Dollar Man is an example of the interaction between medicine
and engineering: whilst essentially science fiction in an otherwise real-life
setting, it details the creation of a cyborg from the badly damaged body of test
pilot Steve Austen. The story is almost a straight secret agent one however,
in that there is no angst or conflict between the organic body, particularly
the mind, and the cybernetic implants and components. This is in contrast
to RoboCop, for example, which can be described as a modern incarnation
of the tale of Frankenstein’s monster. In either case, the implants, which are
designed to augment the physical capabilities of a human being, are beyond
what can currently be achieved. Whilst we have moved a long way from
arguably the most famous prosthetics, the hooks and peg-legs beloved of
pirates, there are still significant limits on what can be achieved.
But people have been using prosthetics for a lot longer than you might
imagine. Vanderwerker (1976) provides ‘A Brief Review of the History of
Amputations and Prostheses’, and states that the first recorded example is that
of Vishpala, and is recorded in the Rig Veda.1 This has all the hallmarks of a
legend, but there are other accounts, slightly more contemporaneous but not
from eye-witnesses, which include Herodotus’ story of Hegesistratus (who
apparently cut off his own foot in order to escape the Spartans, and replaced
it with a wooden one, circa 5th century bce) and Pliny’s description of the
1
It should be noted that there is some dispute over the interpretation, and also that whilst the
story is dated to at least 1200 b ce, the written version of the story dates to at least 900 years
after this. In all probability the events depicted are completely mythic, and it is difficult to
know exactly how much truth underpins the story. From our perspective it doesn’t really
matter.
@seismicisolation
@seismicisolation
136 DOI: 10.1201/9780367822347-13
Biomechanics 137
Roman general Marcus Sergius (who lost his right hand in battle during the
second Punic Wars, 218–201 bce, and had an iron one in its place, to which
his shield was strapped). However, more tangible examples, for which there
is archaeological evidence, include the discovery of an Egyptian mummy
with a prosthetic toe dating to at least the 8th century bce (Nerlich et al.,
2000) and a (crude) prosthetic leg from around 300 bce (e.g. Finch, 2011).
One can debate the definition of ‘modern’, but in general terms modern
prosthetics research begins post-World War II with the work done to help the
massive number of amputees arising from that conflict. Prior to this, a lot of
prosthetics were custom made in their entirety, particularly with respect to the
manner in which they were attached. With the increase in demand, prosthetics
needed to be more robust and cheaper to produce. Whilst prosthetics need
to be tailored to the individual, mass production was required to deliver the
numbers demanded. But in many respects, such prosthetics were still rela-
tively crude: compassionate innovators identified the need for prosthetics
that were more like the real thing: a leg is not a solid piece, and we have
articulated toes for a reason, for example. Such complexity requires the
near simultaneous solution of problems relating to articulation of the joints,
linkages (our muscles and ligaments), and control (replacing nerves with
electronics, or somehow allowing our nerves to control the device). This is
not to mention the issues around the materials to be used.
The replacement of lost limbs is also dependent on the ability to repair
damaged skin and muscle tissue, however. This can be a significant issue
when considering the development and use of prostheses, as the inter-
action between biological and non-biological moieties can be compromised
by scarred tissue, which can be prone to significant chafing. There is there-
fore a whole area of research into the mechanical properties of skin tissue,
and the effect of scarring and skin conditions on these properties, as this can
affect whether or not a prosthetic can actually be used. Further to this point,
there is therefore great interest in being able to define the properties of the
materials used where the prosthetic interfaces with the body in order to pre-
vent chafing (so that the prosthetic can be used for extended periods) and
that the environment at this interface does not become hot and sweaty (and
therefore prone to yeast infections).
So far, this Introduction has only considered prosthetics, but there are of
course all sorts of other bio-mechanical solutions that require an understanding
of mechanical properties to predict long-term performance, including plates
to repair broken bones, hip replacements, stents to open blocked arteries –
the list continues. And of course the consideration of biological systems has
gone beyond just improving repairs to the body. In 1942, the prolific science
fiction author Robert Heinlein published a short story in Astounding Magazine
(under the pseudonym Anson MacDonald). The story, ‘Waldo’, was a classic
example of fiction presenting a hither-to unthought of idea in such detail as to
make it impossible for anyone to be able to claim a patent when the concept
@seismicisolation
@seismicisolation
138 Understanding the Failure of Materials and Structures
2
In fact, the story, which is considered to be a relatively minor one in Heinlein’s canon, and one
which adds elements of magic to an otherwise s.f. story, presents not only the mechanical wal-
does, but also the waterbed.
3
As seen in Burning Chrome by William Gibson. The Encyclopedia of Science Fiction notes that the
term “is used without any further need for explanation”.
4
Something which is important to the potential success of the arcology posited in Oath of Fealty
by Larry Niven and Jerry Pournell.
5
Of great significance, and well ahead of current capability, in Ready Player One, by Ernest Cline.
@seismicisolation
@seismicisolation
Biomechanics 139
6
Effectively the ankle, although noting that in a human this would be directly above the heel
which would come into contact with the ground, whereas canines essentially walk on their
fingers/toes directly and transmit force through the leg to the joint, which will always be at an
angle greater than 90 degrees to the floor.
7
C.f. the question of barrelling observed in compression specimens.
@seismicisolation
@seismicisolation
140 Understanding the Failure of Materials and Structures
@seismicisolation
@seismicisolation
Biomechanics 141
13.6 Summary
Whilst in one sense a little removed from the core focus of this book, i.e.,
understanding the failure of materials and structures, the testing of biological
systems and their interfaces with engineering materials is important for three
reasons which are directly relevant:
@seismicisolation
@seismicisolation
142 Understanding the Failure of Materials and Structures
References
Calandra, I., Gneisinger, W. and Marreiros, J., 2020. A versatile mechanized setup for
controlled experiments in archeology. STAR: Science & Technology of Archaeological
Research, 6(1), pp.30–40.
Finch, J., 2011, February. The ancient origins of prosthetic medicine. The Lancet,
377(9765), pp.548–9.
Gargano, M., Rosina, E., Monticelli, C., Zanelli, A. and Ludwig, N., 2017.
Characterization of aged textile for archeological shelters through thermal,
optical and mechanical tests. Journal of Cultural Heritage, 26, pp.36–43.
Leibinger, A., Forte, A.E., Tan, Z., Oldfield, M.J., Beyrau, F., Dini, D. and Rodriguez y
Baena, F., 2016. Soft tissue phantoms for realistic needle insertion: A compara-
tive study. Annals of Biomedical Engineering, 44, pp.2442–2452.
Nerlich, A. G., Zink, A., Sziemies, U. and Hagedorn, H. G., 2000, December 23/30.
Ancient Egyptian Prosthesis of the Big Toe. The Lancet, 356, pp.2176–2179.
Oldfield, M., Dini, D., Giordano, G. and Rodriguez y Baena, F., 2013. Detailed finite
element modelling of deep needle insertions into a soft tissue phantom using a
cohesive approach. Computer Methods in Biomechanics and Biomedical Engineering,
16(5), pp.530–543.
Shi, X., Chen, T., Zhang, J., Su, B., Cong, Q. and Tian, W., 2021. A review of bioinspired
vibration control technology. Applied Sciences, 11(22), pp.10584.
Vanderwerker, E.E., 1976. A brief review of the history of amputations and prostheses.
Inter Clinic Information Bulletin, 15, pp.25.
Warzee CC, D.L., 2001. Effect of tibial plateau leveling on cranial and caudal tibial
thrusts in canine cranial cruciate-deficient stifles: An in vitro experimental study.
Veterinary Surgery, pp.278–286.
Wells, K.L., Pardo, A.D., Parrott, M.B. and Wassermann, J.F., 1997. A comparison of the
mechanical properties of two external fixator designs for transarticular stabiliza-
tion of the canine hock. Veterinary and Comparative Orthopaedics and Traumatology,
10(01), pp.54–59.
@seismicisolation
@seismicisolation
14
Non-destructive Evaluation
14.1 Introduction
When undertaking testing under laboratory conditions, or when undertaking
a forensic examination which allows for the extraction of samples that can be
examined under the microscope, or in some other way destroyed to yield an
answer, the collection of key information becomes if not easy, then certainly
much more straightforward.
When looking at a structure in the real world and attempting to estimate
its current health and future performance, extracting samples is only feasible
under very specific circumstances. Most of the time, it is necessary to collect the
information required by observation, or by some non-destructive approach that
may be limited in accuracy because of unknowns relating to the conformation,
the microstructure, or inaccurately recorded changes that affect performance.
It would be lovely to have a handheld device that you could wave over
anything that you have a vague interest in and get a read out of composition,
condition, any defects, perhaps even an estimate of residual life.1 Maybe by
1
Residual life in this context refers to the service life remaining. This is a function of the like
lihood of a significant event occurring, one which exceeds the capacity of the structure or
component, and of the residual strength of the component which in turn is dependent on the
degradation that has occurred. In this respect, then, it will be observed that the residual life
may be predicted based on an assessment of the degradation that has occurred to date, and
of the degradation that will continue to accumulate. Further thought will point out the two
factors that mean that such predictions are helpful but should not be relied on in the long
term. Firstly, obviously, the longer the period of time that the prediction extends out, the more
likely it is that a catastrophic event may occur. This is compounded by the fact that further
degradation will reduce the residual strength further, and therefore the required magnitude of
a ‘killer’ event will decrease. Secondly, such predictions are based on a series of assumptions
around normal operating conditions, and whilst some effort may be made to provide factors
of safety, degradation can occur for different reasons, and these processes can change rate
depending on conditions, which can change considerably over time. In the scenario presented
here, an estimation of residual life or a prediction of future performance would simply be a
case of a suite of algorithms crunching the numbers based on the measurements made by our
little handheld gizmo.
@seismicisolation
@seismicisolation
DOI: 10.1201/9780367822347-14 143
144 Understanding the Failure of Materials and Structures
the time the era of Star Trek rolls around, we’ll be able to achieve a sensor
package dense enough to achieve all the different things that are required. Or
possibly not –the challenges facing such a tool are numerous and range from
the complex to the mundane (see, e.g., Rainer et al., 2017). Different materials
require different approaches to work out what is going on. Further, not all
places are as accessible as others –our handy little tricorder can be fitted
into relatively small spaces, and perhaps even poked through a hole at arm’s
length, where the rest of the body can’t follow. Many current NDE techniques
are less forgiving: they are large, unwieldy, and not necessarily suitable to
be taken on-site. Frequently there is a considerable amount of setting up to
do, not the quick wave of a hand –think an early photographer with their
head under a hood and holding up a magnesium flash, rather than a quick
snap with a mobile phone’s built-in camera. Rainer et al. (2017) carried out a
survey of NDE techniques and attempted to find one suitable for a particular
material in a particular setting. The results were not encouraging: whilst there
were a few techniques which might conceivably be able to find the defects of
concern, using them on site was an impossibility.
Hence, there are two approaches to NDE. In an ideal world, we seek primary
data –can we detect sub-critical defects, track them, and determine if they are
growing in a manner that is likely to cause concern between now and the next
survey? Alternatively, we can seek secondary data: indications that all is not
well with the material that we are interested in. For example, whilst we might
be able to use ultrasonics in the same way that the medical profession uses it
to look at a baby in a womb, or to seek information on some internal ailments,
some materials are too dense, or simply have a microstructure that precludes
this. However, we can use ultrasonics in a secondary manner, driving a pulse
(being careful not to introduce too much energy into the system) that can then
be listened for directly, or as an echo, and the speed of sound in the material
measured. If the speed is not what we expect, then we can deduce that there is a
problem developing. However, some caution needs to be used when applying
this technique, as not every instance of a class of material is the same, and
sometimes quite large ranges for the speed of sound in a particular material
can be observed: significant errors can creep if one assumes the property.
The size of the defect we are hoping to detect will be dependent on:
2
Glass-fibre composite materials, especially where the resin system has been matched for
optical transparency, are a good choice for fundamental experiments investigating the the-
ories underpinning behaviour, because the evolution of cracks can be observed without the
need for dye penetrants or other techniques which are required for carbon-fibre composites,
for example.
@seismicisolation
@seismicisolation
Non-destructive Evaluation 145
3
Consider a crack in a windscreen. The crack is sub-critical but could grow due to additional
stresses arising from a period of cold weather, or through fatigue as the vehicle is being driven,
or when additional load is transmitted when the vehicle gets a jolt from a speedbump or a
pothole. Such a crack can be repaired relatively easily, and often an insurer will arrange for
this to be done practically free of charge, rather than waiting for an inevitable claim when the
windscreen is completely, or at least legally, unusable. However, we also know that whilst we
can see the crack or chip, glass being good for this, there is a minimum size that will lead to a
repairer being dispatched. If the crack is below this, then we just have to live with it, and keep
an eye on it to see if it grows.
@seismicisolation
@seismicisolation
146 Understanding the Failure of Materials and Structures
corroded, to the extent that the webs were little more than lace connecting the
top and bottom flanges. When the train crossed the two halves of the bridge
were effectively decoupled and there was counter-torque as the bridge twisted.
With hindsight, the problem is obvious, and could probably have been
prevented by using a different design. With hindsight, the problem could
have been detected earlier, if the right NDE tools had been used during
inspections.
In the case of Stewarton bridge, the SHM process was limited to visual obser-
vation, and so the SHM process stalled at step 1. However, given the age
of the structure, this is not so surprising, as no monitoring methods were
applied when the structure was manufactured, or when it was repurposed.
Such devices are now considered essential, especially for assets such as
subsea cables, i.e., intrinsically expensive, difficult to place, and extremely
difficult to monitor by normal periodic inspection methods.
4
Here, observation is defined literally (“What do my eyes tell me about the state of this struc
ture?”) and figuratively (“What do the results of this inspection tool tell me about the condi-
tion of this structure?”).
5
This is potentially more problematic when detection arrives through a tool such as a measure
ment of the speed of sound, which indicates that the Young’s modulus has changed, due to the
presence of damage, but provides no further information.
@seismicisolation
@seismicisolation
Non-destructive Evaluation 147
14.5 Summary
No matter how advanced our non-destructive evaluation techniques become,
it’s unlikely we’ll be able to eliminate all lower- rank/
red-shirt fatalities
in Star Trek. However, there is much that can be achieved, beyond simply
looking, and not necessarily believing that what we are looking at is all that
is to be evaluated. Non-destructive evaluation is the way in which we can
collect evidence from structures that are in use, or which are somehow dif-
ficult to assess, although even here, care must be taken to ensure that the
condition of structural elements that are hidden from sight are not forgotten
when predicting future behaviour.
References
McCormick, N. and Lord, J., 2012, November. Digital image correlation for structural
measurements. Proceedings of the Institution of Civil Engineers-Civil Engineering
(Vol. 165, No. 4, pp. 185–190). Thomas Telford Ltd.
RAIB (Rail Accident Investigation Branch), 2010. Derailment of a freight train near
Stewarton, Ayrshire, 27 January 2009.
Rainer, A., Capell, T.F., Clay-Michael, N., Demetriou, M., Evans, T.S., Jesson, D.A.,
Mulheron, M.J., Scudder, L. and Smith, P.A., 2017. What does NDE need to
achieve for cast iron pipe networks?. Infrastructure Asset Management, 4(2),
pp.68–82.
@seismicisolation
@seismicisolation
15
Questions for the Future
15.1 Introduction
Following the heady days of the Space Race, humanity’s forays into the
wider solar system have been limited to remote probes and landers. The last
person to walk on the Moon, Gene Cernan, did so in 1972. The fascination
with returning to the Moon and indeed sending a crewed mission to Mars
and other parts of local space where semi/permanent habitats can be built
is increasing once more and has the potential to move beyond the realms of
science fiction and into reality.
Back on Earth, new materials are being developed based on fungi –no
longer limited to food or a threat to the integrity of a structure, fungi are
finding frequent use as replacement for dense foams, such as those used for
surf boards, but more recently have been developed for uses as diverse as
films (Abhijith, 2018) and construction materials (Xing, 2018).
Aside from intrasolar exploration, humanity has been dealing with the
extremes that nature has been presenting us with: whilst changes in averages
tell one aspect of the story, they can cause us to overlook other parts. Too much
of something at the wrong time is as bad or worse than too little. Excessive
rain and flooding in the winter, too little rain in the summer, high winds at
any time… The structures that we build are facing the upper limits of their
capacity on a more regular basis. How can understanding the mechanisms of
failure help us in preparing for the future?
This book has avoided exam-style questions at the end of each chapter,
as are found in many textbooks: the purpose of this book is not to test the
reader on what they have learned but rather to provide a guide to practical
work that will be undertaken by the reader in the future, and to equip the
analyst, critical friend, or simply the interested, with skills to interrogate the
output of others’ work and to assess the robustness of the work that has been
@seismicisolation
@seismicisolation
148 DOI: 10.1201/9780367822347-15
Questions for the Future 149
i. Do I believe these results? Are they credible with what I know of this
material or structure? How can I verify these results?
ii. Is this the most appropriate test that could have been carried out?
What might other tests tell us? Is the cost of further testing justified
within the scenario?
iii. Given what I am looking at, should I be reviewing other instances of
the use of this material? Has my understanding of how this material
or structure behaves been revised?
The following sections present some thought exercises for the reader.
There are no right or wrong answers to the questions posed, but in pro-
viding an answer the reader may be in a position to provide the answer, a
definitive solution that will influence the area of the scenario considered
going forward.
1. The concept has been around since the 1960s (Myers and Swiatecki,
1966; Meldner, 1967);
2. Breakthroughs are only beginning to be seen now (e.g., Oganessian,
2017); and
3. Who knows how long it will be before there is enough available to be
able to characterise physico-chemical properties?1
1
Predictions can be made, and no doubt will be used to justify funding, or the denial of, projects
leading to the generation and ‘mass-production’ of stable super-heavy elements, but the kinds
of quantities required to begin testing of properties and confirm the value of potential elem-
ents will be of the order of thousands of atoms, assuming excellent isolation and recovery
protocols. The generation of usable quantities really is beyond comprehension at this time.
@seismicisolation
@seismicisolation
150 Understanding the Failure of Materials and Structures
New elements aside, there is still much ado in the world of materials,
including the uptake of fungi, new methods of functionalisation leading to
new products (e.g., ultrathin solar panels; Panagiotopoulos et al., 2023), and
new ways of processing familiar materials (e.g. hierarchical nanostructured
aluminium; Wu et al., 2019).
What opportunities might be available in the future, given the rise of a suitable
new material?
Is there a danger of a new route to failure that is overlooked because we do not know
enough about how the new material behaves?
How do we adapt our existing standards and characterisation processes to deal
with new materials?
@seismicisolation
@seismicisolation
Questions for the Future 151
2
There are three other related materials, glue laminated timber, laminated strand lumber,
and laminated veneer lumber, that are used in construction, but CLT is by far the most com-
monly used.
3
These settlements frequently require the development of new materials that can be used in
the Martian environment, especially where cities are built inside ‘domes’, i.e., tents. Others are
buried deep underground.
4
The terms 3D printing and additive manufacturing are often treated as interchangeable, but
3D printing is only one form of additive manufacturing, whereas additive manufacturing
takes in some seventeen different techniques. There have been several attempts to provide an
overarching standard for additive manufacturing, but these have ultimately failed, because
of the complexity of identifying key features that can be found in all seventeen techniques –
beyond the simple case of building a component by adding material rather than by removing
material from a bulk volume.
@seismicisolation
@seismicisolation
152 Understanding the Failure of Materials and Structures
5
A further advantage here is that robotic printers can be sent in advance of a crewed mission
and a substantial base camp prepared for the arrivals to walk straight into and occupy.
@seismicisolation
@seismicisolation
Questions for the Future 153
15.6 Sustainability
Sustainability is becoming a more significant factor in design and oper-
ation. The term sustainability is in danger of becoming diluted however, for
reasons ranging from overuse to loose definition of what it means, to lack of
understanding. Context plays an important part in understanding if some-
thing is sustainable or not. For example, buying a brand-new car every year
is not sustainable for most people, but we cannot simply say that it is sus-
tainable if the purchase comes round every ten years. It might be financially
feasible, but there are other factors to consider, and then again, it might not
be financially feasible because of a move to a larger house. What kind of car?
How big? How much is it used?
In the context of materials, we can say definitively that there is no such
thing as a sustainable material. However, we can talk about the sustainable
use of materials. This requires us to think about the whole life cycle, from
the extraction of the raw materials, their formation into a usable product, its
disposal, and its potential for reuse. Materials from nuclear powerplants will
typically need to be sequestered, so we could think in terms of the potential
for carbon capture, although the embodied energy is effectively lost to the
system.
In the context of failure, what can sustainability teach us, and what infor-
mation do we need to be able to provide, to help people assess the sustain-
ability of a project?
In the context of failure of materials and structures, interest is most likely to focus
on the ‘use phase’: what do we need to do to better understand the lifetime of a struc-
ture or component?
Given that sustainability is the interaction of social, financial, and environmental
concerns, how can these factors be balanced to bring about the best income?
In the context of sustainability, can failure ever be a good thing?
@seismicisolation
@seismicisolation
154 Understanding the Failure of Materials and Structures
References
Abhijith, R., Ashok, A. and Rejeesh, C.R., 2018. Sustainable packaging applications
from mycelium to substitute polystyrene: A review. Materials Today: Proceedings,
5(1), pp.2139–2145.
Gardner, L., Kyvelou, P., Herbert, G. and Buchanan, C., 2020. Testing and initial verifi-
cation of the world’s first metal 3D printed bridge. Journal of Constructional Steel
Research, 172, pp.106233.
ISO/
ASTM 52939:2023, Additive manufacturing for construction— Qualification
principles—Structural and infrastructure elements, ISO Committee TC 261.
Meldner, H., 1967. 10. Predictions of new magic regions and masses for super heavy
nuclei from calculations with realistic shell-model single-particle Hamiltonians’:’.
Nuclear Chemistry Annual Report 1966, pp.157–160.
Myers, W.D. and Swiatecki, W.J., 1966. Nuclear masses and deformations. Nuclear
Physics, 81(1), pp.1–60
Oganessian, Y., 2017. Discovery of the Island of Stability for Super Heavy Elements.
Proc. IPAC’17, pp.4848–4851.
@seismicisolation
@seismicisolation
Questions for the Future 155
Wu, G., Liu, C., Sun, L., Wang, Q., Sun, B., Han, B., Kai, J.J., Luan, J., Liu, C.T., Cao,
K. and Lu, Y., 2019. Hierarchical nanostructured aluminum alloy with ultrahigh
strength and large plasticity. Nature Communications, 10(1), pp.5099.
Xing, Y., Brewer, M., El-Gharabawy, H., Griffith, G. and Jones, P., 2018, February.
Growing and testing mycelium bricks as building insulation materials. In IOP
Conference Series: Earth and Environmental Science (Vol. 121, pp. 022032). IOP
Publishing.
@seismicisolation
@seismicisolation
@seismicisolation
@seismicisolation
157
Index
G N
Gimli Glider, 1–2, 8 Nanocrystal structures, 41
Grain boundaries, 41 Neutral axis, 84
Graphite flake, 78 Non-linear elasticity, 51
Grid method, 107 Normal distribution, 115–118
Notch sensitivity, 59
H Null hypothesis, 124
Haptic devices, 138
P
I Paris-Erdogan Law, see Paris Law
Paris Law, 75–76
ImageJ, 103
Plane strain, 87, 89
Imhotep, 12–13
Plane stress, 87–89
Island of stability, 149
Plastic deformation, 47, 54
Plastic flow, 51
J
Plyscraper, see Plywood skyscraper
J-integral, 65–68 Plywood skyscraper, 150
Poisson’s contractions, 47, 54–55
K Poisson’s ratio, 54–55, 87, 104
Poisson’s reactions, see Poisson’s
Keyhole surgery, 140
contractions
Prestressed, 81
L
Pre-tension, 81
LEFM, see Linear elastic fracture Process of design, 13
mechanics Prostheses, see prosthetics
@seismicisolation
@seismicisolation
Index 159
@seismicisolation
@seismicisolation
@seismicisolation
@seismicisolation