Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Science and Engineering of Cutting: The Mechanics and Processes of Separating, Scratching and Puncturing Biomaterials, Metals and Non-metals
The Science and Engineering of Cutting: The Mechanics and Processes of Separating, Scratching and Puncturing Biomaterials, Metals and Non-metals
The Science and Engineering of Cutting: The Mechanics and Processes of Separating, Scratching and Puncturing Biomaterials, Metals and Non-metals
Ebook923 pages17 hours

The Science and Engineering of Cutting: The Mechanics and Processes of Separating, Scratching and Puncturing Biomaterials, Metals and Non-metals

Rating: 0 out of 5 stars

()

Read preview

About this ebook

The materials mechanics of the controlled separation of a body into two or more parts – cutting – using a blade or tool or other mechanical implement is a ubiquitous process in most engineering disciplines. This is the only book available devoted to the cutting of materials generally, the mechanics of which (toughness, fracture, deformation, plasticity, tearing, grating, chewing, etc.) have wide ranging implications for engineers, medics, manufacturers, and process engineers, making this text of particular interest to a wide range of engineers and specialists.
  • The only book to explain and unify the process and techniques of cutting in metals AND non-metals. The emphasis on biomaterials, plastics and non-metals will be of considerable interest to many, while the transfer of knowledge from non-metals fields offers important benefits to metal cutters
  • Comprehensive, written with this well-known author’s lightness of touch, the book will attract the attention of many readers in this underserved subject
  • The clarity of the text is further enhanced by detailed examples and case studies, from the grating of cheese on an industrial scale to the design of scalpels
LanguageEnglish
Release dateJul 15, 2009
ISBN9780080942452
The Science and Engineering of Cutting: The Mechanics and Processes of Separating, Scratching and Puncturing Biomaterials, Metals and Non-metals
Author

Tony Atkins

Dr. Atkins was Professor Emeritus of Mechanical Engineering at the University of Reading and Visiting Professor at Imperial College, London. He taught and researched in the general field of large deformation flow and fracture of all sorts of materials, including biomaterials. He was a Fellow of the Royal Academy of Engineering, a Fellow of the Institution of Mechanical Engineers, and of the Institute of Materials 3. He sat on various professional institution committees, editorial boards of journals and book series, and has been an advisor on NATO’s Science for Peace Programme.

Related to The Science and Engineering of Cutting

Related ebooks

Technology & Engineering For You

View More

Related articles

Reviews for The Science and Engineering of Cutting

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Science and Engineering of Cutting - Tony Atkins

    Table of Contents

    Cover Image

    Dedication

    Copyright

    Preface

    Acknowledgements

    Chapter 1. Controlled and Uncontrolled Separation of Parts

    Chapter 2. Fracture Mechanics and Friction

    2.1. Introduction

    2.2. Fracture Mechanics

    2.3. Friction in Cutting

    2.4. Muscles

    2.5. Impact Mechanics and Hammering

    2.6. Work of Formation of New Surfaces

    Chapter 3. Simple Orthogonal Cutting of Floppy, Brittle and Ductile Materials

    3.1. Introduction

    3.2. Floppy Offcuts

    3.3. Different Types of Offcut Formation

    3.4. Brittle Offcuts

    3.5. Ductile Offcuts

    3.6. Offcut Formation by Shear

    3.7. Finite Element Simulations

    3.8. Cutting Through the Thickness of Ductile Sheets and Plates: Shearing and Cropping

    Chapter 4. Types of Chip

    4.1. Introduction

    4.2. Energy Scaling

    4.3. Variations in Depth of Cut and Rake Angle

    4.4. Cutting with a Built-up Edge

    4.5. Sawtooth Profile Chips

    4.6. Classification of Chips

    4.7. Wood

    Chapter 5. Slice–Push Ratio

    5.1. Introduction

    5.2. Floppy Materials

    5.3. Offcut Formed in Shear by Oblique Tool

    5.4. Guillotining Edges

    5.5. Drills, Augers and Pencil Sharpeners

    Chapter 6. Cutting with More Than One Edge

    6.1. Introduction

    6.2. Scratching of Low ER/k2 Solids

    6.3. Scratching of High ER/k2 Solids

    6.4. Grinding and Abrasive Papers

    6.5. Scratch Hardness/Scratch Resistance

    6.6. Scratching of Thin Films and Coatings: Pencil Hardness

    6.7. Erosion

    6.8. Definitions of the Coefficient of Friction

    6.9. Engraving, Writing Tablets and Polishing

    Chapter 7. Sawing, Chisels and Files

    7.1. Introduction

    7.2. Knives and Chisels

    7.3. Saw Teeth

    7.4. Files

    Chapter 8. Punching Holes

    8.1. Introduction

    8.2. Quasi-static Piercing with a Pointed Tool

    8.3. Quasi-Static Circular Punching

    8.4. Hollow Punches

    8.5. Arms and Armour

    8.6. Penetration and Perforation of Armour

    Chapter 9. Sharpness and Bluntness: Absolute or Relative?

    9.1. Introduction

    9.2. Tool Materials

    9.3. Manufacture and Sharpening

    9.4. Geometry of the Cutting Edge

    9.5. Measurement of Sharpness

    9.6. Retention of Sharpness: Tool Wear and Machinability/Cuttability

    9.7. Effect of Bluntness and Clearance Face Rubbing on Cutting Forces, FC vs t Intercepts and Subsurface Deformation

    9.8. Cutting Edge Sharpness and Workpiece Critical Crack Tip Opening Displacement

    9.9. Compensation for Bluntness

    9.10. Wiggly Crack Paths Produced by Very Blunt Edges

    Chapter 10. Unrestrained and Restrained Workpieces

    10.1. Introduction

    10.2. Minimum Speed for Cutting Unrestrained Workpieces

    10.3. Comb Cutters

    10.4. Optimum Shape for a Curved Blade

    10.5. Cylinder Lawnmowers

    10.6. Rotary Mowers and Strimmers

    Chapter 11. Cutting in Biology, Palaeontology and Medicine

    11.1. Introduction

    11.2. Biology

    11.3. Palaeontology

    11.4. Medicine

    Chapter 12. Food and Food-Cutting Devices and Wire Cutting

    12.1. Introduction

    12.2. Properties

    12.3. Food Texture

    12.4. The Delicatessen Slicer

    12.5. Wire Cutting

    Chapter 13. Teeth as Cutting Tools

    13.1. Introduction

    13.2. Jaws and Bite Force

    13.3. Occlusion and Contact Mechanics

    13.4. Sharpness and Wear of Teeth

    13.5. Attack and Defence

    13.6. Scaling

    Chapter 14. Burrowing in Soils, Digging and Ploughing

    14.1. Properties of Soils

    14.2. Roots

    14.3. Earths, Sands and Burrows: Hole-making and Scraping by Animals

    14.4. Earthmoving and Ploughing

    14.5. Picks and Crampons

    Chapter 15. Unintentional and Accidental Cutting

    15.1. Introduction

    15.2. Grounding and Collision of Ships: Diverging and Converging Tears

    15.3. Progressive Dynamic Fracture

    15.4. Accidents Involving Cables

    15.5. The Twin Towers

    Appendix 1. Friction Forces on a Wedge-shaped Tool Cutting Orthogonally

    Appendix 2. Friction in Cutting

    References

    INDEX

    Dedication

    For Philip, Richard and Ruth

    Misce stultitiam consiliis brevem:

    Dulce est desipere in loco

    Horace 65–8 BC

    ODES (book 4, poem 12 (a poem on the pleasures of Spring), lines 27–8)

    Mix a little foolishness with your serious moments

    Silliness in its place is charming

    (i.e don't be po-faced all the time)

    Copyright

    Linacre House, Jordan Hill, Oxford OX2 8DP, UK

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands

    First edition 2009

    No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher

    Permissions may be sought directly from Elsevier’s Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions and selecting Obtaining permission to use Elsevier material

    Notice

    No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made

    British Library Cataloguing in Publication Data

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

    Library of Congress Cataloging-in-Publication Data

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

    ISBN: 978-0-7506-8531-3

    For information on all Butterworth-Heinemann publications visit our web site at books.elsevier.com

    Printed and bound in Great Britain

    09 10 11 12 13 10 9 8 7 6 5 4 3 2 1

    Preface

    This book is about the deliberate or accidental cutting of all sorts of materials from ‘soft, compliant and weak’ to ‘hard, stiff and strong’. Examples are drawn from the engineering, physical and biological sciences, history and archaeology, palaeontology, medicine, veterinary medicine and dentistry, and food technology. It is a very broad canvas over which I attempt to demonstrate that separation of one part from another and the factors controlling separation – the mechanics of cutting – are basically the same, whatever the field and whatever the material. There is much to learn by looking at how other people do similar tasks. I am an engineering scientist, interested in why things break, with limited knowledge in biological and other fields, but I have been fortunate for many years to collaborate with members of the Centre for Biomimetics at the University of Reading. Biomechanics is the study of the physical properties of all biological materials (including foodstuffs), and their employment in understanding biological design and function. The subject should not be confused with the complementary field of biomedical engineering whose employment of physical properties of human body parts relates to specific replacements of limbs and organs. Biomimetics (‘mimicking nature’) is inspired by biology and applies biomechanical knowledge to the manufacture of new engineering materials and devices. It may be said, perhaps, that biomimetics was invented at Reading by my predecessor, Jim Gordon, and carried on by George Jeronimidis, Julian Vincent (now at the University of Bath) and Richard Bonser. Professor Gordon also built up contacts with archaeologists, classicists and historians, and that has continued in Engineering at Reading with our group of archaeometallurgists (Henry Blyth, Eddie Cheshire, David Sim and Alan Williams). Having been fortunate to study under Charles Gurney in Cardiff and David Tabor in Philip Bowden’s Laboratory for the Physics & Chemistry of Solids in the Cavendish Laboratory at Cambridge, I am receptive to this sort of wide thinking. The pioneering work on metal cutting at the University of Michigan, Ann Arbor, also rubbed off on me during my tenure in Mechanical Engineering there, as did discussions with Wallace Hirst and Gerry Hamilton at Reading on contact mechanics and wear, for which I am grateful.

    One difficulty in writing an interdisciplinary text is that what may be elementary and well known to a worker in one field is completely strange to others. I have had to have a dictionary next to me when reading non-engineering papers, and have pestered very many colleagues across the world for clarification of lots of topics. I owe a great debt to all those who have advised me on the disparate topics covered in this book and corrected many of my misconceptions. Geoff Rowe once said that I had a habit of ‘setting crossword puzzles for tired minds’. At one stage it seemed that the list of acknowledgements would be longer than the text. Even so, any mistakes in this book are entirely down to me. Even within engineering, some branches have a different language from another: the subject of soil mechanics (geomechanics), for example, developed separately from solid mechanics. Again, I have had to face the problem that different fields use different symbols and different nomenclature for identical things.

    Some ideas about cutting require mathematics for explanation, and biological scientists and others do not always have the mathematical background of physical scientists. The mathematics in this book is not advanced but, to help those to whom modelling and mathematical derivations may appear mysterious, I have tried to be as basic as possible. Detailed workings may be found in original papers. Equally, specialists in the biological and other fields may no doubt consider that I have over-egged the pudding in many places regarding their subjects. The author’s own inadequacies will be evident to all.

    Sir Charles Inglis, in the Proceedings of the Institution of Mechanical Engineers in 1947, said that:

    ‘… Mathematics [required by engineers] though it must be sound and incisive as far as it goes, need not be of that artistic and exalted quality which calls for the mentality of a real mathematician. It can be termed mathematics of the tin-opening variety, and in contrast to real mathematicians, engineers are more interested in the contents of the tin than in the elegance of the tin-opener employed …’.

    In this book we hope to discuss how the tin-opener itself works with such simple mathematics. The reader will become aware that there is, perhaps, less maths in some of the later sections of the book that deal with biology, palaeontology, medicine, food, etc. With notable exceptions, there are fewer workers performing instrumented experiments to provide data to assess the role of cutting and its interaction with biological microstructure. Sometimes the level of taxonomic sampling is not yet wide enough to determine the role played by cutting in biological design and function. Many exciting experiments are waiting to be performed and analysed.

    Unfortunately, there is no space for interesting topics from literature, mythology, art and so on that involve cutting of various sorts: driving of stakes through the chests of Dracula’s victims; King Arthur’s magical sword Excalibur; St George slaying the dragon; Beowulf’s cutting off the head of Grendel’s mother; the sword of Damocles; Odysseus killing the sleeping giant Polyphemus by driving a stake through his eye; Hercules' second labour was to kill the Hydra, a monster with nine heads where, when one was chopped off, two grew back in its place; the oldest of the classical Greek Fates Atropus cut the thread of Lachesis’s life with her shears (Lachesis was the Fate who spun life’s thread and determined its length); the courtly love of Lancelot and Guinevere (King Arthur’s wife) by Chretien de Troyes from the twelfth century involved crawling over a bridge made of swords. A knight with his ‘sword and buckler’ is found in poetry. A buckler is a shield with a boss (from the French bocle, a boss). The word swashbuckler to describe a swaggering bully comes from swash, the noise made by swords clashing or a sword beating on a shield. In Kipling’s The Glory of the Garden, we read that ‘… better men than we go out and start their working lives at grubbing weeds from gravel paths with broken dinner knives …’.

    Museums around the world have collections of cutting instruments or illustrate tools and cutting in various guises: the knife grinder in the Octagonal Room (Tribuna) at the Uffizi Gallery in Florence; among the lozenge-shaped panels from the old campanile now in the Cathedral museum in Florence, the panel entitled Logic by Gino Micheli da Castello shows shears, and an image of David and Goliath appears in relief in the door of the Duomo. A billhook representing winter cutting of wood may be found in the pediment over Minerva’s temple at the Roman baths in Bath. Among paintings that show cutting is David Teniers the Younger’s Interior, Old Woman Peeling Apples at the Fitzwilliam Museum in Cambridge. A strange pair of scissors that will not function is to be found in Salvador Dali’s painting entitled Guillaume Tell: the blades of the scissors in his left hand cannot close up because the thumb and forefinger are already together when the blades are still open. Caillabotte’s Les Raboteurs de Parquet shows workmen finishing a wooden floor by scraping. A lot of arms, armour and decapitations (Medusa; John the Baptist) have been painted and sculpted over the centuries. In an interactive exhibition of modern art in the 1960s Yoko Ono invited people to take part and cut bits off her dress. Man Ray’s Cadeau at the Pompidou Centre in Paris has fourteen carpet tacks stuck in line along the axis of the lower surface of a ‘Gendarme’ flat iron.

    It is curious that teeth in a number of sculptures of heads are not always represented accurately. In the Louvre, for example, there is a stone lion dating from 350 BC from a cemetery at Glyphada (near modern Athens): the teeth are ‘human’ dentition, not animal. The same is true of a number of ceremonial lions guarding doors in China. There is also an oil painting entitled Surprised! by Henri Rousseau at the National Gallery in London showing a tiger in the jungle. Are the teeth correct?

    The cutting of metals is commercially important and there are many admirable books on the subject. There are also books devoted to cutting of wood and plastics (polymers). They contain much practical information and detail of industrial processes that it has not been possible to include here. Apart from numerous empirical formulae for cutting forces, cutting energy and so on, most models of cutting for different materials in such monographs follow what is to be found in the metal-cutting monographs, by and large, namely that the work required for cutting comprises two components: (i) plastic or other irreversible work in forming the offcut or chip; and (ii) work done against friction. Cutting is different from other deformation problems in elasticity and plasticity since after cutting, a single starting body has been separated into a number of entirely separate bodies that are no longer ‘attached’ to the parent body. The work for separation is absent in traditional models of metal cutting because it was believed that it is insignificant. That view is challenged in this book, based on tracing the history of the assumption in original papers (the use of the chemical surface free energy rather than the fracture toughness). Central to the current theme is the idea of separation of parts and that cutting is a branch of elastoplastic fracture mechanics. The cutting of floppy and brittle materials principally concerns fracture: the specific work of separation is not negligible, and calculations for forces and power consumed employ the fracture toughness of the material, not the surface free energy. When significant work of separation is incorporated in analyses for the cutting of ductile materials, as well as the customary plasticity and friction, a number of experimental observations for which the traditional treatment has no explanation, now make sense. Because the work of separation takes place in thin boundary layers contiguous with the cut surfaces, the separation work and remote work are essentially uncoupled, which is why flow fields in the cutting of ductile materials may be estimated without reference to fracture work. However, when cutting forces and power are to be determined, consideration of separation work is necessary.

    The classification of materials into ductile and brittle is based on the behaviour of laboratory-size testpieces. Yet brittle glass can be machined at micrometre depths of cut and ductile steel behaves in a brittle fashion in large sizes. Such behaviour reinforces the concept that cutting is a branch of elastoplastic fracture mechanics because it is a manifestation of the cube-square scaling inherent in fracture mechanics, where there is competition between energies dependent on volume and dependent on area. In cutting a given material, ductile chips are produced at very small depths, but as the thickness of the slice is increased the behaviour becomes less ductile, so that eventually splits form. This is common experience in wood. Ductility (called tenacity in old papers), as well as a material’s strength or hardness, had always been known to influence the ease or difficulty of cutting, but its role was never properly quantified. The use of fracture toughness, as well as yield strength, simply makes that connexion. Both these mechanical properties may be altered independently of each other by different thermomechanical or other treatments so that different samples of the same material may have the same hardness but quite different toughnesses. It is to be expected that they will cut differently.

    Energy methods are employed throughout this book. They permit solutions of problems that might otherwise be difficult: for example, the explanation why cutting with a knife is always easier when sideways motion is coupled with simple pressing-down, and hence the importance of the ‘slice–push ratio’ in cutting. Materials are considered as continua having reproducible mechanical properties, but microstructure is discussed where appropriate. It comes down to the ‘magnification’ at which a material is being looked. Concrete and salami to the naked eye are coarse heterogeneous microstructures; other solids require inspection under the microscope to see the constituents. All biological materials are hierarchical composites of one sort or another and, in trying to understand behaviour, it is important to know the level at which particular properties are controlled.

    The subject and sources of information are scattered over many disciplines and published in a bewildering number of journals. Whole books have been written on just parts of the subject. While I hope I have recorded important papers, I expect that I have missed a number, and can only apologize to the authors. There are far too many references to quote even a small selection. My choices of papers from a given author/school of work are, to an extent, arbitrary, but it is possible to trace other publications from the papers that are referred to. In addition to historically interesting references, some of the ‘working’ references are quite old. I make no apology for that. Early researchers thought carefully about what they did and experimental work was carefully and painstakingly done, often under difficult circumstances: insensitive load and displacement measurement devices; no image recognition schemes for flow fields; no data acquisition and manipulation software; algebra rather than finite element methods (FEM), and so on. It seems to be a disappointing trend that some young researchers are not familiar with the old literature, and know only about things that can be downloaded from a search engine. In consequence, they sometimes reinvent the wheel, and often have the view that if they have used FEM or similar techniques, then it must be all the better for it. What is possible with modern elastoplastic computational models is, of course, truly remarkable, but FEM is not a substitute for thinking and experimentation. Furthermore, FEM requires physical property inputs and they come from experiments. Calibration of FEM models has to be done with care: while computational simulations can explain the results of experiments too complicated to be modelled by simple algebra, the real success comes when FEM is able to predict events ‘blind’.

    Acknowledgements

    Many people have helped and advised me on this book. I must mention David Wyeth, Eddie Cheshire, and all former research and project students; Richard Bonser, Richard Chaplin, George Jeronimidis, Tony Pretlove and other colleagues at Reading; John Frew is especially thanked for experimental assistance over many years. Peter Lillford, Peter Lucas, Gordon Sanson and Julian Vincent have patiently fielded e-mails from me seeking clarification on the biological side. Gordon Sanson has read drafts, made the most helpful comments and prevented my making many biological howlers. What is written down though is entirely my responsibility and I hope that there are few errors.

    Other people to whom I owe thanks are:

    Julian Allwood, Hilary Arnold-Baker, Daniel Balint, Dick Bassett, Roger Bentley, Henry Blyth, Malcolm Bolton, Roy Brigden, Brian Briscoe, Andy Brunner, Tim Burns, Byron Byrne, Peter Chamberlain, Maria Charalambides, Chen Zhong, Tom Childs, Brian Cotterell, Matt Davies, John Dempsey, Coen Dijkman, Peter Dunn, Caroline Ellick, Bill Endres, Roland Ennos, David Felbeck, Paul Fenne, Tony Gee, Giacomo Goli, Roger Hamby, Bryan Harris, Linda Holland, Ian Horsfall, Ian Hutchings, Norman Jones, Dirk Keeley, Kevin Kendall, Tony Kinloch, Raja Kountanya, Hans Kruuk, Brian Lawn, Ming Li, Ken Ludema, Yiu-Wing Mai, Remy Marchal, Adrian Marshall, Paulo Martins, Shelagh McKay, Roy Moore, Sue Mott, Barbara Murray, John Nairn, Kazimierz Orlowski, Andrew Palmer, Lucy Peltz, Gill Pittman, Tracey Popowics, Charles Preston, Tony Pretlove, Richard Rahdon, Jenny Read, Steve Reid, Pedro Reis, Peter Roberts, Liz Robertson, Benoit Roman, Pedro Rosa, Marco Rossi, David Sim, Gerhard Sinn, Stefi Stanzl-Tschegg, Roger Stewart, David Stirling, Hew Strachan, Frank Tallett, Bernard Thibaut, Michael Thouless, Chris Tufnel, John Videler, Julian Vincent, Stephen Walley, Celia Watson, Shelley Wiederhorn, Tomasz Wierzbicki, Alan Williams, Gordon Williams, John Williams, Xianzhong Xu and John Yeo.

    Chapter 1. Controlled and Uncontrolled Separation of Parts

    Cutting, Scraping and Spreading

    The design of structures and components in nature and in engineering usually aims to avoid fracture – at least during life – but there are circumstances where separation of parts is required. These range all the way from the beast of prey tearing apart its victim with teeth and claws, to the manufacture of a precision surface in metal using special cutting tools. Some processes of separation rely on pulling, bending or twisting an object at regions remote from where the object breaks; others load right at the zone of fracture and this includes cutting. Processes of ‘separation’ that are not cutting include pulling corks out of bottles. Some processes that are thought to be cutting are really not: it is a common misconception that ice breakers cut ice fields by splitting; rather, they ride up on the edge of the ice sheet and break pieces off by bending fracture.

    Separation of materials is all around us: in the kitchen (e.g. carving meat, coring apples, grating cheese, peeling vegetables), when eating (on the dinner plate, in the mouth), in carpentry and building (e.g. sawing, planing and drilling wood; cutting bricks and paving stones), in the office (e.g. paper guillotining and shredding, pencil sharpening), in manufacturing (e.g. all metal-cutting operations), in agriculture (e.g. ploughing, harvesting of crops, sheep shearing), in medicine and dentistry (surgery; the drilling of teeth), in nature (hunters, raptors, their prey and defences; teeth and chewing), in shaving, in opening packaging and in war (arms and armour: a spear through ancient armour, depleted uranium missiles through modern tank armour). While, usually, the cutting tool remains undeformed, in the latter field both cutter and target are deformed. ‘Cutting’ is interpreted very broadly in this book, but even so we do not consider flame cutting, liquid jets, abrasive water jet cutting, laser cutting, plasma arc cutting, electrodischarge machining and electrochemical machining.

    Different materials respond differently when cut with a knife, well illustrated by the wide variety of foodstuffs that includes mashed potatoes, boiled potatoes, uncooked potatoes; cooked and uncooked vegetables; stringy vegetables like celery; squidgy food like blancmange or tofu; boiled sweets, fudge and sugar; easy-to-chew high-quality meat, or poor-quality meat with lots of gristle; soft puddings like icecream, or hard puddings like toffee; some are mixtures of hard and soft (crème brûlée); chocolates may have a hard case with a soft inside. Properties may change with time and storage: some fruit has to be stored after picking before it becomes ripe enough to eat. Fresh food and stale food behave differently: when freshly harvested, foods such as carrot or celery are hard and stiff owing to the turgor pressure that pressurizes the composite structure from within (from the Latin for ‘to swell’); turgor pressure is a plant’s internal stressing to keep it erect, among other things. Turgor pressure decreases with time after harvesting, making fruits and vegetables flaccid (from the Latin for flabby) when they become rubbery and bendy. Loss of turgor pressure is why flowers wilt. The condition of food affects how they are dealt with on the plate, their ‘mouth feel’ and how we bite and masticate food.

    A wide variety of different types of implement is found, ranging from butcher’s knives to cheese graters. Kitchen shops offer strange and ingenious gadgets with pointy bits for doing special cutting jobs in the preparation of food. One of the most exotic, perhaps, is a foie-gras cutter. The Swiss have different slicers for potatoes (to prepare roesti) and for apples (for muesli). A mandolin is a device like a wood-plane over the blade of which foodstuffs are sliced, grated or shred depending on the blade. A hachoir is a rocking device for cutting up herbs (from the French hache for axe; hence hatchet). In antique shops may be found old devices such as sugar cleavers (in Victorian times, sugar used to come in big lumps), mechanical apple peelers, nutmeg grinders and so on.

    Experiments in the kitchen can be very instructive about cutting, and the reader is encouraged to do so and get a feel for stiff/compliant, strong/weak, tough, etc., materials. For example, scrape a carrot with a knife and notice the difference depending on the angle of the blade. What controls the depth of cut in a potato peeler? Why is peeling with a knife more wasteful? Can you skin an orange with a potato peeler? Indeed, can you shave with a potato peeler? Are there differences depending on whether the fruit is hard and stiff, or soft and squidgy? What determines the ease of scraping up a portion of butter on to a knife from a block, or a scoop of icecream? What are those serrations that appear on the back of the butter after scraping? What determines the ‘spreadability’ of butter on toast? What is the best way to take the top off a boiled egg? What are those cracks that appear having scraped the back of a spoon across the surface of a table jelly? These are not flippant questions or suggestions: the answers are central to understanding of the mechanics of cutting.

    When we eat with the aid of a knife, fork, spoon, chopsticks or fingers, we often separate (fracture) food into smaller pieces to fit the mouth, where further deformation and fracture takes place before swallowing. Why is it easier to cut when we ‘slice’ across the food as well as simply ‘press down’? Food on the plate will have been previously prepared from larger pieces and/or cooked to make eatable and digestible. Cooking alters the properties of food and distinguishes humans from other creatures. To tell whether potatoes or other vegetables have been cooked for the requisite time, we stab the vegetable in the saucepan with a knife and see how easy it is to pierce, or see whether it can be suspended from the knife. The altered properties revealed by the knife must connect with perception in the mouth and what, for example, al dente means. Similarly, to get food from plate into mouth, we often pierce, indent or perforate the food with the prongs of a fork, the mechanics of which are similar to nailing a piece of wood.

    Cutting may concern big pieces being separated into two or more still-big pieces (sawing logs of timber into planks, slitting metal sheets off rolling mills, cutting wedges from ‘rounds’ of cheese, cutting fruit into segments). In other examples, thin slices or chips are removed from the surface of a larger piece (peeling potatoes, whittling wood, lathe cutting, carving). Sometimes the piece cut off is important (wood veneer, microtomed sections for histological examination); at other times the piece left behind is important (true of most manufacturing processes where the offcut ‘swarf’ is scrap, trimming the edges of bound books); and sometimes both are important (the division of paper sheets into smaller sizes or the slicing and dicing of semiconductor wafers). Sometimes the quality of the resulting surfaces is of particular concern (limits and fits in engineering assembly) but sometimes it does not matter (chopping firewood). Sometimes the same mechanism of cutting may be both undesirable in one situation, yet beneficial in another (erosion versus sandblasting)

    Different types of cutting include:

    • cutting layers or slices from the surface or edge of a body

    • cutting a groove in the surface of the workpiece

    • dividing a workpiece into sections by cutting through the thickness

    • making profiles (e.g. round shapes on a lathe)

    • making some sort of hole down into, or though, the thickness by penetration and perforation.

    Some separation processes using tools are on the borderline between being under control and not (such as in cutting toenails, where offcuts sometimes fly around the bathroom; and in nut cracking, where the aim is to preserve the kernel uncrushed or unbroken, but where the fate of the shell is unimportant). The flexibility of whatever is holding both the cutter and the workpiece can be important: try spreading butter with a very springy knife; observe the deformation of a loaf of bread on cutting. Sometimes there is little control at all (in crushing/ball mills or chipping timber to make wood pulp). ‘Loose tools’ are bullets and shrapnel, or the stream of grits in erosion, abrasive cleaning and sandstorms. Cutting during fault or accident conditions is often uncontrolled. Accidental and unintentional cutting, in the form of scratching, piercing, perforating and tearing, occurs when a nail punctures a tyre, and when supermarket plastic bags are torn by sharp objects. A ship getting holed after hitting a rock is a larger scale example of the same thing, as is the defeat of armour by a weapon but where the process is intentional. An understanding of piercing, cutting and perforation enables better armour to be designed that will defeat the weapon, and vice versa.

    A characteristic feature of some of the controlled cutting considered in this book is that the cutting tool or blade is not deformed when cutting, and remains ready for reuse. The only significant deterioration may be wear and blunting; unless corrected, the quality of the cut deteriorates. Progress in production engineering from the onset of the Industrial Revolution depended on having satisfactory tool and die materials, and heavy rigid machine tools. Wilkinson’s boring machine of 1775 that made possible the manufacture of Watt’s steam engine was a vital development in the history of machine tool manufacture that encompassed lathes and machines to perform drilling, milling, boring, grinding, shaping, planing and so on. Benjamin Thompson (Count Rumford of the Holy Roman Empire) realized the connexion between mechanical work done and heat generated from observing cannon being bored in Bavaria. (Thompson was American but supported the British in the American Revolution and subsequently fled to Europe. He designed the Englische Garten in Munich.)

    ‘The Multifarious Perforating Machine’ is illustrated in Figure 1-1.

    An interesting question is how hard a tool should be to avoid itself being deformed in cutting. Mutual cutting is possible where both tool and workpiece deform (bullets and the target). When it is difficult to insert a woodscrew, the high torque will distort the blade of a poor-quality screwdriver and also cut slivers from the side of the slot in the head of the screw. Cutting tools are usually thought of as hard, stiff solid objects. But when one ship collides with another, both hollow structures deform and may fracture, and this is another example of mutual cutting.

    Many tools can be resharpened and used again. What ‘sharpness’ may, or may not, mean and whether it is an absolute concept is explored later in the book. Sometimes tools are used once only (disposable scalpels) or thrown away when blunt (disposable razors, indexable tool inserts). Hollow needles that pierce the skin and through which liquids may be inserted into the body (hypodermics) or removed (cannulae) may sometimes be reused depending on conditions. Improvements in tool material qualities, and reductions in cost, mean that it is often uneconomic to resharpen tools (few craftsmen these days sharpen and ‘set’ the teeth of wood-cutting handsaws, as saws are relatively cheap at DIY shops).

    Some weapons can be reused (swords, spears, cannon balls, the stones of slingshots); others not (bullets, shells). Sometimes the cutting tool is sacrificial (a bee sting). Sometimes a broken or otherwise defunct tool cannot be repaired or replaced and there must be consequences. What was the effect of their teeth being ground down by sand, picked up with food, on Ancient Egyptians? Animals who lose their teeth in combat or in old age may no longer be able to feed and they die. This raises the whole ‘chicken or egg’ question of the evolution of teeth and animal diet: which led to which?

    It is possible to separate a given solid into pieces by methods not involving cutting tools, by pulling, bending, twisting and so on. A sheet of paper may be torn down the middle as well as cut with scissors. A cotton thread can be snapped in tension by jerking, or cut by scissors or even with the teeth. A plank of wood can be snapped in two by bending, or alternatively cut with a saw. Facial hair can be plucked out with tweezers, but is most often shaved off, except for women’s eyebrows. Holes are usually drilled in wood, but if a drill is not available, they can be made by burning through with a red-hot poker from the fire. Sometimes items separated by tearing will not be exactly the same as those cut into nominally the same shape. The edges of a torn sheet of paper will be rough compared with an edge cut by scissors; burrs may form along the edges of a thin sheet of metal torn into two, which will be absent when cut by shears – unless the shears are in bad condition, when the sheet may fold down between the blades even if they have sharp edges. Whether such differences are important depends on what is going to be done with the separated pieces. Many modern tin cans have ring-pull tops and may be opened simply by pulling, but food cans still exist that have to be opened with a can opener which indents, pierces and then progressively propagates the initial hole by levering around the rim of the tin, thus opening the metal lid. The quality of the edge in either case probably does not matter; its sharpness does.

    Sometimes cutting is not the preferred option (the best long bows are made by splitting the yew, even if an irregular cross-section results, in order to ensure continuity of the wood fibres; spoke-shaven bows have ‘exposed’ grain-ends from which cracks may propagate during flexure). Similarly for split-cane fishing rods: when the nodes are machined off bamboo, the fibres become discontinuous. The best rods have the nodes flattened, so fibres lie in the line of the rod (Vincent, 2008). It is possible to cut the feathers from chickens, turkeys and other birds in preparation for cooking, but that would leave parts of the rachises (Ancient Greek for the spines of feathers) in the flesh, so feathers are plucked off instead (Bonser, 2008).

    In the cutting of solid bodies, the blade creates new surfaces that are exposed from within the bulk of the body. In layered materials, or glued joints or welds, interfaces already exist and, although the mechanics is the same, the specific work of separation along a pre-existing interface will be different from the bulk material. Adhesion between the pages of a book, or between the sides of plastic bags, has to be overcome to permit separation; adhesion is why gauge blocks (Johansson blocks) have to be wrung apart in the workshop. When the adhesion between a deposit and substrate is not too strong, separation is sometimes achieved by scraping (debris adhered on kitchen counter tops, ice on car windscreens, dirt removed from under the fingernails, mud on dock floors, scrapers in sewage plants and in ore dressing buddles). Errors in writing by pen on parchment or paper can, with care, be scratched off and the surface returned to its original state (it is the origin of ‘to gloss over’). Hygienists at the dentist scrape plaque from the surface of teeth. The Roman strigil was an instrument with a curved blade used to scrape sweat and dirt from skin in the bath-house. The scraper used to remove films of paint from window glass is a tool that can be used in both directions, as can a hoe when weeding. Is there a difference in performance depending on direction? When a pile of soil just sits on a concrete base, the effort of a bulldozer or scraper is principally to move soil about with little effort for soil-to-concrete separation; in contrast, more work is required when the machine digs into a pile of soil to cause soil-to-soil separation. Many processes, which may even employ cutting tools, are more processes of ‘dislodgement’ rather than cutting. Is removal of dog hair from sofas by tearing off with sticky tape, or leg hair by waxing, a cutting process? A pin extracting a winkle from its shell is dislodging it rather than cutting, but the pin does have to penetrate the winkle to give a grip. When eating snails, they are held by a plier-like clamp and then are wiggled out with a two-pronged small fork. How the interface is detached (all at once or in stages) is of interest in these situations. Is the device with rollers rotating in different directions, to split open the skins of grapes before pressing, a ‘cutting’ device? Probably not.

    Is a toothbrush (or sweeping or scrubbing brush) a cutting tool? Is a wire brush a cutting tool when taking rust off a steel component? Strongly adhered debris requires more work to flick away, and this is achieved with stiffer bristles and more pressing down of the brush into the surface: the forces deflecting the bristles increase and the bristles store more elastic strain energy before the instant at which the force for detachment is attained in the deflected bristle. Is the tongue a tool when licking icecream from a cone? Certainly an icecream scoop cuts, as does the device for making whorls of butter: it is controlled chip formation.

    Instead of cutting or scraping (in which material is removed), a tool may push ahead a standing wave of material, resulting in a more even spreading of material over a substrate (butter on toast, plastering a wall, painting by brush). Some scraping damages the surface. Wax transferred to clothing from blowing out candles (rather than snuffing them) is difficult to remove: a trick is to use the edge of a piece of ice as a stiff tool to scrape wax away without damage. A painter’s palette knife has varying thickness and hence varying flexibility and is used both for mixing colours and for applying paint to canvas: in action it is the same as a butter knife.

    The way in which words such as cut, scrape and spread are used is not always consistent. The hand scraping of journal bearings is certainly ‘cutting’ into the bearing metal. Engineers' blue is applied on both halves of a split plain bearing and high spots are taken away by the scraping tool that is shaped like an isosceles-triangle screwdriver. Its three edges are very sharp and make excellent pencil sharpeners (Dunn, 2008). Again, a cabinet maker’s wood scraper actually cuts at the very small depths at which it is used, as does a holystone used for scouring the wooden decks of ships.

    So why and when is cutting preferred over other methods? Important reasons for using cutting tools is that control can be introduced into the process of separation, and that there is some precision in the resulting dimensions of the cut piece and its surface finish. The end of a piece of wood broken in bending is rough and splintered; the surface obtained by sawing or planing is much better. Cutting is performed at the location required to give a required size, whereas a snapped piece of wood may break uncontrollably and unpredictably at any location. Furthermore, the forces required for pulling, bending or twisting apart may be impracticably large. Trees in a forest could be felled by bending over and snapping, but the forces would be ridiculously large since the whole tree would have to be loaded and there would be no control. Hence the use of an axe or saw, where the tool concentrates the effort into a relatively small region of material, leaving the majority of the tree unstressed. Limiting the zone in which separation occurs is an important feature of cutting. The sharper the blade the smaller the region where cutting is concentrated and the more effective the effort. Hence paper is cut with scissors if a prescribed path has to be followed. Tearing along a prescribed path is only possible if the material has been weakened along that path by perforating with small holes (postage stamps, lavatory paper) or by creasing; but curved creases are difficult to make. Maps that have been continually opened and refolded eventually fail by fatigue. The ‘teeth’ on sticky tape dispensers make perforations across the width of the tape and permit a chosen length to be unpeeled and removed from the roll. Sharpness is important, as seen when opening envelopes with a paper knife or a finger, with different neatnesses of opening.

    Some solids (called anisotropic materials) have properties that are different in different directions (e.g. paper is not isotropic and attempts to tear across a sheet often result in tears that curve into the machine direction along the manufactured length; try it with newsprint). A well-known example is timber, which is easily split down the grain but virtually impossible to split across the grain. A woven fabric is another example of an anisotropic material: it is easily torn down the warp or weft directions, but impossible to tear at any other angle (if your trousers get caught on a nail, the tear is always L-shaped, and a missed cue gouging the baize in billiards similarly results in an L-shaped tear). So, while the haberdasher may rip fabric to length when selling by the metre off the roll, the dressmaker must use scissors to cut in other directions (known as cutting ‘on the cross’ or ‘on the bias’): tearing, as an alternative to cutting, is a non-starter. The shop assistant may start the rip with a short cut or nick with scissors, which helps to initiate the tear, in the same way that people bite packaging or rolls of sticky tape with their teeth. This suggests once more that the use of sharp tools to concentrate stresses and strains into a local region, rather than diffusely throughout a body, is an important feature in cutting. Other highly extensible materials such as rubber sheet are similarly difficult just to tear, and some thin sheets of plastic used in rubbish bags and packaging for magazines sent through the post just stretch and stretch. Once again, however, nicking the edge often, but not always, enables tearing to take place easily.

    Even when cutting is the preferred method of separating parts, there can be different tools and devices available to perform ostensibly the same task. For example, the Swiss army knife has a variety of blades. Which is the ‘correct’ one for the job? A sheet of metal may be cut by a bench-mounted guillotine, or with hand shears, or by sawing. It may be possible to perform a given operation with different devices; or the same implement can be employed for different cutting tasks. When ‘correct’ cutting techniques are not followed, it may not matter: hedges trimmed with a chainsaw rather than a hedgecutter may still bud out. Choice of tool may relate to the power available. What provides the required effort? Sometimes it will come from the hands and muscles directly: the surgeon in delicate operations; the cabinet-maker with a chisel; the farm hand with a sickle. Feet and muscle power enable digging to be done with a spade in the garden. Sometimes hand forces are increased through the mechanical advantage provided by levers, as in scissors or garden loppers. Various mechanical aids are now available to help aged or infirm people to garden with conventional tools. Aboriginal spear throwers, which project weapons at higher velocity than possible merely by throwing by hand, rely on similar ideas. Work (the same as energy) is done when a force F moves its point of application through some distance δ, where work = Fδ (newton-metres, Nm, called joules with symbol J). The same work is performed when F is high and δ small, or vice versa. That is why a lightweight person at the extreme end of a seesaw can balance someone heavier sitting much closer to the pivot. Thus, by employing a machine (and the garden shears or any system of levers is a machine), a small force at the handle may be made into a big force at the point of cutting, by making the hand force move through a large distance. This will be reacted by a large force over a small distance where cutting is to occur. It is an example of mechanical advantage or changing gear. Inefficiencies in the machine, principally friction in hinges, pivots and bearings, will reduce the available work, as is well known to gardeners who have left tools out all winter in the rain. In a paper guillotine where the long cutting blade is pivoted at one end, cutting is done progressively along the edge rather than all in one go. To cut through the thickness all at once, a big force moving through a small distance (the thickness of the stack of paper) would be required to perform the necessary work, but the same work is done by a smaller force applied at the handle moving through a much greater distance.

    The idea of ‘force times distance’ is the reason why tools are sometimes struck by hammers or mallets to get a job done. It could be that the force required to push a chisel, say, by hand into a workpiece was just too large even with sustained great exertion. Perhaps a weight could be lifted into the air and dropped on to the chisel held against the workpiece. The chisel would enter the workpiece a smaller or greater distance depending on the resistance to penetration. If the weight were too large, or lifted too high, or the resistance were too small, the chisel might go right through in an uncontrolled fashion. Nevertheless, the work done by muscles in lifting a weight over some distance and giving it potential energy has been made to do the work of cutting over a much smaller distance. That is not to say that all the energy has been transmitted upon impact (see Section 2.5). When a hammer is used, muscles in the body accelerate the hammer head as it descends, so that more energy is available than if the hammer were simply allowed to fall on the job. How muscles work, and the power that may be expected from them, is discussed in Section 2.4 and in French (1988).

    When cutting tools are driven other than by muscle power, the drive may just remove the human effort (electric hedge trimmers, powered lawnmowers, powered log splitters, where the jobs could just as well be done by hand tools). Human effort is much reduced by powered devices but the machine has to be purchased and its running costs covered. Animal-hauled ploughs had greater power (i.e. the rate of doing work with units of joules/second, J/s, called watts, W) to cut a deep furrow than did ancient simple scratch ploughs. A driven tool also often speeds things up. Thus a chainsaw is far speedier than a reciprocating saw at felling a tree. The same job done by hand may require a comparable amount of work but takes a far greater time, as human effort is limited by the way muscles work.

    In addition to eliminating human effort and speeding up cutting, the introduction of power drives may have eased problems of dimensional control of products. Thus steam-driven and later electrically driven circular saws replaced reciprocating hand saws and the saw pit. It is remarkable to recall that wood veneer was once made by sawing by hand, instead of as nowadays peeling from a log in a machine.

    Machine tools in a sense deskill hand operations but quicken up time. Where once parts were made by cold chiselling and filing, they are now manufactured by milling on computer-controlled machining centres. Nevertheless, we should always remember the tremendous skill required, and the limited tools then available, to produce the first screw threads and first gauge blocks by hand.

    Opening packaging can be troublesome and sometimes there is a need to stab at it before the contents can be separated, rather like the ‘jab’ of a medical injection. The properties of materials alter when loaded at different speeds (silly putty is an extreme example); they also alter when the temperature changes (the proverbial hot knife through butter) and when the environment changes (a brittle dry biscuit compared with a soggy biscuit after being dipped in a cup of tea). Strain rates and temperatures in some sorts of cutting are high, much greater than in the usual sort of material property testing. They can have subsurface effects. In grinding steel, for example, there are high local temperatures well above 1100°C regardless of cooling. Changes occur to the microstructure within 5–10μm of the surface. In contrast, lapping, polishing and other slow methods of superfinishing affect the subsurface down only to 1μm because the temperatures and forces are low.

    Other interesting questions arise about the speed at which cutting is performed. Beyond the property changes, something else happens with speed of cutting. Even the sharpest scythe will not cut hay or corn if the implement is merely pressed up against the stalks. There is a minimum speed above which cutting will occur. Why? Again, despite the bluntness of its cutting string, a strimmer impacting blades of grass will cut at a sufficiently high speed: battery-driven strimmers fail to cut at low speeds when the battery runs down. Sometimes energy can be stored in a cutting device before doing a job, such as in a catapult or bow and arrow. Flywheels prevent machines from stalling when energy is taken out of the system to perform some operation.

    It appears that in the early days of trying to understand the parameters that influenced cutting, it was believed that the forces and power required depended on some sort of ‘cutting stress’. For example, in a series of machining tests performed by the Manchester Association of Engineers in 1903, cutting forces were related to the area of the uncut chip without, it seems, any worry about the combination of width of cut and depth of cut to produce that area. Furthermore, there was an understanding that ‘tenacity’ (what we now call toughness) was as important in cutting as strength or hardness (they are proportional to one another). People did not know how best to quantify the idea or how to incorporate it in models of cutting. Bengough (in the discussion of Turner, 1909) said that

    ‘… there was the assumption by engineers that the difficulty of cutting or working a metal was a measure of its hardness. Of course, high carbon steel of great penetration hardness was more difficult to machine than a medium carbon steel of smaller hardness; but hardness was not the only factor which entered into the question of difficulty of machining. Another was its toughness, that was the amount of deformation a metal would undergo before fracture, and the toughness was of course measured by the elongation in the tensile machine. A tough metal of low penetration hardness, such as manganese steel, might be more difficult to machine than a steel of greater penetration hardness but less toughness. Boynton had put that point well when he showed experimentally that more energy was used in boring (with a rotating diamond) cementite than corundum, since the former had some slight ductility not possessed by the latter. Within certain limits and with certain steels it so happened that ease of machining and penetration hardness varied directly with the tensile strength, and inversely with the elongation. With very soft steels difficulty of machining increased simultaneously with loss of penetration hardness, owing to the increase of toughness. With hard steels difficulty of machining might increase with increase of penetration hardness and loss of toughness …’.

    Whatever the method of separation into separate parts, the process concerns fracture, and the mechanics of cutting, perforation and so on is a branch of elastoplastic fracture mechanics. Not only yield properties at very high rates of strain (Drucker, 1949), but now also fracture toughness may be determined from cutting experiments, Atkins (2005). The rates and temperatures achievable in cutting can be beyond the range of strain rates achievable by special apparatus such as the split Kolsky–Hopkinson bar. Furthermore, cutting at different rates and temperatures is continuous, in steady state. Therefore, there is time for steady forces to be measured and videos to be taken, at steady rates and steady temperatures. In contrast, most mechanical property measurements are determined in experiments that are transient.

    Chapter 2. Fracture Mechanics and Friction

    Muscles, Impact and New Surfaces

    2.1. Introduction

    This book is aimed at a wide audience. Experience of teaching materials mechanics to students of biology, food science and biomimetics tells me that help is required on mathematical analysis and modelling of problems. The first part of this chapter aims to cover the basics of stress analysis, fracture mechanics and ideas on friction, as these are all central to the ideas in the book. The mathematics should be familiar to engineers, materials and physical scientists, but they, in turn, require help from biologists to understand how biomechanics relates to biological design and function.

    Later parts of the chapter deal with how muscles function, since the force and work necessary for cutting are often supplied by hand or foot when animals attack prey and afterwards chew food in the mouth. Forces for cutting may be provided as a slow push or a fast blow. In turn, that takes us to hammering and impact.

    All cutting generates new surfaces and the final part of the chapter discusses whether this process requires significant work or not.

    2.2. Fracture Mechanics

    2.2.1. Load–deformation curves and stress–strain curves

    To determine the mechanical (strength) properties of materials, a body (testpiece) may be deformed by applying known loads to it, for example by hanging on weights to the suspended body. Deformation may be pulling, compressing, twisting, bending and so on, including combinations of these different ways of loading. The resulting deformation (extension in tension, reduction in height in compression, angle of twist, rotation in bending) can be measured. Alternatively the same body can be deformed by a known amount, and the loads required to achieve the deformation measured.

    Figure 2-1(A) shows various types of load–extension diagrams that may result from such experiments under increasing load or deformation. The simplest is a straight line whose stiffness (the slope of the line) is constant. Alternatively, the lines may be curved up or down with uniformly changing slopes and thus increasing or decreasing stiffnesses. The curve with increasing stiffness (a so-called J-shaped curve) is characteristic of extensible biological tissues. The sigmoid-shape curve (for some rubbers) is stiff at small extensions, less stiff at intermediate extensions and very stiff before fracture. Non-linear load–deformation curves may be caused by the material itself having non-proportional behaviour, or result from certain geometric configurations in structures made of linear materials (shells and domes at large deflexions, for example). How much deformation is possible before something breaks depends upon the material itself and whether there are defects/flaws in the body, which may range from visible holes in the body to microcracks.

    The action of deforming a body requires work to be done. Work (also called energy) is given by the product of force × displacement, so the energy dissipated by friction, when a brick is slid across a surface by constant force F over distance δ, is given by Fδ. During the deformations shown in Figure 2-1(A), the force is not constant as bigger loads are required to produce bigger deformations. The work done must be determined incrementally and is given by ∫Fdδ, i.e. by the area under the load–deformation curve. When an algebraic relation is known between F and δ, the integration can be performed. For a linear material F = (constant)δ, so incremental work done is Fdδ = (constant) δdδ = (constant)δ²/2 = 0.5Fδ, the area of the triangle between the linear stiffness line and the δ-axis. Some stiffening and softening curves may be represented by F = (another constant)δn, where n > 1 for a stiffening (J-shaped) curve and n < 1 for a softening curve. Then ∫Fdδ = (another constant)δ(n + 1)/(n + 1) = Fδ/(n + 1).

    If bodies are unloaded somewhere before fracture, various responses are possible. Whenever load–deformation during unloading retraces exactly the same path as for loading, the behaviour is elastic (reversible). Engineers have been interested mostly in linear behaviour (Hooke’s seventeenth century law for springs is still relevant for returning computer keyboard keys) but reversibility is possible for all the types of non-linear behaviour shown in Figure 2-1(A), as common experience with rubber bands tells us. In these circumstances, all the work put in to deform the body, and stored in the body under load, is recovered. What happens to the energy should fracture occur before unloading is the vital question in fracture mechanics. When fracture occurs within the reversible range of deformation, the material is said to be brittle (so an elastic band is just as brittle as glass or rock in this sense, although brittle is normally associated with ‘hard and stiff’ solids).

    When different paths are followed during loading and unloading, the classification of behaviour depends on whether load–deformation returns to the origin (no deformation at no load), or whether there is some permanent deformation at zero load. In the first case (Figure 2-1B), the strain energy stored when loaded to point B is given by the area OCBEO. The strain energy recovered is given by area ODBEO, so the work represented by area OCBDO has been lost and means hysteresis, i.e. irreversible behaviour where energy has been dissipated even though the body returns to its original state. Friction between cutting tool and workpiece can generate hysteresis loops (pulling out an axe stuck in

    Enjoying the preview?
    Page 1 of 1