ebook img

Load-Bearing Fibre Composites PDF

281 Pages·1980·13.467 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Load-Bearing Fibre Composites

INTERNATIONAL SERIES ON THE STRENGTH AND FRACTURE OF MATERIALS AND STRUCTURES General Editor. D. M. R. Taplin, D.Sc, D.Phil., F.I.M. Other Titles in the Series EASTERLING Mechanisms of Deformation and Fracture HAASEN AND GEROLD Strength of Metals and Alloys (ICSMA 5) -3 Volumes MILLER AND SMITH Mechanical Behaviour of Materials (ICM 3) - 3 Volumes SMITH Fracture Mechanics-Current Status, Future Prospects TAPLIN Advances in Research on the Strength and Fracture of Materials-6 Volumes Related Pergamon Journals Engineering Fracture Mechanics Fatigue of Engineering Materials and Structures NOTICE TO READERS Dear Reader If your library is not already a standing order customer or subscriber to this series, may we recommend that you place a standing or subscription order to receive immediately upon publication all new issues and volumes published in this valuable series. Should you find that these volumes no longer serve your needs your order can be cancelled at any time without notice. The Editors and the Publisher will be glad to receive sugges­ tions or outlines of suitable titles, reviews or symposia for consideration for rapid publication in this series. ROBERT MAXWELL Publisher at Pergamon Press i ' 'ΐίΐ ''''■. '''",; ""»'",'ν''*«' ;.;i{f,',·/f.'."...? " '"''/. ml fWhs· immmmt. ??Ew5*r » Z^'-^^r, ■ ;ll^^^9i0" mmm Frontispiece. A recent application of the first glass-polymer: the 107 m long radome on top of the Toronto 550 m CN Tower (Courtesy of CN Tower Ltd.) Load-Bearing Fibre Composites BY MICHAEL R. PIGGOTT A.R.C.S., Ph.D, P.Eng., F.lnst.P. Centre for the Study of Materials, University of Toronto PERGAMON PRESS OXFORD ■ NEW YORK ■ BEIJING ■ FRANKFURT SAO PAULO ■ SYDNEY · TOKYO ■ TORONTO U.K. Pergamon Press, Headington Hill Hall, Oxford 0X3 OBW, England U.S.A. Pergamon Press, Maxwell House, Fairview Park, Elmsford, New York 10523, U.S.A. PEOPLE'S REPUBLIC Pergamon Press, Room 4037, Qianmen Hotel, Beijing, OF CHINA People's Republic of China FEDERAL REPUBLIC Pergamon Press, Hammerweg 6, OF GERMANY D-6242 Kronberg, Federal Republic of Germany Pergamon Editora, Rue Epa de Queiros, 346, BRAZIL CEP 04011, Paraiso, Säo Paulo, Brazil Pergamon Press Australia, P.O. Box 544, AUSTRALIA Potts Point, N.S.W. 2011, Australia Pergamon Press, 8th Floor, Matsuoka Central Building, JAPAN 1-7-1 Nishishinjuku, Shinjuku-ku, Tokyo 160, Japan Pergamon Press Canada, Suite No 271, CANADA 253 College Street, Toronto, Ontario, Canada M5T 1R5 Copyright © 1980 M. R. Piggott All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or otherwise, without permission in writing from the publishers. First edition 1980 Reprinted 1987 British Library Cataloging in Publication Data Piggott, M. R. Load bearing fibre composites (Pergamon international library: international series on the strength and fracture of materials and structures). 1. Fibrous composites 620.1 1 TA418.9.C6 79-40951 ISBN 0-08-024230-8 (Hard Cover) ISBN 0-08-024231-6 (Flexicover) Printed in Great Britain by A. Wheaton & Co. Ltd, Exeter Preface THIS book is based on a course of lectures given to upper-year and graduate students in Materials Science and Engineering. It is designed to present a unified view of the whole field of fibre (and platelet) composites, rather than going into any aspect of the subject in great detail. The reader who wants to go more deeply into any aspect is referred, at the end of each chapter, to more specialized texts and reviews, or to key papers. The reader is also recommended to look at the literature acknowledged in the subtitles to the figures. The field has developed rapidly over the last 20 years, and no materials scientist should now be without a working knowledge of it. Metallurgists should be aware of the competition from reinforced polymers and ceramics, and be able to appreciate their strengths and weaknesses. Designers need to be able to make a rational choice of which material to use in any situation. This book is aimed at these audiences, but, in addition, indicates areas where our knowledge is not as complete as it should be. Thus I hope also to inspire those inclined to do research. A mathematical development of the subject cannot be avoided if composites are to be properly understood and correctly used. The interactions between fibres and matrix are quite complex, and are the subject of multitudes of erudite papers. Here, however, complicated mathematical expressions are eschewed, and wherever simple ones can be used with adequate precision these are presented in their place. All equations are developed from basic principles, and the reader only needs acquaintance with algebra, calculus, and geometry at first-year University level. The author is grateful for assistance from many sources. Pictures and information have been generously provided, and the author has been kindly received in many laboratories where important work is going on, and benefited greatly from what he has seen and discussed there. The writer is especially grateful for assistance while on study leave at the University of Bath. The University provided material assistance, and Professor Bryan Harris unstint- ingly provided original pictures, comments, encouragement, and assistance. 1 Introduction THIS book describes the basic ideas in the relatively new field of fibre reinforced composites, and provides recent data on their load-bearing capabilities. In addition, selected data is given for more traditional materials, so that the reader can see where composite materials fit into the hierarchy of materials. The discussion starts at a basic level, and such important properties as strength, creep resistance, and fatigue resistance are described. The connection between chemical bond strength and tensile strength is discussed, and the role of imperfections is shown to be of overwhelming importance. The reasons for the good mechanical properties of fibres are set forth, and an account is given of the production of strong and stiff fibres. This is followed by a simplified description of the mechanics of the processes by which fibres can contribute strength, stiffness, and toughness to weak matrices. Finally a brief sketch of the properties of composites is given, classified according to whether the matrix is polymeric, metallic, or ceramic (including cementitious). Emphasis is given, throughout the book, to load- bearing properties and this is followed by a discussion of some of the diverse end uses of these materials. The term composite has come to mean a material made by dispersing particles, of one or more materials in another material, which forms a substantially continuous network around them. The properties of the composite may bear little relation to those of the components, even though the components retain their integrity within the composite. The components can be randomly arranged, or organized in some sort of pattern. Generally the arrangement will have a large effect on the properties. Further, they can have roughly spherical shapes, e.g. stones in concrete, or can have some very distinctive shape such as the iron carbide laminae found in some steels, or long thin fibres, such as the cellulose fibres in wood. The particle shape also has a very profound effect on the properties of the composite. Since this book is concerned with fibre-reinforced composites, it will largely restrict itself to the discussion of two-component composites, where one of the components is a long thin fibre or whisker, and where the composite has to have good load-carrying capacity. Also reinforcement by thin platelets will be described. Fibre-reinforced materials have been used by man for a very long time. The first to be used were naturally occurring composites, such as wood, but man also found out, long ago, that there were advantages to be gained from using artificial mixtures of materials with one component fibrous, such as straw in clay, for bricks, or horse hair in plaster of paris for ceilings. Recently, with the advent of cheap and strong glass fibres, and with the discovery of a 1 2 Load-Bearing Fibre Composites number of new fibre-forming materials with better properties than anything available heretofore, the interest in fibre reinforced materials has increased rapidly, and is still accelerating. Fibre-reinforced polymers are replacing metals in a whole host of situations where load-carrying capacity is important. More efficient aircraft, turbine engines, and cars can be produced with fibre composites, and worthwhile new applications for these materials are being found almost every day. In order to appreciate the potential benefits to be gained from fibre reinforced materials, however, it is necessary first to be aware of what can be done with more traditional materials. The introductory section of this text therefore starts with a review of the important properties required for load-bearing materials, and discusses traditional materials in this context. Then follows a brief statement of the basic ideas of isotropic elasticity theory, and its extension to non-isotropic cases of interest for fibre reinforcement. 1.1. Conventional Materials The important properties for load-bearing materials fall naturally into two groups: mechanical and non-mechanical. Mechanical properties include stiffness, strength, ductility, hardness, toughness, fatigue, and creep. These will be discussed first. Non- mechanical properties, which will be discussed later, include density, temperature resistance, corrosion resistance (including stress corrosion and hydrogen embrittlement), and cost. 1.1.1. Stijfness The stiffness of a material, or its resistance to reversible deformation under load, is a very important mechanical property. In order to characterize the effect of the load on the material, it is normally converted to a stress, that is the force per unit area of cross-section acted on by the load. Expressed in these terms, it is then an indication of the forces experienced by the individual atoms in the material as a result of applying the load. The response of the atoms is to change their positions slightly. Hence, under a stretching, or tensile stress, the atoms move apart in the direction of the force. The distance moved, divided by the original distance, is called the strain. The movement of the atoms under the action of applied loads can be observed and measured using X-ray diffraction or neutron diffraction. The combined movement of the atoms constituting the whole specimen being stressed can usually be measured quite easily using a sensitive distance gauge. A stiff material is one which deforms very little. Young's modulus is normally used as a measure of this, and is defined as the ratio of the tensile stress to the strain produced. The unit normally used for Young's modulus is the Pascal (Pa). (These units are described in Appendix B.) Other stiffness parameters will be discussed in the second part of this chapter. Few materials are isotropic, and so measurements of Young's modulus made in different directions will give different results. These differences are particularly marked with single crystals of non-metals. However, polycrystalline metals, polymers, ceramics are generally sufficiently isotropic that they can be assigned a single value for the Young's Introduction 3 modulus. Young's moduli range from about 1 TPa for a favourable direction in a diamond crystal through 100 GPa for a metal, 1 GPa for a polymer, and 10 MPa for a rubber. Table 1.1 gives the moduli of a representative selection for metals, ceramics, and polymers. The modulus of well-made specimens of crystalline elements or compounds is a measure of the deformability of the bonds between the atoms, and hence will be the same for different samples of the material. This is not quite the case with metal alloys, while with some polymers quite large variations between the modulus of different specimens of chemically similar materials can be found. 1.1.2. Strength The strength of a material is the stress required to break it. It differs markedly from the modulus, in being determined as much by the method of manufacture and previous history of the specimen, as by the nature of the atoms and their arrangement. Pure, annealed, metals, and non-fibrous polymers are weak in both tension and compression. Ceramics and hard materials are generally much weaker in tension than in compression. Some polymer fibres are very strong in tension, while modern alloys, developed for use as structural members, are generally strong in both tension and compression. Typical tensile strengths are given in Table 1.1 for a variety of materials, excluding fibres and wires, some properties of which are given in Table 3.1. TABLE 1.1. Strength and Stiffness of Various Materials Young's Modulus Tensile Material (GPa) strength (MPa) Metals: Aluminium (pure, annealed) 71 60 High-strength aluminium alloy (Al-Zn-Cu-Mg) 71 650 Iron (cast) 152 360 High-strength iron alloy (marag- ing steel) 212 2000 Magnesium alloy (Mg-Zn-Zr) 45 340 Titanium alloy (Ti-Zn-Al-Mo-Si) 120 1400 Tungsten 411 1800 Zirconium alloy (zircaloy 2) 97 590 Ceramics: Alumina (high density) 400 280 Concrete 50 3.5 Glass (sheet) 70 70 Polymers: Epoxy resin 2.5 60 Polycarbonate 2.5 65 Polyethylene (branched) 0.2 10 Rubber (natural) 0.018 32 Rubber (fluorocarbon) 0.002 7 Wood (Douglas fir) 14 34t Wood (white pine) 7.6 lot t Parallel to the grain; long-term strength is half this. 4 Load-Bearing Fibre Composites 1.13. Ductility, Hardness, and Toughness Figure 1.1 shows the stress-strain curves obtained with two different metals. Here the stress is calculated using the initial cross-section of the material, and the strain is calculated directly from the change in length of the test section. 600 STEEL Region Near Origin Strain Axis Expanded ^400 Έ υ 0.1 0.2 0.3 Strain FIG. 1.1. Stress-strain curves for steel and brass. With many metals the curve has three regions. First there is a linear region where the material is elastic, and on removal of the stress at any part of the line, the line is retraced back to the origin. The slope of the line gives the Young's modulus of the material. The second part of the curve starts at the end of the linear region (this point is the yield point) and, except for a slight fall at the yield point in the case of some steels, the stress increases monotonically up to a maximum value, called the ultimate tensile strength. In the third region the stress decreases monotonically with increasing strain until final failure occurs. In this region the specimen is not deforming uniformly; a region of reduced cross-section is formed (necking is occurring) and failure occurs at the neck. (The maximum stress at the neck can be considerably greater than the stress in the rest of the material, so that the true stress at failure is much greater than that indicated by the stress axis. These two failure stresses are sometimes distinguished by calling the lower stress the engineering stress at failure, and the higher stress the true stress at failure. The true stress-true strain curve does not have a peak; the true stress increases monotonically to the failure point.) In the second and third regions of the stress-strain curve, removal of the stress does not result in the material retracing the curve back to the origin. Instead, with decreasing stress the strain decreases at a rate governed by the modulus of the material, i.e. the line has a slope which is the same as that near the origin (Fig. 1.2). When the stress has been reduced to zero, there is now still some strain remaining. This strain is the origin of ductility. A stress-strain curve which is entirely linear up to the breaking-point, such as that obtained with a hard steel (e.g. razor blade steel) indicates very little ductility. A curve with a very long region after the yield point indicates a material with great ductility. With pure metals, the length of the specimen can sometimes be doubled before failure occurs. Introduction 5 Strain —► Residual Strain FIG. 1.2. Residual strain in a metal. Figure 1.3 shows the stress-strain curves for two polymers. It can be seen that they show similar features to those observed with metals, except that yielding is generally not so sharp, and the curve does not usually have a monotonically decreasing region. Some polymers (e.g. polyesters) have very little ductility. Note that the yield point, or elastic limit, occurs at much higher strains for plastics than it does for most metals. 60 POLYCARBONATE 40 CO 20 POLYETHYLENE 0.5 1.0 1.5 Strain FIG. 1.3. Stress-strain curves for polyethylene and polycarbonate. Hardness is a measure of the resistance of materials to plastic deformation, and hence ductile materials are usually not hard. It is determined by indenting the material with a hard ball, or a diamond with a pyramid-shaped tip, and measuring the size of indentation for a given load. Diamonds will leave indentations in glass, showing that plastic deformation, albeit small, is possible with glass. Table 1.2 gives some typical hardness values. They can be converted to roughly equivalent yield stresses in MPa by multiplying by three. Note, however, that with any metal or alloy, a wide range of hardness values can usually be obtained by heat treatment and cold working.

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.