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

Only $11.99/month after trial. Cancel anytime.

Aircraft Structures for Engineering Students
Aircraft Structures for Engineering Students
Aircraft Structures for Engineering Students
Ebook1,894 pages11 hours

Aircraft Structures for Engineering Students

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Aircraft Structures for Engineering Students, Sixth Edition, is the leading self-contained aircraft structures course text. It covers all fundamental subjects, including elasticity, structural analysis, airworthiness and aeroelasticity. Now in its sixth edition, the author has expanded the book’s coverage of analysis and design of composite materials for use in aircraft, and has added new, real-world and design-based examples, along with new end-of-chapter problems of varying complexity.

  • Expanded coverage of composite materials and structures
  • New practical and design-based examples and problems throughout the text aid understanding and relate concepts to real world applications
  • Updated and additional Matlab examples and exercises support use of computational tools in analysis and design
  • Available online teaching and learning tools include downloadable Matlab code, solutions manual, and image bank of figures from the book
LanguageEnglish
Release dateOct 17, 2016
ISBN9780081009987
Aircraft Structures for Engineering Students
Author

T.H.G. Megson

T.H.G. Megson is a professor emeritus with the Department of Civil Engineering at Leeds University (UK). For Elsevier he has written the market leading Butterworth Heinemann textbooks Aircraft Structures for Engineering Students and Introduction to Aircraft Structural Analysis (a briefer derivative of the aircraft structures book), as well as the text/ref hybrid Structural and Stress Analysis.

Read more from T.H.G. Megson

Related to Aircraft Structures for Engineering Students

Related ebooks

Aviation & Aeronautics For You

View More

Related articles

Reviews for Aircraft Structures for Engineering Students

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

    Aircraft Structures for Engineering Students - T.H.G. Megson

    9780081009987_FC

    Aircraft Structures for engineering students

    Sixth Edition

    T.H.G. Megson

    Table of Contents

    Cover image

    Title page

    Copyright

    Dedication

    Preface to the Sixth Edition of Aircraft Structures

    Preface

    Part A: Fundamentals of Structural Analysis

    Section A1: Elasticity

    Chapter 1: Basic elasticity

    Abstract

    1.1 Stress

    1.2 Notation for forces and stresses

    1.3 Equations of equilibrium

    1.4 Plane stress

    1.5 Boundary conditions

    1.6 Determination of stresses on inclined planes

    1.7 Principal stresses

    1.8 Mohr's circle of stress

    1.9 Strain

    1.10 Compatibility equations

    1.11 Plane strain

    1.12 Determination of strains on inclined planes

    1.13 Principal strains

    1.14 Mohr's circle of strain

    1.15 Stress–strain relationships

    1.16 Experimental measurement of surface strains

    Chapter 2: Two-dimensional problems in elasticity

    Abstract

    2.1 Two-dimensional problems

    2.2 Stress functions

    2.3 Inverse and semi-inverse methods

    2.4 St. Venant's principle

    2.5 Displacements

    2.6 Bending of an end-loaded cantilever

    Chapter 3: Torsion of solid sections

    Abstract

    3.1 Prandtl stress function solution

    3.2 St. Venant warping function solution

    3.3 The membrane analogy

    3.4 Torsion of a narrow rectangular strip

    Section A2: Virtual work, energy, and matrix methods

    Chapter 4: Virtual work and energy methods

    Abstract

    4.1 Work

    4.2 Principle of virtual work

    4.3 Applications of the principle of virtual work

    Chapter 5: Energy methods

    Abstract

    5.1 Strain energy and complementary energy

    5.2 Principle of the stationary value of the total complementary energy

    5.3 Application to deflection problems

    5.4 Application to the solution of statically indeterminate systems

    5.5 Unit load method

    5.6 Flexibility method

    5.7 Total potential energy

    5.8 Principle of the stationary value of the total potential energy

    5.9 Principle of superposition

    5.10 Reciprocal theorem

    5.11 Temperature effects

    Chapter 6: Matrix methods

    Abstract

    6.1 Notation

    6.2 Stiffness matrix for an elastic spring

    6.3 Stiffness matrix for two elastic springs in line

    6.4 Matrix analysis of pin-jointed frameworks

    6.5 Application to statically indeterminate frameworks

    6.6 Matrix analysis of space frames

    6.7 Stiffness matrix for a uniform beam

    6.8 Finite element method for continuum structures

    Section A3: Thin plate theory

    Chapter 7: Bending of thin plates

    Abstract

    7.1 Pure bending of thin plates

    7.2 Plates subjected to bending and twisting

    7.3 Plates subjected to a distributed transverse load

    7.4 Combined bending and in-plane loading of a thin rectangular plate

    7.5 Bending of thin plates having a small initial curvature

    7.6 Energy method for the bending of thin plates

    Section A4: Structural instability

    Chapter 8: Columns

    Abstract

    8.1 Euler buckling of columns

    8.2 Inelastic buckling

    8.3 Effect of initial imperfections

    8.4 Stability of beams under transverse and axial loads

    8.5 Energy method for the calculation of buckling loads in columns

    8.6 Flexural–torsional buckling of thin-walled columns

    Chapter 9: Thin plates

    Abstract

    9.1 Buckling of thin plates

    9.2 Inelastic buckling of plates

    9.3 Experimental determination of the critical load for a flat plate

    9.4 Local instability

    9.5 Instability of stiffened panels

    9.6 Failure stress in plates and stiffened panels

    9.7 Tension field beams

    Section A5: Vibration of structures

    Chapter 10: Structural vibration

    Abstract

    10.1 Oscillation of mass–spring systems

    10.2 Oscillation of beams

    10.3 Approximate methods for determining natural frequencies

    Part B: Analysis of Aircraft Structures

    Section B1: Principles of stressed skin construction

    Chapter 11: Materials

    Abstract

    11.1 Aluminum alloys

    11.2 Steel

    11.3 Titanium

    11.4 Plastics

    11.5 Glass

    11.6 Composite materials

    11.7 Properties of materials

    Chapter 12: Structural components of aircraft

    Abstract

    12.1 Loads on structural components

    12.2 Function of structural components

    12.3 Fabrication of structural components

    12.4 Connections

    Section B2: Airworthiness and airframe loads

    Chapter 13: Airworthiness

    Abstract

    13.1 Factors of safety-flight envelope

    13.2 Load factor determination

    Chapter 14: Airframe loads

    Abstract

    14.1 Aircraft inertia loads

    14.2 Symmetric maneuver loads

    14.3 Normal accelerations associated with various types of maneuver

    14.4 Gust loads

    Chapter 15: Fatigue

    Abstract

    15.1 Safe life and fail-safe structures

    15.2 Designing against fatigue

    15.3 Fatigue strength of components

    15.4 Prediction of aircraft fatigue life

    15.5 Crack propagation

    Section B3: Bending, shear and torsion of thin-walled beams

    Chapter 16: Bending of open and closed, thin-walled beams

    Abstract

    16.1 Symmetrical bending

    16.2 Unsymmetrical bending

    16.3 Deflections due to bending

    16.4 Calculation of section properties

    16.5 Applicability of bending theory

    16.6 Temperature effects

    Chapter 17: Shear of beams

    Abstract

    17.1 General stress, strain, and displacement relationships for open and single-cell closed section thin-walled beams

    17.2 Shear of open section beams

    17.3 Shear of closed section beams

    Chapter 18: Torsion of beams

    Abstract

    18.1 Torsion of closed section beams

    18.2 Torsion of open section beams

    Chapter 19: Combined open and closed section beams

    Abstract

    19.1 Bending

    19.2 Shear

    19.3 Torsion

    Chapter 20: Structural idealization

    Abstract

    20.1 Principle

    20.2 Idealization of a panel

    20.3 Effect of idealization on the analysis of open and closed section beams

    20.4 Deflection of open and closed section beams

    Section B4: Stress analysis of aircraft components

    Chapter 21: Wing spars and box beams

    Abstract

    21.1 Tapered wing spar

    21.2 Open and closed section beams

    21.3 Beams having variable stringer areas

    Chapter 22: Fuselages

    Abstract

    22.1 Bending

    22.2 Shear

    22.3 Torsion

    22.4 Cut-outs in fuselages

    Chapter 23: Wings

    Abstract

    23.1 Three-boom shell

    23.2 Bending

    23.3 Torsion

    23.4 Shear

    23.5 Shear center

    23.6 Tapered wings

    23.7 Deflections

    23.8 Cut-outs in wings

    Chapter 24: Fuselage frames and wing ribs

    Abstract

    24.1 Principles of stiffener/web construction

    24.2 Fuselage frames

    24.3 Wing ribs

    Chapter 25: Laminated composite structures

    Abstract

    25.1 Elastic constants of a simple lamina

    25.2 Stress–strain relationships for an orthotropic ply (macro approach)

    25.3 Laminates

    25.4 Thin-walled composite beams

    Section B5: Structural and loading discontinuities

    Chapter 26: Closed section beams

    Abstract

    26.1 General aspects

    26.2 Shear stress distribution at a built-in end of a closed section beam

    26.3 Thin-walled rectangular section beam subjected to torsion

    26.4 Shear lag

    Chapter 27: Open section beams

    Abstract

    27.1 I-Section beam subjected to torsion

    27.2 Torsion of an arbitrary section beam

    27.3 Distributed torque loading

    27.4 Extension of the theory to allow for general systems of loading

    27.5 Moment couple (bimoment)

    Section B6: Introduction to aeroelasticity

    Chapter 28: Wing problems

    Abstract

    28.1 Types of problem

    28.2 Load distribution and divergence

    28.3 Control effectiveness and reversal

    28.4 Introduction to flutter

    Appendix: Design of a rear fuselage

    A.1 Specification

    A.2 Data

    A.3 Initial calculations

    A.4 Balancing out calculations

    A.5 Fuselage loads

    A.6 Fuselage design calculations

    Index

    Copyright

    Butterworth-Heinemann is an imprint of Elsevier

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    Copyright © 2017, 2013, 2007, 1999 T.H.G. Megson. Published by Elsevier Ltd. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher's permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability 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.

    MATLAB® is a trademark of The MathWorks, Inc. and is used with permission. The MathWorks does not warrant the accuracy of the text or exercises in this book. This book's use or discussion of MATLAB® software or related products does not constitute endorsement or sponsorship by The MathWorks of a particular pedagogical approach or particular use of the MATLAB® software.

    Library of Congress Cataloging-in-Publication Data

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

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library.

    ISBN: 978-0-08-100914-7

    For information on all Butterworth-Heinemann publications visit our website at https://www.elsevier.com

    fm01-9780081009147

    Publisher: Todd Green

    Acquisition Editor: Steve Merken

    Editorial Project Manager: Nate McFadden

    Production Project Manager: Stalin Viswanathan

    Designer: Greg Harris

    Dedication

    To the memory of my darling wife, Margaret.

    Preface to the Sixth Edition of Aircraft Structures

    T.H.G. Megson

    The publication of a sixth edition has enabled me to review some of the topics included in the fifth edition and to provide additional worked-out examples and end-of-chapter exercises of a more practical nature than previously. It has also given me the opportunity to carefully examine the text and correct the printing errors which had, unfortunately, crept into the fifth edition.

    The layout of the book remains the same as that in the fifth edition with the first seven chapters remaining unchanged. In Chapter 8 on structural instability of columns, I have included an additional example and an additional end-of-chapter exercise illustrating the application of the reduced modulus theory, while in Chapter 9 the work on the stability of thin-walled columns has been extended to the determination of average failure stresses.

    In Chapter 21 on the stress analysis of wing spars and box beams, I have illustrated the effects of sweep in a wing which, of course, is particularly relevant to modern high-speed jet aircraft.

    I have revised the theory presented in Chapter 23 for the determination of stresses in fuselage panels containing cut-outs and provided an illustrative practical example together with an end-of-chapter exercise.

    The major modification in the sixth edition is the extension of the work on composite materials and structures presented in Chapter 25. In the fifth edition, the theory was restricted to single-ply laminates; this has now been extended to a consideration of multi-ply laminates. In this, the method of specifying different ply lay-ups is presented together with the effects of symmetry and reinforcement orientation. The calculation of the equivalent elastic constants of a laminate is presented for the case of in-plane loading only since this is normally the situation in the thin skins of aircraft structures. The calculation of the distribution of stresses across the thickness of a laminate is illustrated by an example and the strength of laminates investigated using the maximum stress theory. Several additional examples are included as well as end-of-chapter exercises.

    Finally, in Chapter 26, I have included a practical example on the effect of shear lag on the loads in a wing panel containing a cut-out for an undercarriage bay.

    Preface

    The idea of a textbook on aircraft structures for students of aeronautical engineering was born during the early part of my career teaching the subject. I felt at that time that the books available were either out of date or too specialised to fulfil the requirements of an undergraduate textbook. My aim, therefore, was to fill this gap and provide a self-contained course in aircraft structures which included not only the fundamentals of elasticity and aircraft structural analysis but also the associated topics of airworthiness and aeroelasticity.

    Developments in aircraft construction led me, in subsequent years, to re-examine the contents of the book and introduce modifications. In the second edition I reorganised some of the text and expanded the work on composite materials. I also introduced the analysis of more realistic structures such as fuselages and wings containing cut-outs for door openings and undercarriage bays together with the analysis of fuselage frames and wing ribs. The third edition saw the publication of an accompanying solutions manual and a reorganisation of the contents into two parts as opposed, previously, to three. For the fourth edition I decided that a major overhaul would be beneficial in the light of developments in the aircraft industry and the fact that students were wanting more worked examples and end of chapter exercises. I also felt that some of the chapters were too long so I therefore broke them down into shorter, more digestible, ones. At the same time I rearranged the material to emphasize the application of the fundamentals of structural analysis, contained in Part A, to the analysis of aircraft structures in Part B. I also expanded the application of the powerful principle of virtual work to a complete chapter and extended the work on tension field beams to include post-buckling behaviour. The design of riveted connections was added and the work on crack propagation extended. The study of composite structures was expanded and, finally, the actual design of part of the rear fuselage of a trainer/semi-aerobatic aircraft was presented in an Appendix.

    The fifth edition has been produced in response to requests for more worked examples and end of chapter exercises and also for computer based solutions (MATLAB). The organisation of the contents is the same as in the fourth edition. Part A, Fundamentals of Structural Analysis, comprises basic elasticity, virtual work and energy methods, thin plate theory, structural instability and the vibration of structures. Part B, Analysis of Aircraft Structures, includes the principles of stressed skin construction from the points of view of materials and structural components, airworthiness and airframe loads and a detailed study of fatigue including the prediction of aircraft fatigue life. Part B also includes the analysis of thin-walled beams subjected to bending, shear and torsional loads and the method of idealising a structure into one more amenable to analysis. This then leads into the stress analysis of aircraft components such as wing spars, fuselages, wings, fuselage frames and wing ribs. Composite structures are considered and the effects on the analysis of structural and loading discontinuities investigated. An introduction to aeroelasticity is given. Finally the design of a portion of the rear fuselage of a trainer/semi-aerobatic aircraft is presented. I would like to thank Patrick Lewis in the Design Exploration Group at Brigham Young University for providing examples, exercises and solutions using the MATLAB(r) program. I would also like to thank Joe Hayton and Jeff Freeland of Elsevier for their sterling work on the production of the book.

    Supporting Material TO ACCOMPANY This Book

    For instructors using this text in their course, a full set of worked solutions and electronic images of the figures in the text are available by registering at: www.textbooks.elsevier.com

    Part A

    Fundamentals of Structural Analysis

    Section A1

    Elasticity

    Chapter 1

    Basic elasticity

    Abstract

    The basic concepts of elasticity are introduced. Stresses and strains are defined and the notation established. The equations of equilibrium and the boundary conditions for a three-dimensional body are derived and then reduced for the two-dimensional case of plane stress. Complex stress systems are considered and expressions for principal stresses determined. A graphical method (Mohr's circle) is introduced for the solution of complex stress system problems. Strain-displacement relationships are established and equations of compatibility derived for three- and two-dimensional cases. Equations for the strains produced by complex stress systems are derived as are equations for principal strains. Mohr's circle for the graphical solution of strain problems is introduced. Stress-strain relationships are derived and the elastic constants for a material defined. The effect of temperature variation is examined and the experimental measurement of surface strains described.

    Keywords

    Stress; Strain; Stress–strain relationships; Complex stress; Complex strain; Mohr's circle; Temperature effects; Experimental measurement of strain

    We consider, in this chapter, the basic ideas and relationships of the theory of elasticity. The treatment is divided into three broad sections: stress, strain, and stress–strain relationships. The third section is deferred until the end of the chapter to emphasize the fact that the analysis of stress and strain, for example, the equations of equilibrium and compatibility, does not assume a particular stress–strain law. In other words, the relationships derived in Sections 1.1–1.14 are applicable to nonlinear as well as linearly elastic bodies.

    1.1 Stress

    Consider the arbitrarily shaped, three-dimensional body shown in Fig. 1.1. The body is in equilibrium under the action of externally applied forces P1, P2, … and is assumed to constitute a continuous and deformable material, so that the forces are transmitted throughout its volume. It follows that, at any internal point O, there is a resultant force δP. The particle of material at O subjected to the force δP is in equilibrium, so that there must be an equal but opposite force δP (shown dotted in Fig. 1.1) acting on the particle at the same time. If we now divide the body by any plane nn containing O, then these two forces δP may be considered as being uniformly distributed over a small area δA of each face of the plane at the corresponding point O, as in Fig. 1.2. The stress at O is defined by the equation

    si1_e    (1.1)

    f01-01-9780081009147

    Figure 1.1 Internal Force at a Point in an Arbitrarily Shaped Body

    f01-02-9780081009147

    Figure 1.2 Internal Force Components at the Point O

    The directions of the forces δP in Fig. 1.2 are such as to produce tensile stresses on the faces of the plane nn. It must be realized here that, while the direction of δP is absolute, the choice of plane is arbitrary, so that, although the direction of the stress at O is always in the direction of δP, its magnitude depends upon the actual plane chosen, since a different plane has a different inclination and therefore a different value for the area δA. This may be more easily understood by reference to the bar in simple tension in Fig. 1.3. On the cross-sectional plane mm, the uniform stress is given by P/A, while on the inclined plane mm′ the stress is of magnitude P/A′. In both cases, the stresses are parallel to the direction of P.

    f01-03-9780081009147

    Figure 1.3 Values of Stress on Different Planes in a Uniform Bar

    Generally, the direction of δP is not normal to the area δA, in which case, it is usual to resolve δP into two components: one, δPn, normal to the plane and the other, δPs, acting in the plane itself (see Fig. 1.2). Note that, in Fig. 1.2, the plane containing δP is perpendicular to δA. The stresses associated with these components are a normal or direct stress defined as

    si2_e    (1.2)

    and a shear stress defined as

    si3_e    (1.3)

    The resultant stress is computed from its components by the normal rules of vector addition, i.e.:

    si4_e

    Generally, however, as indicated previously, we are interested in the separate effects of σ and τ.

    However, to be strictly accurate, stress is not a vector quantity for, in addition to magnitude and direction, we must specify the plane on which the stress acts. Stress is therefore a tensor, its complete description depending on the two vectors of force and surface of action.

    1.2 Notation for forces and stresses

    It is usually convenient to refer the state of stress at a point in a body to an orthogonal set of axes Oxyz. In this case we cut the body by planes parallel to the direction of the axes. The resultant force δP acting at the point O on one of these planes may then be resolved into a normal component and two in-plane components, as shown in Fig. 1.4, thereby producing one component of direct stress and two components of shear stress.

    f01-04-9780081009147

    Figure 1.4 Components of Stress at a Point in a Body

    The direct stress component is specified by reference to the plane on which it acts, but the shear stress components require a specification of direction in addition to the plane. We therefore allocate a single subscript to direct stress to denote the plane on which it acts and two subscripts to shear stress, the first specifying the plane, the second direction. Therefore, in Fig. 1.4, the shear stress components are τzx and τzy acting on the z plane and in the x and y directions, respectively, while the direct stress component is σz.

    We may now completely describe the state of stress at a point O in a body by specifying components of shear and direct stress on the faces of an element of side δx, δy, δz, formed at O by the cutting planes as indicated in Fig. 1.5.

    f01-05-9780081009147

    Figure 1.5 Sign Conventions and Notation for Stresses at a Point in a Body

    The sides of the element are infinitesimally small, so that the stresses may be assumed to be uniformly distributed over the surface of each face. On each of the opposite faces there will be, to a first simplification, equal but opposite stresses.

    We now define the directions of the stresses in Fig. 1.5 as positive, so that normal stresses directed away from their related surfaces are tensile and positive; opposite compressive stresses are negative. Shear stresses are positive when they act in the positive direction of the relevant axis in a plane on which the direct tensile stress is in the positive direction of the axis. If the tensile stress is in the opposite direction, then positive shear stresses are in directions opposite to the positive directions of the appropriate axes.

    Two types of external force may act on a body to produce the internal stress system we have already discussed. Of these, surface forces such as P1, P2, …, or hydrostatic pressure, are distributed over the surface area of the body. The surface force per unit area may be resolved into components parallel to our orthogonal system of axes, and these are generally given the symbols si5_e si6_e and si7_e The second force system derives from gravitational and inertia effects, and the forces are known as body forces. These are distributed over the volume of the body and the components of body force per unit volume are designated X, Y, and Z.

    1.3 Equations of equilibrium

    Generally, except in cases of uniform stress, the direct and shear stresses on opposite faces of an element are not equal, as indicated in Fig. 1.5, but differ by small amounts. Therefore if, say, the direct stress acting on the z plane is σz, then the direct stress acting on the z + δz plane is, from the first two terms of a Taylor's series expansion, σz + (σz/∂zz.

    We now investigate the equilibrium of an element at some internal point in an elastic body where the stress system is obtained by the method just described.

    In Fig. 1.6, the element is in equilibrium under forces corresponding to the stresses shown and the components of body forces (not shown). Surface forces acting on the boundary of the body, although contributing to the production of the internal stress system, do not directly feature in the equilibrium equations.

    f01-06-9780081009147

    Figure 1.6 Stresses on the Faces of an Element at a Point in an Elastic Body

    Taking moments about an axis through the center of the element parallel to the z axis,

    si8_e

    which simplifies to

    si9_e

    dividing through by δxδyδz and taking the limit as δx and δy approach zero.

    si10_e

       (1.4)

    We see, therefore, that a shear stress acting on a given plane (τxy, τxz, τyz) is always accompanied by an equal complementary shear stress yx, τzx, τzy) acting on a plane perpendicular to the given plane and in the opposite sense.

    Now, considering the equilibrium of the element in the x direction,

    si11_e

    which gives

    si12_e

    Or, writing τxy = τyx and τxz = τzx from Eq. (1.4),

    si13_e

       (1.5)

    The equations of equilibrium must be satisfied at all interior points in a deformable body under a three-dimensional force system.

    1.4 Plane stress

    Most aircraft structural components are fabricated from thin metal sheet, so that stresses across the thickness of the sheet are usually negligible. Assuming, say, that the z axis is in the direction of the thickness, then the three-dimensional case of Section 1.3 reduces to a two-dimensional case in which σz, τxz, and τyz are all zero. This condition is known as plane stress; the equilibrium equations then simplify to

    si14_e    (1.6)

    1.5 Boundary conditions

    The equations of equilibrium (1.5)—and also (1.6), for a two-dimensional system—satisfy the requirements of equilibrium at all internal points of the body. Equilibrium must also be satisfied at all positions on the boundary of the body, where the components of the surface force per unit area are si5_e si6_e and si7_e The triangular element of Fig. 1.7 at the boundary of a two-dimensional body of unit thickness is then in equilibrium under the action of surface forces on the elemental length AB of the boundary and internal forces on internal faces AC and CB.

    f01-07-9780081009147

    Figure 1.7 Stresses on the Faces of an Element at the Boundary of a Two-Dimensional Body

    Summation of forces in the x direction gives

    si18_e

    which, by taking the limit as δx approaches zero and neglecting second order terms, becomes

    si19_e

    The derivatives dy/ds and dx/ds are the direction cosines l and m of the angles that a normal to AB makes with the x and y axes, respectively. It follows that

    si20_e

    and in a similar manner

    si21_e

    A relatively simple extension of this analysis produces the boundary conditions for a three-dimensional body, namely,

    si22_e    (1.7)

    where l, m, and n become the direction cosines of the angles that a normal to the surface of the body makes with the x, y, and z axes, respectively.

    1.6 Determination of stresses on inclined planes

    The complex stress system of Fig. 1.6 is derived from a consideration of the actual loads applied to a body and is referred to a predetermined, though arbitrary, system of axes. The values of these stresses may not give a true picture of the severity of stress at that point, so that it is necessary to investigate the state of stress on other planes on which the direct and shear stresses may be greater.

    We restrict the analysis to the two-dimensional system of plane stress defined in Section 1.4.

    Figure 1.8(a) shows a complex stress system at a point in a body referred to axes Ox, Oy. All stresses are positive, as defined in Section 1.2. The shear stresses τxy and τyx were shown to be equal in Section 1.3. We now, therefore, designate them both τxy. The element of side δx, δy and of unit thickness is small, so that stress distributions over the sides of the element may be assumed to be uniform. Body forces are ignored, since their contribution is a second-order term.

    f01-08-9780081009147

    Figure 1.8 (a) Stresses on a Two-Dimensional Element; (b) Stresses on an Inclined Plane at the Point

    Suppose that we need to find the state of stress on a plane AB inclined at an angle θ to the vertical. The triangular element EDC formed by the plane and the vertical through E is in equilibrium under the action of the forces corresponding to the stresses shown in Fig. 1.8(b), where σn and τ are the direct and shear components of the resultant stress on AB. Then, resolving forces in a direction perpendicular to ED, we have

    si23_e

    Dividing through by ED and simplifying,

    si24_e

       (1.8)

    Now, resolving forces parallel to ED,

    si25_e

    Again, dividing through by ED and simplifying,

    si26_e

       (1.9)

    Example 1.1

    A cylindrical pressure vessel has an internal diameter of 2 m and is fabricated from plates 20 mm thick. If the pressure inside the vessel is 1.5 N/mm² and, in addition, the vessel is subjected to an axial tensile load of 2500 kN, calculate the direct and shear stresses on a plane inclined at an angle of 60° to the axis of the vessel. Calculate also the maximum shear stress.

    The expressions for the longitudinal and circumferential stresses produced by the internal pressure may be found in any text on stress analysis¹ and are

    si27_esi28_e

    The direct stress due to the axial load will contribute to σx and is given by

    si29_e

    A rectangular element in the wall of the pressure vessel is then subjected to the stress system shown in Fig. 1.9. Note that no shear stresses act on the x and y planes; in this case, σx and σy form a biaxial stress system.

    f01-09-9780081009147

    Figure 1.9 Element of Example 1.1

    The direct stress, σn, and shear stress, τ, on the plane AB, which makes an angle of 60° with the axis of the vessel, may be found from first principles by considering the equilibrium of the triangular element ABC or by direct substitution in Eqs. (1.8) and (1.9). Note that, in the latter case, θ = 30° and τxy = 0. Then,

    si30_e

    The negative sign for τ indicates that the shear stress is in the direction BA and not AB.

    From Eq. (1.9), when τxy = 0,

    si31_e    (i)

    The maximum value of τ therefore occurs when sin2θ is a maximum, that is, when sin2θ = 1 and θ = 45°. Then, substituting the values of σx and σy in Eq. (i),

    si32_e

    Example 1.2

    A cantilever beam of solid, circular cross-section supports a compressive load of 50 kN applied to its free end at a point 1.5 mm below a horizontal diameter in the vertical plane of symmetry together with a torque of 1200 Nm (Fig. 1.10). Calculate the direct and shear stresses on a plane inclined at 60° to the axis of the cantilever at a point on the lower edge of the vertical plane of symmetry.

    f01-10-9780081009147

    Figure 1.10 Cantilever Beam of Example 1.2.

    The direct loading system is equivalent to an axial load of 50 kN together with a bending moment of 50 × 10³ × 1.5 = 75,000 Nmm in a vertical plane. Therefore, at any point on the lower edge of the vertical plane of symmetry, there are compressive stresses due to the axial load and bending moment that act on planes perpendicular to the axis of the beam and are given, respectively, by Eqs. (1.2) and (16.9); that is,

    si33_esi34_e

    The shear stress, τxy, at the same point due to the torque is obtained from Eq. (iv) in Example 3.1; that is,

    si35_e

    The stress system acting on a two-dimensional rectangular element at the point is shown in Fig. 1.11. Note that, since the element is positioned at the bottom of the beam, the shear stress due to the torque is in the direction shown and is negative (see Fig. 1.8).

    f01-11-9780081009147

    Figure 1.11 Stress System on a Two-Dimensional Element of the Beam of Example 1.2

    Again, σn and τ may be found from first principles or by direct substitution in Eqs. (1.8) and (1.9). Note that θ = 30°, σy = 0, and τxy = –28.3 N/mm², the negative sign arising from the fact that it is in the opposite direction to τxy in Fig. 1.8.

    Then,

    si36_esi37_e

    Different answers are obtained if the plane AB is chosen on the opposite side of AC.

    1.7 Principal stresses

    For given values of σx, σy, and τxy, in other words, given loading conditions, σn varies with the angle θ and attains a maximum or minimum value when dσn/dθ = 0. From Eq. (1.8),

    si38_e

    Hence,

    si39_e

    or

    si40_e    (1.10)

    Two solutions, θ and θ + π/2, are obtained from Eq. (1.10), so that there are two mutually perpendicular planes on which the direct stress is either a maximum or a minimum. Further, by comparison of Eqs. (1.9) and (1.10), it will be observed that these planes correspond to those on which there is no shear stress. The direct stresses on these planes are called principal stresses and the planes themselves, principal planes.

    From Eq. (1.10),

    si41_e

    and

    si42_e

    Rewriting Eq. (1.8) as

    si43_e

    and substituting for {sin2θ, cos2θ} and {sin2(θ + π/2), cos2(θ + π/2)} in turn gives

    si44_e

       (1.11)

    and

    si45_e

       (1.12)

    where σI is the maximum or major principal stress and σII is the minimum or minor principal stress. Note that σI is algebraically the greatest direct stress at the point while σII is algebraically the least. Therefore, when σII is negative, that is, compressive, it is possible for σII to be numerically greater than σI.

    The maximum shear stress at this point in the body may be determined in an identical manner. From Eq. (1.9),

    si46_e

    giving

    si47_e    (1.13)

    It follows that

    si48_e

    Substituting these values in Eq. (1.9) gives

    si49_e

       (1.14)

    Here, as in the case of principal stresses, we take the maximum value as being the greater algebraic value.

    Comparing Eq. (1.14) with Eqs. (1.11) and (1.12), we see that

    si50_e    (1.15)

    Equations (1.14) and (1.15) give the maximum shear stress at the point in the body in the plane of the given stresses. For a three-dimensional body supporting a two-dimensional stress system, this is not necessarily the maximum shear stress at the point.

    Since Eq. (1.13) is the negative reciprocal of Eq. (1.10), the angles 2θ given by these two equations differ by 90° or, alternatively, the planes of maximum shear stress are inclined at 45° to the principal planes.

    1.8 Mohr's circle of stress

    The state of stress at a point in a deformable body may be determined graphically by Mohr's circle of stress.

    In Section 1.6, the direct and shear stresses on an inclined plane were shown to be given by

    si51_e

    and

    si52_e

    respectively. The positive directions of these stresses and the angle θ are defined in Fig. 1.12(a). Equation (1.8) may be rewritten in the form

    si43_e

    or

    si54_ef01-12-9780081009147

    Figure 1.12 (a) Stresses on a Triangular Element; (b) Mohr's Circle of Stress for the Stress System Shown in (a)

    Squaring and adding this equation to Eq. (1.9), we obtain

    si55_e

    which represents the equation of a circle of radius si56_e and having its center at the point [(σx – σy)/2, 0].

    The circle is constructed by locating the points Q1 (σx, τxy) and Q2 (σy, – τxy) referred to axes Oστ, as shown in Fig. 1.12(b). The center of the circle then lies at C, the intersection of Q1Q2 and the Oσ axis; clearly C is the point [(σx – σy)/2, 0] and the radius of the circle is si57_e as required. CQ′ is now set off at an angle 2θ (positive clockwise) to CQ1, Q′ is then the point (σn, – τ), as demonstrated next. From Fig. 1.12(b), we see that

    si58_e

    or, since OC = (σx + σy)/2, CN = CQ′ cos(β – 2θ), and CQ′ = CQ1, we have

    si59_e

    But,

    si60_e

    Hence,

    si61_e

    which, on rearranging, becomes

    si62_e

    as in Eq. (1.8). Similarly, it may be shown that

    si63_e

    as in Eq. (1.9). Note that the construction of Fig. 1.12(b) corresponds to the stress system of Fig. 1.12(a), so that any sign reversal must be allowed for. Also, the Oσ and Oτ axes must be constructed to the same scale or the equation of the circle is not represented.

    The maximum and minimum values of the direct stress, that is, the major and minor principal stresses σI and σII, occur when N (and Q′) coincide with B and A, respectively. Thus,

    si64_e

    or

    si65_e

    and, in the same fashion,

    si66_e

    The principal planes are then given by 2θ = β(σI) and 2θ = β + π(σII).

    Also, the maximum and minimum values of shear stress occur when Q′ coincides with D and E at the upper and lower extremities of the circle.

    At these points, Q′N is equal to the radius of the circle, which is given by

    si67_e

    Hence, si68_e as before. The planes of maximum and minimum shear stress are given by 2θ = β + π/2 and 2θ = β + 3π/2, these being inclined at 45° to the principal planes.

    Example 1.3

    Direct stresses of 160 N/mm² (tension) and 120 N/mm² (compression) are applied at a particular point in an elastic material on two mutually perpendicular planes. The principal stress in the material is limited to 200 N/mm² (tension). Calculate the allowable value of shear stress at the point on the given planes. Determine also the value of the other principal stress and the maximum value of shear stress at the point. Verify your answer using Mohr's circle.

    The stress system at the point in the material may be represented as shown in Fig. 1.13 by considering the stresses to act uniformly over the sides of a triangular element ABC of unit thickness. Suppose that the direct stress on the principal plane AB is σ. For horizontal equilibrium of the element,

    si69_e

    which simplifies to

    si70_e    (i)

    f01-13-9780081009147

    Figure 1.13 Stress System for Example 1.3

    Considering vertical equilibrium gives

    si71_e

    or

    si72_e    (ii)

    Hence, from the product of Eqs. (i) and (ii),

    si73_e

    Now, substituting the values σx = 160 N/mm², σy = –120 N/mm², and σ = σ1 = 200 N/mm², we have

    si74_e

    Replacing cotθ in Eq. (ii) with 1/tanθ from Eq. (i) yields a quadratic equation in σ:

    si75_e

       (iii)

    The numerical solutions of Eq. (iii) corresponding to the given values of σx, σy, and τxy are the principal stresses at the point, namely,

    si76_e

    Having obtained the principal stresses, we now use Eq. (1.15) to find the maximum shear stress, thus

    si77_e

    The solution is rapidly verified from Mohr's circle of stress (Fig. 1.14). From the arbitrary origin O, OP1, and OP2 are drawn to represent σx = 160 N/mm² and σy = –120 N/mm². The mid-point C of P1P2 is then located. Next, OB = σ1 = 200 N/mm² is marked out and the radius of the circle is then CB. OA is the required principal stress. Perpendiculars P1Q1 and P2Q2 to the circumference of the circle are equal to ± τxy (to scale), and the radius of the circle is the maximum shear stress.

    f01-14-9780081009147

    Figure 1.14 Solution of Example 1.3 Using Mohr's Circle of Stress

    Example 1.3

    MATLAB®

    Repeat the derivations presented in Example 1.3 using the Symbolic Math Toolbox in MATLAB®. Do not recreate Mohr's circle.

    Using the element shown in Fig. 1.13, derivations of the principal stresses and maximum shear stress are obtained through the following MATLAB file:

    % Declare any needed symbolic variables

    syms sig tau_xy sig_x sig_y theta AB BC AC

    % Define known stress values

    sig_x = sym(160);

    sig_y = sym(-120);

    sig_val = sym(200);

    % Define relationships between AB, BC, and AC

    BC = AB*cos(theta);

    AC = AB*sin(theta);

    % For horizontal equalibrium of the element

    eqI = sig*AB*cos(theta)-sig_x*BC-tau_xy*AC;

    % For vertical equalibrium of the element

    eqII = sig*AB*sin(theta)-sig_y*AC-tau_xy*BC;

    % Solve eqI and eqII for tau_xy

    tau_xyI = solve(eqI,tau_xy);

    tau_xyII = solve(eqII,tau_xy);

    % Take the square-root of tau_xyI times tau_xyII to get tau_xy

    tau_xy_val = sqrt(tau_xyI*tau_xyII);

    % Substitite the given value of sig into tau_xy

    tau_xy_val = subs(tau_xy_val,sig,sig_val);

    % Solve eqI for theta and substitute into eqII

    eqI = simplify(eqI/cos(theta));

    theta_I = solve(eqI,theta);

    eqIII = subs(eqII,theta,theta_I);

    % Substitute the value of tau_xy into eqIII and solve for the principal stresses (sig_p)

    sig_p = solve(subs(eqIII,tau_xy,tau_xy_val),sig);

    sig_I = max(double(sig_p));

    sig_II = min(double(sig_p));

    % Calculate the maximum shear stress using Eq. (1.15)

    tau_max = (sig_I-sig_II)/2;

    % Output tau_xy, the principal stresses, and tau_max to the Command Window

    disp([‘tau_xy = +/-’ num2str(double(tau_xy_val)) ‘N/mmˆ2’])

    disp([‘sig_I =’ num2str(sig_I) ‘N/mmˆ2’])

    disp([‘sig_II =’ num2str(sig_II) ‘N/mmˆ2’])

    disp([‘tau_max =’ num2str(tau_max) ‘N/mmˆ2’])

    The Command Window outputs resulting from this MATLAB file are as follows:

    tau_xy = +/- 113.1371 N/mmˆ2

    sig_I = 200 N/mmˆ2

    sig_II = -160 N/mmˆ2

    tau_max = 180 N/mmˆ2

    1.9 Strain

    The external and internal forces described in the previous sections cause linear and angular displacements in a deformable body. These displacements are generally defined in terms of strain. Longitudinal or direct strains are associated with direct stresses σ and relate to changes in length, while shear strains define changes in angle produced by shear stresses. These strains are designated, with appropriate suffixes, by the symbols ɛ and γ, respectively, and have the same sign as the associated stresses.

    Consider three mutually perpendicular line elements OA, OB, and OC at a point O in a deformable body. Their original or unstrained lengths are δx, δy, and δz, respectively. If, now, the body is subjected to forces that produce a complex system of direct and shear stresses at O, such as that in Fig. 1.6, then the line elements deform to the positions O′A′, O′B′, and O′C′ shown in Fig. 1.15.

    f01-15-9780081009147

    Figure 1.15 Displacement of Line Elements OA, OB, and OC

    The coordinates of O in the unstrained body are (x, y, z) so that those of A, B, and C are (x + δx, y, z), (x, y + δy, z), and (x, y, z + δz). The components of the displacement of O to O′ parallel to the x, y, and z axes are u, v, and w. These symbols are used to designate these displacements throughout the book and are defined as positive in the positive directions of the axes. We again employ the first two terms of a Taylor's series expansion to determine the components of the displacements of A, B, and C. Thus, the displacement of A in a direction parallel to the x axis is u + (∂u/∂xx. The remaining components are found in an identical manner and are shown in Fig. 1.15.

    We now define direct strain in more quantitative terms. If a line element of length L at a point in a body suffers a change in length ΔL, then the longitudinal strain at that point in the body in the direction of the line element is

    si78_e

    The change in length of the element OA is (O′A′ – OA), so that the direct strain at O in the x direction is obtained from the equation

    si79_e    (1.16)

    Now,

    si80_e

    or

    si81_e

    which may be written, when second-order terms are neglected, as

    si82_e

    Applying the binomial expansion to this expression, we have

    si83_e    (1.17)

    in which squares and higher powers of ∂u/∂x are ignored. Substituting for O′A′ in Eq. (1.16), we have

    si84_e

       (1.18)

    The shear strain at a point in a body is defined as the change in the angle between two mutually perpendicular lines at the point. Therefore, if the shear strain in the xz plane is γxz, then the angle between the displaced line elements O′A′ and O′C′ in Fig. 1.15 is π/2 – γxz radians.

    Now, cosA′O′C′ = cos(π/2 – γxz) = sinγxz and as γxz is small, cosA′O′C′ = γxz. From the trigonometrical relationships for a triangle,

    si85_e

       (1.19)

    We showed in Eq. (1.17) that

    si83_e

    Similarly,

    si87_e

    But, for small displacements, the derivatives of u, v, and w are small compared with l, so that, as we are concerned here with actual length rather than change in length, we may use the approximations

    si88_e

    Again, to a first approximation,

    si89_e

    Substituting for O′A′, O′C′, and A′C′ in Eq. (1.19), we have

    si90_e

    Expanding and neglecting fourth-order powers gives

    si91_e

    or,

    si92_e

       (1.20)

    It must be emphasized that Eqs. (1.18) and (1.20) are derived on the assumption that the displacements involved are small. Normally, these linearized equations are adequate for most types of structural problem, but in cases where deflections are large, for example, types of suspension cable, the full, nonlinear, large deflection equations, given in many books on elasticity, must be employed.

    1.10 Compatibility equations

    In Section 1.9, we expressed the six components of strain at a point in a deformable body in terms of the three components of displacement at that point, u, v, and w. We supposed that the body remains continuous during the deformation, so that no voids are formed. It follows that each component, u, v, and w, must be a continuous, single-valued function or, in quantitative terms,

    si93_e

    If voids are formed, then displacements in regions of the body separated by the voids are expressed as different functions of x, y, and z. The existence, therefore, of just three single-valued functions for displacement is an expression of the continuity or compatibility of displacement, which we presupposed.

    Since the six strains are defined in terms of three displacement functions, they must bear some relationship to each other and cannot have arbitrary values. These relationships are found as follows. Differentiating γxy from Eq. (1.20) with respect to x and y gives

    si94_e

    or, since the functions of u and v are continuous,

    si95_e

    which may be written, using Eq. (1.18), as

    si96_e    (1.21)

    In a similar manner,

    si97_e    (1.22)

    si98_e    (1.23)

    If we now differentiate γxy with respect to x and z and add the result to γxz, differentiated with respect to y and x, we obtain

    si99_e

    or

    si100_e

    Substituting from Eqs. (1.18) and (1.21) and rearranging,

    si101_e

       (1.24)

    Similarly,

    si102_e

       (1.25)

    and

    si103_e

       (1.26)

    Equations (1.21)–(1.26) are the six equations of strain compatibility which must be satisfied in the solution of three-dimensional problems in elasticity.

    1.11 Plane strain

    Although we derived the compatibility equations and the expressions for strain for the general three-dimensional state of strain, we shall be concerned mainly with the two-dimensional case described in Section 1.4. The corresponding state of strain, in which it is assumed that particles of the body suffer displacements in one plane only, is known as plane strain. We shall suppose that this plane is, as for plane stress, the xy plane. Then, ɛz, γxz, and γyz become zero and Eqs. (1.18) and (1.20) reduce to

    si104_e    (1.27)

    and

    si105_e    (1.28)

    Further, by substituting ɛz = γxz = γyz = 0 in the six equations of compatibility and noting that ɛx, ɛy, and γxy are now purely functions of x and y, we are left with Eq. (1.21), namely,

    si106_e

    as the only equation of compatibility in the two-dimensional or plane strain case.

    1.12 Determination of strains on inclined planes

    Having defined the strain at a point in a deformable body with reference to an arbitrary system of coordinate axes, we may calculate direct strains in any given direction and the change in the angle (shear strain) between any two originally perpendicular directions at that point. We shall consider the two-dimensional case of plane strain described in Section 1.11.

    An element in a two-dimensional body subjected to the complex stress system of Fig. 1.16(a) distorts into the shape shown in Fig. 1.16(b). In particular, the triangular element ECD suffers distortion to the shape E′C′D′ with corresponding changes in the length FC and angle EFC. Suppose that the known direct and shear strains associated with the given stress system are ɛx, ɛy, and γxy (the actual relationships are investigated later) and we are required to find the direct strain ɛn in a direction normal to the plane ED and the shear strain γ produced by the shear stress acting on the plane ED.

    f01-16-9780081009147

    Figure 1.16 (a) Stress System on a Rectangular Element; (b) Distorted Shape of the Element Due to Stress System in (a)

    To a first order of approximation,

    si107_e    (1.29)

    where ɛn + π/2 is the direct strain in the direction ED. From the geometry of the triangle E′C′D′ in which angle E′C′D′ = π/2 – γxy,

    si108_e

    or, substituting from Eqs. (1.29),

    si109_e

    Noting that (ED)² = (CD)² + (CE)² and neglecting squares and higher powers of small quantities, this equation may be rewritten as

    si110_e

    Dividing through by 2(ED)² gives

    si111_e

       (1.30)

    The strain ɛn in the direction normal to the plane ED is found by replacing the angle θ in Eq. (1.30) by θ – π/2. Hence,

    si112_e

       (1.31)

    Turning our attention to the triangle C′F′E′, we have

    si113_e

       (1.32)

    in which

    si114_e

    Substituting for C′E′, C′F′, and F′E′ in Eq. (1.32) and writing cos(π/2 – γ) = sinγ, we find

    si115_e

       (1.33)

    All the strains are assumed to be small, so that their squares and higher powers may be ignored. Further, sinγ ≈ γ and Eq. (1.33) becomes

    si116_e

    From Fig. 1.16(a), (CE)² = (CF)² + (FE)² and the preceding equation simplifies to

    si117_e

    Dividing through by 2(CE)² and transposing,

    si118_e

    Substitution of ɛn and ɛn+ π/2 from Eqs. (1.31) and (1.30) yields

    si119_e    (1.34)

    1.13 Principal strains

    If we compare Eqs. (1.31) and (1.34) with Eqs. (1.8) and (1.9), we observe that they may be obtained from Eqs. (1.8) and (1.9) by replacing σn with ɛn, σx by ɛx, σy by ɛy, τxy by γxy/2, and τ by γ/2. Therefore, for each deduction made from Eqs. (1.8) and (1.9) concerning σn and τ, there is a corresponding deduction from Eqs. (1.31) and (1.34) regarding ɛn and γ/2.

    Therefore, at a point in a deformable body, there are two mutually perpendicular planes on which the shear strain γ is zero and normal to which the direct strain is a maximum or minimum. These strains are the principal strains at that point and are given (from comparison with Eqs. (1.11) and (1.12)) by

    si120_e

       (1.35)

    and

    si121_e

       (1.36)

    If the shear strain is zero on these planes, it follows that the shear stress must also be zero; and we deduce, from Section 1.7, that the directions of the principal strains and principal stresses coincide. The related planes are then determined from Eq. (1.10) or from

    si122_e    (1.37)

    In addition, the maximum shear strain at the point is

    si123_e    (1.38)

    or

    si124_e    (1.39)

    (compare with Eqs. (1.14) and (1.15)).

    1.14 Mohr's circle of strain

    We now apply the arguments of Section 1.13 to the Mohr's circle of stress described in Section 1.8. A circle of strain, analogous to that shown in Fig. 1.12(b), may be drawn when σx, σy, etc., are replaced by ɛx, ɛy, etc., as specified in Section 1.13. The horizontal extremities of the circle represent the principal strains, the radius of the circle, half the maximum shear strain, and so on.

    1.15 Stress–strain relationships

    In the preceding sections, we developed, for a three-dimensional deformable body, three equations of equilibrium (Eqs. (1.5)) and six strain-displacement relationships (Eqs. (1.18) and (1.20)). From the latter, we eliminated displacements, thereby deriving six auxiliary equations relating strains. These compatibility equations are an expression of the continuity of displacement, which we have assumed as a prerequisite of the analysis. At this stage, therefore, we have obtained nine independent equations toward the solution of the three-dimensional stress problem. However, the number of unknowns totals 15, comprising six stresses, six strains, and three displacements. An additional six equations are therefore necessary to obtain a solution.

    So far we have made no assumptions regarding the force–displacement or stress–strain relationship in the body. This will, in fact, provide us with the required six equations, but before these are derived, it is worthwhile considering some general aspects of the analysis.

    The derivation of the equilibrium, strain–displacement, and compatibility equations does not involve any assumption as to the stress–strain behavior of the material of the body. It follows that these basic equations are applicable to any type of continuous, deformable body, no matter how complex its behavior under stress. In fact, we shall consider only the simple case of linearly elastic, isotropic materials, for which stress is directly proportional to strain and whose elastic properties are the same in all directions. A material possessing the same properties at all points is said to be homogeneous.

    Particular cases arise where some of the stress components are known to be zero and the number of unknowns may then be no greater than the remaining equilibrium equations which have not identically vanished. The unknown stresses are then found from the conditions of equilibrium alone and the problem is said to be statically determinate. For example, the uniform stress in the member supporting a tensile load P in Fig. 1.3 is found by applying one equation of equilibrium and a boundary condition. This system is therefore statically determinate.

    Statically indeterminate systems require the use of some, if not all, of the other equations involving strain–displacement and stress–strain relationships. However, whether the system be statically determinate or not, stress–strain relationships are necessary to determine deflections. The role of the six auxiliary compatibility equations will be discussed when actual elasticity problems are formulated in Chapter 2.

    We now proceed to investigate the relationship of stress and strain in a three–dimensional, linearly elastic, isotropic body.

    Experiments show that the application of a uniform direct stress, say σx, does not produce any shear distortion of the material and that the direct strain ɛx is given by the equation

    si125_e    (1.40)

    where E is a constant known as the modulus of elasticity or Young's modulus. Equation (1.40) is an expression of Hooke's law. Further, ɛx is accompanied by lateral strains

    si126_e    (1.41)

    in which ν is a constant termed Poisson's ratio.

    For a body subjected to direct stresses σx, σy, and σz, the direct strains are, from Eqs. (1.40) and (1.41) and the principle of superposition (see Chapter 5, Section 5.9),

    si127_e    (1.42)

    Equations (1.42) may be transposed to obtain expressions for each stress in terms of the strains. The procedure adopted may be any of the standard mathematical approaches and gives

    si128_e

       (1.43)

    si129_e

       (1.44)

    si130_e

       (1.45)

    in which

    si131_e

    See Eq. (1.53).

    For the case of plane stress in which σz = 0, Eqs. (1.43) and (1.44) reduce to

    si132_e    (1.46)

    si133_e    (1.47)

    Suppose now that, at some arbitrary point in a material, there are principal strains ɛI and ɛII corresponding to principal stresses σI and σII. If these stresses (and strains) are in the direction of the coordinate axes x and y, respectively, then τxy = γxy = 0 and, from Eq. (1.34), the shear strain on an arbitrary plane at the point inclined at an angle θ to the principal planes is

    si134_e    (1.48)

    Using the relationships of Eqs. (1.42) and substituting in Eq. (1.48), we have

    si135_e

    or

    si136_e    (1.49)

    Using Eq. (1.9) and noting that for this particular case τxy = 0, σx = σI, and σy = σII,

    si137_e

    from which we may rewrite Eq. (1.49) in terms of τ as

    si138_e    (1.50)

    The term E/2(1 + ν) is a constant known as the modulus of rigidity G. Hence,

    si139_e

    and the shear strains γxy, γxz, and γyz are expressed in terms of their associated shear stresses as follows:

    si140_e

       (1.51)

    Equations (1.51), together with Eqs. (1.42), provide the additional six equations required to determine the 15 unknowns in a general three-dimensional problem in elasticity. They are, however, limited in use to a linearly elastic, isotropic body.

    For the case of plane stress, they simplify to

    si141_e    (1.52)

    It may be seen from the third of Eqs. (1.52) that the conditions of plane stress and plane strain do not necessarily describe identical situations. See Ex. 1.1.

    Changes in the linear dimensions of a strained body may lead to a change in volume. Suppose that a small element of a body has dimensions δx, δy, and δz. When subjected to a three-dimensional stress system, the element sustains a volumetric strain e (change in volume/unit volume) equal to

    si142_e

    Neglecting products of small quantities in the expansion of the right-hand side of this equation yields

    si143_e    (1.53)

    Substituting for ɛx, ɛy, and ɛz from Eqs. (1.42), we find, for a linearly elastic, isotropic body,

    si144_e

    or

    si145_e

    In the case of a uniform hydrostatic pressure, σx = σy = σz = –p and

    si146_e    (1.54)

    The constant E/3(1 – 2ν) is known as the bulk modulus or modulus of volume expansion and is often given the symbol K.

    An examination of Eq. (1.54) shows that ν ≤ 0.5, since a body cannot increase in volume under pressure. Also, the lateral dimensions of a body subjected to uniaxial tension cannot increase, so that ν > 0. Therefore, for an isotropic material 0 ≤ ν ≤ 0.5 and for most isotropic materials, ν is in the range 0.25–0.33 below the elastic limit. Above the limit of proportionality, ν increases and approaches 0.5.

    Example 1.4

    A rectangular element in a linearly elastic, isotropic material is subjected to tensile stresses of 83 and 65 N/mm² on mutually perpendicular planes. Determine the strain in the direction of each stress and in the direction perpendicular to both stresses. Find also the principal strains, the maximum shear stress, the maximum shear strain, and their directions at the point. Take E = 200,000 N/mm² and v = 0.3.

    If we assume that σx = 83 N/mm² and σy = 65 N/mm², then from Eqs (1.52),

    si147_e

    In this case, since there are no shear stresses on the given planes, σx and σy are principal stresses, so that ɛx and ɛy are the principal strains and are in the directions of σx and σy. It follows from Eq. (1.15) that the maximum shear stress (in the plane of the stresses) is

    si148_e

    acting on planes at 45° to the principal planes.

    Further, using Eq. (1.50), the maximum shear strain is

    si149_e

    so that γmax = 1.17 × 10−4 on the planes of maximum shear stress.

    Example 1.5

    At a particular point in a structural member, a two-dimensional stress system exists where σx = 60 N/mm², σy = –40 N/mm², and τxy = 50 N/mm². If Young's modulus E = 200,000 N/mm² and Poisson's ratio ν = 0.3, calculate the direct strain in the x and y directions and the shear strain at the point. Also calculate the principal strains at the point and their inclination to the plane on which σx acts; verify these answers using a graphical method.

    From Eqs. (1.52),

    si150_esi151_e

    From Eq. (1.50), the shear modulus, G, is given by

    si152_e

    Hence, from Eqs. (1.52),

    si153_e

    Now substituting in Eq. (1.35) for ɛx, ɛy, and γxy,

    si154_e

    which gives

    si155_e

    Similarly, from Eq. (1.36),

    si156_e

    From Eq. (1.37),

    si157_e

    Therefore,

    si158_e

    so that

    si159_e

    The values of ɛI, ɛII, and θ are verified using Mohr's circle of strain (Fig. 1.17). Axes Oɛ and Oγ are set up and the points Q1(360 × 10−6, ½ × 650 × 10−6) and Q2(–290 × 10−6, –½ × 650 × 10−6) located. The center C of the circle is the intersection of Q1Q2 and the Oɛ axis. The circle is then drawn with radius CQ1 and the points B(ɛI) and A(ɛII) located. Finally, angle Q1CB = 2θ and angle Q1CA = 2θ + π.

    f01-17-9780081009147

    Figure 1.17 Mohr's Circle of Strain for Example 1.5

    1.15.1 Temperature effects

    The stress–strain relationships of Eqs. (1.43)–(1.47) apply to a body or structural member at a constant uniform temperature. A temperature rise (or

    Enjoying the preview?
    Page 1 of 1