ChapterPDF Available

Classification of Composite Materials

Authors:

Abstract

Fibre reinforced polymer composite systems have become increasingly important in a variety of engineering fields. Naturally, the rapid growth in the use of composite materials for structural elements has motivated the extension of existing theories in structural mechanics, therein.
HolmAltenbach· JohannesAltenbach
WolfgangKissing
Mechanics of
Composite
Structural
Elements
Second Edition
Mechanics of Composite Structural Elements
Holm Altenbach Johannes Altenbach
Wolfgang Kissing
Mechanics of Composite
Structural Elements
Second Edition
123
Holm Altenbach
Institut für Mechanik
Otto-von-Guericke-Universität Magdeburg
Magdeburg, Saxony-Anhalt
Germany
Johannes Altenbach
Magdeburg
Germany
Wolfgang Kissing
Bad Kleinen, Mecklenburg-Vorpommern
Germany
ISBN 978-981-10-8934-3 ISBN 978-981-10-8935-0 (eBook)
https://doi.org/10.1007/978-981-10-8935-0
Library of Congress Control Number: 2018939885
1st edition: ©Springer-Verlag Berlin Heidelberg 2004
2nd edition: ©Springer Nature Singapore Pte Ltd. 2018
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microlms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specic statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional afliations.
Printed on acid-free paper
This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
part of Springer Nature
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
This is the second edition of the textbook Mechanics of Composite Structural El-
ements published first in 2004. Since that time the course has been delivered at
several universities in Germany and abroad. Throughout the past years the authors
received a lot of suggestions for improvements from students and colleagues alike.
In addition, the textbook is recommended as the basic reading material in a relevant
course at the Otto-von-Guericke-Universit¨at Magdeburg.
In 2016 the first author was invited by Prof. Andreas ¨
Ochsner to present a course
with the same title at the Griffith University (Gold Coast campus) for third and
fourth year students of the bachelor program in the departments of Mechanical Engi-
neering and Civil Engineering. The two weeks’ course included 60 hours of lectures
and tutorials. Finally, the course was concluded with a written exam and a project.
Our special thanks are due to Dr. Christoph Baumann and Springer who provided
personal copies of the first edition of the book for the attendants of the course. As
the result of the discussions with the students the idea was born to prepare a second
edition.
By and large the preliminaries of the first edition remain unchanged: the presen-
tation of the mechanics of composite materials is based on the knowledge of the
first and second year of the bachelor program in Engineering Mechanics (or in other
countries the courses of General Mechanics and Strength of Materials). The focus of
the students will be directed to the elementary theory as the starting point of further
advanced courses.
There are some changes in the second edition in comparison with the first one:
some problems are added or clarified (and we hope now better understandable),
Chapter 11 is slightly shortened (some details are no more important),
some details were adopted considering the developments of Springer’s templates.
Some references for further reading, but also some original sources are added and
the tables with material data are improved. Of course, we hope you will now find
fewer misprints and typos.
We have to acknowledge Dr.-Ing. Heinz K ¨oppe (Otto-von-Guericke-Universit¨at
Magdeburg) and Dipl.-Ing. Christoph Kammer (formaly at Otto-von-Guericke-
v
Preface to the 2nd Edition
vi Preface
Universit¨at Magdeburg) for finding a lot of typos. In addition, we have to thank
Dr. Christoph Baumann (Executive Editor Engineering, Springer Nature Singapore)
for the permanent support of the project. We appreciate for any comment and sug-
gestion for improvements which should be sent to holm.altenbach@ovgu.de.
Magdeburg and Bad Kleinen, Holm Altenbach
March 2018 Johannes Altenbach
Wolfgang Kissing
Preface to the 1st Edition
Laminate and sandwich structures are typical lightweight elements with rapidly ex-
panding application in various industrial fields. In the past, these structures were
used primarily in aircraft and aerospace industries. Now, they have also found ap-
plication in civil and mechanical engineering, in the automotive industry, in ship-
building, the sport goods industries, etc. The advantages that these materials have
over traditional materials like metals and their alloys are the relatively high specific
strength properties (the ratio strength to density, etc). In addition, the laminate and
sandwich structures provide good vibration and noise protection, thermal insulation,
etc. There are also disadvantages - for example, composite laminates are brittle, and
the joining of such elements is not as easy as with classical materials. The recycling
of these materials is also problematic, and a viable solution is yet to be developed.
Since the application of laminates and sandwiches has been used mostly in new
technologies, governmental and independent research organizations, as well as big
companies, have spent a lot of money for research. This includes the development
of new materials by material scientists, new design concepts by mechanical and
civil engineers as well as new testing procedures and standards. The growing de-
mands of the industry for specially educated research and practicing engineers and
material scientists have resulted in changes in curricula of the diploma and master
courses. More and more universities have included special courses on laminates and
sandwiches, and training programs have been arranged for postgraduate studies.
The concept of this textbook was born 10 years ago. At that time, the first edition
of ”Einf ¨uhrung in die Mechanik der Laminat- und Sandwichtragwerke”, prepared
by H. Altenbach, J. Altenbach and R. Rikards, was written for German students
only. The purpose of that book consisted the following objectives:
to provide a basic understanding of composite materials like laminates and sand-
wiches,
to perform and engineering analysis of structural elements like beams and plates
made from laminates and sandwiches,
to introduce the finite element method for the numerical treatment of composite
structures and
vii
viii Preface
to discuss the limitations of analysis and modelling concepts.
These four items are also included in this textbook. It must be noted that between
1997 and 2000, there was a common education project sponsored by the European
Community (coordinator T. Sadowski) with the participation of colleagues from
U.K., Belgium, Poland and Germany. One of the main results was a new created
course on laminates and sandwiches, and finally an English textbook ”Structural
Analysis of Laminate and Sandwich Beams and Plates” written by H. Altenbach, J.
Altenbach and W. Kissing.
The present textbook follows the main ideas of its previous versions, but has been
significantly expanded. It can be characterized by the following items:
The textbook is written in the style of classical courses of strength of materials (or
mechanics of materials) and theory of beams, plates and shells. In this sense the
course (textbook) can be recommended for master students with bachelor degree
and diploma students which have finished the second year in the university. In
addition, postgraduates of various levels can find a simple introduction to the
analysis and modelling of laminate and sandwich structures.
In contrast to the traditional courses referred to above, two extensions have been
included. Firstly, consideration is given to the linear elastic material behavior of
both isotropic and anisotropic structural elements. Secondly, the case of inhomo-
geneous material properties in the thickness direction was also included.
Composite structures are mostly thin, in which case a dimension reduction of the
governing equations is allowed in many applications. Due to this fact, the one-
dimensional equations for beams and the two-dimensional equations for plates
and shells are introduced. The presented analytical solutions can be related to the
in-plane, out-of-plane and coupled behavior.
Sandwiches are introduced as a special case of general laminates. This results in
significant simplifications because sandwiches with thin or thicker faces can be
modelled and analyzed in the frame of laminate theories of different order and
so a special sandwich theory is not necessary.
All analysis concepts are introduced for the global structural behavior. Local
effects and their analysis must be based on three-dimensional field equations
which can usually be solved with the help of numerical methods. It must be
noted that the thermomechanical properties of composites on polymer matrix at
high temperatures can be essentially different from those at normal temperatures.
In engineering applications generally three levels of temperature are considered
normal or room temperature (10–30C)
elevated temperatures (30–200C)
high temperatures (>200C)
High temperatures yield an irreversible variation of the mechanical properties,
and thus are not included in modelling and analysis. All thermal and moisture ef-
fects are considered in such a way that the mechanical properties can be assumed
unchanged.
Preface ix
Finite element analysis is only briefly presented. A basic course in finite elements
is necessary for the understanding of this part of the book. It should be noted
that the finite element method is general accepted for the numerical analysis of
laminate and sandwich structures. This was the reason to include this item in the
contents of this book.
The textbook is divided into 11 chapters and several appendices summarizing the
material properties (for matrix and fiber constituents, etc) and some mathematical
formulas:
In the first part (Chaps. 1–3) an introduction into laminates and sandwiches as
structural materials, the anisotropic elasticity, variational methods and the basic
micromechanical models are presented.
The second part (Chaps. 4–6) can be related to the modelling from single laminae
to laminates including sandwiches, the improved theories and simplest failure
concepts.
The third part (Chaps. 7–9) discusses structural elements (beams, plates and
shells) and their analysis if they are made from laminates and sandwiches. The
modelling of laminated and sandwich plates and shells is restricted to rectan-
gular plates and circular cylindrical shells. The individual fiber reinforced lam-
inae of laminated structured elements are considered to be homogeneous and
orthotropic, but the laminate is heterogeneous through the thickness and gener-
ally anisotropic. An equivalent single layer theory using the classical lamination
theory, and the first order shear deformation theory are considered. Multilayered
theories or laminate theories of higher order are not discussed in detail.
The fourth part (Chap. 10) includes the modelling and analysis of thin-walled
folded plate structures or generalized beams. This topic is not normally consid-
ered in standard textbooks on structural analysis of laminates and sandwiches, but
it was included here because it demonstrates the possible application of Vlasov’s
theory of thin-walled beams and semi-membrane shells on laminated structural
elements.
Finally, the fifth part (Chap. 11) presents a short introduction into the finite el-
ement procedures and developed finite classical and generalized beam elements
and finite plate elements in the frame of classical and first order shear defor-
mation theory. Selected examples demonstrate the possibilities of finite element
analysis.
This textbook is written for use not only in engineering curricula of aerospace, civil
and mechanical engineering, but also in material science and applied mechanics. In
addition, the book may be useful for practicing engineers, lectors and researchers in
the mechanics of structures composed of composite materials.
The strongest feature of the book is its use as a textbook. No prior knowledge of
composite materials and structures is required for the understanding of its content.
It intends to give an in-depth view of the problems considered and therefore the
number of topics considered is limited. A large number of solved problems are
included to assess the knowledge of the presented topics. The list of references at
the end of the book focuses on three groups of suggested reading:
x Preface
Firstly, a selection of textbooks and monographs of composite materials and
structures are listed, which constitute the necessary items for further reading.
They are selected to reinforce the presented topics and to provide information
on topics not discussed. We hope that our colleagues agree that the number of
recommended books for a textbook must be limited and we have given priority
to newer books available in university libraries.
Some books on elasticity, continuum mechanics, plates and shells and FEM are
recommended for further reading, and a deeper understanding of the mathemati-
cal, mechanical and numerical topics.
A list of review articles shall enable the reader to become informed about the
numerous books and proceedings in composite mechanics.
The technical realization of this textbook was possible only with the support of
various friends and colleagues. Firstly, we would like to express our special thanks
to K. Naumenko and O. Dyogtev for drawing most of the figures. Secondly, Mrs. B.
Renner and T. Kumar performed many corrections of the English text. At the same
time Mrs. Renner checked the problems and solutions. We received access to the
necessary literature by Mrs. N. Altenbach. Finally, the processing of the text was
done by Mrs. S. Runkel. We would also like to thank Springer Publishing for their
service.
Any comments or remarks are welcome and we kindly ask them to be sent to
holm.altenbach@iw.uni-halle.de.
June 2003
Halle Holm Altenbach
Magdeburg Johannes Altenbach
Wismar Wolfgang Kissing
Contents
Part I Introduction, Anisotropic Elasticity, Micromechanics
1 Classification of Composite Materials ............................ 3
1.1 Definition and Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Significance and Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Material Characteristics of the Constituents . . . . . . . . . . . . . . . . . . . . . 14
1.5 Advantages and Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
References..................................................... 18
2 Linear Anisotropic Materials .................................... 19
2.1 Generalized Hooke’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.1 Stresses, Strains, Stiffness, and Compliances . . . . . . . . . . . . . 21
2.1.2 Transformation Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1.3 Symmetry Relations of Stiffness and Compliance Matrices . 32
2.1.3.1 Monoclinic or Monotropic Material Behavior . . . . 32
2.1.3.2 Orthotropic Material Behavior . . . . . . . . . . . . . . . . . 34
2.1.3.3 Transversely Isotropic Material Behavior . . . . . . . . 35
2.1.3.4 Isotropic Material Behavior. . . . . . . . . . . . . . . . . . . . 36
2.1.4 Engineering Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.1.4.1 Orthotropic Material Behavior . . . . . . . . . . . . . . . . . 36
2.1.4.2 Transversally-Isotropic Material Behavior . . . . . . . 40
2.1.4.3 Isotropic Material Behavior. . . . . . . . . . . . . . . . . . . . 42
2.1.4.4 Monoclinic Material Behavior . . . . . . . . . . . . . . . . . 43
2.1.5 Two-Dimensional Material Equations . . . . . . . . . . . . . . . . . . . 45
2.1.6 Curvilinear Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2.1.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.2 Fundamental Equations and Variational Solution Procedures . . . . . . 59
2.2.1 Boundary and Initial-Boundary Value Equations . . . . . . . . . . 59
2.2.2 Principle of Virtual Work and Energy Formulations . . . . . . . . 63
xi
xii Contents
2.2.3 Variational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.2.3.1 Rayleigh-Ritz Method . . . . . . . . . . . . . . . . . . . . . . . . 69
2.2.3.2 Weighted Residual Methods . . . . . . . . . . . . . . . . . . . 73
2.2.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
References..................................................... 84
3 Effective Material Moduli for Composites . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.1 Elementary Mixture Rules for Fibre-Reinforced Laminae . . . . . . . . . 86
3.1.1 Effective Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.1.2 Effective Longitudinal Modulus of Elasticity . . . . . . . . . . . . . 88
3.1.3 Effective Transverse Modulus of Elasticity . . . . . . . . . . . . . . . 89
3.1.4 Effective Poisson’s Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.1.5 Effective In-Plane Shear Modulus . . . . . . . . . . . . . . . . . . . . . . 91
3.1.6 Discussion on the Elementary Mixture Rules . . . . . . . . . . . . . 92
3.2 Improved Formulas for Effective Moduli of Composites . . . . . . . . . . 93
3.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Part II Modelling of a Single Laminae, Laminates and Sandwiches
4 Elastic Behavior of Laminate and Sandwich Composites . . . . . . . . . . . . 103
4.1 Elastic Behavior of Laminae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.1.1 On-Axis Stiffness and Compliances of UD-Laminae . . . . . . . 104
4.1.2 Off-Axis Stiffness and Compliances of UD-Laminae . . . . . . 109
4.1.3 Stress Resultants and Stress Analysis . . . . . . . . . . . . . . . . . . . . 118
4.1.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.2 Elastic Behavior of Laminates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
4.2.1 General Laminates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.2.2 Stress-Strain Relations and Stress Resultants . . . . . . . . . . . . . 135
4.2.3 Laminates with Special Laminae Stacking Sequences . . . . . . 142
4.2.3.1 Symmetric Laminates . . . . . . . . . . . . . . . . . . . . . . . . 143
4.2.3.2 Antisymmetric Laminates . . . . . . . . . . . . . . . . . . . . . 148
4.2.3.3 Stiffness Matrices for Symmetric and
Unsymmetric Laminates in Engineering
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.2.4 Stress Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.2.5 Thermal and Hygroscopic Effects . . . . . . . . . . . . . . . . . . . . . . 158
4.2.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.3 Elastic Behavior of Sandwiches. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
4.3.1 General Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.3.2 Stress Resultants and Stress Analysis . . . . . . . . . . . . . . . . . . . . 170
4.3.3 Sandwich Materials with Thick Cover Sheets . . . . . . . . . . . . . 172
4.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Contents xiii
5 Classical and Improved Theories .................................177
5.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
5.2 Classical Laminate Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
5.3 Shear Deformation Theory for Laminates and Sandwiches . . . . . . . . 188
5.4 Layerwise Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
5.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
6 Failure Mechanisms and Criteria ................................201
6.1 Fracture Modes of Laminae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
6.2 Failure Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
6.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
Part III Analysis of Structural Elements
7 Modelling and Analysis of Beams ................................227
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
7.2 Classical Beam Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
7.3 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
7.4 Sandwich Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
7.4.1 Stresses and Strains for Symmetrical Cross-Sections . . . . . . . 250
7.4.2 Stresses and Strains for Non-Symmetrical Cross-Sections . . 254
7.4.3 Governing Sandwich Beam Equations . . . . . . . . . . . . . . . . . . . 255
7.5 Hygrothermo-Elastic Effects on Beams . . . . . . . . . . . . . . . . . . . . . . . . 259
7.6 Analytical Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
8 Modelling and Analysis of Plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
8.2 Classical Laminate Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
8.3 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
8.4 Sandwich Plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
8.5 Hygrothermo-Elastic Effects on Plates . . . . . . . . . . . . . . . . . . . . . . . . . 299
8.6 Analytical Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
8.6.1 Classical Laminate Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
8.6.1.1 Plate Strip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
8.6.1.2 Navier Solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
8.6.1.3 N´adai-L´evy Solution . . . . . . . . . . . . . . . . . . . . . . . . . 312
8.6.2 Shear Deformation Laminate Theory . . . . . . . . . . . . . . . . . . . . 316
8.6.2.1 Plate Strip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
8.6.2.2 Navier Solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
8.6.2.3 N´adai-L´evy Solution . . . . . . . . . . . . . . . . . . . . . . . . . 322
8.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
xiv Contents
9 Modelling and Analysis of Circular Cylindrical Shells . . . . . . . . . . . . . . 341
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
9.2 Classical Shell Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
9.2.1 General Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
9.2.2 Specially Orthotropic Circular Cylindrical Shells
Subjected by Axial Symmetric Loads . . . . . . . . . . . . . . . . . . . 346
9.2.3 Membrane and Semi-Membrane Theories. . . . . . . . . . . . . . . . 350
9.3 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352
9.4 Sandwich Shells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
9.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
Part IV Modelling and Analysis of Thin-Walled Folded Plate Structures
10 Modelling and Analysis of Thin-walled Folded Structures . . . . . . . . . . 367
10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
10.2 Generalized Beam Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
10.2.1 Basic Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
10.2.2 Potential Energy of the Folded Structure . . . . . . . . . . . . . . . . . 375
10.2.3 Reduction of the Two-dimensional Problem . . . . . . . . . . . . . . 376
10.2.4 Simplified Structural Models . . . . . . . . . . . . . . . . . . . . . . . . . . 381
10.2.4.1 Structural Model A . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
10.2.4.2 Structural Model B . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
10.2.4.3 Structural Model C . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
10.2.4.4 Structural Model D . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
10.2.4.5 Structural Model E . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
10.2.4.6 Further Special Models by Restrictions of the
Cross-Section Kinematics . . . . . . . . . . . . . . . . . . . . . 385
10.2.5 An Efficient Structure Model for the Analysis of General
Prismatic Beam Shaped Thin-walled Plate Structures . . . . . . 387
10.2.6 Free Eigen-Vibration Analysis, Structural Model A . . . . . . . . 388
10.3 Solution Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
10.3.1 Analytical Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
10.3.2 Transfer Matrix Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392
10.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
Part V Finite Classical and Generalized Beam Elements, Finite Plate
Elements
11 Finite Element Analysis .........................................409
11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
11.1.1 FEM Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
11.1.2 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414
11.2 Finite Beam Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
11.2.1 Laminate Truss Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416
Contents xv
11.2.2 Laminate Beam Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
11.2.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
11.3 Finite Plate Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
11.3.1 Classical Laminate Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
11.3.2 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
11.4 Generalized Finite Beam Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
11.4.1 Foundations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
11.4.2 Element Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
11.4.3 Element Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
11.4.4 System Equations and Solution. . . . . . . . . . . . . . . . . . . . . . . . . 444
11.4.5 Equations for the Free Vibration Analysis . . . . . . . . . . . . . . . . 445
11.5 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 446
11.5.1 Examples for the Use of Laminated Shell Elements . . . . . . . . 447
11.5.1.1 Cantilever Beam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
11.5.1.2 Laminate Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
11.5.1.3 Sandwich Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451
11.5.1.4 Buckling Analysis of a Laminate Plate . . . . . . . . . . 452
11.5.2 Examples of the Use of Generalized Beam Elements . . . . . . . 456
Part VI Appendices
A Matrix Operations .............................................463
A.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463
A.2 Special Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465
A.3 Matrix Algebra and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
B Stress and Strain Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
C Differential Operators for Rectangular Plates . . . . . . . . . . . . . . . . . . . . . 473
C.1 Classical Plate Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
C.2 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
D Differential Operators for Circular Cylindrical Shells . . . . . . . . . . . . . . 477
D.1 Classical Shell Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 477
D.2 Shear Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479
E Krylow-Functions as Solution Forms of a Fourth Order Ordinary
Differential Equation ..........................................481
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 482
F Material’s Properties ...........................................483
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
xvi Contents
G References ....................................................489
G.1 Comprehensive Composite Materiala . . . . . . . . . . . . . . . . . . . . . . . . . . 489
G.2 Selected Textbooks and Monographs on Composite Mechanics . . . . 490
G.3 Supplementary Literature for Further Reading . . . . . . . . . . . . . . . . . . 493
G.4 Selected Review Articles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 494
Index .............................................................497
Part I
Introduction, Anisotropic Elasticity,
Micromechanics
In the first part (Chaps. 1–3) an introduction into laminates and sandwiches as
structural materials, the anisotropic elasticity, variational methods and the basic mi-
cromechanical models are presented.
The laminates are introduced as layered structures each of the layers is a fibre-
reinforced material composed of high-modulus, high-strength fibers in a polymeric,
metallic, or ceramic matrix material. Examples of fibers used include graphite, glass,
boron, and silicon carbide, matrix materials are epoxies, polyamide, aluminium, ti-
tanium, and aluminium. A sandwich is a special class of composite materials consist
of two thin but stiff skins and a lightweight but thick core.
The anisotropic elasticity is an extension of the isotropic elasticity. The geomet-
rical relations are assumed to be linear. The constitutive equations contain more
than two material parameters. In addition, the transition from the general three-
dimensional equations to the special two-dimensional equations results in more
complicated constrains. At the same time the introduction of reduced stiffness and
compliance parameters result in a powerful tool for the analysis of laminates.
The variational methods are the base of many numerical solution techniques (for
example, the finite element method). Here only the classical principles and methods
are briefly discussed.
There are many, partly sophisticated micromechanical approaches. They are the
base of a better understanding of the local behavior. In the focus of this textbook is
the global structural analysis. Thats way the micromechanical models are presented
only in the elementary form.
Chapter 1
Classification of Composite Materials
Fibre reinforced polymer composite systems have become increasingly important in
a variety of engineering fields. Naturally, the rapid growth in the use of composite
materials for structural elements has motivated the extension of existing theories in
structural mechanics, therein. The main topics of this textbook are
a short introduction into the linear mechanics of deformable solids with an-
isotropic material behavior,
the mechanical behavior of composite materials as unidirectional reinforced sin-
gle layers or laminated composite materials, the analysis of effective moduli,
some basic mechanisms and criteria of failure,
the modelling of the mechanical behavior of laminates and sandwiches, gen-
eral assumptions of various theories, classical laminate theory (CLT), effect of
stacking of the layers of laminates and the coupling of stretching, bending and
twisting, first order shear deformation theory (FOSDT), an overview on refined
equivalent single layer plate theories and on multilayered plate modelling,
modelling and analysis of laminate and sandwich beams, plates and shells, prob-
lems of bending, vibration and buckling and
modelling and analysis of fibre reinforced long thin-walled folded-plate struc-
tural elements.
The textbook concentrates on a simple unified approach to the basic behavior of
composite materials and the structural analysis of beams, plates and circular cylin-
drical shells made of composite material being a laminate or a sandwich. The in-
troduction into the modelling and analysis of thin-walled folded structural elements
is limited to laminated elements and the CLT. The problems of manufacturing and
recycling of composites will be not discussed, but to use all benefits of the new
young material composite, an engineer has to be more than a material user as for
classical materials as steel or alloys. Structural engineering qualification must in-
clude knowledge of material design, manufacturing methods, quality control and
recycling.
In Chap. 1 some basic questions are discussed, e.g. what are composites and how
they can be classified, what are the main characteristics and significance, micro-
3
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_1
4 1 Classification of Composite Materials
and macro-modelling, why composites are used, what are the advantages and the
limitations. The App. F contains some values of the material characteristics of the
constituents of composites.
1.1 Definition and Characteristics
Material science classifies structural materials into three categories
metals,
ceramics and
polymers.
It is difficult to give an exact assessment of the advantages and disadvantages of
these three basic material classes, because each category covers whole groups of
materials within which the range of properties is often as broad as the differences
between the three material classes. But at the simplistic level some obvious charac-
teristic properties can be identified:
Mostly metals are of medium to high density. They have good thermal stability
and can be made corrosion-resistant by alloying. Metals have useful mechanical
characteristics and it is moderately easy to shape and join. For this reason metals
became the preferred structural engineering material, they posed less problems
to the designer than either ceramic or polymer materials.
Ceramic materials have great thermal stability and are resistant to corrosion,
abrasion and other forms of attack. They are very rigid but mostly brittle and
can only be shaped with difficulty.
Polymer materials (plastics) are of low density, have good chemical resistance
but lack thermal stability. They have poor mechanical properties, but are eas-
ily fabricated and joined. Their resistance to environmental degradation, e.g. the
photomechanical effects of sunlight, is moderate.
A material is called homogeneous if its properties are the same at every point and
therefore independent of the location. Homogeneity is associated with the scale of
modelling or the so-called characteristic volume and the definition describes the
average material behavior on a macroscopic level. On a microscopic level all ma-
terials are more or less homogeneous but depending on the scale, materials can be
described as homogeneous, quasi-homogeneous or inhomogeneous. A material is
inhomogeneous or heterogeneous if its properties depend on location. But in the av-
erage sense of these definitions a material can be regarded as homogeneous, quasi-
homogeneous or heterogeneous if the scale decreases.
A material is isotropic if its properties are independent of the orientation, they do
not vary with direction. Otherwise the material is anisotropic. A general anisotropic
material has no planes or axes of material symmetry, but in Sect. 2.1.3 some special
kinds of material symmetries like orthotropy, transverse isotropy, etc., are discussed
in detail.
1.1 Definition and Characteristics 5
Furthermore, a material can depend on several constituents or phases, single
phase materials are called monolithic. The above three mentioned classes of conven-
tional materials are on the macroscopic level more or less monolithic, homogeneous
and isotropic.
The group of materials which can be defined as composite materials is extremely
large. Its boundaries depend on definition. In the most general definition we can
consider a composite as any material that is a combination of two or more materi-
als, commonly referred to as constituents, and have material properties derived from
the individual constituents. These properties may have the combined characteristics
of the constituents or they are substantially different. Sometimes the material prop-
erties of a composite material may exceed those of the constituents. This general
definition of composites includes natural materials like wood, traditional structural
materials like concrete, as well as modern synthetic composites such as fibre or par-
ticle reinforced plastics which are now an important group of engineering materials
where low weight in combination with high strength and stiffness are required in
structural design.
In the more restrictive sense of this textbook a structural composite consists of
an assembly of two materials of different nature. In general, one material is dis-
continuous and is called the reinforcement, the other material is mostly less stiff
and weaker. It is continuous and is called the matrix. The properties of a composite
material depends on
The properties of the constituents,
The geometry of the reinforcements, their distribution, orientation and concen-
tration usually measured by the volume fraction or fiber volume ratio,
The nature and quality of the matrix-reinforcement interface.
In a less restricted sense, a structural composite can consist of two or more phases
on the macroscopic level. The mechanical performance and properties of compos-
ite materials are superior to those of their components or constituent materials taken
separately. The concentration of the reinforcement phase is a determining parameter
of the properties of the new material, their distribution determines the homogeneity
or the heterogeneity on the macroscopic scale. The most important aspect of com-
posite materials in which the reinforcement are fibers is the anisotropy caused by the
fiber orientation. It is necessary to give special attention to this fundamental charac-
teristic of fibre reinforced composites and the possibility to influence the anisotropy
by material design for a desired quality.
Summarizing the aspects defining a composite as a mixture of two or more dis-
tinct constituents or phases it must be considered that all constituents have to be
present in reasonable proportions that the constituent phases have quite different
properties from the properties of the composite material and that man-made com-
posites are produced by combining the constituents by various means. Figure 1.1
demonstrates typical examples of composite materials. Composites can be classi-
fied by their form and the distribution of their constituents (Fig. 1.2). The reinforce-
ment constituent can be described as fibrous or particulate. The fibres are continuous
(long fibres) or discontinuous (short fibres). Long fibres are arranged usually in uni-
6 1 Classification of Composite Materials
ab
s s ss
s
s
s
sts
s s s
s
s
cd
efg
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
h
Fig. 1.1 Examples of composite materials with different forms of constituents and distributions
of the reinforcements. aLaminate with uni- or bidirectional layers, birregular reinforcement with
long fibres, creinforcement with particles, dreinforcement with plate strapped particles, erandom
arrangement of continuous fibres, firregular reinforcement with short fibres, gspatial reinforce-
ment, hreinforcement with surface tissues as mats, woven fabrics, etc.
or bidirectional, but also irregular reinforcements by long fibres are possible. The
arrangement and the orientation of long or short fibres determines the mechani-
cal properties of composites and the behavior ranges between a general anisotropy
to a quasi-isotropy. Particulate reinforcements have different shapes. They may be
spherical, platelet or of any regular or irregular geometry. Their arrangement may be
random or regular with preferred orientations. In the majority of practical applica-
tions particulate reinforced composites are considered to be randomly oriented and
unidirectional
reinforced
bidirectional
reinforced
spatial
reinforced
random
orientation preferred orientation
continous fibre reinforced
(long fibres)
discontinous fibre reinforced
(short fibres)
random
orientation
preferred
orientation
fibre reinforced particle reinforced
Composite
Fig. 1.2 Classification of composites
1.1 Definition and Characteristics 7
the mechanical properties are homogeneous and isotropic (for mor details Chris-
tensen, 2005; Torquato, 2002). The preferred orientation in the case of continuous
fibre composites is unidirectional for each layer or lamina. Fibre reinforced compos-
ites are very important and in consequence this textbook will essentially deal with
modelling and analysis of structural elements composed of this type of composite
material. However, the level of modelling and analysis used in this textbook does not
really differentiate between unidirectional continuous fibres, oriented short-fibres or
woven fibre composite layers, as long as material characteristics that define the layer
response are used. Composite materials can also be classified by the nature of their
constituents. According to the nature of the matrix material we classify organic,
mineral or metallic matrix composites.
Organic matrix composites are polymer resins with fillers. The fibres can be min-
eral (glass, etc.), organic (Kevlar, etc.) or metallic (aluminium, etc.).
Mineral matrix composites are ceramics with metallic fibres or with metallic or
mineral particles.
Metallic matrix composites are metals with mineral or metallic fibres.
Structural composite elements such as fibre reinforced polymer resins are of par-
ticular interest in this textbook. They can be used only in a low temperature range
up to 200to 300C. The two basic classes of resins are thermosets and thermo-
plastics. Thermosetting resins are the most common type of matrix system for com-
posite materials. Typical thermoset matrices include Epoxy,Polyester,Polyamide
(Thermoplastics) and Vinyl Ester, among popular thermoplastics are Polyethylene,
Polystyrene and Polyether-ether-ketone (PEEK) materials. Ceramic based compos-
ites can also be used in a high temperature range up to 1000C and metallic matrix
composites in a medium temperature range.
In the following a composite material is constituted by a matrix and a fibre re-
inforcement. The matrix is a polyester or epoxy resin with fillers. By the addition
of fillers, the characteristics of resins will be improved and the production costs
reduced. But from the mechanical modelling, a resin-filler system stays as a ho-
mogeneous material and a composite material is a two phase system made up of a
matrix and a reinforcement.
The most advanced composites are polymer matrix composites. They are charac-
terized by relatively low costs, simple manufacturing and high strength. Their main
drawbacks are the low working temperature, high coefficients of thermal and mois-
ture expansion and, in certain directions, low elastic properties. Most widely used
manufacturing composites are thermosetting resins as unsaturated polyester resins
or epoxy resins. The polyester resins are used as they have low production cost.
The second place in composite production is held by epoxy resins. Although epoxy
is costlier than polyester, approximately five time higher in price, it is very popu-
lar in various application fields. More than two thirds of polymer matrices used in
aerospace industries are epoxy based. Polymer matrix composites are usually rein-
forced by fibres to improve such mechanical characteristics as stiffness, strength,
etc. Fibres can be made of different materials (glass, carbon, aramid, etc.). Glass fi-
bres are widely used because their advantages include high strength, low costs, high
8 1 Classification of Composite Materials
chemical resistance, etc., but their elastic modulus is very low and also their fatigue
strength. Graphite or carbon fibres have a high modulus and a high strength and
are very common in aircraft components. Aramid fibres are usually known by the
name of Kevlar, which is a trade name. Summarizing some functional requirements
of fibres and matrices in a fibre reinforced polymer matrix composite
fibres should have a high modulus of elasticity and a high ultimate strength,
fibres should be stable and retain their strength during handling and fabrication,
the variation of the mechanical characteristics of the individual fibres should be
low, their diameters uniform and their arrangement in the matrix regular,
matrices have to bind together the fibres and protect their surfaces from damage,
matrices have to transfer stress to the fibres by adhesion and/or friction and
matrices have to be chemically compatible with fibres over the whole working
period.
The fibre length, their orientation, their shape and their material properties are main
factors which contribute to the mechanical performance of a composite. Their vol-
ume fraction usually lies between 0.3 and 0.7. Although matrices by themselves
generally have poorer mechanical properties than compared to fibres, they influence
many characteristics of the composite such as the transverse modulus and strength,
shear modulus and strength, thermal resistance and expansion, etc.
An overview of the material characteristics is given in Sect. 1.4. One of the most
important factors which determines the mechanical behavior of a composite material
is the proportion of the matrix and the fibres expressed by their volume or their
weight fraction. These fractions can be established for a two phase composite in the
following way. The volume Vof the composite is made from a matrix volume Vm
and a fibre volume Vf(V=Vf+Vm). Then
vf=Vf
V,vm=Vm
V(1.1.1)
with
vf+vm=1,vm=1vf
are the fibre and the matrix volume fractions. In a similar way the weight or mass
fractions of fibres and matrices can be defined. The mass Mof the composite is
made from Mfand Mm(M=Mf+Mm) and
mf=Mf
M,mm=Mm
M(1.1.2)
with
mf+mm=1,mm=1mf
are the mass fractions of fibres and matrices. With the relation between volume,
mass and density
ρ
=M/V, we can link the mass and the volume fractions
1.2 Significance and Objectives 9
ρ
=M
V=Mf+Mm
V=
ρ
fVf+
ρ
mVm
V
=
ρ
fvf+
ρ
mvm=
ρ
fvf+
ρ
m(1vf)
(1.1.3)
Starting from the total volume of the composite V=Vf+Vmwe obtain
M
ρ
=Mf
ρ
f
+Mm
ρ
m
and
ρ
=1
mf
ρ
f
+mm
ρ
m
(1.1.4)
with
mf=
ρ
f
ρ
vf,mm=
ρ
m
ρ
vm
The inverse relation determines
vf=
ρ
ρ
f
mf,vm=
ρ
ρ
m
mm(1.1.5)
The density
ρ
is determined by Eqs. (1.1.3) or (1.1.4). The equations can be easily
extended to multi-phase composites.
Mass fractions are easier to measure in material manufacturing, but volume frac-
tions appear in the theoretical equations for effective moduli (Sect. 3.1). Therefore,
it is helpful to have simple expressions for shifting from one fraction to the other.
The quality of a composite material decreases with an increase in porosity. The
volume of porosity should be less than 5 % for a medium quality and less than 1 %
for a high quality composite. If the density
ρ
exp is measured experimentally and
ρ
theor is calculated with (1.1.4), the volume fraction of porosity is given by
vpor =
ρ
theor
ρ
exp
ρ
theor
(1.1.6)
1.2 Significance and Objectives
Development and applications of composite materials and structural elements com-
posed of composite materials have been very rapid in the last decades. The mo-
tivations for this development are the significant progress in material science and
technology of the composite constituents, the requirements for high performance
materials is not only in aircraft and aerospace structures, but also in the develop-
ment of very powerful experimental equipments and numerical methods and the
availability of efficient computers. With the development of composite materials
a new material design is possible that allows an optimal material composition in
connection with the structural design. A useful and correct application of compos-
ite materials requires a close interaction of different engineering disciplines such
10 1 Classification of Composite Materials
as structural design and analysis, material science, mechanics of materials, process
engineering, etc. Summarizing the main topics of composite material research and
technology are
investigation of all characteristics of the constituent and the composite materials,
material design and optimization for the given working conditions,
development of analytical modelling and solution methods for determining ma-
terial and structural behavior,
development of experimental methods for material characteristics, stress and de-
formation states, failure,
modelling and analysis of creep, damage and life prediction,
development of new and efficient fabrication and recycling procedures among
others.
Within the scope of this book are the first three items.
The most significant driving force in the composite research and application was
weight saving in comparison to structures of conventional materials such as steel,
alloys, etc. However, to have only material density, stiffness and strength in mind
when thinking of composites is a very narrow view of the possibilities of such ma-
terials as fibre-reinforced plastics because they often may score over conventional
materials as metals not only owing to their mechanical properties. Fibre reinforced
plastics are extremely corrosion-resistant and have interesting electromagnetic prop-
erties. In consequence they are used for chemical plants and for structures which
require non-magnetic materials. Further carbon fibre reinforced epoxy is used in
medical applications because it is transparent to X-rays.
With applications out of aerospace or aircraft, cost competitiveness with con-
ventional materials became important. More recently requirements such as quality
assurance, reproducibility, predictability of the structure behavior over its life time,
recycling, etc. became significant.
Applications of polymer matrix composites range from the aerospace industry to
the industry of sports goods. The military aircraft industry has mainly led the field
in the use of polymer composites when compared to commercial airlines which
has used composites, because of safety concerns more restrictively and frequently
limited to secondary structural elements. Automotive applications, sporting goods,
medical devices and many other commercial applications are examples for the appli-
cation of polymer matrix composites. Also applications in civil engineering are now
on the way but it will take some time to achieve wide application of composites in
civil engineering as there are a lot of prescribed conditions to guarantee the reliabil-
ity of structures. But it is clear that over the last decades considerable advances have
been made in the use of composite materials in construction and building industries
and this trend will continue.
1.3 Modelling 11
1.3 Modelling
Composite materials consist of two or more constituents and the modelling, analy-
sis and design of structures composed of composites are different from conventional
materials such as steel. There are three levels of modelling. At the micro-mechanical
level the average properties of a single reinforced layer (a lamina or a ply) have to
be determined from the individual properties of the constituents, the fibres and ma-
trix. The average characteristics include the elastic moduli, the thermal and mois-
ture expansion coefficients, etc. The micro-mechanics of a lamina does not consider
the internal structure of the constituent elements, but recognizes the heterogene-
ity of the ply. The micro-mechanics is based on some simplifying approximations.
These concern the fibre geometry and packing arrangement, so that the constituent
characteristics together with the volume fractions of the constituents yield the av-
erage characteristics of the lamina. Note that the average properties are derived by
considering the lamina to be homogeneous. In the frame of this textbook only the
micro-mechanics of unidirectional reinforced laminates are considered (Sect. 3).
The calculated values of the average properties of a lamina provide the basis
to predict the macrostructural properties. At the macro-mechanical level, only the
averaged properties of a lamina are considered and the microstructure of the lamina
is ignored. The properties along and perpendicular to the fibre direction, these are
the principal directions of a lamina, are recognized and the so-called on-axis stress-
strain relations for a unidirectional lamina can be developed. Loads may be applied
not only on-axis but also off-axis and the relationships for stiffness and flexibility,
for thermal and moisture expansion coefficients and the strength of an angle ply can
be determined. Failure theories of a lamina are based on strength properties. This
topic is called the macro-mechanics of a single layer or a lamina (Sect. 4.1).
A laminate is a stack of laminae. Each layer of fibre reinforcement can have
various orientation and in principle each layer can be made of different materi-
als. Knowing the macro-mechanics of a lamina, one develops the macro-mechanics
of the laminate. Average stiffness, flexibility, strength, etc. can be determined for
the whole laminate (Sect. 4.2). The structure and orientation of the laminae in pre-
scribed sequences to a laminate lead to significant advantages of composite materi-
als when compared to a conventional monolithic material. In general, the mechan-
ical response of laminates is anisotropic. One very important group of laminated
composites are sandwich composites. They consist of two thin faces (the skins or
sheets) sandwiching a core (Fig. 1.3). The faces are made of high strength materials
having good properties under tension such as metals or fibre reinforced laminates
while the core is made of lightweight materials such as foam, resins with special
fillers, called syntactic foam, having good properties under compression. Sandwich
composites combine lightness and flexural stiffness. The macro-mechanics of sand-
wich composites is considered in Sect. 4.3.
When the micro- and macro-mechanical analysis for laminae and laminates are
carried out, the global behavior of laminated composite materials is known. The last
step is the modelling on the structure level and to analyze the global behavior of a
structure made of composite material. By adapting the classical tools of structural
12 1 Classification of Composite Materials
foam core balsa wood core
foam core with fillers balsa wood core with holes
folded plates core honeycomb core
Fig. 1.3 Sandwich materials with solid and hollow cores
analysis on anisotropic elastic structure elements the analysis of simple structures
as beams or plates may be achieved by analytical methods, but for more general
boundary conditions and/or loading and for complex structures, numerical methods
are used.
The composite structural elements in the restricted view of this textbook are lam-
inated or sandwich composites. The motivation for sandwich composites are two-
fold:
If a beam is bent, the maximum stresses occur at the top and the bottom surface.
So it makes sense using high strength materials only for the sheets and using low
and lightweight materials in the middle.
1.3 Modelling 13
The resistance to bending of a rectangular cross-sectional beam is proportional
to the cube of the thickness. Increasing the thickness by adding a core in the
middle increases the resistance. The shear stresses have a maximum in the mid-
dle of a sandwich beam requiring the core to support the shear. This advantage
of weight and bending stiffness makes sandwich composites more attractive for
some applications than other composite or conventional materials.
The most commonly used face materials are aluminium alloys or fibre reinforced
laminates and most commonly used core materials are balsa wood, foam and hon-
eycombs (Fig. 1.3). In order to guarantee the advantages of sandwich composites, it
is necessary to ensure that there is perfect bonding between the core and the sheets.
For laminated composites, assumptions are necessary to enable the mathematical
modelling. These are an elastic behavior of fibres and matrices, a perfect bonding
between fibres and matrices, a regular fibre arrangement in regular or repeating ar-
rays, etc.
Summarizing the different length scales of mechanical modelling structure ele-
ments composed of fibre reinforced composites it must be noted that, independent
of the different possibilities to formulate beam, plate or shell theories (Chaps. 7–9),
three modelling levels must be considered:
The microscopic level, where the average mechanical characteristics of a lamina
have to be estimated from the known characteristics of the fibres and the ma-
trix material taking into account the fibre volume fracture and the fibre packing
arrangement. The micro-mechanical modelling leads to a correlation between
constituent properties and average composite properties. In general, simple mix-
ture rules are used in engineering applications (Chap. 3). If possible, the aver-
age material characteristics of a lamina should be verified experimentally. On
the micro-mechanical level a lamina is considered as a quasi-homogeneous or-
thotropic material.
The macroscopic level, where the effective (average) material characteristics of a
laminate have to be estimated from the average characteristics of a set of laminae
taking into account their stacking sequence. The macro-mechanical modelling
leads to a correlation between the known average laminae properties and effec-
tive laminate properties. On the macro-mechanical level a laminate is consid-
ered generally as an equivalent single layer element with a quasi-homogeneous,
anisotropic material behavior (Chap. 4).
The structural level, where the mechanical response of structural members like
beams, plates, shells etc. have to be analyzed taking into account possibilities to
formulate structural theories of different order (Chap. 5).
In the recent years in the focus of the researchers is an additional level - the
nanoscale level. There are two reasons for this new direction:
composites reinforced by nanoparticles and
nanosize structures.
Both directions are beyond this elementary textbook. Partly new concepts should
be introduced since bulk effects are no more so important and the influence of sur-
14 1 Classification of Composite Materials
face effects is increasing. In this case the classical continuum mechanics approaches
should be extended. More details are presented in Altenbach and Eremeyev (2015);
Cleland (2003).
1.4 Material Characteristics of the Constituents
The optimal design and the analysis of structural elements requires a detailed knowl-
edge of the material properties. They depend on the nature of the constituent mate-
rials but also on manufacturing.
For conventional structure materials such as metals or concrete, is available much
research and construction experience over many decades, the codes for structures
composed of conventional materials have been revised continuously and so design
engineers pay less attention to material problems because there is complete docu-
mentation of the material characteristics.
It is quite an another situation for structures made of composites. The list of
composite materials is numerous but available standards and specifications are very
rare. The properties of each material used for both reinforcements and matrices of
composites are very much diversified. The experiences of nearly all design en gineers
in civil or mechanical engineering with composite materials, are insufficient. So it
should be borne in mind that structural design based on composite materials requires
detailed knowledge about the material properties of the singular constituents of the
composite for optimization of the material in the frame of structural applications
and also detailed codes for modelling and analysis are necessary.
The following statements are concentrated on fibre reinforced composites with
polymer resins. Material tests of the constituents of composites are in many cases a
complicated task and so the material data in the literature are limited. In engineering
applications the average data for a lamina are often tested to avoid this problem and
in order to use correct material characteristics in structural analysis. But in the area
of material design and selection, it is also important to know the properties of all
constituents.
The main properties for the estimation of the material behavior are
density
ρ
,
Young’s modulus1E,
ultimate strength
σ
uand
thermal expansion coefficient
α
.
The material can be made in bulk form or in the form of fibres. To estimate proper-
ties of a material in the form of fibres, the fibre diameter dcan be important.
Table F.1 gives the specific performances of selected material made in bulk form.
Traditional materials, such as steel, aluminium alloys, or glass have comparable
1Thomas Young (13 June 1773 Milverton, Somersetshire - †10 May 1829 London) - polymath
and physician, notable scientific contributions to the fields of light, mechanics (elastic material
parameter, surface tension), energy among others
1.5 Advantages and Limitations 15
specific moduli E/
ρ
but in contrast the specific ultimate stress
σ
u/
ρ
of glass is
significantly higher than that of steel and of aluminium alloys. Table F.2 presents the
mechanical characteristics of selected materials made in the form of fibres. It should
be borne in mind that the ultimate strength measured for materials made in bulk form
is remarkably smaller than the theoretical strengths. This is attributed to defects or
micro-cracks in the material. Making materials in the form of fibres with a very
small diameter of several microns decreases the number of defects and the values
of ultimate strength increases. Table F.3 gives material properties for some selected
matrix materials and core materials of sandwich composites. Tables F.4 and F.5
demonstrate some properties of unidirectional fibre reinforced composite materials:
ELis the longitudinal modulus in fibre direction, ETthe transverse modulus, GLT the
in-plane shear modulus,
ν
LT and
ν
TL are the major and the minor Poisson’s ratio2,
σ
Lu,
σ
Tu,
σ
LTu the ultimate stresses or strengths,
α
Land
α
Tthe longitudinal and the
transverse thermal expansion coefficients.
Summarizing the reported mechanical properties, which are only a small selec-
tion, a large variety of fibres and matrices are available to design a composite ma-
terial with high modulus and low density or other desired qualities. The impact of
the costs of the composite material can be low for applications in the aerospace
industry or high for applications such as in automotive industry. The intended per-
formance of a composite material and the cost factors play an important role and
structural design with composite materials has to be compared with the possibilities
of conventional materials.
1.5 Advantages and Limitations
The main advantage of polymer matrix composites in comparison with conventional
materials, such as metals, is their low density. Therefore two parameters are com-
monly used to demonstrate the mechanical advantages of composites:
1. The specific modulus E/
ρ
is the Young’s modulus per unit mass or the ratio
between Young’s modulus and density.
2. The specific strength
σ
u/
ρ
is the tensile strength per unit mass or the ratio be-
tween strength and density
The benefit of the low density becomes apparent when the specific modulus and the
specific strength are considered. The two ratios are high and the higher the specific
parameters the more weight reduction of structural elements is possible in relation
to special loading conditions. Therefore, even if the stiffness and/or the strength
performance of a composite material is comparable to that of a conventional alloy,
the advantages of high specific stiffness and/or specific strength make composites
2Sim´eon Denis Poisson (21 June 1781 Pithiviers - †25 April 1840 Paris) - mathematician, ge-
ometer, and physicist, with contributions to mechanics, after Poisson an elastic material parameter
is named
16 1 Classification of Composite Materials
more attractive. Composite materials are also known to perform better under cyclic
loads than metallic materials because of their fatigue resistance.
The reduction of mass yields reduced space requirements and lower material and
energy costs. The mass reduction is especially important in moving structures. Be-
ware that in some textbooks the specific values are defined as E/
ρ
gand
σ
u/
ρ
g,
where gis the acceleration due to the gravity. Furthermore it should be noted that
a single performance indicator is insufficient for the material estimation and that
comparison of the specific modulus and the specific strength of unidirectional com-
posites to metals gives a false impression. Though the use of fibres leads to large
gains in the properties in fibre direction, the properties in the two perpendicular di-
rections are greatly reduced. Additionally, the strength and stiffness properties of
fibre-reinforced materials are poor in another important aspect. Their strength de-
pends critically upon the strength of the fibre, matrix interface and the strength of the
matrix material, if shear stresses are being applied. This leads to poor shear proper-
ties and this lack of good shear properties is as serious as the lack of good transverse
properties. For complex structure loadings, unidirectional composite structural ele-
ments are not acceptable and so-called angle-ply composite elements are necessary,
i.e. the structural components made of fibre-reinforced composites are usually lam-
inated by using a number of layers. This number of fibre-reinforced layers can vary
from just a few to several hundred. While generally the majority of the layers in the
laminate have their fibres in direction of the main loadings, the other layers have
their fibres oriented specifically to counter the poor transverse and shear properties.
Additional advantages in the material performances of composites are low ther-
mal expansion, high material damping, generally high corrosion resistance and elec-
trical insulation. Composite materials can be reinforced in any direction and the
structural elements can be optimized by material design or material tailoring.
There are also limitations and drawbacks in the use of composite materials:
1. The mechanical characterization of composite materials is much more complex
than that of monolithic conventional material such as metal. Usually composite
material response is anisotropic. Therefore, the material testing is more compli-
cated, cost and time consuming. The evaluation and testing of some composite
material properties, such as compression or shearing strengths, are still in discus-
sion.
2. The complexity of material and structural response makes structural modelling
and analysis experimentally and computationally more expensive and compli-
cated in comparison to metals or other conventional structural materials. There
is also limited experience in the design, calculating and joining composite struc-
tural elements.
Additional disadvantages are the high cost of fabrication, but improvements in pro-
duction technology will lower the cost more and more, further the complicated re-
pair technology of composite structures, a lot of recycling problems, etc.
Summarizing, it can be said that the application of composite materials in struc-
ture design beyond the military and commercial aircraft and aerospace industry and
some special fields of automotive, sporting goods and medical devices is still in the
1.6 Problems 17
early stages. But the advancing of technology and experience yields an increasing
use of composite structure elements in civil and mechanical engineering and pro-
vides the stimulus to include composite processing, modelling, design and analysis
in engineering education.
1.6 Problems
1. What is a composite and how are composites classified?
2. What are the constituents of composites?
3. What are the fibre and the matrix factors which contribute to the mechanical
performance of composites?
4. What are polymer matrix, metal matrix and ceramic matrix composites, what are
their main applications?
5. Define isotropic, anisotropic, homogeneous, nonhomogeneous.
6. Define lamina, laminate, sandwich. What is micro-mechanical and macro-
mechanical modelling and analysis?
7. Compare the specific modulus, specific strength and coefficient of thermal ex-
pansion of glas fibre, epoxy resin and steel.
Exercise 1.1. A typical CFK plate (uni-directional reinforced laminae composed of
carbon fibres and epoxy matrix) has the size 300 mm ×200 mm ×0.5 mm (length
l×bright b×thickness d). Please show that the fibre volume fraction and the fibre
mass fraction are not the same. The estimate should be based on the following data:
density of the carbon fibres
ρ
f=1,8 g/cm3,
diameter of the fibres is df=6
µ
m
fibre volume fraction vf=0.8 and
density of the epoxy
ρ
m=1,1 g/cm3.
Solution 1.1. Let us assume that the fibres are parallel to the longer plate side. In
this case the volume of one fibre V1fis
V1f=
π
d2
f
4l=
π
(6
µ
m)2
4·300 mm =8,48 ·103mm3
The volume of the plate is
Vplate =300 ·200 ·0,5 mm3=3·104mm3
With the fibre volume fraction vf=0.8 we can calculate the matrix volume fraction
vm
vm=1vf=10.8=0.2
The volume of all fibres is
Vf=vfVplate =0,8·3·104mm3=2,4·104mm3
18 1 Classification of Composite Materials
which is equivalent to the following number of fibres nf
nf=Vf
V1f
=2,4·104mm3
8,48 ·103mm3=0,28 ·107
The fibre mass fraction can be computed as it follows
mf=Mf
M=Mf
Mf+Mm
=
ρ
fVf
ρ
fVf+
ρ
mVm
With
Vm=Vplate Vf=3·104mm32,4·104mm3=0,6·104mm3
fibre mass fraction is
mf=1,8g/cm3·2,4·104mm3
1,8g/cm3·2,4·104mm3+1,1g/cm3·0,6·104mm3=0,867
This value is slightly grater than the assumed fibre volume fraction.
References
Altenbach H, Eremeyev VA (2015) On the theories of plates and shells at the
nanoscale. In: Altenbach H, Mikhasev GI (eds) Shell and Membrane Theories
in Mechanics and Biology: From Macro- to Nanoscale Structures, Springer Inter-
national Publishing, Cham, pp 25–57
Christensen RM (2005) Mechanics of Composite Materials. Dover, Mineola (NY)
Cleland AN (2003) Foundations of Nanomechanics. Advanced Texts in Physics,
Springer, Berlin, Heidelberg
Torquato S (2002) Random Heterogeneous Materials - Microstructure and Macro-
scopic Properties, Interdisciplinary Applied Mathematics, vol 16. Springer, New
York
Chapter 2
Linear Anisotropic Materials
The classical theory of linear elastic deformable solids is based on the following
restrictions to simplify the modelling and analysis:
The body is an ideal linear elastic body.
All strains are small.
The material of the constituent phases is homogeneous and isotropic.
These assumptions of classical theory of elasticity guarantee a satisfying quality
of modelling and analysis of structure elements made of conventional monolithic
materials. Structural analysis of elements composed of composite materials is based
on the theory of anisotropic elasticity, the elastic properties of composites depend
usually on the direction and the deformable solid is anisotropic. In addition, now the
composite material is not homogeneous at all. It must be assumed that the material
is piecewise homogeneous or quasi-homogeneous.
The governing equations of elastic bodies are nearly the same for isotropic and
anisotropic material response. There are equilibrium equations, which describe the
static or dynamic equilibrium of forces acting on an elastic body. The kinematic
equations describe the strain-displacement relations and the compatibility equations
guarantee a unique solution to the equations relating strains and displacements. All
these equations are independent of the elastic properties of the material. Only the
material relations (so-called constitutive equations), which describe the relations
between stresses and strains are very different for an isotropic and an anisotropic
body. This difference in formulating constitutive equations has a great influence
on the model equations in the frame of the isotropic and the anisotropic theory of
elasticity. Note that in many cases the material behavior of the constituents can
assumed to be homogeneous and isotropic.
The governing equations, as defined above, including so-called initial-boundary
conditions for forces/stresses and/or displacements, yield the basic model equations
for linear elastic solids such as differential equations or variational and energy for-
mulations, respectively. All equations for structural elements which are given in this
textbook, are founded on these general equations for the theory of elasticity of linear
elastic anisotropic solids.
19
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_2
20 2 Linear Anisotropic Materials
The objective of this chapter is to review the generalized Hooke’s law1, the con-
stitutive equations for anisotropic elastic bodies, and to introduce general relations
for stiffness and strains including transformation rules and symmetry relations. The
constitution of a unidirectional composite material and simplified approaches for
so-called effective moduli result in an engineering formulation of constitutive equa-
tions for fibre reinforced composites and will be considered in Chap. 3.
The theory of anisotropic elasticity presented in Sect. 2.1 begins with the most
general form of the linear constitutive equations, and passes from all specific cases
of elastic symmetries to the classical Hooke’s law for an isotropic body. The only
assumptions are
all elastic properties are the same in tension and compression,
the stress tensor is symmetric,
an elastic potential exists and is an invariant with respect to linear orthogonal
coordinate transformation.
In addition to the general three-dimensional stress-strain relationships, the plane
stress and plane strain cases are derived and considered for an anisotropic body and
for all the derived specific cases of elastic symmetries. The type of anisotropy con-
sidered in Sects. 2.1.1–2.1.5 can be called as rectilinear anisotropy, i.e. the homo-
geneous anisotropic body is characterized by the equivalence of parallel directions
passing through different points of the body. Another kind of anisotropy, which can
be interesting to some applications, e.g. to modelling circular plates or cylindrical
tubes, is considered in a comprehensive formulation in Sect. 2.1.6. If one chooses
a system of curvilinear coordinates in such a manner that the coordinate directions
coincide with equivalent directions of the elastic properties at different points of the
body, the elastic behavior is called curvilinear.
The chapter ends with the derivation of the fundamental equations of anisotropic
elasticity and the formulation of variational solution methods. In Sect. 2.2 the dif-
ferential equations for boundary and initial boundary problems are considered. The
classical and generalized variational principles are formulated and approximate an-
alytical solution methods based on variational principles are discussed.
2.1 Generalized Hooke’s Law
The phenomenological modelling neglects the real on the microscopic scale discon-
tinuous structure of the material. On the macroscopic or phenomenological scale the
material models are assumed to be continuous and in general homogeneous. In the
case of fibre reinforced composites, the heterogeneity of the bulk material is a con-
sequence of the two constituents, the fibres and the matrix, but generally there exists
a representative volume element of the material on a characteristic scale at which the
1Robert Hooke (28greg./18jul.July 1635 Freshwater, Isle of Wight - †3 March 1703 London) -
natural philosopher, architect and polymath, first constitutive law for elasticity published as an
anagram ceiiinosssttuu which is in Latain ut tensio sic vis and means as the extension, so the force
2.1 Generalized Hooke’s Law 21
properties of the material can be averaged to a good approximation. The composite
material can be considered as macroscopically homogeneous and the problem of
designing structural elements made from of composite materials can be solved in an
analogous manner as for conventional materials with the help of the average mate-
rial properties or the so-called effective moduli. Chapter 3 explains the calculation
of effective moduli in detail.
Unlike metals or polymeric materials without reinforcements or reinforced by
stochastically distributed and orientated particles or short fibres, the material behav-
ior of an off-axis forced unidirectional lamina is anisotropic. In comparison to con-
ventional isotropic materials, the experimental identification of the material param-
eters is much more complicated in the case of anisotropic materials. But anisotropic
material behavior also has the advantage of material tailoring to suit the main load-
ing cases.
2.1.1 Stresses, Strains, Stiffness, and Compliances
In preparation for the formulation of the generalized Hooke’s law, a one-dimen-
sional problem will be considered. The deformations of an elastic body can be char-
acterized by displacements or by strains:
Dilatational or extensional strains
ε
: The body changes only its volume but not
its shape.
Shear strains
γ
: The body changes only its shape but not its volume.
Figure 2.1 demonstrates extensional and shear strains for a simple prismatic body
loaded by forces Fand T, normal and tangential to the cross-section, respectively.
Assuming a uniform distribution of the forces Fand Ton the cross-section, the
elementary one-dimensional definitions for stresses and strains are given by (2.1.1)
σ
=F
A0
normal stress
σ
,
ε
=ll0
l0
=l
l0
extensional strain
ε
,
τ
=T
A0
shear stress
τ
,
γ
tan
γ
=u
l0
shear strain
γ
(2.1.1)
The last one definition is restricted by small strain assumption. This assumption
can be accepted for many composite material applications and will be used for both
types of strains.
The material or constitutive equations couple stresses and strains. In linear elas-
ticity the one-to-one transformation of stresses and strains yield Hooke’s law (2.1.2)
22 2 Linear Anisotropic Materials
F
A
A0
l0l=l0+l
l
F
l0
u
TA0
T
γ
Fig. 2.1 Extensional strain
ε
and shear strain
γ
of a body with the length l0and the cross-section
area A0
σ
=E
ε
,E=
σ
ε
,Eis the elasticity or Young’s modulus ,
τ
=G
γ
,G=
τ
γ
,Gis the shear modulus
(2.1.2)
For a homog eneous material Eand Gare parameters with respect to the coordinates.
For the extensionally strained prismatic body (Fig. 2.1) the phenomenon of con-
traction in a direction normal to the direction of the tensile loading has to be con-
sidered. The ratio of the contraction, expressed by the transverse strain
ε
t, to the
elongation in the loaded direction, expressed by
ε
, is called Poisson’s ratio
ν
ε
t=
νε
,
ν
=
ε
t
ε
(2.1.3)
For an isotropic bar with an extensional strain
ε
>0 it follows that
ε
t<02.
Hooke’s law can be written in the inverse form
ε
=E1
σ
=S
σ
(2.1.4)
S=E1is the inverse modulus of elasticity or the flexibility/compliance modulus.
For homogeneous material, Sis an elastic parameter.
Consider a tensile loaded prismatic bar composed of different materials
(Fig. 2.2). Since
σ
=F/Aand
σ
=E
ε
then
σ
A=F=EA
ε
and
ε
= (EA)1F.
EA is the tensile stiffness and (EA)1the tensile flexibility or compliance. The dif-
2The class of auxetic materials that have a negative Poisson’s ratio that means when stretched,
they become thicker perpendicular to the applied force, is not in the focus of the present book.
2.1 Generalized Hooke’s Law 23
l
l
F
F
C1
C1
Cn
Cn
Ai
li
Fig. 2.2 Tensile bar with stiffness Ci=EiAiarranged in parallel and in series
ferent materials of the prismatic bar in Fig. 2.2 can be arranged in parallel or in
series. In the first case we have
F=
n
i=1
Fi,A=
n
i=1
Ai,
ε
=
ε
i(2.1.5)
Fiare the loading forces on Aiand the strains
ε
iare equal for the total cross-section.
With
F=EA
ε
,Fi=EiAi
ε
,
n
i=1
Fi=F=
n
i=1
EiAi
ε
(2.1.6)
follow the coupling equations for the stiffness EiAifor a parallel arrangement
EA =
n
i=1
EiAi,(EA)1=1
n
i=1
EiAi
(2.1.7)
This equal strain treatment is often described as a Voigt3model which is the upper-
bound stiffness.
In the other case, we have
l=
n
i=1li
and F=Fi, the elongation lof the bar is obtained by addition of the liof the
different parts of the bar with the lengths liand the tensile force is equal for all
cross-sectional areas. With
l=l
ε
=l(EA)1F,li=li
ε
i=li(EiAi)1F
and
n
i=1li="n
i=1
li(EiAi)1#F(2.1.8)
3Woldemar Voigt (2 September 1850 Leipzig - †13 December 1919 G ¨ottingen) - physicist,
worked on crystal physics, thermodynamics and electro-optics, the word tensor in its current mean-
ing was introduced by him in 1898, Voigt notation
24 2 Linear Anisotropic Materials
follow the coupling equations for the stiffness EiAiarranged in series
EA =l
n
i=1
li(EiAi)1
,(EA)1=
n
i=1
li(EiAi)1
l(2.1.9)
This equal stress treatment is described generally as a Reuss4model which is the
lower bound stiffness. The coupling equations illustrate a first clear insight into a
simple calculation of effective stiffness and compliance parameters for two com-
posite structures.
The three-dimensional state of stress or strain in a continuous solid is completely
determined by knowing the stress or strain tensor. It is common practice to repre-
sent the tensor components acting on the faces of an infinitesimal cube with sides
parallel to the reference axes (Fig. 2.3). The sign convention is defined in Fig. 2.3.
Positive stresses or strains act on the faces of the cube with an outward vector in
the positive direction of the axis of the reference system and vice versa. Using the
tensorial notation for the stress tensor
σ
i j and the strain tensor
ε
i j for the stresses
and the strains we have normal stresses or extensional strains respectively for i=j
and shear stresses or shear strains for i6=j.
ε
i j with i6=jare the tensor shear coor-
dinates and 2
ε
i j =
γ
i j,i6=jthe engineering shear strains. The first subscript of
σ
i j
and
ε
i j indicates the plane xi=const on which the load is acting and the second
subscript denotes the direction of the loading. Care must be taken in distinguish-
ing in literature the strain tensor
ε
i j from the tensor ei j which is the tensor of the
relative displacements, ei j =
ui/
xj. An application of shear stresses
σ
i j and
σ
ji
produces in the i j-plane of the infinitesimal cube (Fig. 2.3) angular rotations of the
i- and j-directions by ei j and eji. These relative displacements represent a combina-
tion of strain (distorsion) and rigid body rotation with the limiting cases ei j =eji,
e
e
e1
e
e
e2
e
e
e3
σ
11
σ
12
σ
13
σ
22
σ
21
σ
23
σ
33
σ
32
σ
31
ε
11
ε
13
ε
12
ε
22
ε
21
ε
23
ε
33
ε
32
ε
31
Fig. 2.3 Stress and strain components on the positive faces of an infinitesimal cube in a set of axis
e
e
e1,e
e
e2,e
e
e3
4Andr´as (Endre) Reuss (1 July 1900 Budapest - †10 May 1968 Budapest) - mechanical engineer,
contributions to the theory of plasticity
2.1 Generalized Hooke’s Law 25
j
jj
i
ii
ei j
ei j
eji
eji
γ
i j
γ
i j
ei j =eji =
ε
i j =1
2
γ
i j
ω
i j =0
ei j =eji
ε
i j =
γ
i j =0
ei j =2
ε
ji =
γ
i j
eji =0
abc
Fig. 2.4 Examples of distorsions and rigid body rotation. aPure shear, bpure rotation, csimple
shear
.
i.e. no rotation, and ei j =ej i, i.e. no distorsion (Fig. 2.4). From the figure follows
that simple shear is the sum of pure shear and rigid rotation. ei j is positive when
it involves rotating the positive j-direction towards the positive i-direction and vice
versa. Writing the tensor ei j as the sum of symmetric and antisymmetric tensors
ei j =1
2(ei j +eji) + 1
2(ei j eji) =
ε
i j +
ω
i j (2.1.10)
where
ε
i j is the symmetric strain tensor and
ω
i j is the antisymmetric rotation ten-
sor. For normal strains, i.e. i=j, there is ei j =
ε
i j, however for i6=jwe have
γ
i j =2
ε
i j =ei j +eji with the engineering shear strains
γ
i j and the tensorial shear
strains
ε
i j. Careful note should be taken of the factor of two related engineering and
tensorial shear strains,
γ
i j is often more convenient for practical use but tensor oper-
ations such as rotations of the axis, Sect. 2.1.2, must be carried out using the tensor
notation
ε
i j.
The stress and the strain tensors are symmetric tensors of rank two. They can be
represented by the matrices
σ
σ
σ
=
σ
11
σ
12
σ
13
σ
12
σ
22
σ
23
σ
13
σ
23
σ
33
,
ε
ε
ε
=
ε
11
ε
12
ε
13
ε
12
ε
22
ε
23
ε
13
ε
23
ε
33
(2.1.11)
The symmetry of the tensors (2.1.11) reduces the number of unknown components
for defining these tensors to six components. For this reason, an engineering matrix
notation can be used by replacing the matrix table with nine values by a column
matrix or a vector with six components. The column matrices (stress and strain
vector) are written in Eqs. (2.1.12) in a transposed form
26 2 Linear Anisotropic Materials
[
σ
11
σ
22
σ
33
σ
23
τ
23
σ
13
τ
13
σ
12
τ
12]T,
[
ε
11
ε
22
ε
33 2
ε
23
γ
23 2
ε
13
γ
13 2
ε
12
γ
12]T(2.1.12)
The stress and strain states are related by a material law which is deduced from
experimental observations. For a linear elastic anisotropic material, the generalized
Hooke’s law relates the stress and the strain tensor
σ
i j =Cij kl
ε
kl (2.1.13)
Ci jkl are the material coefficients and define the fourth rank elasticity tensor which
in general case contains 81 coordinates. Due to the assumed symmetry of
σ
i j =
σ
ji
and
ε
i j =
ε
ji the symmetry relations follow for the material tensor
Ci jkl =Cjikl ,Ci jkl =Cij lk (2.1.14)
and reduce the number of coordinates to 36. Introducing a contracted single-
subscript notation for the stress and strain components and a double-subscript nota-
tion for the elastic parameters, the generalized relation for stresses and strains can
be written in vector-matrix form
[
σ
i] = [Cij ][
ε
j];i,j=1,2,...,6; Ci j 6=Cj i;i6=j(2.1.15)
At this stage we have 36 independent material coefficients, but a further reduction
of the number of independent values is possible because we have assumed the exis-
tence of an elastic potential function.
The elastic strain energy is defined as the energy expended by the action of exter-
nal forces in deforming an elastic body: essentially all the work done during elastic
deformations is stored as elastic energy. The strain energy per unit volume, i.e. the
strain energy density function, is defined as follows
W=1
2
σ
i j
ε
i j (2.1.16)
or in a contracted notation
W(
ε
i) = 1
2
σ
i
ε
i=1
2Ci j
ε
j
ε
i(2.1.17)
With
W
ε
i
=
σ
i,
2W
ε
i
ε
j
=Ci j ,
2W
ε
j
ε
i
=Cji
and
2W
ε
i
ε
j
=
2W
ε
j
ε
i
follow the symmetry relations
Ci j =Cj i;i,j=1,2,...,6 (2.1.18)
2.1 Generalized Hooke’s Law 27
and the number of the independent material coefficients is reduced to 21. The gen-
eralized relations for stresses and strains of an anisotropic elastic body written again
in a contracted vector-matrix form have a symmetric matrix for Ci j
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
=
C11 C12 C13 C14 C15 C16
C22 C23 C24 C25 C26
C33 C34 C35 C36
C44 C45 C46
S Y M C55 C56
C66
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
(2.1.19)
The transformation rules for the contraction of the subscripts of
σ
i j,
ε
i j and Ci jkl of
(2.1.13) are given in Tables 2.1 and 2.2.
The elasticity equation (2.1.19) can be written in the inverse form as follows
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
S11 S12 S13 S14 S15 S16
S22 S23 S24 S25 S26
S33 S34 S35 S36
S44 S45 S46
S Y M S55 S56
S66
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.20)
with
[Ci j ][Sjk ] = [
δ
ik] = 1i=k,
0i6=k,i,j,k=1,...,6
In a condensed symbolic or subscript form, Eqs. (2.1.19) and (2.1.20) are (summa-
tion on double subscripts)
σ
i=Ci j
ε
j,
ε
i=Si j
σ
j;i,j=1,...,6
σ
σ
σ
=C
C
C
ε
ε
ε
,
ε
ε
ε
=S
S
S
σ
σ
σ
(2.1.21)
C
C
C[Ci j ]is the stiffness matrix and S
S
S[Si j]the compliance or flexibility matrix. Ci j
and Si j are only for homogeneou s anisotropic materials constant material parameters
Table 2.1 Transformation of
the tensor coordinates
σ
i j and
ε
i j to the vector coordinates
σ
pand
ε
p
σ
i j
σ
p
ε
i j
ε
p
σ
11
σ
1
ε
11
ε
1
σ
22
σ
2
ε
22
ε
2
σ
33
σ
3
ε
33
ε
3
σ
23 =
τ
23
σ
42
ε
23 =
γ
23
ε
4
σ
31 =
τ
31
σ
52
ε
31 =
γ
31
ε
5
σ
12 =
τ
12
σ
62
ε
12 =
γ
12
ε
6
Table 2.2 Transformation
of the tensor coordinates
Ci jkl to the matrix coordi-
nates Cpq
Ci jkl Cpq
i j : 11, 22, 33 p: 1, 2, 3
23, 31, 12 4, 5, 6
kl : 11, 22, 33 q: 1, 2, 3
23, 31, 12 4, 5, 6
28 2 Linear Anisotropic Materials
with respect to the coordinates. Their values depend on the reference coordinate
system. A change of the reference system yields a change of the constant values.
Summarizing the stiffness and the compliance relations, it can be seen that for
a linear elastic anisotropic material 21 material parameters have to be measured
experimentally in the general case. But in nearly all engineering applications there
are material symmetries and the number of material parameters can be reduced.
Section 2.1.2 describes some transformation rules for C
C
Cand S
S
Sfollowing from the
change of reference system and the symmetric properties of anisotropic materials
discussed in Sect. 2.1.3. Furthermore the way that the material parametersCij and Si j
are related to the known engineering elastic parameters Ei,Gij and
ν
i j is considered.
2.1.2 Transformation Rules
If we have a reference system which is characterized by the orthonormal basic unit
vectors e
e
e1,e
e
e2,e
e
e3and another reference system with the vector basis e
e
e
1,e
e
e
2,e
e
e
3. Both
systems are related by a rotation of the coordinate axis (Fig. 2.5), the transformation
rules are
e
e
e
i=Ri je
e
ej,e
e
ei=Rjie
e
ej,
Ri j cos(e
e
e
i,e
e
ej),Rji cos(e
e
ei,e
e
ej),i,j=1,2,3 (2.1.22)
These relationships describe a general linear orthogonal coordinate transformation
and can be expressed in vector-matrix notation
e
e
e=R
R
Re
e
e,e
e
e=R
R
R1e
e
e=R
R
RTe
e
e(2.1.23)
R
R
Ris the transformation or rotation matrix. In the case of an orthogonal set of axes
such as given in Fig. 2.5 the matrix R
R
Ris symmetric and unitary. This means the
determinant of this matrix is unity (Det R
R
R=|Ri j|=1 and the inverse matrix R
R
R1
is identical to the transposed matrix (R
R
R1=R
R
RT). In the special case of a rotation
φ
Fig. 2.5 Rotation of a refer-
ence system with the basic
vectors e
e
eiinto a system with
the basic vectors e
e
e
i
x1
x2
x3
x1
x2
x3
e
e
e1e
e
e
1
e
e
e2
e
e
e
2
e
e
e3
e
e
e
3
2.1 Generalized Hooke’s Law 29
about the direction e
e
e3, the rotation matrix R
R
Rand the inverse matrix R
R
R1are
3
Ri j=
c s 0
s c 0
0 0 1
,3
Ri j1
=3
Ri jT
=
cs0
s c 0
001
,(2.1.24)
and the transformation rules are
e
e
e
1
e
e
e
2
e
e
e
3
=
c s 0
s c 0
0 0 1
e
e
e1
e
e
e2
e
e
e3
,
e
e
e1
e
e
e2
e
e
e3
=
cs0
s c 0
001
e
e
e
1
e
e
e
2
e
e
e
3
with c=cos
φ
,s=sin
φ
. For rotations
ψ
or
θ
about the directions e
e
e2or e
e
e1the
rotations matrices [2
Ri j]and [1
Ri j]are
2
Ri j=
c0s
0 1 0
s0c
,2
Ri j1
=2
Ri jT
=
c0s
0 1 0
s0c
,
1
Ri j=
100
0c s
0s c
,1
Ri j1
=1
Ri jT
=
1 0 0
0cs
0s c
with c=cos
ψ
or cos
θ
and s=sin
ψ
or sin
θ
for rotations about e
e
e2or e
e
e1, respec-
tively.
The transformation rule (2.1.22) can be interpreted as a rule for vectors or first-
rank tensors. The generalization to second-rank tensors yields e.g. for the stress
tensor
σ
i j =RikRjl
σ
kl ,
σ
i j =RkiRl j
σ
kl (2.1.25)
For the following reflections the transformation rules for the contracted notation are
necessary. The nine tensor coordinates
σ
i j are shifted to six vector coordinates
σ
p.
The transformations
σ
p=T
σ
pq
σ
q,
σ
p=T
σ
pq1
σ
q,p,q=1,...,6 (2.1.26)
are not tensor transformation rules. The transformation matrices T
σ
pq and (T
σ
pq)1
follow by comparison of Eqs. (2.1.25) and (2.1.26). In the same manner we can find
the transformation rules for the strains
ε
p=T
ε
pq
ε
q,
ε
p=T
ε
pq1
ε
q,p,q=1,...,6 (2.1.27)
The elements of the transformation matrices [T
σ
pq]and [T
ε
pq]are defined in App. B.
Summarizing, the transformation rules for stresses and strains in a condensed
vector-matrix notation as follows
σ
σ
σ
=T
T
T
σσ
σ
σ
,
ε
ε
ε
=T
T
T
εε
ε
ε
,
σ
σ
σ
= (T
T
T
σ
)1
σ
σ
σ
,
ε
ε
ε
= (T
T
T
ε
)1
ε
ε
ε
(2.1.28)
30 2 Linear Anisotropic Materials
The comparison of
σ
i j =RkiRl j
σ
kl with
σ
p= (T
σ
pq)1
σ
q
and
ε
i j =RkiRl j
ε
kl with
ε
p= (T
ε
pq)1
ε
q
yields an important result on the linkage of inverse and transposed stress and strain
transformation matrices
(T
T
T
σ
)1= (T
T
T
ε
)T,(T
T
T
ε
)1= (T
T
T
σ
)T(2.1.29)
The transformation relations for the stiffness and the compliance matrices C
C
Cand
S
S
Scan be obtained from the known rules for stresses and strains. With
σ
σ
σ
=C
C
C
ε
ε
ε
and
σ
σ
σ
=C
C
C
ε
ε
ε
, it follows that
(T
T
T
σ
)1
σ
σ
σ
=
σ
σ
σ
=C
C
C
ε
ε
ε
=C
C
C(T
T
T
ε
)1
ε
ε
ε
,
σ
σ
σ
=T
T
T
σ
C
C
C(T
T
T
ε
)1
ε
ε
ε
=C
C
C
ε
ε
ε
,
T
T
T
σσ
σ
σ
=
σ
σ
σ
=C
C
C
ε
ε
ε
=C
C
CT
T
T
εε
ε
ε
,
σ
σ
σ
= (T
T
T
σ
)1C
C
CT
T
T
εε
ε
ε
=C
C
C
ε
ε
ε
,
(2.1.30)
respectively.
Considering (2.1.29) the transformation relations for the stiffness matrix are
C
C
C=T
T
T
σ
C
C
C(T
T
T
σ
)T,C
C
C= (T
T
T
ε
)TC
C
CT
T
T
ε
(2.1.31)
or in index notation
C
i j =T
σ
ik T
σ
jl Ckl ,Ci j =T
ε
ik T
ε
jlC
kl (2.1.32)
The same procedure yields the relations for the compliance matrix. With
ε
ε
ε
=S
S
S
σ
σ
σ
,
ε
ε
ε
=S
S
S
σ
σ
σ
(2.1.33)
it follows
(T
T
T
ε
)1
ε
ε
ε
=
ε
ε
ε
=S
S
S
σ
σ
σ
=S
S
S(T
T
T
σ
)1
σ
σ
σ
,
ε
ε
ε
=T
T
T
ε
S
S
S(T
T
T
σ
)1
σ
σ
σ
=S
S
S
σ
σ
σ
,
T
T
T
εε
ε
ε
=
ε
ε
ε
=S
S
S
σ
σ
σ
=S
S
ST
T
T
σσ
σ
σ
,
ε
ε
ε
= (T
T
T
ε
)1S
S
ST
T
T
σσ
σ
σ
=S
S
S
σ
σ
σ
,
(2.1.34)
i.e.
ε
ε
ε
=T
T
T
ε
S
S
S(T
T
T
σ
)1
σ
σ
σ
,
ε
ε
ε
= (T
T
T
ε
)1S
S
ST
T
T
σσ
σ
σ
(2.1.35)
The comparison leads to the transformation equations for S
S
Sand S
S
S
S
S
S=T
T
T
ε
S
S
S(T
T
T
σ
)1,S
S
S= (T
T
T
ε
)1S
S
ST
T
T
σ
(2.1.36)
or taking into account (2.1.29)
S
S
S=T
T
T
ε
S
S
S(T
T
T
ε
)T,S
S
S= (T
T
T
σ
)TS
S
ST
T
T
σ
,(2.1.37)
2.1 Generalized Hooke’s Law 31
respectively, in subscript notation
S
i j =T
ε
ik T
ε
jl Skl ,Si j =T
σ
ik T
σ
jl S
kl (2.1.38)
In the special case of a rotation
φ
about the e
e
e3-direction (Fig. 2.6) the coordinates
of the transformation matrices T
T
T
σ
and T
T
T
ε
are given by the (2.1.39) and (2.1.40)
3
T
σ
pq=
c2s20 0 0 2cs
s2c20 0 0 2cs
0 0 1 0 0 0
0 0 0 cs0
0 0 0 s c 0
cs cs 0 0 0 c2s2
,3
T
σ
pq1
="3
T
ε
pq#T
(2.1.39)
"3
T
ε
pq#=
c2s20 0 0 cs
s2c20 0 0 cs
0 0 1 0 0 0
0 0 0 cs0
0 0 0 s c 0
2cs 2cs 0 0 0 c2s2
,"3
T
ε
pq#1
=3
T
σ
pqT
(2.1.40)
By all rules following from a rotation of the reference system the stresses, strains,
stiffness and compliance parameters in the rotated system are known. They are sum-
marized in symbolic notation (2.1.41)
σ
σ
σ
=T
T
T
σσ
σ
σ
,
ε
ε
ε
=T
T
T
εε
ε
ε
,
σ
σ
σ
= (T
T
T
ε
)T
σ
σ
σ
,
ε
ε
ε
= (T
T
T
σ
)T
ε
ε
ε
,
C
C
C=T
T
T
σ
C
C
C(T
T
T
σ
)T,S
S
S=T
T
T
ε
S
S
S(T
T
T
ε
)T,
C
C
C= (T
T
T
ε
)TC
C
CT
T
T
ε
,S
S
S= (T
T
T
σ
)TS
S
ST
T
T
σ
(2.1.41)
Fig. 2.6 Rotation about the
e
e
e3-direction x1
x2
x3,x
3
x
1
x
2
e
e
e3,e
e
e
3
e
e
e
1
e
e
e2
e
e
e
2
e
e
e1
φ
φ
32 2 Linear Anisotropic Materials
For special cases of a rotation about a direction e
e
eithe general transformation ma-
trices T
T
T
σ
and T
T
T
ε
are substituted by
i
T
T
T
σ
or
i
T
T
T
ε
. The case of a rotation about the
e
e
e1-direction yields the coordinates of the transformation matrices T
T
T
σ
and T
T
T
ε
which
are given in App. B.
2.1.3 Symmetry Relations of Stiffness and Compliance Matrices
In the most general case of the three-dimensional generalized Hooke’s law the stiff-
ness and the compliance matrices have 36 non-zero material parameters Ci j or Si j
but they are each determined by 21 independent parameters. Such an anisotropic
material is called a triclinic material, it has no geometric symmetry properties. The
experimental tests to determine 21 independent material parameters would be diffi-
cult to realize in engineering applications. So it is very important that the majority
of anisotropic materials has a structure that exhibits one or more geometric sym-
metries and the number of independent material parameters needed to describe the
material behavior can be reduced.
In the general case of 21 independent parameters, there is a coupling of each
loading component with all strain states and the model equations for structure el-
ements would be very complicated. The reduction of the number of independent
material parameters results therefore in a simplifying of the modelling and analysis
of structure elements composed of composite materials and impact the engineering
applications. The most important material symmetries are:
2.1.3.1 Monoclinic or Monotropic Material Behavior
A monoclinic material has one symmetry plane (Fig. 2.7). It is assumed that the
Fig. 2.7 Symmetry plane
(x1x2)of a monoclinic
material. All points of a
body which are symmetric
to this plane have identical
values of Ci j and Si j. Mirror
transformation (x1=x
1,
x2=x
2,x3=x
3)
x1
x2
x3
e
e
e3
e
e
e1e
e
e2
x
3
2.1 Generalized Hooke’s Law 33
symmetry plane is the (x1x2)plane. The structure of the stiffness or compliance
matrix must be in that way that a change of a reference system carried out by a sym-
metry about this plane does not modify the matrices, i.e. that the material properties
are identical along any two rays symmetric with respect to the (x1x2)plane. The
exploitation of the transformation rules leads to a stiffness matrix with the following
structure in the case of monoclinic material behavior
[Ci j]MC =
C11 C12 C13 0 0 C16
C12 C22 C23 0 0 C26
C13 C23 C33 0 0 C36
0 0 0 C44 C45 0
0 0 0 C45 C55 0
C16 C26 C36 0 0 C66
(2.1.42)
The compliance matrix has the same structure
[Si j]MC =
S11 S12 S13 0 0 S16
S12 S22 S23 0 0 S26
S13 S23 S33 0 0 S36
000S44 S45 0
000S45 S55 0
S16 S26 S36 0 0 S66
(2.1.43)
The number of non-zero elements Ci j or Si j reduces to twenty, the number of in-
dependent elements to thirteen. The loading-deformation couplings are reduced.
Consider for example the stress component
σ
6
τ
12. There is a coupling with the
extensional strains
ε
1,
ε
2,
ε
3and the shear strain
ε
6
γ
12 but the shear stress
σ
4or
σ
5produces only shear strains.
If an anisotropic material has the plane of elastic symmetry x1x3then it can be
shown that
[Ci j]MC =
C11 C12 C13 0C15 0
C12 C22 C23 0C25 0
C13 C23 C33 0C35 0
0 0 0 C44 0C46
C15 C25 C35 0C55 0
0 0 0 C46 0C66
(2.1.44)
and for the plane of elastic symmetry x2x3
[Ci j]MC =
C11 C12 C13 C14 0 0
C12 C22 C23 C24 0 0
C13 C23 C33 C34 0 0
C14 C24 C34 C44 0 0
0 0 0 0 C55 C56
0 0 0 0 C56 C66
(2.1.45)
34 2 Linear Anisotropic Materials
x1
x1
x2
x2
x3
x3
a b
Fig. 2.8 Orthotropic material behavior. aSymmetry planes (x1x2)and (x2x3),badditional
symmetry plane (x1x3)
The monoclinic compliance matrices [Si j ]MC have for both cases the same structure
again as the stiffness matrices [Ci j]MC.
2.1.3.2 Orthotropic Material Behavior
An orthotropic material behavior is characterized by three symmetry planes that are
mutually orthogonal (Fig. 2.8). It should be noted that the existence of two orthog-
onal symmetry planes results in the existence of a third. The stiffness matrix of an
orthotropic material has the following structure
[Ci j]O=
C11 C12 C13 0 0 0
C12 C22 C23 0 0 0
C13 C23 C33 0 0 0
0 0 0 C44 0 0
0 0 0 0 C55 0
0 0 0 0 0 C66
(2.1.46)
The compliance matrix has the same structure. An orthotropic material has 12 non-
zero and 9 independent material parameters. The stress-strain coupling is the same
as for isotropic material behavior. Normal stresses give rise to only extensional
strains and shear stresses only shear strains. Orthotropic material behavior is typ-
ical for unidirectional laminae with on-axis loading.
2.1 Generalized Hooke’s Law 35
2.1.3.3 Transversely Isotropic Material Behavior
A material behavior is said to be transversely isotropic if it is invariant with respect
to an arbitrary rotation about a given axis. This material behavior is of special im-
portance in the modelling of fibre-reinforced composite materials with coordinate
axis in the fibre direction and an assumed isotropic behavior in cross-sections or-
thogonal to the fibre direction. This type of material behavior lies between isotropic
and orthotropic.
If x1is the fibre direction, x2and x3are both rectangular to the fibre direction
and assuming identical material properties in these directions is understandable.
The structure of the stiffness matrix of a transversely isotropic material is given
in (2.1.47)
[Ci j]TI =
C11 C12 C12 0 0 0
C12 C22 C23 0 0 0
C12 C23 C22 0 0 0
0001
2(C22 C23)0 0
0 0 0 0 C55 0
0 0 0 0 0 C55
(2.1.47)
If x2is the fibre direction, x1and x3are both rectangular to the fibre direction and
assuming identical material properties in these directions then
[Ci j]TI =
C11 C12 C13 0 0 0
C12 C22 C12 0 0 0
C13 C12 C11 0 0 0
000C44 0 0
0 0 0 0 1
2(C11 C13)0
0 0 0 0 0 C44
(2.1.48)
If x3is the fibre direction, x1and x2are both rectangular to the fibre direction and
assuming identical material properties in these directions then
[Ci j]TI =
C11 C12 C13 0 0 0
C12 C11 C13 0 0 0
C13 C13 C33 0 0 0
000C44 0 0
0 0 0 0 C44 0
000001
2(C11 C12)
(2.1.49)
The compliance matrix has the same structure. For example for the case (2.1.47)
[Si j]TI =
S11 S12 S12 0 0 0
S12 S22 S23 0 0 0
S12 S23 S22 0 0 0
0 0 0 2(S22 S23)0 0
0 0 0 0 S55 0
0 0 0 0 0 S55
(2.1.50)
36 2 Linear Anisotropic Materials
The number of non-zero elements reduces again to twelve but the independent pa-
rameters are only five.
2.1.3.4 Isotropic Material Behavior
A material behavior is said to be isotropic if its properties are independent of the
choice of the reference system. There exist no preferred directions, i.e. the material
has an infinite number of planes and axes of material symmetry. Most conventional
materials satisfy this behavior approximately on a macroscopic scale.
The number of independent elasticity parameters is reduced to two and this leads
to the following stiffness matrix in the case of isotropic material behavior
[Ci j]I=
C11 C12 C12 000
C12 C11 C12 000
C12 C12 C11 000
0 0 0 C0 0
0 0 0 0 C0
0 0 0 0 0 C
(2.1.51)
with C=1
2(C11 C12). The compliance matrix has the same structure but with
diagonal terms 2(S11 S12 )instead of 1
2(C11 C12). Tables 2.3 and 2.4 summarize
the stiffness and compliance matrices for all material models described above.
2.1.4 Engineering Parameters
2.1.4.1 Orthotropic Material Behavior
The coordinates Cij and Si j of the stiffness and compliance matrix are mathematical
symbols relating stresses and strains. For practising engineers, a clear understand-
ing of each material parameter is necessary and requires a more mechanical meaning
by expressing the mathematical symbols in terms of engineering parameters such as
moduli Ei,Gi j and Poisson’s ratios
ν
i j. The relationships between mathematical and
engineering parameters are obtained by basic mechanical tests and imaginary math-
ematical experiments. The basic mechanical tests are the tension, compression and
torsion test to measure the elongation, the contraction and the torsion of a specimen.
In general, these tests are carried out by imposing a known stress and measuring the
strains or vice versa.
It follows that the compliance parameters are directly related to the engineering
parameters, simpler than those of the stiffness parameters. In the case of orthotropic
materials the 12 engineering parameters are Young’s moduli E1,E2,E3, the shear
moduli G23,G13,G12 and Poisson’s ratios
ν
i j,i,j=1,2,3(i6=j). For orthotropic
materials one can introduce the contracted engineering notation
2.1 Generalized Hooke’s Law 37
Table 2.3 Three-dimensional stiffness matrices for different material symmetries
Material model Elasticity matrix [Ci j ]
Anisotropy:
21 independent
material parameters
C11 C12 C13 C14 C15 C16
C22 C23 C24 C25 C26
C33 C34 C35 C36
C44 C45 C46
S Y M C55 C56
C66
Monoclinic:
13 independent
material parameters
Symmetry planex3=0
C14 =C15 =C24 =C25 =C34 =C35 =C46 =C56 =0
Symmetry planex2=0
C14 =C16 =C24 =C26 =C34 =C36 =C45 =C56 =0
Symmetry planex1=0
C15 =C16 =C25 =C26 =C35 =C36 =C45 =C56 =0
Orthotropic:
9 independent
material parameters
3 planes of symmetry x1=0,x2=0,x3=0
C14 =C15 =C16 =C24 =C25 =C26 =C34
=C35 =C36 =C45 =C46 =C56 =0
Transversely isotropic:
5 independent
material parameters
Plane of isotropy x3=0
C11 =C22,C23 =C13,C44 =C55,C66 =1
2(C11 C12)
Plane of isotropy x2=0
C11 =C33,C12 =C23,C44 =C66,C55 =1
2(C33 C13)
Plane of isotropy x1=0
C22 =C33,C12 =C13,C55 =C66,C44 =1
2(C22 C23)
all otherCi j like orthotropic
Isotropy:
2 independent
material parameters
C11 =C22 =C33,C12 =C13 =C23,
C44 =C55 =C66 =1
2(C11 C12)
all otherCi j =0
σ
1=E1
ε
1,
σ
2=E2
ε
2,
σ
3=E3
ε
3,
σ
4=E4
ε
4,
σ
5=E5
ε
5,
σ
6=E6
ε
6
(2.1.52)
with G23 E4,G13 E5,G12 E6.
The generalized Hooke’s law in the form (2.1.19) and (2.1.20) leads, for example,
to the relations
38 2 Linear Anisotropic Materials
Table 2.4 Three-dimensional compliance matrices for different material symmetries
Material model Compliance matrix [Si j ]
Anisotropy:
21 independent
material parameters
S11 S12 S13 S14 S15 S16
S22 S23 S24 S25 S26
S33 S34 S35 S36
S44 S45 S46
S Y M S55 S56
S66
Monoclinic:
13 independent
material parameters
Symmetry plane x3=0
S14 =S15 =S24 =S25 =S34 =S35 =S46 =S56 =0
Symmetry plane x2=0
S14 =S16 =S24 =S26 =S34 =S36 =S45 =S56 =0
Symmetry plane x1=0
S15 =S16 =S25 =S26 =S35 =S36 =S46 =S45 =0
Orthotropic:
9 independent
material parameters
3 planes of symmetry x1=0,x2=0,x3=0
S14 =S15 =S16 =S24 =S25 =S26 =S34
=S35 =S36 =S45 =S46 =S56 =0
Transversely isotropic:
5 independent
material parameters
Plane of isotropy x3=0
S11 =S22,S23 =S13,S44 =S55 ,S66 =2(S11 S12)
Plane of isotropy x2=0
S11 =S33,S12 =S23,S44 =S66 ,S55 =2(S33 S13)
Plane of isotropy x1=0
S22 =S33,S13 =S12,S55 =S66 ,S44 =2(S22 S23)
all other Si j like orthotropic
Isotropy:
2 independent
material parameters
S11 =S22 =S33,S12 =S13 =S23 ,
S44 =S55 =S66 =2(S11 S12)
all other Si j =0
ε
1=S11
σ
1+S12
σ
2+S13
σ
3,
ε
4=S44
σ
4,
ε
2=S12
σ
1+S22
σ
2+S23
σ
3,
ε
5=S55
σ
5,
ε
3=S13
σ
1+S23
σ
2+S33
σ
3,
ε
6=S66
σ
6
(2.1.53)
For uniaxial tension in xi-direction,
σ
16=0,
σ
i=0,i=2,...,6. This reduces (2.1.53)
to
ε
1=S11
σ
1,
ε
2=S12
σ
1,
ε
3=S13
σ
1,
ε
4=
ε
5=
ε
6=0,(2.1.54)
and the physical tensile tests provides the elastic parameters E1,
ν
12,
ν
13
2.1 Generalized Hooke’s Law 39
E1=
σ
1
ε
1
=1
S11
,
ν
12 =
ε
2
ε
1
=S12E1,
ν
13 =
ε
3
ε
1
=S13E1(2.1.55)
Analogous relations resulting from uniaxial tension in x2- and x3-directions and all
Si j are related to the nine measured engineering parameters (3 Young’s moduli and
6 Poisson’s ratios) by uniaxial tension tests in three directions x1,x2and x3.
From the symmetry of the compliance matrix one can conclude
ν
12
E1
=
ν
21
E2
,
ν
23
E2
=
ν
32
E3
,
ν
31
E3
=
ν
13
E1
or
ν
i j
Ei
=
ν
ji
Ej
,
ν
i j
ν
ji
=Ei
Ej
,i,j=1,2,3(i6=j)(2.1.56)
Remember that the first and the second subscript in Poisson’s ratios denote stress
and strain directions, respectively. Equations (2.1.56) demonstrate that the nine en-
gineering parameters are n ot independent parameters and that in addition to the three
tension tests, three independent shear tests are necessary to find the equations
ε
4=S44
σ
4,
ε
5=S55
σ
5,
ε
6=S66
σ
6,
which yield the relations
S44 =1
G23
=1
E4
,S55 =1
G13
=1
E5
,S66 =1
G12
=1
E6
(2.1.57)
Now all Si j in (2.1.20) can be substituted by the engineering parameters
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
E1
ν
12
E1
ν
13
E1
0 0 0
1
E2
ν
23
E2
0 0 0
1
E3
0 0 0
1
E4
0 0
S Y M 1
E5
0
1
E6
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.58)
As seen above, the relations between compliances Si j and engineering parameters
are very simple. This, however, is not the case for the relations between the stiffness
and engineering parameters. First we need to invert the compliance matrix S
S
Sand
to express the stiffness Ci j as a function of the compliances as follows. The shear
relations are uncoupled, and we obtain
C44 =1
S44
=G23,C55 =1
S55
=G13,C66 =1
S66
=G12 (2.1.59)
40 2 Linear Anisotropic Materials
So only a symmetric [3x3]-matrix must be inverted. The general formula is
Ci j =S1
i j =(1)i+jUi j
Det[Si j],Det[Si j ] =
S11 S12 S13
S12 S22 S23
S13 S23 S33
,(2.1.60)
where Ui j are the submatrices of S
S
Sto the element Si j , and yielding the relations
C11 =S22S33 S2
23
Det[Si j],C12 =S13S23 S12S33
Det[Si j],
C22 =S33S11 S2
13
Det[Si j],C23 =S12S13 S23S11
Det[Si j],
C33 =S11S22 S2
12
Det[Si j],C13 =S12S23 S13S22
Det[Si j]
(2.1.61)
Substituting the relations between Si j and engineering parameters given above in
(2.1.58), we obtain
C11 =(1
ν
23
ν
32)E1
,C12 =(
ν
12 +
ν
13
ν
32)E2
,
C22 =(1
ν
31
ν
13)E2
,C23 =(
ν
23 +
ν
21
ν
13)E3
,
C33 =(1
ν
21
ν
12)E3
,C13 =(
ν
13 +
ν
12
ν
23)E3
(2.1.62)
with
=1
ν
21
ν
12
ν
32
ν
23
ν
13
ν
31 2
ν
21
ν
13
ν
32. Considering Ei/
Ei,
1/SiEiwe finally get two subsystems
σ
1
σ
2
σ
3
=
(1
ν
23
ν
32)E1(
ν
12 +
ν
13
ν
32)E2(
ν
13 +
ν
12
ν
23)E3
(1
ν
31
ν
13)E2(
ν
23 +
ν
21
ν
13)E3
SYM (1
ν
21
ν
12)E3
ε
1
ε
2
ε
3
,
σ
4
σ
5
σ
6
=
E40 0
E50
SYM E6
ε
4
ε
5
ε
6
(2.1.63)
2.1.4.2 Transversally-Isotropic Material Behavior
It should be noted that in the case of transversely isotropic material with the
(x2x3)-plane of isotropy
E2=E3,
ν
12 =
ν
13,G12 =G13 ,G23 =E2
2(1+
ν
23),(2.1.64)
and we get
2.1 Generalized Hooke’s Law 41
S11 =1
E1
,S12 =S13 =
ν
12
E1
,S44 =1
G23
=2(1+
ν
23)
E2
,
S22 =S33 =1
E2
,S23 =
ν
23
E2
,S55 =S66 =1
G12
,
C11 =(1
ν
2
23)E1
,C12 =C13 =
ν
21(1+
ν
23)E1
,
C22 =C33 =(1
ν
12
ν
21)E2
,C23 =(
ν
23 +
ν
21
ν
12)E2
,
C44 =G23 =E2
2(1+
ν
23),C55 =C66 =G12
with
= (1+
ν
23)(1
ν
23 2
ν
21
ν
12).
Similar expressions one gets with the (x1x2)-plane of isotropy and considering
E1=E2,
ν
13 =
ν
23,G13 =G23 ,G12 =E1
2(1+
ν
12)(2.1.65)
or with the (x1x3)-plane of isotropy and
E1=E3,
ν
12 =
ν
23,G12 =G23 ,G13 =E1
2(1+
ν
13)(2.1.66)
The Young’s modulus and the Poissons ratio in the plane of isotropy often will be
designated as ETand
ν
TT.ETcharacterizes elongations or contractions of a trans-
versely isotropic body in the direction of the applied load in any direction of the
plane of isotropy and
ν
TT characterizes contractions or elongations of the body in
the direction perpendicular to the applied load, but parallel to the plane of isotropy.
The shear modulus GTT characterizes the material response under shear loading in
the plane of isotropy and takes the form
GTT =ET
2(1+
ν
TT),
i.e. any two of the engineering parameters ET,
ν
TT and GTT can be used to fully
describe the elastic properties in the plane of isotropy. A third independent primary
parameter should be EL. This Young’s modulus characterizes the tension respec-
tively compression response for the direction perpendicular to the plane of isotropy.
The fourth primary parameter should be the shear modulus GLT in the planes per-
pendicular to the plane of isotropy. As a fifth primary parameter can be chosen
ν
LT
or
ν
TL, which characterize the response in the plane of isotropy under a load in L-
direction or the response in the L-direction under a load in the plane of isotropy. The
stress-strain relations for an transversely isotropic body, if (x2x3) is the plane of
isotropy and with the reciprocity relations
ν
LT
EL
=
ν
TL
ET
42 2 Linear Anisotropic Materials
can be used in the following matrix form
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
EL
ν
TL
ET
ν
TL
ET
0 0 0
1
ET
ν
TT
ET
0 0 0
1
ET
0 0 0
1
GTT
0 0
S Y M 1
GLT
0
1
GLT
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.67)
with the engineering parameters
E1=EL,E2=E3=ET,E4=G23 =GTT =ET
2(1+
ν
TT),
E5=G13 =E6=G12 =GLT,
ν
12 =
ν
13 =
ν
LT,
ν
23 =
ν
TT,
ν
LT
EL
=
ν
TL
ET
With the (x1x2)-plane of isotropy the engineering parameters are
E1=E2=ET,E3=EL,E4=G23 =E5=G13 =GTL,
E6=G12 =GTT =ET
2(1+
ν
TT),
ν
13 =
ν
23 =
ν
TL,
ν
12 =
ν
TT
Notice that in literature the notations of Poisson’s ratios
ν
LT and
ν
TL can be corre-
spond to the opposite meaning.
2.1.4.3 Isotropic Material Behavior
For an isotropic material behavior in all directions, the number of independent en-
gineering parameters reduces to two
E1=E2=E3=E,
ν
12 =
ν
23 =
ν
13 =
ν
,
G12 =G13 =G23 =G=E
2(1+
ν
)
(2.1.68)
S11 =S22 =S33 =1
E,S12 =S13 =S23 =
ν
E,
S44 =S55 =S66 =1
G=2(1+
ν
)
E,C11 =C22 =C33 =(1
ν
)E
∗∗ ,
C12 =C13 =C23 =
ν
E
∗∗ ,C44 =C55 =C66 =G=E
2(1+
ν
)
2.1 Generalized Hooke’s Law 43
assuming
∗∗ = (1+
ν
)(12
ν
).
With this, all three-dimensional material laws for various material symmetries
interesting in engineering applications of composites are known. The relations be-
tween Si j,Ci j and engineering parameters are summarized in Table 2.5.
Consider that the elastic properties of an isotropic material are determined by two
independent parameters. The elastic parameters Young’s modulus Eand Poisson’s
ratio
ν
are generally used because they are determined easily in physical tests. But
also the so-called Lam ´e coefficients5
λ
and
µ
, the shear modulus Gor the bulk
modulus Kcan be used if it is suitable. There are simple relations between the
parameters, e.g. as a function of E,
ν
λ
=E
ν
(1+
ν
)(12
ν
),
µ
=E
2(1+
ν
)=G,K=E
3(12
ν
),
ν
=
λ
2(
λ
+
µ
),E=
µ
(3
λ
+2
µ
)
λ
+
µ
,K=
λ
+2
3
µ
(2.1.69)
Summarizing the constitutive equations for isotropic, transversely isotropic and
orthotropic materials, which are most important in the engineering applications of
composite structural mechanics one can find that the common features of the re-
lationships between stresses and strains for these material symmetries are that the
normal stresses are not couplet with shear strains and shear stresses are not coupled
with the normal strains. Each shear stress is only related to the corresponding shear
strain. These features are not retained in the more general case of an monoclinic or
a general anisotropic material.
2.1.4.4 Monoclinic Material Behavior
In the case of monoclinic materials we have 13 mutually independent stiffness or
compliances. Therefore we have in comparison with orthotropic materials to intro-
duce four additional engineering parameters and keeping in mind, that the mono-
clinic case must comprise the orthotropic case, we should not change the engineer-
ing parameters of orthotropic material behavior. Assuming that (x1x2) is the plane
of elastic symmetry, the additional parameters are related to the compliance matrix
components S16,S26,S36 and S46.
The first three pair normal strains
ε
1,
ε
2,
ε
3to the shear stress
σ
6and vice versa
the shear strain
ε
6to the normal stresses
σ
1,
σ
2,
σ
3. The fourth one couple the shear
strain
ε
4to the shear stress
σ
5and vice versa the shear strain
ε
5to the shear stress
σ
4.
In a compact notation the strain-stress relations for an anisotropic material having
one plane of elastic symmetry (x1x2) are
5Gabriel eon Jean Baptiste Lam´e (22 July 1795 Tours - †1 May 1870 Paris) - mathematician,
who contributed to the mathematical theory of elasticity (Lam´e’s parameters in elasticity and fluid
mechanics)
44 2 Linear Anisotropic Materials
Table 2.5 Relationships between Si j,Ci j and the engineering parameters for orthotropic, trans-
versely-isotropic and isotropic material
Orthotropic material
S11 =E1
1,S12 =S21 =
ν
12E1
1,S44 =G1
23 =E1
4
S22 =E1
2,S13 =S31 =
ν
13E1
1,S55 =G1
13 =E1
5
S33 =E1
3,S23 =S32 =
ν
23E1
2,S66 =G1
12 =E1
6
ν
i j =
ν
ji(Ei/Ej),Ei= (
ν
i j/
ν
ji)Eji,j=1,2,3
C11 =S22S33 S2
23
det[Si j]=(1
ν
23
ν
32)E1
,C44 =1
S44
=G23 =E4
C22 =S33S11 S2
13
det[Si j]=(1
ν
31
ν
13)E2
,C55 =1
S55
=G31 =E5
C33 =S11S22 S2
12
det[Si j]=(1
ν
21
ν
12)E3
,C66 =1
S66
=G12 =E6
C12 =S13S23 S12S33
det[Si j]=(
ν
12 +
ν
32
ν
13)E2
=(
ν
21 +
ν
31
ν
23)E1
=C21
C13 =S12S23 S13S22
det[Si j]=(
ν
13 +
ν
12
ν
23)E3
=(
ν
31 +
ν
21
ν
32)E1
=C31
C23 =S12S13 S23S11
det[Si j]=(
ν
23 +
ν
21
ν
13)E3
=(
ν
32 +
ν
12
ν
31)E2
=C32
=1
ν
12
ν
21
ν
23
ν
32
ν
31
ν
13 2
ν
21
ν
13
ν
32
Transversely-isotropic material
(x2x3)-plane of isotropy
E1,E2=E3,E5=E6,
ν
12 =
ν
13,E4=E2
2(1+
ν
23)
(x1x2)-plane of isotropy
E1=E2,E3,E4=E5,
ν
13 =
ν
23,E6=E1
2(1+
ν
12)
(x1x3)-plane of isotropy
E1=E3,E2,E4=E6,
ν
12 =
ν
23,E5=E3
2(1+
ν
13)
Isotropic material
E1=E2=E3=E,
ν
12 =
ν
21 =
ν
13 =
ν
31 =
ν
23 =
ν
32 =
ν
,
E4=E5=E6=G=E
2(1+
ν
)
2.1 Generalized Hooke’s Law 45
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
E1
ν
21
E2
ν
31
E3
0 0
η
61
E6
ν
12
E1
1
E2
ν
32
E3
0 0
η
62
E6
ν
13
E1
ν
23
E2
1
E3
0 0
η
63
E6
0 0 0 1
E4
µ
54
E5
0
0 0 0
µ
45
E4
1
E5
0
η
16
E1
η
26
E2
η
36
E3
0 0 1
E6
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.70)
with the following reciprocal relations
η
61
E6
=
η
16
E1
,
η
62
E6
=
η
26
E2
,
η
63
E6
=
η
36
E3
,
µ
54
E5
=
µ
45
E4
(2.1.71)
and the compliance components
S16 =
η
61
E6
,S26 =
η
62
E6
,S36 =
η
63
E6
,S45 =
µ
54
E5
(2.1.72)
η
61,
η
62 and
η
63 are extension-shear coupling coefficients indicating normal strains
induced by shear stress
σ
6and
η
16,
η
26 and
η
36 the shear-extension coupling co-
efficients characterizing shear strain
ε
6caused by normal stresses.
µ
45 and
µ
54 are
shear-shear coupling coefficients.
The stiffness matrix for the monoclinic material can be found as the inverse of
the compliance matrix, but the expressions are unreasonable to present in an explicit
form. However, the inverse of a matrix can be easily calculated using standard nu-
merical procedures. Also for a generally anisotropic material the compliance can be
formulated with help of eight additional coupling parameters but the stiffness matrix
should be calculated numerically.
2.1.5 Two-Dimensional Material Equations
In most structural applications the structural elements are simplified models by
reducing the three-dimensional state of stress and strain approximately to a two-
dimensional plane stress or plane strain state. A thin lamina for instance can be
considered to be under a condition of plane stress with all stress components in
the out-of-plane direction being approximately zero. The different conditions for a
plane stress state in the planes (x1x2),(x2x3)and (x1x3)are demonstrated in
Fig. 2.9.
In the following, the plane stress state with respect to the (x1x2)plane (Fig.
2.9a) is considered. In addition, since in the case of unidirectional long-fibre rein-
forced laminae the most general type of symmetry of the material properties is the
46 2 Linear Anisotropic Materials
x1
x1
x1
x2
x2
x2
x3
x3
x3
σ
2
σ
1
σ
6
σ
3
σ
2
σ
4
σ
1
σ
3
σ
5
abc
Fig. 2.9 Plane stress state. a(x1x2)-plane,
σ
3=
σ
4=
σ
5=0, b(x2x3)-plane,
σ
1=
σ
5=
σ
6=0, c(x1x3)-plane,
σ
2=
σ
4=
σ
6=0
monoclinic one the generalized Hooke’s law (2.1.20) is reduced taking into account
(2.1.43) to
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
S11 S12 S13 0 0 S16
S22 S23 0 0 S26
S33 0 0 S36
SS44 S45 0
YS55 0
MS66
σ
1
σ
2
0
0
0
σ
6
(2.1.73)
That means
σ
3=
σ
4=
σ
5=0, and we have three in-plane constitutive equations
ε
1=S11
σ
1+S12
σ
2+S16
σ
6
ε
2=S12
σ
1+S22
σ
2+S26
σ
6,Si j =Sji
ε
6=S16
σ
1+S26
σ
2+S66
σ
6
(2.1.74)
and an additional equation for strain
ε
3in x3-direction
ε
3=S13
σ
1+S23
σ
2+S36
σ
6(2.1.75)
The other strains
ε
4,
ε
5are equal to 0 considering the monoclinic material behavior.
If the plane stress assumptions are used to simplify the generalized stiffness equa-
tions (2.1.19) taking into account (2.1.42), the result is
σ
1
σ
2
0
0
0
σ
6
=
C11 C12 C13 0 0 C16
C22 C23 0 0 C26
C33 0 0 C36
SC44 C45 0
YC55 0
MC66
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
(2.1.76)
or again three in-plane equations
2.1 Generalized Hooke’s Law 47
σ
1=C11
ε
1+C12
ε
2+C13
ε
3+C16
ε
6,
σ
2=C12
ε
1+C22
ε
2+C23
ε
3+C26
ε
6,Ci j =Cj i
σ
6=C16
ε
1+C26
ε
2+C36
ε
3+C66
ε
6
(2.1.77)
There are another three equations. At first we obtain
σ
4=S44
ε
4+S45
ε
5=0,
σ
5=S45
ε
4+S55
ε
5=0
Since S44,S45 ,S55 are arbitrary but the stiffness matrix should be positive definite it
is obvious that
S44S55 S2
45 >0
and from the
σ
4=
σ
5=0 condition it follows that
ε
4,
ε
5must be equal to 0. Taking
into account the last condition
σ
3=C13
ε
1+C23
ε
2+C33
ε
3+C36
ε
6=0
the strain
ε
3can be obtained
ε
3=1
C33
(C13
ε
1+C23
ε
2+C36
ε
6)(2.1.78)
and eliminated and substituted in Eqs. (2.1.77). After substituting
ε
3using Eq.
(2.1.78) Eqs. (2.1.77) leads to
σ
i=Ci j Ci3Cj3
C33
ε
j,i,j=1,2,6,(2.1.79)
respectively
σ
i=Qi j
ε
j,i,j=1,2,6 (2.1.80)
The Qi j are the reduced stiffness. For the three cases in Fig. 2.9 we obtain
σ
i=Qi j
ε
j,Qi j =Ci j Ci3Cj3
C33
,i,j=1,2,6,(x1x2)plane of symmetry,
σ
i=Qi j
ε
j,Qi j =Ci j Ci1Cj1
C11
,i,j=2,3,4,(x2x3)plane of symmetry,
σ
i=Qi j
ε
j,Qi j =Ci j Ci2Cj2
C22
,i,j=1,3,5,(x1x3)plane of symmetry
The number of unknown independent parameters of each of the matrices Si j,Ci j or
Qi j is six. It is very important to note that the elements in the plane stress compliance
matrix are simply a subset of the elements from the three-dimensional compliance
matrix and their numerical values are identical. On the other hand, the elements
of the reduced stiffness matrix involve a combination of elements from the three-
dimensional stiffness matrix and the numerical values of the Qi j differ from their
counterparts Ci j , i.e. they are actually less than the numerical values for Ci j . In
order to keep consistency with the generalized Hooke’s law, Eq. (2.1.78) should be
48 2 Linear Anisotropic Materials
used when calculating the transverse normal strain
ε
3and the general case leads to
consistent relations for the transverse shear strains
ε
4and
ε
5.
For an orthotropic material behavior under plane stress and on-axis orientation of
the reference system there are four independent parameters and for isotropic behav-
ior there are only two. The mathematical notations Sij ,Ci j or Qi j can be shifted to
the engineering notation. Tables 2.6 and 2.7 summarize the compliance and stiffness
matrices for the plane stress state.
Considering a plane strain state in the (x1x2)plane we have the three non-zero
strains
ε
1,
ε
2and
ε
6but the four nonzero stress components
σ
1,
σ
2,
σ
3,
σ
6. Analo-
gous to the plane stress state, here the stress
σ
3normal to the (x1x2)plane is not
an independent value and can be eliminated
ε
3=S13
σ
1+S23
σ
2+S33
σ
3+S36
σ
6=0,
σ
3=1
S33
(S13
σ
1+S23
σ
2+S36
σ
6)
Therefore in the case of plane strain, the Cij ,i,j=1,2,6 can be taken directly from
the three-dimensional elasticity law and instead of Si j reduced compliances Vi j have
to be used
Table 2.6 Compliance matrices for various material models, plane stress state
Material model
ε
=S
σ
ε
=S
σ
ε
=S
σ
Anisotropy: Compliances Si j
6 independentmaterial parameters
ε
1
ε
2
ε
6
=
S11 S12 S16
S22 S26
S66
σ
1
σ
2
σ
6
Orthotropy: S16 =S26 =0
4 independent material parameters S11 =1
E1
,S22 =1
E2
S12 =
ν
12
E1
=
ν
21
E2
Reference system: on-axis S66 =1
G12
Isotropy: S16 =S26 =0
2 independent material parameters S11 =S22 =1
E,S12 =
ν
E,
Reference system: as you like S66 =2(S11 S12) = 2(1+
ν
)
E=1
G
2.1 Generalized Hooke’s Law 49
Table 2.7 Stiffness matrices for various material models, plane stress state
Material model
σ
=Q
ε
σ
=Q
ε
σ
=Q
ε
Anisotropy: Reduced stiffness Qij
6 independent material parameters
σ
1
σ
2
σ
6
=
Q11 Q12 Q16
Q22 Q26
Q66
ε
1
ε
2
ε
3
Orthotropy: Q16 =Q26 =0,Q66 =1
S66
=G12
4 independent material parameters Q11 =S22
=E1
1
ν
12
ν
21
Q22 =S11
=E2
1
ν
12
ν
21
Reference system: on-axis Q12 =S12
=
ν
12E2
1
ν
12
ν
21
=S11S22 S2
12
Isotropy: Q16 =Q26 =0
2 independent material parameters Q11 =Q22 =E
1
ν
2
Reference system: as you like Q12 =
ν
E
1
ν
2,Q66 =E
2(1+
ν
)=G
σ
i=Ci j
ε
j,i,j=1,2,6,
ε
i=Vij
σ
j,Vij =Si j Si3Sj3
S33
,i,j=1,2,6
Table 2.8 summarizes for the three-dimensional states and the plane stress and strain
states the number of non-zero and of independent material parameters.
In the two-dimensional equations of anisotropic elasticity, either reduced stiff-
ness or reduced compliances are introduced into the material laws. These equations
are most important in the theory of composite single or multilayered elements, e.g.
of laminae or laminates. The additional transformations rules which are necessary
in laminae and laminate theories are discussed in more detail in Chap. 3. Tables 2.6
and 2.7 above shows the relationship between stress and strain through the compli-
ance [Si j ]or the reduced stiffness [Qi j ]matrix for the plane stress state and how the
Si j and Qi j are related to the engineering parameters. For a unidirectional lamina the
engineering parameters are:
50 2 Linear Anisotropic Materials
Table 2.8 Stiffness and compliance parameters for stress and strain equations
σ
σ
σ
=C
C
C
ε
ε
ε
,
ε
ε
ε
=S
S
S
σ
σ
σ
,C
C
C=S
S
S1, plane stress state
σ
σ
σ
=Q
Q
Q
ε
ε
ε
,
ε
ε
ε
=S
S
S
σ
σ
σ
,Q
Q
Q=S
S
S1, plane strain state -
σ
σ
σ
=C
C
C
ε
ε
ε
,
ε
ε
ε
=V
V
V
σ
σ
σ
,C
C
C=V
V
V1,Qi j and Vi j are the reduced stiffness and compliances
Material model Number of non-zero Number of independent
parameters parameters
Three-dimensional Cij ;Si j Ci j;Si j
stress- or i,j=1,...,6i,j=1,...,6
strain state
Anisotropic 36 21
Monotropic 20 13
Orthotropic 12 9
Transversely isotropic 12 5
Isotropic 12 2
Plane stress state Qi j;Si j Qij ;Si j
(x1x2)-plane i,j=1,2,6i,j=1,2,6
Anisotropic 9 6
Orthotropic 5 4
Isotropic 5 2
Plane strain state Ci j;Vi j Cij ;Vij
(x1x2)-plane i,j=1,2,6i,j=1,2,6
Anisotropic 9 6
Orthotropic 5 4
Isotropic 5 2
E1longitudinal Young’s modulus in the principal direction 1 (fibre direction)
E2transverse Young’s modulus in direction 2 (orthogonal to the fibre direction)
ν
12 major Poisson’s ratio as the ratio of the negative normal strain in direction 2
to the normal strain in direction 1 only if normal load is applied in direction 1
G12 in-plane shear modulus for (x1x2)plane
The four independent engineering elastic parameters are experimentally measured
as follows:
Pure tensile load in direction 1:
σ
16=0,
σ
2=0,
σ
6=0
With
ε
1=S11
σ
1,
ε
2=S12
σ
1,
ε
6=0 are
E1=
σ
1
ε
1
=1
S11
,
ν
12 =
ε
2
ε
1
=S12
S11
Pure tensile load in direction 2:
σ
1=0,
σ
26=0,
σ
6=0
With
ε
1=S12
σ
2,
ε
2=S22
σ
2,
ε
6=0 are
E2=
σ
2
ε
2
=1
S22
,
ν
21 =
ε
1
ε
2
=S12
S22
ν
21 is usually called the minor Poisson’s ratio and we have the reciprocal rela-
tionship
ν
12/E1=
ν
21/E2.
Pure shear stress in the (x1x2)plane:
σ
1=
σ
2=0,
σ
66=0
With
ε
1=
ε
2=0,
ε
6=S66
σ
6is
2.1 Generalized Hooke’s Law 51
G12 =
σ
6
ε
6
=1
S66
With the help of Tables 2.6 and 2.7, the relating equations of stresses and strains
are given through any of the following combinations of four parameters: (Q11,Q12 ,
Q22,Q66 ),(S11,S12,S22,S66),(E1,E2,
ν
12,G12 ).
In Chap. 3 the evaluation of the four engineering elastic parameters is given ap-
proximately by averaging the fibre-matrix material behavior. There are different ap-
proaches for determining effective elastic moduli, e.g. in a simple way with ele-
mentary mixture rules, with semi-empirical models or an approach based on the
elementary theory.
2.1.6 Curvilinear Anisotropy
The type of anisotropy considered above was characterized by the equivalence of
parallel directions passing through different points of the homogeneous anisotropic
body and we can speak of a rectilinear anisotropy. Another kind of anisotropy is the
case, if one chooses a system of curvilinear coordinates in such a manner that the
coordinate directions coincide with equivalent directions of elastic properties at dif-
ferent points of an anisotropic body. The elements of the body, which are generated
by three pairs of coordinate surfaces possess identical elastic properties and we can
speak of a curvilinear anisotropy.
In the frame of this textbook we limit the considerations to cylindrical anisotropy,
which is also the most common case of this type of anisotropy. The generalized
Hooke’s law equations (2.1.21) are now considered in cylindrical coordinates x1=r,
x2=
θ
,x3=zand we have the stress and strain vectors in the contracted single
subscript notation
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6T=
σ
r
σθσ
z
σθ
z
σ
rz
σ
r
θ
T,
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6T=
ε
r
εθε
z
εθ
z
ε
rz
ε
r
θ
T(2.1.81)
In the specific cases of material symmetries the general constitutive equation
in cylindrical coordinates can be simplified analogous to the case of rectilinear
anisotropy.
In the specific case of an orthotropic cylindrical response there are three orthog-
onal planes of elastic symmetries. One plane is perpendicular to the axis z, another
one is tangential to the surface (
θ
z) and the third one is a radial plane (Fig. 2.10).
Another case of material symmetry in possible practical situation is a transversely
isotropic cylinder or cylindrical tube with the plane of isotropy (r
θ
). In this case
we obtain, as analog to (2.1.67)
52 2 Linear Anisotropic Materials
z
x
r
θ
z-plane
rz-plane
θ
z-plane
z
x
θ
r
Fig. 2.10 Cylindrical orthotropic material symmetry
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
ET
ν
TT
ET
ν
LT
EL
0 0 0
1
ET
ν
LT
EL
0 0 0
1
EL
0 0 0
1
GLT
0 0
S Y M 1
GLT
0
1
GTT
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
,(2.1.82)
where the index T is associated with the coordinate directions rand
θ
and the index
L with the coordinate direction zand the reciprocal relations are
ν
LT
EL
=
ν
TL
ET
The stress-strain equations for the orthotropic case of cylindrical anisotropy are ob-
tained using engineering parameters
2.1 Generalized Hooke’s Law 53
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
Er
νθ
r
E
θ
ν
zr
Er
0 0 0
1
E
θ
ν
z
θ
Er
0 0 0
1
Er
0 0 0
1
G
θ
z
0 0
S Y M 1
Grz
0
1
Gr
θ
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.83)
The indices r,
θ
and zof the engineering parameters are associated with the indices
1, 2 and 3 and the strain-stress equations may also be written in a different way by
using the numerical subscripts. Further notice the reciprocal relations
Ei
ν
ji =Ej
ν
i j,i,j=r,
θ
,z
and the G
θ
z,Grz and Gr
θ
may be written in the more general form E4,E5,E6.
There are two practical situations for a monoclinic material behavior. The first
case can be one plane of elastic symmetry (r
θ
) which is rectilinear to the z-axis.
The case is interesting when considering composite discs or circular plates. The
stress-strain equations follow from (2.1.70) after substituting the subscripts 1,2 and
3 by the engineering parameters to r,
θ
and zand the shear moduli E4,E5and E6
by G
θ
z,Grz and Gr
θ
. The second case can be one plane of elastic symmetry (
θ
z)
as a cylindrical surface with the axis rperpendicular to this surface. This situation
is of practical interest when considering e.g. filament wound cylindrical shells and
we get the strain-stress relations which couple all three normal strains to the shear
strain
ε
4and both shear strains
ε
5,
ε
6to both shear stresses
σ
5,
σ
6
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
=
1
E1
ν
21
E2
ν
31
E3
η
41
E4
0 0
1
E2
ν
32
E3
η
42
E4
0 0
1
E3
η
43
E4
0 0
1
E4
0 0
S Y M 1
E5
µ
65
E6
1
E6
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
(2.1.84)
The subscripts 1, 2, 3 of the Young’s moduli and the Poisson’s ratios will be shifted
to r,
θ
,zand the moduli E4,E5,E6to G
θ
z,Grz,Gr
θ
. There are as above reciprocal
relations for
ν
i j ,
η
i j and
µ
i j
54 2 Linear Anisotropic Materials
η
41
E4
=
η
14
E1
,
η
42
E4
=
η
24
E2
,
η
43
E4
=
η
34
E3
,
µ
65
E6
=
µ
56
E5
,
ν
i j
Ei
=
ν
ji
Ej
,i,j=1,2,3
(2.1.85)
2.1.7 Problems
Exercise 2.1. Calculate for the tensile bar consisting of three parts (Fig. 2.2) the
elongation l, the strain
ε
and the stress
σ
as functions of A,land F:
1. The stiffness are arranged in parallel:
E1=E3=70 GPa, E2=3 GPa, A1=A3=0,1A,A2=0,8A.
2. The stiffness are arranged in series:
E1=E3=70 GPa, E2=3 GPa, l1=l3=0,1l,l2=0,8l.
Solution 2.1. The solution can be obtained considering the basic assumptions of the
iso-strain and the iso-stress models.
1. Assumptions
ε
i=
ε
,li=l,i=1,2,3,A=
3
i=1
Ai,F=
3
i=1
Fi
From
σ
=E
ε
follows
σ
1=E1
ε
=70GPa
ε
,
σ
2=E2
ε
=3GPa
ε
,
σ
3=E3
ε
=70GPa
ε
With F=
σ
A=EA
ε
and Fi=EiAi
ε
follows
F=
3
i=1
(EiAi)
ε
=16,4 GPa
ε
A,E=16,4 GPa,
ε
=F
EA =l
l
yields
l=Fl
EA =1
16,4
Fl
A(GPa)1
and the solutions are
l=1
16,4
Fl
A(GPa)1=l(F,l,A),
ε
=l
l=1
16,4
F
A(GPa)1=
ε
(F,A),
σ
i=Ei
ε
=1
16,4
EiF
A(GPa)1=
σ
i(F,A),i=1,2,3
2.1 Generalized Hooke’s Law 55
2. Assumptions
l=
3
i=1li,Fi=F,
ε
i=li
li
,i=1,2,3
From l=3
i=1li=3
i=1
li
ε
iand F=EA
ε
follows
l=
3
i=1li
EiF
A=0,1
70 +0,8
3+0,1
70 Fl
A(GPa)1,
l=
ε
l=1
E
Fl
A=0,2695 Fl
A(GPa)1,E=3,71 GPa
The functions l,
ε
and
σ
are
l=0,2695 Fl
A(GPa)1=l(F,l,A),
ε
=0,2695 F
A(GPa)1=
ε
(F,A),
σ
=E
ε
=F
A=
σ
(F,A)
Exercise 2.2. The relationship between the load Fand the elongation lof a tensile
bar (Fig. 2.1) is
F=EA0
l0l=Kl
Eis the Young’s modulus of the material, A0the cross-sectional area of the bar and
l0is the length. The factor K=E A0/l0is the stiffness per length and characterizes
the mechanical performance of the tensile bar. In the case of two different bars
with Young’s moduli E1,E2, densities
ρ
1,
ρ
2, the cross-sectional areas A1,A2and
the lengths l1,l2the ratios of the stiffness K1and K2per length and the mass of the
bars are K1
K2
=E1A1
E2A2
l2
l1
,m1
m2
=l1A1
ρ
1
l2A2
ρ
2
Verify that for l1=l2and m1=m2the ratio K1/K2only depends on the ratio of the
specific Young’s moduli E1/
ρ
1and E2/
ρ
2.
Solution 2.2. Introducing the densities
ρ
1/
ρ
2into the stiffness ratio K1/K2yields
K1
K2
=E1/
ρ
1
E2/
ρ
2
m1
m2l2
l12
and with m1=m2,l2=l1
K1
K2
=E1/
ρ
1
E2/
ρ
2
56 2 Linear Anisotropic Materials
Note 2.1. A material with the highest value of E/
ρ
has the highest tension stiffness.
Exercise 2.3. For a simply supported beam with a single transverse load in the mid-
dle of the beam we have the following equation
F=48 EI
l3f=K f
Fis the load and fis the deflection in the middle of the beam, lis the length of
the beam between the supports and Ithe moment of inertia of the cross-section.
The coefficient K=48E I/l3characterizes the stiffness of the beam. Calculate Kfor
beams with a circle or a square cross-sectional area (radius ror square length a) and
two different materials E1,
ρ
1and E2,
ρ
2but of equal length l, moments of inertia
and masses. Verify that for m1=m2the ratio of the stiffness coefficients K1/K2only
depends on the ratios E1/
ρ
2
1and E2/
ρ
2
2.
Solution 2.3. Moments of inertia and masses of the two beams are
1. circle cross-sectional: I=
π
r4/4,m=r2
π
l
ρ
,
2. square cross-sectional: I=a4/12,m=a2l
ρ
In case 1. we have
K=48EI
l3=48E
l3
π
r4
4=48E/
ρ
2
4l3
m2
π
l2,
K1
K2
=E1/
ρ
2
1
E2/
ρ
2
2m1
m22l2
l15
With l1=l2,m1=m2we obtain
K1
K2
=E1/
ρ
2
1
E2/
ρ
2
2
In case 2. we have
K=48EI
l3=48E
l3
a4
12 =4E
l3
m2
ρ
2l2,
K1
K2
=E1/
ρ
2
1
E2/
ρ
2
2m1
m22l2
l15
With l1=l2,m1=m2we obtain
K1
K2
=E1/
ρ
2
1
E2/
ρ
2
2
Note 2.2. The best material for an optimal bending stiffness of the beam is that with
the highest value of E/
ρ
2.
Exercise 2.4. Formulate explicitly the transformation matrices (
3
T
T
T
σ
σ
σ
)1and (
3
T
T
T
ε
ε
ε
)1
for a rotation about the e
e
e3-direction (Fig. 2.6).
2.1 Generalized Hooke’s Law 57
Solution 2.4. With Eqs. (2.1.29), (2.1.39) and (2.1.40) follows (
3
T
T
T
σ
σ
σ
)1= (
3
T
T
T
ε
ε
ε
)Tand
(
3
T
T
T
ε
ε
ε
)1= (
3
T
T
T
σ
σ
σ
)T
(
3
T
T
T
σ
σ
σ
)1=
c2s20 0 0 2cs
s2c20 0 0 2sc
0 0 1 0 0 0
0 0 0 c s 0
0 0 0 s c 0
cs cs 0 0 0 c2s2
,
(
3
T
T
T
ε
ε
ε
)1=
c2s2000 cs
s2c2000 sc
0 0 1 0 0 0
0 0 0 c s 0
0 0 0 s c 0
2cs 2cs 000c2s2
Exercise 2.5. Consider the coordinate transformation that corresponds with reflec-
tion in the plane x1x2:x
1=x1,x
2=x2,x
3=x3. Define for this case
1. the coordinate transformation matrix [Ri j ]and
2. the stress and strain transformation matrices [
σ
Tpq]and [
ε
Tpq].
Solution 2.5. The solution contains two parts: the coordinate transformation and the
stress/strain transformation.
1. With Ri j =cos(e
i,ej)(2.1.22) follows
R11 =1,R12 =0,R13 =0,R21 =0,R22 =1,R23 =0,
R31 =0,R32 =0,R33 =1
and the transformation matrix takes the form
[Ri j] =
1 0 0
0 1 0
0 0 1
2. With the help of the transformation matrix App. B we can see that both matrices
are diagonal with the following nonzero elements for
σ
Tpq
R2
11 =1,R2
22 =1,R2
33 =1,
R22R33 +R23R32 =1,R11R33 +R13R31 =1,R11 R22 +R12R21 =1
and for
ε
Tpq
R2
11 =1,R2
22 =1,R2
33 =1,
R22R33 +R23R22 =1,R11R33 +R13R31 =1,R11 R22 +R12R21 =1
The transformation matrices take the form
58 2 Linear Anisotropic Materials
[
σ
Tpq] =
1 0 0 0 0 0
0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
= [
ε
Tpq]
Exercise 2.6. The engineering material parameters for an orthotropic material are
given by
E1=173GPa,E2=33,1GPa,E3=5,17GPa,
E4=3,24GPa,E5=8,27GPa,E6=9,38GPa,
ν
12 =0,036,
ν
13 =0,25,
ν
23 =0,171
Calculate the stiffness matrix C
C
Cand the compliance matrix S
S
S.
Solution 2.6. With Table 2.5 we find the Si j and the Ci j
S11 =E1
1=5,780 103GPa1,
S12 =S21 =
ν
12E1
1=0,208 103GPa1,
S22 =E1
2=30,211 103GPa1,
S13 =S31 =
ν
13E1
1=1,445 103GPa1,
S33 =E1
3=193,424 103GPa1,
S23 =S32 =
ν
23E1
2=5,166 103GPa1,
S44 =E1
4=308,642 103GPa1,
S55 =E1
5=120,919 103GPa1,
S66 =E1
6=106,610 103GPa1,
=1
ν
12
ν
21
ν
23
ν
32
ν
31
ν
13 2
ν
21
ν
13
ν
32,
ν
21 =
ν
12(E2/E1) = 0,0069,
ν
31 =
ν
13(E3/E1) = 0,0075,
ν
32 =
ν
23(E3/E2) = 0,027,=0,993,¯
Ei=Ei/,i=1,2,3
C11 = (1
ν
23
ν
32)¯
E1=173,415GPa,
C22 = (1
ν
31
ν
13)¯
E2=33,271GPa,
C33 = (1
ν
12
ν
21)¯
E3=5,205GPa,
C12 = (
ν
12 +
ν
13
ν
32)¯
E2=1,425GPa,
C13 = (
ν
13 +
ν
12
ν
23)¯
E3=1,334GPa,
C23 = (
ν
23 +
ν
21
ν
13)¯
E3=0,899GPa,
C44 =E4,C55 =E5,C66 =E6
With the values for Ci j and Si j the stiffness matrix C
C
Cand the compliance matrix S
S
S
can be written
2.2 Fundamental Equations and Variational Solution Procedures 59
C
C
C=
173,415 1,425 1,334 0 0 0
1,425 33,271 0,899 0 0 0
1,334 0,899 5,205 0 0 0
0 0 0 3,24 0 0
0 0 0 0 8,27 0
0 0 0 0 0 9,38
GPa,
S
S
S=
5,780 0,208 1,445 0 0 0
0,208 30,211 5,166 0 0 0
1,445 5,166 193,424 0 0 0
0 0 0 308,642 0 0
0 0 0 0 120,919 0
0 0 0 0 0 106,610
103GPa1
2.2 Fundamental Equations and Variational Solution Procedures
Below we discuss at first the fundamental equations of the anisotropic elasticity for
rectilinear coordinates. The system of equations can be divided into two subsys-
tems: the first one is material independent that means we have the same equations
as in the isotropic case. To this subsystem belong the equilibrium equations (static
or dynamic) and the kinematic equations (the strain-displacement equations and the
compatibility conditions). To this subsystem one has to add the constitutive equa-
tions. In addition, we must introduce the boundary and, may be, the initial conditions
to close the initial-boundary problem. At second, considering that closed solutions
are impossible in most of the practical cases approximative solution techniques are
briefly discussed. The main attention will be focussed on variational formulations.
2.2.1 Boundary and Initial-Boundary Value Equations
The fundamental equations of anisotropic elasticity can be formulated and solved
by a displacement, a stress or a mixedapproach. In all cases the starting point are the
following equations:
The static or dynamic equilibrium equations formulated for an infinitesimal cube
of the anisotropic solid which is subjected to body forces and surface forces char-
acterized by force density per unit surface. In Fig. 2.11 the stress and the volume
force components are shown in the x1-direction. Assuming the symmetry of the
stress tensor, three static equations link six unknown stress components. In the
case of dynamic problems the inertia forces are expressed through displacements,
therefore the equations of motion contain both, six unknown stress and three un-
known displacement components
60 2 Linear Anisotropic Materials
Fig. 2.11 Infinitesimal cube
with lengths dx1,dx2,dx3:
stress and volume force com-
ponents in x1-direction
x1
x2
x3
dx1
dx2
dx3
σ
1+d
σ
1
p1
σ
6+d
σ
6
σ
1
σ
5+d
σ
5
σ
6
σ
5
σ
1
x1
+
σ
6
x2
+
σ
5
x3
+p1=0,
σ
6
x1
+
σ
2
x2
+
σ
4
x3
+p2=0,
σ
5
x1
+
σ
4
x2
+
σ
3
x3
+p3=0,
static equations (2.2.1)
σ
1
x1
+
σ
6
x2
+
σ
5
x3
+p1=
ρ
2u1
t2,
σ
6
x1
+
σ
2
x2
+
σ
4
x3
+p2=
ρ
2u2
t2,
σ
5
x1
+
σ
4
x2
+
σ
3
x3
+p3=
ρ
2u3
t2
dynamic equations (2.2.2)
The inertial terms in (2.2.2) are dynamic body forces per unit volume. The den-
sity
ρ
for unidirectional laminae can be calculated e.g. using the rule of mixtures
(Sect. 3.1.1).
The kinematic equations that are six strain-displacement relations and six com-
patibility conditions for strains. For linear small deformation theory, the six
stress-displacement equations couple six unknown strains and three unknown
displacements. Figure 2.12 shows the strains of an infinitesimal cube in the
(x1x2)-plane and we find the relations
ε
1
u1
x1
,
ε
2
u2
x2
,
α
=
u2
x1
,
β
=
u1
x22
ε
12
γ
12
u1
x2
+
u2
x1
ε
6
and analogous relations for the (x2x3)- and (x1x3)-planes yield
2.2 Fundamental Equations and Variational Solution Procedures 61
x1
x1
x2
x2
u1u1+
u1
x1
dx1
u2
u2+
u2
x2
dx2
dx1
dx2
A
A
A
A
C
C
C
C
BB
B
B
u1
x2
dx2
u2
x1
dx1
a b
α
β
γ
Fig. 2.12 Strains of the infinitesimal cube shown for the (x1x2)-plane. aextensional strains,
bshear strains
ε
1=
u1
x1
,
ε
2=
u2
x2
,
ε
3=
u3
x3
,
ε
4=
u3
x2
+
u2
x3
,
ε
5=
u3
x1
+
u1
x3
,
ε
6=
u2
x1
+
u1
x2
(2.2.3)
The displacement field of the body corresponding to a given deformation state is
unique, the components of the strain tensor must satisfy the following six com-
patibility conditions
2
ε
1
x2
2
+
2
ε
2
x2
1
=
2
ε
6
x1
x2
,
x3
ε
4
x1
+
ε
5
x2
ε
6
x3=2
2
ε
3
x1
x2
,
2
ε
2
x2
3
+
2
ε
3
x2
2
=
2
ε
4
x2
x3
,
x1
ε
5
x2
+
ε
6
x3
ε
4
x1=2
2
ε
1
x2
x3
,
2
ε
3
x2
1
+
2
ε
1
x2
3
=
2
ε
5
x1
x3
,
x2
ε
6
x3
+
ε
4
x1
ε
5
x2=2
2
ε
2
x3
x1
In the two-dimensional case the compatibility conditions reduce to a single equa-
tion
2
ε
1
x2
2
+
2
ε
2
x2
12
2
ε
6
x1
x2
=0
The material or constitutive equations which are described in Sect. 2.1 are
ε
i=Si j
σ
j,
σ
i=Ci j
ε
j(2.2.4)
The generalized Hooke’s law yields six equations relating in each case six un-
known stress and strain components. The elements of the stiffness matrix C
C
Cand
the compliance matrix S
S
Sare substituted with respect to the symmetry conditions
of the material.
62 2 Linear Anisotropic Materials
Summarizing all equations, we have 15 independent equations for 15 unknown com-
ponents of stresses, strains and displacements. In the displacement approach, the
stresses and strains are eliminated and a system of three coupled partial differential
equations for the displacement components are left.
In the static case we have a boundary-value problem, and we have to include
boundary conditions. In the dynamic case the system of partial differential equations
defines an initial-boundary-value problem and we have additional initial conditions.
A clear symbolic formulation of the fundamental equations in displacements is
given in vector-matrix notation. With the transposed vectors
σ
σ
σ
T,
ε
ε
ε
Tand u
u
uT
σ
σ
σ
T= [
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6],
ε
ε
ε
T= [
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6],u
u
uT= [u1u2u3](2.2.5)
the transformation and the differential matrices T
T
Tand D
D
D(n
n
nis the surface normal
unit vector)
T
T
T=
n10 0
0n20
0 0 n3
0n3n2
n30n1
n2n10
,D
D
D=
10 0
0
20
0 0
3
0
3
2
30
1
2
10
,
ni=cos(n
n
n,xi)
i=
xi
i=1,2,3
(2.2.6)
and the stiffness m atrix C
C
Cwe get
D
D
DT
σ
σ
σ
+p
p
p=0
0
0Vstatic equilibrium equations,
D
D
DT
σ
σ
σ
+p
p
p=
ρ
2u
u
u
t2Vdynamic equilibrium equations,
T
T
TT
σ
σ
σ
=q
q
qAqsurface equilibrium equations,
ε
ε
ε
=D
D
Du
u
uVkinematic equations,
σ
σ
σ
=C
C
C
ε
ε
ε
constitutive equations
(2.2.7)
Vis the volume and Aqthe surface of the body with surface forces q
q
q.
Eliminating the stresses and the strains leads to the differential equations for the
displacements
Boundary-value problem - elastostatics
D
D
DTC
C
CD
D
Du
u
u=p
p
pVequilibrium for the volume element V,
u
u
u=u
u
uAuprescribed displacements u
u
uon Au,
T
T
TTC
C
CD
D
Du
u
u=q
q
qAqprescribed surface forces q
q
qon Aq
(2.2.8)
Initial boundary-value problem - elastodynamics
D
D
DTC
C
CD
D
Du
u
u
ρ
¨
u
u
u=p
p
pVequilibrium equation,
u
u
u=u
u
uAu,T
T
TTC
C
CD
D
Du
u
u=q
q
qAqboundary conditions,
u
u
u(x
x
x,0) = u
u
u(x
x
x,0),˙
u
u
u(x
x
x,0) = ˙
u
u
u(x
x
x,0)initial conditions
(2.2.9)
In the general case of material anisotropic behavior the three-dimensional equa-
tions are very complicated and analytical solutions are only possible for some spe-
2.2 Fundamental Equations and Variational Solution Procedures 63
cial problems. This is independent of the approach to displacements or stresses.
Some elementary examples are formulated in Sect. 2.2.4. The equations for beams
and plates are simplified with additional kinematic and/or static hypotheses and the
equations are deduced separately in Chaps. 7 and 8. The simplified structural equa-
tions for circular cylindrical shells and thin-walled folded structures are given in
Chaps. 9 and 10.
Summarizing the fundamental equations of elasticity we have introduced stresses
and displacements as static and kinematic field variables. A field is said to be stati-
cally admissible if the stresses satisfy equilibrium equations (2.2.1) and are in equi-
librium with the surface traction ¯
q
q
qon the body surface Aq, where the traction are
given. A field is referred to as kinematically admissible if displacements and strains
are restricted by the strain-displacement equations (2.2.3) and the displacement sat-
isfies kinematic boundary conditions on the body surface Au, where the displace-
ments are prescribed. Admissible field variables are considered in principles of vir-
tual work and energy formulations, Sect. 2.2.2. The mutual correspondence between
static and kinematic field variables is established through the constitutive equations
(2.2.4).
2.2.2 Principle of Virtual Work and Energy Formulations
The analytical description of the model equations of anisotropic elasticity may re-
alized as above by a system of partial differential equations but also by integral
statements which are equivalent to the governing equations of Sect. 2.2.1 and based
on energy or variational formulations. The utility of variational formulations is in
general twofold. They yield convenient methods for the derivation of the governing
equations of problems in applied elasticity and provide the mathematical basis for
consistent approximate theories and solution procedures. There are three variational
principles which are used mostly in structural mechanics. There are the principles of
virtual work, the principle of complementary virtual work, Reissner’s6variational
theorem and the related energy principles.
Restricting ourselves to static problems, extremal principles formulated for the
total elastic potential energy of the problem or the complementary potential energy
are very useful in the theory of elasticity and in modelling and analysis of structural
elements. The fundamental equations and boundary conditions given beforehand
can be derived with the extremal principles and approximate solutions are obtained
by direct variational methods. Both extremal principles follow from the principle of
virtual work.
If an elastic body is in equilibrium, the virtual work
δ
Wof all acting forces in
moving through a virtual displacement
δ
u
u
uis zero
δ
W
δ
Wa+
δ
Wi=0 (2.2.10)
6Eric (Max Erich) Reissner (January 5th, 1913 in Aachen; November 1st, 1996 La Jolla, USA)
- engineer, contributions to the theory of beams, plates and shells
64 2 Linear Anisotropic Materials
δ
Wis the total virtual work,
δ
Wathe external virtual work of body or volume and
surface forces and
δ
Withe internal virtual work of internal stresses, for the forces
associated with the stress field of a body move the body points through virtual dis-
placements
δ
u
u
ucorresponding to the virtual strain field
δε
ε
ε
.
A displacement is called virtual, if it is infinitesimal, and satisfies the geometric
constraints (compatibility with the displacement-strain equations and the boundary
conditions) and all forces are fixed at their values. These displacements are called
virtual because they are assumed to be infinitesimal while time is held constant. The
symbol
δ
is called a variational operator and in the mathematical view a virtual dis-
placement is a variation of the displacement function. To use variational operations
in structural mechanics only the following operations of the
δ
-operator are needed
δ
dnf
dxn=dn
dxn(
δ
f),
δ
(fn) = n f n1
δ
fn1,
δ
Zfdx=Z
δ
fdx
For a deformable body, the external and the internal work are given in Eqs.
(2.2.11) and (2.2.12), respectively,
δ
Wa=Z
V
pk
δ
ukdV+Z
Aq
qk
δ
ukdA,(2.2.11)
δ
Wi=Z
V
σ
k
δε
kdV(2.2.12)
pkare the components of the actual body force vector p
p
pper unit volume and qk
the components of the actual surface force vector q
q
q(surface traction per unit area).
Aqdenotes the portion of the boundary on which surface forces are specified.
σ
k
and
ε
kare the components of the stress and the strain vector. The negative sign in
(2.2.12) indicates that the inner forces oppose the inner virtual displacements, e.g.
if the virtual displacement
δ
u1=
δε
1dx1is subjected an inner force (
σ
1dx2dx3)the
inner work is (
σ
1dx2dx3)
δε
1dx1. The vectors p
p
p,q
q
qand u
u
uhave three components
but the vectors
σ
σ
σ
and
ε
ε
ε
have six components. The double subscript kin pk
δ
ukand
qk
δ
ukmeans the summarizing on 1 to 3 but in
σ
k
δε
kon 1 to 6.
The general formulation of the principle of virtual work for a deformable body
δ
Wa+
δ
Wi
δ
W=0
or Z
V
pk
δ
ukdV+Z
Aq
qk
δ
ukdAZ
V
σ
k
δε
kdV=0 (2.2.13)
is independent of the constitutive equations. For the three-dimensional boundary-
value problem of a deformable body the principle can be formulated as follow:
Theorem 2.1 (Principle of virtual work). The sum of virtual work done by the
internal and external forces in arbitrary virtual displacements satisfying the pre-
2.2 Fundamental Equations and Variational Solution Procedures 65
scribed geometrical constraints and the strain-displacement relations is zero, i.e.
the arbitrary field variables
δ
ukare kinematically admissible.
An important case is restricted to linear elastic anisotropic bodies and is known as
the principle of minimum total potential energy. The external virtual work
δ
Wais
stored as virtual strain energy
δ
Wf=
δ
Wi, i.e. there exists a strain energy density
function
Wf(
ε
ε
ε
) = 1
2
σ
k
ε
k=1
2Ckl
ε
k
ε
l
Assuming conservative elasto-static problems, the principle of virtual work takes
the form
δΠ
(u
u
u)
δΠ
a(u
u
u) +
δΠ
i(
ε
ε
ε
)0,
ε
ε
ε
=
ε
ε
ε
(u
u
u)(2.2.14)
with the total potential energy function
Π
(u
u
u)of the elastic body.
Π
a(u
u
u)and
Π
i(
ε
ε
ε
)
are the potential functions of the external and the internal forces, respectively,
Π
i=
Π
(
ε
k) = 1
2Z
V
Ckl
ε
k
ε
ldV,
Π
a=
Π
a(uk) = Z
V
pkukdVZ
Aq
qkukdA
(2.2.15)
The principle of minimum total potential energy may be stated for linear elastic
bodies with the constraints
σ
σ
σ
=C
C
C
ε
ε
ε
(u
u
u)as follows:
Theorem 2.2 (Principle of minimum of the total potential energy). Of all the ad-
missible displacement functions satisfying strain-stress relations and the prescribed
boundary conditions, those that satisfy the equilibrium equations make the total po-
tential energy an absolute minimum.
The Euler7-Lagrange8equations of the variational problem yield the equilibrium
and mechanical boundary conditions of the problem. The minimum total potential
energy is widely used in solutions to problems of structural mechanics.
The principle of virtual work can be formulated in a complementary statement.
Then virtual forces are introduced and the displacements are fixed. In analogy to
(2.2.13) we have the principle of complementary virtual work as
δ
W
a+
δ
W
i
δ
W=0
with the complimentary external and internal virtual works
δ
W
a=Z
Au
uk
δ
qkdA,
δ
W
i=Z
V
ε
k
δσ
kdV(2.2.16)
7Leonhard Euler (15 April 1707 Basel - †7jul./18greg.September 1783 St. Petersburg) - mathe-
matician, physicist, astronomer, logician and engineer, introducing beam theory elements and the
equations of motion
8Joseph-Louis Lagrange, born Giuseppe Lodovico Lagrangia or Giuseppe Ludovico De la Grange
Tournier, also reported as Giuseppe Luigi Lagrange or Lagrangia (25 January 1736 Turin - †10
April 1813 Paris) - mathematician and astronomer, variational calculations, general mechanics
66 2 Linear Anisotropic Materials
Audenotes the portion of the boundary surface on which displacements are speci-
fied.
With the complementary stress energy density function
W
f(
σ
σ
σ
) = 1
2Skl
σ
k
σ
l,
δ
W
f(
σ
σ
σ
) = Skl
σ
l
δσ
k=
ε
k
δσ
k
and assuming conservative elasto-static problems, the principle of complementary
work can be formulated as principle of minimum total complementary energy
δΠ
=
δΠ
i+
δΠ
a
or
δ
Z
V
W(
σ
k)dVZ
Au
ukqkdA
=0 (2.2.17)
The principle of minimum total complementary energy may be stated for linear
elastic bodies with constraints
ε
ε
ε
=S
S
S
σ
σ
σ
as follows:
Theorem 2.3 (Principle of minimum total complementary energy). Of all admis-
sible stress states satisfying equilibrium equations and stress boundary conditions,
those which are kinematically admissible make the total complementary energy an
absolute minimum.
The Euler-Lagrange equations of the variational statement yield now the compati-
bility equations and the geometrical boundary conditions.
The both well-known principles of structure mechanics, the principle of vir-
tual displacements (displacement method) and the theorem of Castigliano (principle
of virtual forces, force method) correspond to the principle of minimum potential
energy and complementary energy. The principle of minimum potential energy is
much more used in solution procedures, because it is usually far easier to formulate
assumptions about functions to represent admissible displacements as to formulate
admissible stress functions that ensure stresses satisfying mechanical boundary con-
ditions and equilibrium equations. It should be kept in mind that from the two prin-
ciples considered above no approximate theory can be obtained in its entirety. One
must either satisfy the strain-displacement relations and the displacement bound-
ary conditions exactly and get approximate equilibrium conditions or vice versa.
Both principles yield the risk to formulate approximate theories or solution proce-
dures which may be mathematically inconsistent. Reissner’s variational statement
yields as Euler-Lagrange equations both, the equilibrium equations and the strain-
displacement relations, and has the advantage that its use would yield approximate
theories and solution procedures which satisfy both requirements to the same de-
gree and would be consistent. Reissner’s variational theorem (Reissner, 1950) can
be formulated as follows:
Theorem 2.4 (Reissner’s variational theorem, 1950). Of all sets of stress and
displacement functions of an elastic body
ε
ε
ε
=C
C
C
σ
σ
σ
which satisfy the boundary
2.2 Fundamental Equations and Variational Solution Procedures 67
conditions, those which also satisfy the equilibrium equations and the stress-
displacements relations correspond to a minimum of the functional
Ψ
Rdefined as
Ψ
R(u
u
u,
σ
σ
σ
) = Z
V
[
σ
k
ε
kWf(
σ
k)]dVZ
V
pkukdVZ
Aq
qkukdA(2.2.18)
Wf(
σ
k)is the strain energy density function in terms of stresses only,
Ψ
Ris Reissner’s
functional.
It should be noted that all stress and strain components must be varied while pk
and qkare prescribed functions and therefore fixed. The variation of the functional
Ψ
R(u
u
u,
σ
σ
σ
)yields
δΨ
R=Z
V
σ
k
δε
k+
ε
k
δσ
k
Wf
σ
k
δσ
kdVZ
V
pk
δ
ukdV
Z
Aq
qk
δ
ukdA,
(2.2.19)
where
ε
kis determined by (2.2.3).
δΨ
R(u
u
u,
σ
σ
σ
)can be rearranged and we obtain finally
δΨ
R=Z
V
ε
ε
ε
Wf
σ
σ
σ
δσ
σ
σ
TD
D
DT
σ
σ
σ
+p
p
p
δ
u
u
uTdVZ
Aq
q
q
q
δ
u
u
uTdA(2.2.20)
Since
δσ
σ
σ
and
δ
u
u
uare arbitrary variations
δΨ
R=0 is satisfied only if
ε
i j =
Wf(
σ
i j )
σ
i j
,
σ
i j
xj
+pi=0 (2.2.21)
Summarizing we have considered two dual energy principles with ukor
σ
kas admis-
sible functions which have to be varied and one generalized variational principle,
where both, ukand
σ
k, have to be varied. The considerations are limited to lin-
ear problems of elasto-statics, i.e. the generalized Hooke’s law describes the stress-
strain relations.
Expanding the considerations on dynamic problems without dissipative forces
following from external or inner damping effects the total virtual work has in the
sense of the d’Alambert principle an additional term which represents the inertial
forces
δ
W=Z
V
ρ
¨uk
δ
ukdVZ
V
σ
k
δε
kdV+Z
V
pk
δ
ukdV +Z
Aq
qk
δ
ukdA(2.2.22)
Equation (2.2.22) represents an extension of the principle of virtual work from stat-
ics to dynamics.
ρ
is the density of the elastic body.
68 2 Linear Anisotropic Materials
For conservative systems of elasto-dynamics, the Hamilton9principle replaces
the principle of the minimum of the total potential energy
δ
t2
Z
t1
(T
Π
)dt
δ
t2
Z
t1
L(uk)dt=0,T=1
2Z
V
ρ
˙uk˙ukdV(2.2.23)
Π
(u
u
u)is the potential energy given beforehand and T(u
u
u)is the so-called kinetic
energy. L=T
Π
is the Lagrangian function.
Theorem 2.5 (Hamilton’s principle for conservative systems). Of all possible
paths between two points at time interval t1and t2along which a dynamical sys-
tem may move, the actual path followed by the system is the one which minimizes
the integral of the Lagrangian function.
In the contracted vector-matrix notation we can summarize:
Conservative elasto-static problems
Π
(u
u
u) = 1
2Z
V
σ
σ
σε
ε
ε
TdVZ
V
p
p
pu
u
uTdVZ
Aq
q
q
qu
u
uTdA,
δΠ
=Z
V
σ
σ
σδε
ε
ε
TdVZ
V
p
p
p
δ
u
u
uTdVZ
Aq
q
q
q
δ
uTdA=0
(2.2.24)
Conservative elasto-dynamic problems
L(u
u
u) = T(u
u
u)
Π
(u
u
u),T(u
u
u) = 1
2Z
V
ρ
˙
u
u
uT˙
u
u
udV,
δ
t2
Z
t1
L(u
u
u)dt=0
(2.2.25)
All variations are related to the displacement vector u
u
u. For the stress and the strain
vector we have to take into consideration that
σ
σ
σ
=
σ
σ
σ
(
ε
ε
ε
) =
σ
σ
σ
[
ε
ε
ε
(u
u
u)] and for the time
integrations
δ
u
u
u(x
x
x,t1) =
δ
u
u
u(x
x
x,t2) = 0 (2.2.26)
For non-conservative systems of elastodynamics, the virtual work
δ
Wincludes an
approximate damping term
Z
V
µ
˙
u
u
uT
δ
u
u
udV
with
µ
as a damping parameter and Eq. (2.2.22) is substituted by
9William Rowan Hamilton (4 August 1805 Dublin - †2 September 1865 Dublin) - physicist,
astronomer, and mathematician, who made important contributions to classical mechanics, optics,
and algebra
2.2 Fundamental Equations and Variational Solution Procedures 69
δ
W=Z
V
(
ρ
¨
u
u
uT+
µ
˙
u
u
uT)
δ
u
u
udVZ
V
δ
(
σ
σ
σ
T
ε
ε
ε
)dVZ
V
p
p
p
δ
u
u
udV
Z
Aq
q
q
q
δ
u
u
udA
(2.2.27)
A generalized Hamilton’s principle in conjunction with the Reissner’s variational
statement can be presented as
δ χ
(u
u
u,
σ
σ
σ
) =
δ
t2
Z
t1
[T(u
u
u)
ψ
R(u
u
u,
σ
σ
σ
)]dt=0,
where T(u
u
u)is the kinetic energy as above,
ψ
R(u
u
u,
σ
σ
σ
)the Reissner’s functional
(2.2.18).
2.2.3 Variational Methods
The variational principles can be used to obtain, in a mathematical way, the govern-
ing differential equations and associated boundary conditions as the Euler-Lagrange
equations of the variational statement. Now we consider the use of the variational
principles in the solution of the model equations. We seek in the sense of the clas-
sical variational methods, approximate solutions by direct methods, i.e. the approx-
imate solution is obtained directly by applying the same variational statement that
are used to derive the fundamental equations.
2.2.3.1 Rayleigh-Ritz Method
Approximate methods are used when exact solutions to a problem cannot be derived.
Among the approximation methods, Ritz10 method is a very convenient method
based on a variational approach. The variational methods of approximation de-
scribed in this textbook are limited to Rayleigh11-Ritz method for elasto-statics and
elasto-dynamics problems of anisotropic elasticity theory and to some extent on
weighted-residual methods.
The Rayleigh-Ritz method is based on variational statements, e.g. the principle of
minimum total potential energy, which is equivalent to the fundamental differential
equations as well as to the so-called natural or static boundary conditions including
10 Walter Ritz (February 22nd, 1878, Sion, Switzerland; July 7th , 1909, ottingen) - theoretical
physicist, variational methods
11 John William Strutt, 3rd Baron Rayleigh (November 12th, 1842, Langford Grove, Maldon,
Essex, UK; June 30th, 1919, Terling Place, Witham, Essex, UK) - physicist, Nobel Prize in
Physics winner (1904), variational method
70 2 Linear Anisotropic Materials
force boundary conditions . This variational formulation is known as the weak form
of the model equations. The method was proposed as the direct method by Rayleigh
and a generalization was given by Ritz.
The starting point for elasto-static problems is the total elastic potential energy
functional
Π
=1
2Z
V
ε
ε
ε
TC
C
C
ε
ε
ε
dVZ
V
p
p
pTu
u
udVZ
Aq
q
q
qTu
u
udA
=1
2Z
V
(D
D
Du
u
u)TC
C
CD
D
Du
u
udVZ
V
p
p
pTu
u
udVZ
Aq
q
q
qTu
u
udA
(2.2.28)
The variations are related to the displacements u
u
uand the strains
ε
ε
ε
which have to be
substituted with help of the differential matrix D
D
D, (2.2.6), by the displacements. The
approximate solution is sought in the form of a finite linear combination. Looking
first at a scalar displacement approach, the approximation of the scalar displacement
function u(x1,x2,x2)is given by the Ritz approximation
˜u(x1,x2,x3) =
N
i=1
ai
ϕ
i(x1,x2,x3)
or
˜u(x1,x2,x3) =
N
i=1
ai
ϕ
i(x1,x2,x3) +
ϕ
0(x1,x2,x3)(2.2.29)
The
ϕ
iare known functions chosen a priori, named approximation functions or co-
ordinate functions. The aidenote undetermined constants named generalized coor-
dinates. The approximation ˜uhas to make (2.2.28) extremal
˜
Π
(˜u) = ˜
Π
(ai),
δ
˜
Π
(ai) = 0 (2.2.30)
This approximation is characterized by a relative extremum. From (2.2.30) comes
˜
Π
in form of a function of the constants aiand
δ
˜
Π
(ai) = 0 yields Nstationary
conditions
˜
Π
(ai)
ai
=0,i=1,2,...,N(2.2.31)
˜
Π
may be written as a quadratic form in aiand from Eqs. (2.2.31) follows a system
of Nlinear equations allowing the Nunknown constants aito be determined. In
order to ensure a solution of the system of linear equations and a convergence of the
approximate solution to the true solution as the number Nof the aiis increased, the
ϕ
ivalues have to fulfill the following requirements:
ϕ
0satisfies specified inhomogeneous geometric boundary conditions, the so-
called essential conditions of the variational statement and
ϕ
i,i=1,2,...,Nsat-
isfy the homogeneous form of the geometric boundary conditions.
ϕ
iare continuous as required in the variational formulation, e.g. they should have
a non-zero contribution to ˜
Π
.
ϕ
iare linearly independent and complete.
2.2 Fundamental Equations and Variational Solution Procedures 71
The completeness property is essential for the convergence of the Ritz approxima-
tion. Polynomial and trigonometric functions are selected examples of complete
systems of functions.
Generalizing the considerations to three-dimensional problems and using vector-
matrix notation it follows
˜
u
u
u(x1,x2,x3)
˜u1
˜u2
˜u3
=
a
a
aT
1
ϕ
ϕ
ϕ
1
a
a
aT
2
ϕ
ϕ
ϕ
2
a
a
aT
3
ϕ
ϕ
ϕ
3
ϕ
ϕ
ϕ
1o
o
o o
o
o
o
o
o
ϕ
ϕ
ϕ
2o
o
o
o
o
o o
o
o
ϕ
ϕ
ϕ
3
T
a
a
a1
a
a
a2
a
a
a3
(2.2.32)
or
˜
u
u
u(x1,x2,x3) = G
G
GTa
a
a(2.2.33)
with
G
G
GT=
ϕ
ϕ
ϕ
1o
o
o o
o
o
o
o
o
ϕ
ϕ
ϕ
2o
o
o
o
o
o o
o
o
ϕ
ϕ
ϕ
3
T
=
ϕ
ϕ
ϕ
T
1o
o
oTo
o
oT
o
o
oT
ϕ
ϕ
ϕ
T
2o
o
oT
o
o
oTo
o
oT
ϕ
ϕ
ϕ
T
3
,a
a
a=
a
a
a1
a
a
a2
a
a
a3
G
G
Gis the matrix of the approximation functions,
ϕ
ϕ
ϕ
iand o
o
oare N-dimensional vectors
and a
a
aiare N-dimensional subvectors of the vector a
a
aof the unknown coordinates.
The application of the Ritz method using the minimum principle of elastic potential
energy
Π
has the following steps:
1. Choose the approximation function ˜
u
u
u=G
G
GTa
a
a.
2. Substitute ˜
u
u
uinto
Π
˜
Π
(˜
u
u
u) = 1
2Z
V
(D
D
D˜
u
u
u)TC
C
CD
D
D˜
u
u
udVZ
V
p
p
pT˜
u
u
udVZ
Aq
q
q
qT˜
u
u
udA
=1
2a
a
aTK
K
Ka
a
aa
a
aTf
f
f
(2.2.34)
with
K
K
K=Z
V
(D
D
DG
G
G)TC
C
C(D
D
DG
G
G)dV=Z
V
B
B
BTC
C
CB
B
BdV,
f
f
f=Z
V
G
G
GTp
p
pdV+Z
Aq
G
G
GTq
q
qdA
3. Formulate the stationary conditions of ˜
Π
(a
a
a)
˜
Π
(a
a
a)
a
a
a=o
o
o
i.e. with
a
a
a(a
a
aTK
K
Ka
a
a) = 2K
K
Ka
a
a,
a
a
a(a
a
aTf
f
f) = a
a
a
follows
72 2 Linear Anisotropic Materials
K
K
Ka
a
a=f
f
f(2.2.35)
K
K
Kis called the stiffness matrix, a
a
athe vector of unknowns and f
f
fthe force vector.
These notations are used in a generalized sense.
4. Solve the system of linear equations K
K
Ka
a
a=f
f
f. The vector a
a
aof unknown coeffi-
cients is known.
5. Calculate the approximation solution ˜
u
u
u=a
a
aT
ϕ
ϕ
ϕ
and the ˜
ε
ε
ε
=D
D
D˜
u
u
u,˜
σ
σ
σ
=C
C
C˜
ε
ε
ε
,...
For an increasing number N, the previously computed coefficients of a
a
aremain un-
changed provided the previously chosen coordinate functions are not changed. Since
the strains are calculated from approximate displacements, strains and stresses are
less accurate than displacements.
The Ritz approximation of elasto-dynamic problems is carried out in an analo-
gous manner and can be summarized as follows. For conservative problems we start
with the variational statement (2.2.23). The displacement vector u
u
uis now a function
of x
x
xand tand the a
a
a-vector a function of t. The stationary condition yields
a
a
a1
2a
a
aT(t)K
K
Ka
a
a(t)a
a
aT(t)f
f
f(t) + ¨
a
a
aT(t)M
M
Ma
a
a(t)=0
0
0 (2.2.36)
M
M
M¨
a
a
a(t) + K
K
Ka
a
a(t) = f
f
f(t),(2.2.37)
M
M
M=Z
V
ρ
G
G
GTG
G
GdV,
K
K
K=Z
V
(D
D
DG
G
G)TC
C
C(D
D
DG
G
G)dV,
f
f
f=Z
V
G
G
GTp
p
pdV+Z
Aq
G
G
GTq
q
qdA
The matrix G
G
Gdepends on x
x
x,p
p
pand q
q
qon x
x
xand t.M
M
Mis called the mass matrix.
An direct derivation of a damping matrix from the Ritz approximation analogous
to the K
K
K- and the M
M
M-matrix of (2.2.37) is not possible. In most engineering applica-
tions (2.2.37) has an additional damping term and the damping matrix is formulated
approximately as a linear combination of mass- and stiffness-matrix (modal damp-
ing)
M
M
M¨
a
a
a(t) +C
C
CD˙
a
a
a(t) + K
K
Ka
a
a(t) = f
f
f(t),C
C
CD
α
M
M
M+
β
K
K
K(2.2.38)
In the case of the study of free vibrations, we write the time dependence of a
a
a(t)in
the form
a
a
a(t) = ˆ
a
a
acos(
ω
t+
ϕ
)(2.2.39)
and from (2.2.38) with C
C
CD=0
0
0,f
f
f(t) = 0
0
0 comes the matrix eigenvalue problem (App.
A.3)
(K
K
K
ω
2M
M
M)ˆ
a
a
a=o
o
o,det[K
K
K
ω
2M
M
M] = 0 (2.2.40)
For Ncoordinate functions the algebraic equation (2.2.40) yields Neigenfrequen-
cies of the deformable body.
2.2 Fundamental Equations and Variational Solution Procedures 73
The Rayleigh-Ritz method approximates the continuous deformable body by a
finite number of degree of freedoms, i.e. the approximated system is less flexible
than the actual body. Consequently for the approximated energy potential ˜
Π
Π
.
The energy potential converges from below. The approximate displacements satisfy
the equilibrium equations only in the energy sense and not pointwise, unless the
solution converges to the exact solution. The Rayleigh-Ritz method can be applied
to all mechanical problems since a virtual statement exists, i.e. a weak form of the
model equations including the natural boundary conditions. If the displacements
are approximate, the approximate eigenfrequencies are higher than the exact, i.e.
˜
ω
ω
.
2.2.3.2 Weighted Residual Methods
Finally some brief remarks on weighted residual methods are given. The fundamen-
tal equation in the displacement approach may be formulated in the form
A(u) = f(2.2.41)
Ais a differential operator. We seek again an approximate solution (2.2.29), where
now the constants aiare determined by requiring the residual
RN=A N
i=1
ai
ϕ
i+
ϕ
0!f6=0 (2.2.42)
be orthogonal to Nlinear independent weight function
ψ
i
Z
V
RN
ψ
idx
x
x=0,i=1,2,...,N(2.2.43)
ϕ
0,
ϕ
ishould be linear independent and complete and fulfill all boundary conditions.
Various known special methods follow from (2.2.43). They differ from each other
due to the choice of the weight functions
ψ
i:
Galerkin’s method12
ψ
i
ϕ
i,
Least-squares method
ψ
iA(
ϕ
i),
Collocation method
ψ
i
δ
(x
x
xx
x
xi)(
δ
(x
x
xx
x
xi) = 1 if x
x
x=x
x
xiotherwise 0)
The Galerkin method is a generalization of the Ritz method, if it is not possible to
construct a weak form statement. Otherwise the Galerkin and the the Ritz method for
weak formulations of problems yield the same solution equations, if the coordinate
functions
ϕ
iin both are the same.
12 Boris Grigorjewitsch Galerkin, surname more accurately romanized as Galyorkin (4
Marchgreg./20 Februaryjul.1871 Polozk - 12 July 1945 Leningrad) - mathematician and engineer,
contributions to the theory of approximate solutions of partial differential equations
74 2 Linear Anisotropic Materials
The classical variational methods of Ritz and Galerkin are widely used to solve
problems of applied elasticity or structural mechanics. When applying the Ritz or
Galerkin method to special problems involving, e.g. a two-dimensional functional
Π
[u(x1,x2)] or a two-dimensional differential equation A[u(x1,x2)] = f(x1,x2), an
approximative solution is usually assumed in the form
˜u(x1,x2) =
N
i=1
ai
ϕ
i(x1,x2)or ˜u(x1,x2) =
N
i=1
M
j=1
ai j
ϕ
1i(x1)
ϕ
2j(x2),(2.2.44)
where
ϕ
i(x1,x2)or
ϕ
1i(x1),
ϕ
2j(x2)are a priori chosen trial functions and the aior
ai j are unknown constants. The approximate solution depends very strongly on the
assumed trial functions.
To overcome the shortcoming of these solution methods Vlasov13 and Kan-
torovich14 suggested an approximate solution in the form
˜u(x1,x2) =
N
i=1
ai(x1)
ϕ
i(x1,x2)(2.2.45)
The
ϕ
iare again a priori chosen trial functions but the ai(x1)are unknown coefficient
functions of one of the independent variables. The condition
δΠ
[˜u(x1,x2)] = 0 or
with dA=dx1dx2Z
A
RN(˜u)
ϕ
idA=0,i=1,2,...,N
lead to a system of Nordinary differential equations for the unknown functions
ai(x1). Generally it is advisable to choose if possible the trial functions
ϕ
ias func-
tions of one independent variable, i.e.
ϕ
i=
ϕ
i(x2), since otherwise the system of
ordinary differential equations will have variable coefficients. The approximate so-
lution ˜u(x1,x2)tends in regard of the arbitrariness of the assumed trial function
ϕ
i(x2)to a better solution in the x1-direction. The obtained approximative solution
can be further improved in the x2-direction in the following manner. In a first step
the assumed approximation
˜u(x1,x2) =
N
i=1
ai(x1)
ϕ
i(x2)(2.2.46)
yields the functions ai(x1)by solving the resulting set of ordinary differential
equations with constant coefficients. In the next step, with ai(x1)ai1(x1)and
ϕ
i(x2)ai2(x2), i.e.
˜u[I](x1,x2) =
N
i=1
ai2(x2)ai1(x1),
13 Vasily Zakharovich Vlasov (11greg./24j ul.February 1906 Kareevo - †7 August 1958 Moscow)
- civil engineer
14 Leonid Vitaliyevich Kantorovich (19 January 1912 St. Petersburg - †7 April 1986 Moscow) -
mathematician and economist, Nobel prize winner in economics in 1975
2.2 Fundamental Equations and Variational Solution Procedures 75
the ai1(x1)are the given trial functions and the unknown functions ai2(x2)can be
determined as before the ai1(x1)by solving a set of ordinary differential equations.
After completing the first cycle, which yields ˜u[I](x1,x2), the procedure can be con-
tinued iteratively. This iterative solution procedure is denoted in literature as vari-
ational iteration or extended Vlasov-Kantorowich method. The final form of the
generated solution is independent of the initial choice of the trial function
ϕ
i(x2)
and the iterative procedure converges very rapidly. It can be demonstrated, that the
iterative generated solutions ˜u(x1,x2)agree very closely with the exact analytical
solutions u(x1,x2)even with a single term approximation
˜u(x1,x2) = a1(x1)
ϕ
1(x2)(2.2.47)
In engineering applications, e.g. for rectangular plates, the single term approxima-
tions yield in general sufficient accuracy.
Summarizing it should be said that the most difficult problem in the application
of the classical variational methods or weighted residual methods is the selection of
coordinate functions, especially for structures with irregular domains. The limita-
tions of the classical variational methods can be overcome by numerical methods,
e.g. the finite element method which is discussed in more detail in Chap. 11.
2.2.4 Problems
Exercise 2.7. An anisotropic body is subjected to a hydrostatic pressure
σ
1=
σ
2=
σ
3=p,
σ
4=
σ
5=
σ
6=0.
1. Calculate the strain state
ε
ε
ε
.
2. Calculate the stress state
σ
σ
σ
for a change of the coordinate system obtained by a
rotation T
T
T
σ
.
Solution 2.7. The solution can be presented in two parts:
1. The generalized Hooke’s law yields (Eq. 2.1.20)
ε
1=(S11 +S12 +S13)p,
ε
4=(S14 +S24 +S34)p,
ε
2=(S12 +S22 +S23)p,
ε
5=(S15 +S25 +S35)p,
ε
3=(S13 +S23 +S33)p,
ε
6=(S16 +S26 +S36)p
Note 2.3. A hydrostatic pressure in an anisotropic solid yields extensional and
shear strains.
2. From (2.1.39) follows
σ
1=p(c2+s2) = p,
σ
4=0,
σ
2=p(s2+c2) = p,
σ
5=0,
σ
3=p,
σ
6=p(cs +cs) = 0
76 2 Linear Anisotropic Materials
Exercise 2.8. An anisotropic body has a pure shear stress state
σ
1=
σ
2=
σ
3=
σ
5=
σ
6=0,
σ
4=t.
1. Calculate the strain state
ε
ε
ε
.
2. Compare the strain state for the anisotropic case with the isotropic case.
Solution 2.8. The solution can be presented again in two parts:
1. Equations (2.1.20) yield
ε
1=S14t,
ε
2=S24t,
ε
3=S34t,
ε
4=S44t,
ε
5=S45t,
ε
6=S46t
The anisotropic body has extensional and shear strains in all coordinate planes.
2. In an isotropic body a pure shear stress state yields only shearing strains:
S14 =S24 =S34 =0
Exercise 2.9. Consider a prismatic homogeneous anisotropic bar which is fixed at
one end. The origin of the coordinates x1,x2,x3is placed in the centroid of the
fixed section and the x3-axis is directed along the bar axis, land Aare the length
and the cross-section of the undeformed bar. Assume that the bar at the point
x1=x2=x3=0 has no displacement and torsion:
u1=u2=u3=0,u1,3=u2,3=u2,1u1,2=0
1. A force Facts on the bar on the cross-section x3=land the stress state is de-
termined by
σ
1=
σ
2=
σ
4=
σ
5=
σ
6=0,
σ
3=F/A. Determine the strains, the
displacements and the extension of the axis.
2. The fixed bar is deformed only under its own weight: p1=p2=0, p3=g
ρ
.
Determine the strain state and the displacements and calculate the displacements
in the point (0,0,l).
Solution 2.9. Now one has
1. The generalized Hooke’s law (2.1.20) with
σ
1=
σ
2=
σ
4=
σ
5=
σ
6=0,
σ
3=F/A=
σ
6=0,
gives
ε
1=S13
σ
,
ε
2=S23
σ
,
ε
3=S33
σ
,
ε
4=S34
σ
,
ε
5=S35
σ
,
ε
6=S36
σ
The displacements can be determined by the introduction of the following equa-
tions
S13
σ
=u1,1,S23
σ
=u2,2,S33
σ
=u3,3,
S34
σ
=u3,2+u2,3,S35
σ
=u3,1+u1,3,S36
σ
=u2,1+u1,2
The equations
2.2 Fundamental Equations and Variational Solution Procedures 77
u1=
σ
(S13x1+0.5S36x2),
u2=
σ
(0.5S36x1+S23x2),
u3=
σ
(S35x1+S34x2+S33 x3)
satisfy the displacement differential equations and the boundary conditions
which are prescribed at the point x1=x2=x3=0.
Note 2.4. The stress-strain formulae show that an anisotropic tension bar does not
only lengthens in the force direction x3and contracts in the transverse directions,
but also undergoes shears in all planes parallel to the coordinate planes. The
cross-sections of the bar remain plane. The stress states of an isotropic or an
anisotropic bar are identical, the anisotropy effects the strain state only.
2. The stress in a bar under its own weight is
σ
1=
σ
2=
σ
4=
σ
5=
σ
6=0,
σ
3=
ρ
g(lx3)
The stress-strain displacement equations are
ε
1=u1,1=S13
ρ
g(lx3),
ε
4=u3,2+u2,3=S34
ρ
g(lx3),
ε
2=u2,2=S23
ρ
g(lx3),
ε
5=u3,1+u1,3=S35
ρ
g(lx3),
ε
3=u3,3=S33
ρ
g(lx3),
ε
6=u2,1+u1,2=S36
ρ
g(lx3)
The boundary conditions are identical to case a) and the following displacement
state which satisfies all displacement differential equations and the conditions at
the point x1=x2=x3=0 can be calculated by integration
u1=
ρ
g[0.5S35x2
3+S13x1(lx3) + 0.5S36x2(lx3)],
u2=
ρ
g[0.5S34x2
3+S23x2(lx3) + 0.5S36x1(lx3)],
u3=
ρ
g[0.5S13x2
1+0.5S23x2
2
+0.5S36x1x2+ (S34x2+S35 x1)l+0.5S33x3(2lx3)]
Note 2.5. The cross-section does not remain plane, it is deformed to the shape
of a second-order surface and the bar axis becomes curved. The centroid of the
cross-section x3=lis displaced in all three directions
u1(0,0,l) = 0.5
ρ
gS35l2,
u2(0,0,l) = 0.5
ρ
gS34l2,
u3(0,0,l) = 0.5
ρ
gS33l2
Exercise 2.10. Show that for a composite beam subjected to a distributed continu-
ous load q(x1)the differential equation and the boundary conditions can be derived
using the extremal principle of potential energy.
Solution 2.10. The beam is of the length l, width band height h.q(x1)is the lateral
load per unit length. The average elasticity modulus of the beam is E1. With the
stress-strain-displacement relations from Bernoulli15 beam theory
15 Jakob I. Bernoulli (27 December 1654jul./6 January 1655greg.Basel - †16 August 1705 Basel)
- mathematician, beam theory
78 2 Linear Anisotropic Materials
σ
1=E1
ε
1,
ε
1=x3
d2u3
dx2
1
the strain energy function Wfis seen to be
Wf=1
2
σ
1
ε
1=1
2E1
ε
2
1=1
2E1x2
3d2u3
dx2
12
and
Π
i=
l
Z
0
b/2
Z
b/2
h/2
Z
h/2
E1
2d2u3
dx2
12
x2
3dx3dx2dx1=E1I
2
l
Z
0d2u3
dx2
12
dx1
with inertial moment I=bh3/12 for a rectangular cross-section.
In the absence of body forces, the potential function
Π
aof the external load q(x1)
is
Π
a=
l
Z
0
q(x1)u3(x1)dx1
and the total elastic potential energy is
Π
(u3) = E1I
2
l
Z
0d2u3
dx2
12
dx1
l
Z
0
q(x1)u3(x1)dx1
From the stationary condition
δΠ
(u3) = 0 it follows that
δΠ
(u3) = E1I
2
l
Z
0
δ
d2u3
dx2
12
dx1
l
Z
0
q(x1)
δ
u3(x1)dx1=0
There is no variation of E1Ior q(x1), because they are specified. From the last equa-
tion one gets
δΠ
(u3) = E1I
2
l
Z
0
2d2u3
dx2
1
δ
d2u3
dx2
12
dx1
l
Z
0
q(x1)
δ
u3(x1)dx1=0
The first term can be integrated by parts. The first integration can be performed with
the following substitution
u=d2u3
dx2
1
,u=d3u3
dx3
1
,v=
δ
d2u3
dx2
1,v=
δ
du3
dx1
and yields
2.2 Fundamental Equations and Variational Solution Procedures 79
δΠ
(u3) = E1I
d2u3
dx2
1
δ
du3
dx1l
0
l
Z
0
d3u3
dx3
1
δ
du3
dx1dx1
l
Z
0
q
δ
u3dx1=0
The second integration can be performed with the following substitution
u=d3u3
dx3
1
,u=d4u3
dx4
1
,v=
δ
du3
dx1,v=
δ
u3
and yields
δΠ
(u3) = E1Id2u3
dx2
1
δ
du3
dx1l
0E1Id3u3
dx3
1
δ
u3l
0
+
l
Z
0
E1Id4u3
dx4
1
dx1
l
Z
0
q(x1)
δ
u3dx1=0
Finally we obtain
δΠ
(u3) = E1Id2u3
dx2
1
δ
du3
dx1l
0E1Id3u3
dx3
1
δ
u3l
0
+
l
Z
0E1Id4u3
dx4
1q(x1)
δ
u3dx1=0
Since the variations are arbitrary the equation is satisfied if
E1I(u3)′′′′ =q
and either
E1I(u3)′′ =0
or u
3is specified and
E1I(u3)′′′ =0
or u3is specified at x1=0 and x1=l.
The beam differential equation is the Euler-Lagrange equation of the variational
statement
δΠ
=0, u3,u
3represent essential boundary conditions, and E1Iu′′
3,E1Iu′′′
3
are natural boundary conditions of the problem. Note that the boundary conditions
include the classical conditions of simply supports, clamped and free edges.
Exercise 2.11. The beam of Exercise 2.10 may be moderately thick and the ef-
fects of transverse shear deformation
ε
5and transverse normal stress are taken into
account. Show that the differential equations and boundary condition can be de-
rived using the Reissner’s variational principle. The beam material behavior may be
isotropic.
80 2 Linear Anisotropic Materials
Solution 2.11. To apply Reissner’s variational statement one must assume admis-
sible functions for the displacements u1(x1),u3(x1)and for the stresses
σ
1(x1,x3),
σ
3(x1,x3),
σ
5(x1,x3). For the beam with rectangular cross-section and lateral load-
ing q(x1)follow in the frame of beam theory
u2=0,
σ
2=
σ
4=
σ
6=0
As in the classical beam theory (Bernoulli theory) we assume that the beam cross-
sections undergo a translation and a rotation, the cross-sections are assumed to re-
main plane but not normal to the deformed middle surface (Timoshenko16 theory).
Therefore we can assume in the simplest case
u1=x3
ψ
(x1),u3=w(x1)
and the strain-displacement relations may be written
ε
1=
u1
x1
=x3
ψ
(x1),
ε
3=
u3
x3
=0,
ε
5=
u3
x1
+
u1
x3
=w(x1) +
ψ
(x1)
For the stresses
σ
1,
σ
3and
σ
5the following functions are assumed
σ
1=M
Ix3,I=bh3
12 ,
σ
3=3q
4b"x3
h/2+2
31
3x3
h/22#,
σ
5=3Q
2A"1x3
h/22#,A=bh
The assumed functions for
σ
1and
σ
5are identical to those of the Bernoulli beam
theory and
σ
3may be derived from the stress equation of equilibrium in the thick-
ness direction, Eq. (2.2.1), with
σ
3(+h/2) = q,
σ
1(h/2) = 0
The bending moment Mand the shear force resultant Qwill be defined in the usual
manner
M(x1) =
+h/2
Z
h/2
b
σ
1dx3,Q(x1) =
+h/2
Z
h/2
b
σ
5dx3
Now Reissner’s functional
Ψ
R(u
u
u,
σ
σ
σ
), e.g. (2.2.18), takes with the assumption above
the form
16 Stepan Prokopovich Timoshenko (22 December, 1878 Schpotiwka - 29 May, 1972
Wuppertal-Elberfeld) - engineer, founder of the modern applied mechanics
2.2 Fundamental Equations and Variational Solution Procedures 81
Ψ
R(u
u
u,
σ
σ
σ
) =
l
Z
0
+h/2
Z
h/2
σ
1
ε
1+
σ
5
ε
51
2E
σ
2
1+
σ
2
3+2(1+
ν
)
σ
2
5bdx3dx1
l
Z
0
qwdx1
=
l
Z
0
+h/2
Z
h/2
σ
1x3
ψ
(x1) +
σ
5w(x1) +
ψ
(x1)
1
2E
σ
2
1+
σ
2
3+2(1+
ν
)
σ
2
5bdx1
l
Z
0
qwdx1
Substituting Mand Qand neglecting the term
σ
3which only depends on qand not
on the basic unknown functions
ψ
,wrespectively, Mand Q, and yields no contribu-
tion to the variation
δΨ
we obtain
Ψ
(w,
ψ
,M,Q) =
l
Z
0M
ψ
+Q(w+
ψ
)M2
2EI +6
ν
qM
5EA 3Q2
5GA qwdx1
δΨ
R=
l
Z
0M
δψ
+
ψ
δ
M+Q(
δ
w+
δψ
) + (w+
ψ
)
δ
Q
M
EI
δ
M+6
ν
q
5EA
δ
M6Q
5GA
δ
Qq
δ
wdx1=0
Integration the terms M
δψ
and Q
δ
wby parts and rearranging the equation
δΨ
R= [M
δψ
+Q
δ
w]l
0+
l
Z
0n[Q+M]
δψ
[Qq]
δ
w
+
ψ
M
EI +6
ν
q
5EA
δ
M+
ψ
+w6Q
5GA
δ
Qdx1=0
The first term yields the natural boundary conditions of the variational statement:
1. Either M=0 or
ψ
must be prescribed at x1=0,l.
2. Either Q=0 or wmust be prescribed at x1=0,l.
The variations
δψ
,
δ
w,
δ
Mand
δ
Qare all arbitrary independent functions of x1and
therefore
δΨ
R=0 only if
dM(x1)
dx1
+Q(x1) = 0,dQ(x1)
dx1
+q(x1) = 0,
d
ψ
(x1)
dx1M(x1)
EI +6
ν
q(x1)
5EA =0,dw(x1)
dx1
+
ψ
(x1)6Q(x1)
5GA =0
82 2 Linear Anisotropic Materials
The both equations for the stress resultants are identical with the equations of the
classical Bernoulli’s theory. The term (w+
ψ
)in the fourth equation describes the
change in the angle between the beam cross-section and the middle surface during
the deformation. The term (w+
ψ
)is proportional to the average shear stress Q/A
and so a measure of shear deformation. With GA the shear deformation tends
to zero and
ψ
to was assumed in the Bernoulli’s theory. The third term in the
third equation depends on the lateral load qand Poisson’s ratio
ν
and tends to zero
for
ν
0. This term described the effect of the transverse normal stress
σ
3. which
will be vanish if
ν
=0 as in the classical beam theory.
Substituting the differential equations for
ψ
and winto the differential relations
for Mand Qleads to
EI
ψ
′′(x1) +
ν
6
5
Q(x1)
EA Q(x1) = 0,5
6GA[
ψ
(x1) + w′′(x1)] + q(x1) = 0
or
EI
ψ
′′(x1)
ν
6
5
q(x1)
EA 5
6GA[
ψ
+w] = 0,5
6GA[
ψ
(x1) + w′′(x1)] + q(x1) = 0
Derivation and rearrangement yield a differential equation of 3rd order for
ψ
(x1)
EI
ψ
′′′(x1) = q(x1) +
ν
6
5
q(x1)
EA
With
Q(x1) = dM(x1)
dx1
=E I
ψ
′′(x1) +
ν
6
5
Q(x1)
EA
and
Q(x1) = 5
6
EI
GA [
ψ
(x1) + w(x1)]
follows an equation for w(x1)
w(x1) =
ψ
(x1) + 6
5
EI
GA
ψ
′′(x1) +
ν
6
5
q(x1)
EA
Neglecting with
ν
=0 the effect of the transverse normal stress
σ
3we get the Tim-
oshenko’s beam equation
EI
ψ
′′′(x1) = q(x1),M(x1) = E I
ψ
(x1),Q(x1) = EI
ψ
′′(x1)
w(x1) =
ψ
(x1) + E I
ψ
′′(x1)
ksGA ,ks=5
6
One can note that as Gthe shear deformation tends to vanish as assumed in
classical beam theory, i.e.
ψ
(x1) = w(x1)and the classical beam equations follow
to
EIw′′′′(x1) = q(x1),M(x1) = E Iw′′(x1),Q(x1) = EI w′′′(x1)
2.2 Fundamental Equations and Variational Solution Procedures 83
The derivation of the Timoshenko’s beam equation for laminated beams one can
find in more detail in Sect. 7.3.
Exercise 2.12. Derive the free vibration equations for the moderately thick beam,
Exercise 2.11, using Hamilton’s principle in conjunction with the Reissner varia-
tional theorem.
Solution 2.12. In order to derive the free vibration equations we apply the gener-
alized Hamilton’s principle, i.e. the Hamilton’s principle in conjunction with the
Reissner variational statement
δ χ
(u
u
u,
σ
σ
σ
) =
δ
t2
Z
t1
[T(u
u
u)
Ψ
R(u
u
u,
σ
σ
σ
)]dt=0
T(u
u
u)is the kinetic energy and
Ψ
R(u
u
u,
σ
σ
σ
)the Reissner functional of the moderately
thick beam.
Ψ
R(u
u
u,
σ
σ
σ
)is known from Exercise 2.11 and the kinetic energy for the
beam may be written as
T(u
u
u) =
l
Z
0
h/2
Z
h/2
1
2
ρ
"
u
t2
+
w
t2#bdx3dx1
=
l
Z
0
1
2
ρ
"I
ψ
t2
+A
w
t2#dx1
with
I=
h/2
Z
h/2
bx2
3dx3,A=
h/2
Z
h/2
bdx3
and the mass density
ρ
of the beam material. The substitution of T[u
u
u]above and
Ψ
(w,
ψ
,M,Q)of Exercise 2.11 in the functional
χ
yields
δ χ
(u
u
u,
σ
σ
σ
) =
δ
t2
Z
t1
l
Z
0(1
2
ρ
"I
ψ
t2
+A
w
t2#
M
ψ
+Q(w+
ψ
)M2
2EI +6
ν
qM
5EA 3Q2
5GA qw)dtdx1
=
t2
Z
t1
l
Z
0
ρ
I
2
ψ
t2
δψ
+A
2w
t2
δ
w
M
δψ
ψ
δ
MQ(
δ
w+
δψ
)(w+
ψ
)
δ
Q
+M
EI
δ
M6
ν
q
5EA
δ
M6Q
5GA
δ
Qq
δ
wdtdx1
84 2 Linear Anisotropic Materials
Integration the terms M
δψ
+Q
δ
wby parts
l
Z
0
M
δψ
dx1=M
δψ
l
0
l
Z
0
M
δψ
dx1,
l
Z
0
Q
δ
wdx1=Q
δ
w
l
0
l
Z
0
Q
δ
wdx1,
rearranging the equation
δ χ
and setting
δ χ
=0 yield
t2
Z
t1
(M
δψ
+Q
δ
w)l
0dt+
t2
Z
t1
l
Z
0
ρ
I
2
ψ
t2+QM
δψ
Q+qA
2w
t2
δ
w+
ψ
M
EI +6
ν
q
5EA
δ
M
+
ψ
+w6Q
5GA
δ
Adtdx1=0
and the equations for free vibrations follow with q0
Q
M
x1
+
ρ
I
2
ψ
t2=0,
Q
x1
ρ
A
2w
t2=0,
ψ
x1M
EI =0,
ψ
+
w
x16Q
5GA =0
The underlined terms represent the contribution of rotatory inertia and the effect of
transverse shear deformation.
The system of four equations can be reduced to a system of two equations for the
unknowns wand
ψ
. Substitution of
Q=EI
2
ψ
x2
1
ρ
I
2
ψ
t2,I=bh3
12 ,A=bh
in the second and fourth equation leads
3
ψ
x3
1
ρ
E
3
ψ
t2
x1
ρ
A
E
2w
t2=0,
ψ
+
w
x1h2
10 E
G
2
ψ
x2
1
ρ
G
2
ψ
t2=0
The equations for forced or free vibrations are given in Sect. 7.3 in a more general
form.
References
Reissner E (1950) On a variational theorem in elasticity. Journal of Mathematics
and Physics 29(1-4):90–95
Chapter 3
Effective Material Moduli for Composites
Composite materials have at least two different material components which are
bonded. The material response of a composite is determined by the material moduli
of all constituents, the volume or mass fractions of the single constituents in the
composite material, by the quality of their bonding, i.e. of the behavior of the in-
terfaces, and by the arrangement and distribution of the fibre reinforcement, i.e. the
fibre architecture.
The basic assumptions made in material science approach models of fibre rein-
forced composites are:
The bond between fibres and matrix is perfect.
The fibres are continuous and parallel aligned in each ply, they are packed regu-
larly, i.e. the space between fibres is uniform.
Fibre and matrix materials are linear elastic, they follow approximately Hooke’s
law and each elastic modulus is constant.
The composite is free of voids.
Composite materials are heterogeneous, but in simplifying the analysis of compos-
ite structural elements in engineering applications, the heterogeneity of the material
is neglected and approximately overlayed to a homogeneous material. The most
important composites in structural engineering applications are laminates and sand-
wiches. Each single layer of laminates or sandwich faces is in general a fibre rein-
forced lamina. For laminates we have therefore two different scales of modelling:
The modelling of the mechanical behavior of a lamina, is called the micro-
mechanical or microscopic approach of a composite. The micro-mechanical
modelling leads to a correlation between constituent properties and average ef-
fective composite properties. Most simple mixture rules are used in engineering
applications. Whenever possible, the average properties of a lamina should be
verified experimentally by the tests described in Sect. 3.1 or Fig. 3.1.
The modelling of the global behavior of a laminate constituted of several quasi-
homogeneous laminae is called the macroscopic approach of a composite.
Fibre reinforced material is in practice neither monolithic nor homogeneous, but it
is impossible to incorporate the real material structure into design and analysis of
85
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_3
86 3 Effective Material Moduli for Composites
σ
L
σ
L
test
ε
L,
ε
T
EL=
σ
L
ε
L
,
ν
LT =
ε
T
ε
L
σ
T
test
ε
T
σ
T
ET=
σ
T
ε
T
τ
τ
test
γ
GLT =
τ
γ
Fig. 3.1 Experimental testing of the mechanical properties of an UD-layer: EL=E1,ET=E2,
GLT =G12,
ν
LT =
ν
12
composite or any other structural component. Therefore the concept of replacing the
heterogeneous material behavior with an effective material which is both homoge-
neous and monolithic, thus characterized by the generalized Hooke’s law, will be
used in engineering applications. We assume that the local variations in stress and
strain state are very small in comparison to macroscopical measurements of material
behavior.
In the following section some simple approaches to the lamina properties are
given with help of the mixture rules and simple semi-empirical consideration. The
more theoretical modelling in Sect. 3.2 has been developed to establish bounds on
effective properties. The modelling of the average mechanical characteristics of lam-
inates will be considered in Chap. 4.
3.1 Elementary Mixture Rules for Fibre-Reinforced Laminae
In Sect. 1.1 the formulas for volume fraction, mass fraction and density for fibre
reinforced composites are given by (1.1.1) - (1.1.5). The rule of mixtures and the
inverse rule of mixtures is based on the statement that the composite property is
the weighted mean of the properties or the inverse properties of each constituent
multiplied by its volume fraction. In the first case we have the upper-bound effective
property, in the second - the lower-bound. In composite mechanics these bounds are
related to W. Voigt and A. Reuss. In crystal plasticity the bounds were indroduced
by G. Taylor1and O. Sachs2. The notation used is as follows:
1Geoffrey Ingram Taylor (7 March 1886 St. John’s Wood, England - †27 June 1975 Cambridge)
- physicist and mathematician, contributions to the theory of plasticity, fluid mechanics and wave
theory
2Oscar Sachs (5 April 1896 Moscow - †30 October 1960 Syracus, N.Y.) - metallurgist, contribu-
tions to the theory of plasticity
3.1 Elementary Mixture Rules for Fibre-Reinforced Laminae 87
EYoung’s modulus
ν
Poisson’s ratio
GShear modulus
σ
Stress
ε
Strain
V,MVolume, mass
v,mVolume fraction, mass fraction
ACross-section area
ρ
Density
The subscripts f and m refer to fibre and matrix, the subscripts L 1,T2 refer to
the principal direction (fibre direction) and transverse to the fibre direction.
3.1.1 Effective Density
The derivation of the effective density of fibre reinforced composites in terms of
volume fractions is given in Sect. 1.1
ρ
=M
V=Mf+Mm
V=
ρ
fVf+
ρ
mVm
V
=
ρ
fvf+
ρ
mvm=
ρ
fvf+
ρ
m(1vf)
(3.1.1)
In literature we also find vf
φ
for the fibre volume fraction and we have
ρ
=
ρ
f
φ
+
ρ
m(1
φ
)(3.1.2)
In an actual lamina the fibres are randomly distributed over the lamina cross-section
and the lamina thickness is about 1 mm and much higher than the fibre diameter
(about 0,01 mm). Because the actual fibre cross-sections and the fibre packing gen-
erally are not known and can hardly be predicted exactly typical idealized regular
fibre arrangements are assumed for modelling and analysis, e.g. a layer-wise, square
or a hexagonal packing, and the fibre cross-sections are assumed to have circular
form. There exists ultimate fibre volume fractions vfmax , which are less than 1 and
depend on the fibre arrangements:
square or layer-wise fibre packing - vfmax =0.785,
hexagonal fibre packing - vfmax =0.907
For real UD-laminae we have vfmax about 0.50 - 0.65. Keep in mind that a lower
fibre volume fraction results in lower laminae strength and stiffness under tension
in fibre-direction, but a very high fibre volume fraction close to the ultimate values
of vfmay lead to a reduction of the lamina strength under compression in fibre
direction and under in-plane shear due to the poor bending of the fibres.
88 3 Effective Material Moduli for Composites
3.1.2 Effective Longitudinal Modulus of Elasticity
When an unidirectional lamina is acted upon by either a tensile or compression load
parallel to the fibres, it can be assumed that the strains of the fibres, matrix and
composite in the loading direction are the same (Fig. 3.2)
ε
Lf =
ε
Lm =
ε
L=
l
l(3.1.3)
The mechanical model has a parallel arrangement of fibres and matrix (Voigt model,
Sect. 2.1.1) and the resultant axial force FLof the composite is shared by both fibre
and matrix so that
FL=FLf +FLm or FL=
σ
LA=
σ
LfAf+
σ
LmAm(3.1.4)
With Hooke’s law it follows that
σ
L=EL
ε
L,
σ
Lf =ELf
ε
Lf,
σ
Lm =ELm
ε
Lm
or
EL
ε
LA=Ef
ε
LfAf+Em
ε
LmAm(3.1.5)
Since the strains of all phases are assumed to be identical (iso-strain condition),
(3.1.5) reduces to
EL=Ef
Af
A+Em
Am
A(3.1.6)
with Af
A=Afl
Al =Vf
V=vf,Am
A=Aml
Al =Vm
V=vm(3.1.7)
and the effective modulus ELcan be written as follows
EL=Efvf+Emvm=Efvf+Em(1vf) = Ef
φ
+Em(1
φ
)(3.1.8)
Equation (3.1.8) is referred to the Voigt estimate or is more familiarly known as the
rule of mixture. The predicted values of ELare in good agreement with experimental
results. The stiffness in fibre direction is dominated by the fibre modulus. The ratio
of the load taken by the fibre to the load taken by the composite is a measure of the
load shared by the fibre
FL
FL
ll+
l
Fig. 3.2 Mechanical model to calculate the effective Young’s modulus EL
3.1 Elementary Mixture Rules for Fibre-Reinforced Laminae 89
FLf
FL
=ELf
EL
vf(3.1.9)
Since the fibre stiffness is several times greater than the matrix stiffness, the second
term in (3.1.8) may be neglected
ELEfvf(3.1.10)
3.1.3 Effective Transverse Modulus of Elasticity
The mechanical model in Fig. 3.3 has an arrangement in a series of fibre and matrix
(Reuss model, Sect. 2.1.1). The stress resultant FTrespectively the stress
σ
Tis equal
for all phases (iso-stress condition)
FT=FTf =FTm,
σ
T=
σ
Tf =
σ
Tm (3.1.11)
From Fig. 3.3 it follows that
b=
bf+
bm,
ε
T=
b
b=
bf+
bm
b(3.1.12)
and with
b=vfb+ (1vf)b=bf+bm(3.1.13)
and
ε
T=
bf
vfb
vfb
b+
bm
(1vf)b
(1vf)b
b=vf
ε
Tf + (1vf)
ε
Tm (3.1.14)
with
ε
Tf =
bf
vfb,
ε
Tm =
bm
(1vf)b
Using Hooke’s law for the fibre, the matrix and the composite
b
b+
b
FT
FT
Fig. 3.3 Mechanical model to calculate the effective transverse modulus ET
90 3 Effective Material Moduli for Composites
σ
T=ET
ε
T,
σ
Tf =ETf
ε
Tf,
σ
Tm =ETm
ε
Tm (3.1.15)
substituting Eqs. (3.1.15) in (3.1.14) and considering (3.1.11) gives the formula of
ET
1
ET
=vf
Ef
+1vf
Em
=vf
Ef
+vm
Em
or ET=EfEm
(1vf)Ef+vfEm
(3.1.16)
Equation (3.1.16) is referred to Reuss estimate or sometimes called the inverse rule
of mixtures. The predicted values of ETare seldom in good agreement with exper-
imental results. With EmEffollows from (3.1.16) ETEm(1vf)1, i.e. ETis
dominated by the matrix modulus Em.
3.1.4 Effective Poisson’s Ratio
Assume a composite is loaded in the on-axis direction (parallel to the fibres) as
shown in Fig. 3.4. The major Poison’s ratio is defined as the negative of the ratio of
the normal strain in the transverse direction to the normal strain in the longitudinal
direction
ν
LT =
ε
T
ε
L
(3.1.17)
With
ε
T=
ν
LT
ε
L=
b
b=
bf+
bm
b=[vf
ε
Tf + (1vf)
ε
T m],
ν
f=
ε
Tf
ε
Lf
,
ν
m=
ε
Tm
ε
Lm
it follows that
ε
T=
ν
LT
ε
L=vf
ν
f
ε
Lf (1vf)
ν
m
ε
Lm (3.1.18)
The longitudinal strains in the composite, the fibres and the matrix are assumed to be
equal (Voigt model of parallel connection) and the equation for the major Poisson’s
ratio reduces to
b
l
l+
l
b
b
FLFL
Fig. 3.4 Mechanical model to calculate the major Poisson’s ration
ν
LT
3.1 Elementary Mixture Rules for Fibre-Reinforced Laminae 91
ν
LT =vf
ν
f+ (1vf)
ν
m=vf
ν
f+vm
ν
m=
φν
f+ (1
φ
)
ν
m(3.1.19)
The major Poisson’s ratio
ν
LT obeys the rule of mixture. The minor Poisson’s ratio
ν
TL =
ε
L/
ε
Tcan be derived with the symmetry condition or reciprocal relationship
ν
TL
ET
=
ν
LT
EL
,
ν
TL =
ν
LT
ET
EL
= (vf
ν
f+vm
ν
m)EfEm
(vfEm+vmEf)(vfEf+vmEm)(3.1.20)
The values of Poisson’s ratios for fibres or matrix material rarely differ significantly,
so that neither matrix nor fibre characteristics dominate the major or the minor elas-
tic parameters
ν
LT and
ν
TL.
3.1.5 Effective In-Plane Shear Modulus
Apply a pure shear stress
τ
to a lamina as shown in Fig. 3.5. Assuming that the shear
stresses on the fibre and the matrix are the same, but the shear strains are different
γ
m=
τ
Gm
,
γ
f=
τ
Gf
,
γ
=
τ
GLT
(3.1.21)
The model is a connection in series (Reuss model) and therefore
τ
=
τ
f=
τ
m,
=
f+
m,
=ˆ
btan
γ
=
γ
fˆ
bf+
γ
mˆ
bm(3.1.22)
and with
ˆ
b=ˆ
bf+ˆ
bm= (vf+vm)ˆ
b=vfˆ
b+ (1vf)ˆ
b(3.1.23)
it follows that
f=
γ
fvfˆ
b,
m=
γ
m(1vf)ˆ
b(3.1.24)
f
m/2
ˆ
b
fibre
matrix
matrix
γ
m/2
τ
τ
τ
τ
Fig. 3.5 Mechanical model to calculate the effective in-plane shear modulus GLT
92 3 Effective Material Moduli for Composites
Using Hooke’s law we have
τ
/GLT = (
τ
/Gm)vm+ (
τ
/Gf)vfwhich yields
GLT =GmGf
(1vf)Gf+vfGm
=GmGf
(1
φ
)Gf+
φ
Gm
(3.1.25)
or by analogy to (3.1.16)
1
GLT
=vf
Gf
+1vf
Gm
=vf
Gf
+vm
Gm
which is again a Reuss estimate. Note that assuming isotropic fibres and matrix
material one gets
Gf=Ef
2(1+
ν
f),Gm=Em
2(1+
ν
m)(3.1.26)
3.1.6 Discussion on the Elementary Mixture Rules
Summarizing the rule of mixtures as a simple model to predict effective engineering
moduli it must be kept in mind that there is no interaction between fibres and matrix.
There are only two different types of material response: the Voigt or iso-strain model
in which applied strain is the same in both material phases (parallel response) and
the Reuss or iso-stress model in which the applied stress is the same in both material
phases (series response).
For an aligned fibre composite the effective material behavior may be assumed
as transversally isotropic and five independent effective engineering moduli have to
be estimated. With x2x3as the plane of isotropy (Table 2.5) we have
E1=EL,E2=E3=ET,E4=G23 =GTT =ET/[2(1+
ν
TT)],
E5=G13 =E6=G12 =GLT,
ν
12 =
ν
13 =
ν
LT,
ν
23 =
ν
TT,
ν
LTET=
ν
TLEL
If we make choice of EL,ET,GLT,GTT,
ν
LT as the ve independent moduli the rules
of mixture yield
EL=vfEf+ (1vf)Em,
ET=EfEm
vfEm+ (1vf)Ef
,
GLT =GfGm
vfGm+ (1vf)Gf
,
ν
LT =vf
ν
f+ (1vf)
ν
m
(3.1.27)
The shear modulus GTT corresponds to an iso-shear strain model and is analogous
to the axial tensile modulus case
GTT =vfGf+ (1vf)Gm(3.1.28)
3.2 Improved Formulas for Effective Moduli of Composites 93
It may be noted that neither the iso-shear stress nor the iso-shear strain condition
for GLT and GTT estimation are close to the real situation of shearing loaded fibre
reinforced composites. Therefore the equations for GLT and GTT cannot be expected
as very reliable.
If one considers the approximative predictions for the effective moduli ELand
ETas a function of the fibre volume fraction vf, i.e., EL=EL(vf),ET=ET(vf),
and the ratio Ef/Emis fixed, it is clear that reinforcing a matrix by fibres mainly
influences the stiffness in fibre direction (ELis a linear function of vf) and rather
high fibre volume fractions are necessary to obtain a significant stiffness increase in
the transverse direction (ETis a non-linear function of vfand rather constant in the
interval 0 <vf<0,5).
Very often fibres material behavior is transversally isotropic but the matrix ma-
terial is isotropic. For such cases simple alternative relations for the effective engi-
neering moduli of the UD-lamina can be given
EL=vfEf+ (1vf)Em,
ν
LT =vf
ν
LTf + (1vf)
ν
m,
ET=ETfEm
vfEm+ (1vf)ETf
,
ν
T T =
ν
TTf
ν
m
vf
ν
m+ (1vf)
ν
TTf
,
GLT =GLTfGm
vfGm+ (1vf)GLTf
,GT T =vfGf+ (1vf)Gm
(3.1.29)
In Eq. (3.1.29) Em,
ν
m,Gm=Em/2(1+
ν
m)are the isotropic matrix moduli and
Ef,ETf,GLTf,Gf=ETf /2(1+
ν
TTf),
ν
TTf,
ν
LTf are transversally isotropic fibre mod-
uli. Em,
ν
mof the matrix material and Ef,ETf,GLTf,
ν
LTf,
ν
TTf or GTTf of the fibre
material can be chosen as the independent moduli.
3.2 Improved Formulas for Effective Moduli of Composites
Effective elastic moduli related to loading in the fibre direction, such as ELand
ν
LT,
are dominated by the fibres. All estimations in this case and experimental results are
very close to the rule of mixtures estimation. But the values obtained for transverse
Young’s modulus and in-plane shear modulus with the rule of mixtures which can
be reduced to the two model connections of Voigt and Reuss, do not agree well with
experimental results. This establishes a need for better modelling techniques based
on elasticity solutions and variational principle models and includes analytical and
numerical solution methods.
Unfortunately, the theoretical models are only available in the form of compli-
cated mathematical equations and the solution is very limited and needs huge effort.
Semi-empirical relationships have been developed to overcome the difficulties with
purely theoretical approaches.
94 3 Effective Material Moduli for Composites
The most useful of those semi-empirical models are those of Halpin and Tsai3
which can be applied over a wide range of elastic properties and fibre volume frac-
tions. The Halpin-Tsai relationships have a consistent form for all properties. They
are developed as simple equations by curve fitting to results that are based on the
theory of elasticity.
Starting from results obtained in theoretical analysis, Halpin and Tsai proposed
equations that are general and simple in formulation. The moduli of a unidirectional
composite are given by the following equations
ELand
ν
LT by the law of mixtures Eqs. (3.1.8), (3.1.19)
For the other moduli by
M
Mm
=1+
ξ η
vf
1
η
vf
(3.2.1)
Mis the modulus under consideration, e.g. ET,GLT, ...,
η
is a coefficient given
by
η
=(Mf/Mm)1
(Mf/Mm) +
ξ
(3.2.2)
ξ
is called the reinforcement factor and depends on
the geometry of the fibres
the packing arrangement of the fibres
the loading conditions.
The main difficulty in using (3.2.1) is the determination of the factor
ξ
by comparing
the semi-empirical values with analytical solutions or with experimental results.
In addition to the rule of mixtures and the semi-empirical solution of Halpin and
Tsai there are some solutions available which are based on elasticity models, e.g. for
the model of a cylindrical elementary cell subjected to tension. The more compli-
cated formulas for ELand
ν
LT as the formulas given above by the rule of mixtures
yields practically identical values to the simpler formulas and are not useful. But
the elasticity solution for the modulus GLT yields much better results and should be
applied
GLT =Gm
Gf(1+vf) + Gm(1vf)
Gf(1vf) + Gm(1+vf)=Gm
Gf(1+
φ
) + Gm(1
φ
)
Gf(1
φ
) + Gm(1+
φ
)(3.2.3)
Summarizing the results of Sects. 3.1 and 3.2 the following recommendations
may be possible for an estimation of effective elastic moduli of unidirectional lami-
nae
ELand
ν
LT should be estimated by the rule of mixtures
ν
T L follows from the reciprocal condition
GLT should be estimated from (3.2.3) or the Halpin/Tsai formulas (3.2.1) and
(3.2.2)
3Stephen Wei-Lun Tsai (6 July 1929, Beijing) - US-American engineer, strength criteria for
composites, founder of the Journal of Composite Materials
3.3 Problems 95
ETmay be estimated with help of the Halpin/Tsai formulas. But only when reli-
able experimental values of ETand GLT are available for a composite the
ξ
-factor
can be derived for this case and can be used to predict effective moduli for a range
of fibre volume ratios of the same composite. It is also possible to look for nu-
merical or analytical solutions for
ξ
based on elasticity theory. In general,
ξ
may
vary from zero to infinity, and the Reuss and Voigt models are special cases, e.g.
ET=1+
ξ η
vf
1
η
vf
Em
for
ξ
=0 and
ξ
=, respectively. In the case of circular cross-sections of the
fibres
ξ
=2 or
ξ
=1 can be recommended for Halpin-Tsai equation for ETor
GLT. But it is dangerous to use uncritically these values for any given composite.
3.3 Problems
Exercise 3.1. Determine for a glass/epoxy lamina with a 70 % fibre volume fraction
1. the density and the mass fractions of the fibre and matrix,
2. the Young’s moduli E
1ELand E
2ETand determine the ratio of a tensile
load in L-direction taken by the fibres to that of the composite,
3. the major and the minor Poisson’s ratio
ν
LT,
ν
TL,
4. the in-plane shear modulus
The properties of glass and epoxy are taken approximately from Tables F.1 and
F.2 as
ρ
glass
ρ
f=2,5 gcm3,
ν
f=0,7,
ν
m=0,3,
ρ
epoxy
ρ
m=1,35 gcm3,
Ef=70 GPa, Em=3,6 GPa.
Solution 3.1. Taking into account the material parameters and the volume fraction
of the fibres one obtains:
1. Using Eq. (1.1.3) the density of the composite is
ρ
=
ρ
fvf+
ρ
mvm=2,5·0,7 gcm3+1,35 ·0,3 gcm3=2,155 gcm3
Using Eq. (1.1.4) the mass fractions are
mf=
ρ
f
ρ
vf=25
2,1550,7=0,8121,mm=
ρ
m
ρ
vm=1,35
2,1550,3=0,1879
Note that the sum of the mass fractions must be 1
mf+mm=0,8121 +0,1879 =1
2. Using Eqs. (3.1.8), (3.1.16) and (3.1.9) we have
EL=70 ·0,7 GPa +3,6·0,3 GPa =50,08 GPa,
96 3 Effective Material Moduli for Composites
1
ET
=0,7
70GPa +0,3
3,6GPa =0,09333 GPa1,ET=10,71 GPa,
FLf
FL
=ELf
EL
vf=70
50,08 0,7=0,9784
The ratio of the tensile load FLtaken by the fibre is 0,9784.
3. Using Eqs. (3.1.19) and (3.1.20) follows
ν
LT =0.2·0.7+0.3·0.3=0.230,
ν
TL =0.230 10.71
50.08 =0,049
4. Using (3.1.26) and (3.1.25)
Gf=70 GPa
2(1+0.2)=29.17 GPa,Gm=3.6 GPa
2(1+0,3)=1.38 GPa,
GLT =1.38 GPa ·29.17 GPa
29.17 GPa ·0.3+1,38 GPa ·0.7)=4.14 GPa
Using (3.2.3)
GLT =1.38 GPa 29.17 GPa ·1.7+1.38 GPa ·0.3
29.17 GPa ·0.3+1.38 GPa ·1.7=6.22 GPa
Conclusion 3.1. The difference between the both formulae for GLT is significant.
The improved formulae should be used.
Exercise 3.2. Two composites have the same matrix materials but different fibre
material. In the first case Ef/Em=60, in the second case Ef/Em=30. The fibre
volume fraction for both cases is vf=0,6. Compare the stiffness values ELand ET
by EL/ET,EL/Em,ET/Em.
Solution 3.2. First case
EL1 =Ef1vf+Em(1vf),Ef1
Em
=60,Ef1 =60Em,
EL1 =36Em+0,4Em=36,4Em,ET1 =vf
Ef1
+1vf
Em1
=2,439Em,
EL1
ET1
=14,925,EL1
Em
=36,4,ET1
Em
=2,439
Second case
EL2 =Ef2vf+Em(1vf),Ef2 =30Em,
EL2 =18Em+0,4Em=18,4Em,ET2 =vf
Ef2
+1vf
Em1
=2,3810Em,
EL2
ET2
=7,728,EL2
Em
=18,4,ET2
Em
=2,381
Conclusion 3.2. The different fibre material has a significant influence on the
Young’s moduli in the fibre direction. The transverse moduli are nearly the same.
3.3 Problems 97
Exercise 3.3. For a composite material the properties of the constituents are Ef=90
GPa,
ν
f=0,2,Gf=35 GPa, Em=3,5 GPa,
ν
m=0,3, Gm=1,3 GPa. The volume
fraction vf=
φ
=60%. Calculate EL,ET,GLT,
ν
LT,
ν
T L with the help of the rule of
mixtures and also GLT with the improved formula.
Solution 3.3.
EL=Efvf+Em(1vf) = 55,4 GPa,ET=vf
Ef
+vm
Em1
=8,27 GPa,
ν
LT =
ν
fvf+
ν
m(1vf) = 0,24,
ν
TL =
ν
LTET
EL
=0,0358,
GLT =GmGf
(1vf)Gf+vfGm
=3,0785 GPa
The improved formula (3.2.3) yields
GLT =Gm
(1+vf)Gf+ (1vf)Gm
(1vf)Gf+ (1+vf)Gm
=4,569 GPa
Conclusion 3.3. The difference in the GLT values calculated using the rule of mix-
tures and the improved formula is again significant.
Exercise 3.4. A unidirectional glass/epoxy lamina is composed of 70% by volume
of glass fibres in the epoxy resin matrix. The material properties are Ef=85 GPa,
Em=3,4 GPa.
1. Calculate ELusing the rule of mixtures.
2. What fraction of a constant tensile force FLis taken by the fibres and by the
matrix?
Solution 3.4. Withe the assumed material parameters and fibre volume fraction one
obtains
1. EL=Efvf+Em(1vf) = 60,52 GPa,
2. FL=
σ
LA=EL
ε
A,Ff=
σ
fA=Ef
ε
Af,Fm=
σ
mA=Em
ε
Am
With FL=Ff+Fmit follows that
EL
ε
A=Ef
ε
Af+Em
ε
AmEL=0,7Ef+0,3Em,
60,52 =0,7·85 +3·3,4=59,5+1,02
and therefore FL=60,52 N, Ff=59,5 N, Fm=1,02 N.
Conclusion 3.4. The fractions of a constant tensile load in the fibres and the ma-
trix are: Fibres: 98,31 %, Matrix: 1,69 %
Exercise 3.5. The fibre and the matrix characteristics of a lamina are Ef=220 GPa,
Em=3,3 GPa, Gf=25 GPa, Gm=1,2 GPa,
ν
f=0,15,
ν
m=0,37. The fibre is
transversally isootropic and transverse Young’s modulus is ETf =22 GPa. The fibre
volume fraction is vf=0,56. The experimentally measured effective moduli are
EL=125 GPa, ET=9,1 GPa, GLT =5 GPa,
ν
LT =0,34.
98 3 Effective Material Moduli for Composites
1. compare the experimental values with predicted values based on rule of mixtures.
2. Using the Halpin-Tsai approximate model for calculating ETand GLT, what value
of
ξ
must be used in order to obtain moduli that agree with experimental values?
Solution 3.5. The comparison can be be performed by results from the mixture rules
and improved equations.
1. The rule of mixtures (3.1.27) yields
EL=0,56 ·220 GPa + (10,56)·3,3 GPa =124,65 GPa,
ET=220 GPa ·3,3 GPa
0,56 ·3,3 GPa + (10,56)·220 GPa =7,36 GPa,
GLT =25 GPa ·1,2 GPa
0,56 ·1,2 GPa + (10,56)·25 GPa =2,57 GPa,
ν
LT =0,56 ·0,15 + (10,56)·0,37 =0,25
It is seen that the fibre dominated modulus ELis well predicted by the rule of
mixtures, while ET,GLT and
ν
LT are not exactly predicted.
2. The Halpin-Tsai approximation yields with ETf =22 GPa,
ξ
=2 for ET
ET=1+
ξ η
vf
1
η
vf
Em,
η
=ETf/Em1
ETf/Em+
ξ
=0,6538,ET=9,02 GPa
With a value
ξ
=2.5 follows for GLT
GLT =1+
ξ η
vf
1
η
vf
Gm,
η
=Gf/Gm1
Gf/Gm+
ξ
=0,8339,GLT =4,88 GPa
The predicted value for ETis nearly accurate for
ξ
=2 which is the recom-
mended value in literature but the recommended value of
ξ
=1 for GLT would
underestimate the predicted value significantly
GLT =1+
η
vf
1
η
vf
Gm,
η
=Gf/Gm1
Gf/Gm+1=0,9084,GLT =3,68 GPa
It can be seen that it is dangerous to accept these values uncritically without
experimental measurements.
Exercise 3.6. Let us assume the following material parameters for the shear mod-
ulus of the fibre and the matrix: Gf=2,5·104Nmm2,Gm=1,2·103Nmm2.
The experimental value of the effective shear modulus is Gexp
LT =5·103Nmm2, the
Poisson’s ratios are
ν
f=0,15,
ν
m=0,37 and the fibre volume fraction is vf=0,56.
For the improvement in the sense of Halpin-Tsai we assume
ξ
=1...2 for a circular
cross-section of the fibres. Compute:
1. the effective shear modulus,
2. the improved effective shear modulus,
3.3 Problems 99
3. the improved effective shear modulus in the sense of Halpin-Tsai
Solution 3.6. The three problems have the following solutions:
1. the effective shear modulus after Eq. (3.1.25)
GLT =1,2·103Nmm2·2,5·104Nmm2
(10,56)·2,5·104Nmm2+0,56 ·1,2·103Nmm2
=2,57 ·103Nmm2
2. the improved effective shear modulus after Eq. (3.2.3)
GLT =1,2·103N
mm2
2,5·104Nmm2(1+0,56) + 1,2·103Nmm2(10,56)
2,5·104Nmm2(10,56) + 1,2·103Nmm2(1+0,56)
=3,68 ·103Nmm2
3. For the improved effective shear modulus in the sense of Halpin-Tsai at first
should be estimated
η
with Eq. (3.2.2)
η
=(2,5·104Nmm2/1,2·103Nmm2)1
(2,5·104Nmm2/1,2·103Nmm2) +
ξ
In the case of
ξ
=1 we get
η
=0,9084, if
ξ
=2 we get
η
=0,8686. The shear
modulus itself follows from Eq. (3.2.1)
GLT =1+
ξ η
·0,15
1
η
·0,15 1,2·103Nmm2
Finally we get
GLT(
ξ
=1) = 3,68 ·103Nmm2,GLT(
ξ
=2) = 4,61 ·103Nmm2
The last result is the closest one to the experimental value. Let us compute the
deviation
δ
=Gexp
LT Gcomp
LT
Gexp
LT ·100%
The results are
δ
(case a) = 5·103Nmm22,57 ·103Nmm2
5·103Nmm2·100% =46,6%
δ
(case b) = 5·103Nmm22,66 ·103Nmm2
5·103Nmm2·100% =26,8%
δ
(case c,
ξ
=1) = 5·103Nmm23,68 ·103Nmm2
5·103Nmm2·100% =26,3%
δ
(case c,
ξ
=2) = 5·103Nmm24,61 ·103Nmm2
5·103Nmm2·100% =7,8%
100 3 Effective Material Moduli for Composites
abc
ra
r
h
r
Fig. 3.6 Fibre arrangements. aSquare array, bhexagonal array, clayer-wise array
The result can be improved if the
ξ
-value will be increased.
Exercise 3.7. Calculate the ultimate fibre volume fractions vu
ffor the following fibre
arrangements:
1. square array,
2. hexagonal array,
3. layer-wise array.
Assume circular fibre cross-sections.
Solution 3.7. For the three fibre arrangements one gets:
1. Square array (Fig. 3.6 a)
vu
f=Af
Ac
=4(r2
π
)/4
4r2=
π
4=0.785
2. Hexagonal array (Fig. 3.6 b) with a2r,h3rfollow
Ac=hr,Af=6(1/3)
π
r2+
π
r2
and
vu
f=Af
Ac
=6(1/3)
π
r2+
π
r2)/4
63r2=
π
23=0.9069
3. Layer-wise array (Fig. 3.6 c)
vu
f=Af
Ac
=r2
π
(2r)2=
π
4=0.785
Part II
Modelling of a Single Laminae, Laminates
and Sandwiches
The second part (Chaps. 4–6) can be related to the modelling from single lami-
nae to laminates including sandwiches, the improved theories and simplest failure
concepts.
The single layer (lamina) is modelled with the help of the following assumptions
linear-elastic isotropic behaviour of the matrix and the fibre materials,
the fibres are unidirectional oriented and uniformly distributed
These assumptions result in good stiffness approximations in the longitudinal and
transverse directions. The stiffness under shear is not well-approximated.
After the transfer from the local to the global coordinates the stiffness parame-
ters for the laminate can be estimated. For the classical cases the effective stiffness
parameters of the laminate is a simple sum up over the laminate thickness of the
weighted laminae reduced stiffness parameters transferred into the global coordi-
nate system. Some improved theories are briefly introduced.
The failure concepts are at the moment a research topic characterised by a large
amount of suggestions for new criteria. With respect to this only some classical
concepts are discussed here.
Chapter 4
Elastic Behavior of Laminate and Sandwich
Composites
A lamina has been defined as a thin single layer of composite material. A lamina
or ply is a typical sheet of composite materials, which is generally of a thickness
of the order 1 mm. A laminate is constructed by stacking a number of laminae in
the direction of the lamina thickness. The layers are usually bonded together with
the same matrix material as in the single lamina. A laminate bonded of n(n2)
laminae of nearly the same thickness. A sandwich can be defined as a special case
of a laminate with n=3. Generally, the sandwich is made of a material of low
density for the inner layer, the core or the supporting pith respectively, and of high
strength material for the outer layers, the cover or face sheets. The thickness of the
core is generally much greater than the thickness of the sheets and core and sheets
are bonded to each other at the surfaces.
The design and analysis of structures composed of composite materials demands
knowledge of the stresses and strains in laminates or sandwiches. However, the lam-
inate elements are single laminae and so understanding the mechanical behavior of
a lamina precedes that of a laminate. Section 4.1 introduces elastic behavior of lam-
inae. For in-plane and out-of-plane loading, the stress resultants are formulated and
basic formulae for stress analysis are derived. These considerations are expanded to
laminates and sandwiches in Sects. 4.2 and 4.3. The governing equations of the clas-
sical laminate theory, the shear deformation theory and of a layer-wise theory are
discussed in Chap. 5. A successful design of composite structures requires knowl-
edge of the strength and the reliability of composite materials. Strength failure the-
ories have to be developed in order to compare the actual stress states in a material
to a failure criteria. Chapter 6 gives an overview on fracture modes of laminae. For
laminates, the strength is related to the strength of each individual lamina. Various
failure theories are discussed for laminates or sandwiches based on the normal and
shear strengths of unidirectional laminae.
103
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_4
104 4 Elastic Behavior of Laminate and Sandwich Composites
4.1 Elastic Behavior of Laminae
The macro-mechanical modelling and analysis of a lamina is based on average ma-
terial properties and by considering the lamina to be homogeneous. The methods
to find these average properties based on the individual mechanical values of the
constituents are discussed in Sect. 3.1. Otherwise the mechanical characterization
of laminae can be determined experimentally but it demands special experimental
equipment and is costly and time-consuming. Generally the modelling goal is to
find the minimum of parameters required for the mechanical characterization of a
lamina.
For the considerations on the elastic behavior of laminae in the following Sects.
4.1.1 - 4.1.3 one has to keep in mind that two assumptions are most important to
model the mechanics of fibre reinforced laminae:
The properties of the fibres and the matrix can be smeared into an equiva-
lent homogeneous material with orthotropic behavior. This assumption allows
to develop the stress-strain relations and to formulate the response of a fibre-
reinforced lamina sufficient simply to deal with the structural level response in a
tractable manner.
Three of the six components of stress state are generally much smaller than the
other three, i.e. the plane stress assumption, which is based on the manner in
which fibre-reinforced materials are used in such structural elements as beams,
plates or shells, will be sufficient accurately. With the assumption that the (x1
x2)-plane of the principal co-ordinate system is in-plane stress state, the in-plane
stress components
σ
1,
σ
2,
σ
6are considered to be much larger in value than the
out-of-plane stress components
σ
3,
σ
4,
σ
5and the last ones are set approximately
to zero.
Using the plane stress assumption it has to be in mind that some serious inaccuracies
in the mechanical response of laminates can be occurred, Sect. 4.2.
Therefore, together with the plane stress assumption two major misconceptions
should be avoided:
The stress components
σ
3,
σ
4,
σ
5equated to zero have to be estimated to their
magnitude and effect. Fibre reinforced material is often very poor in resisting
stresses transversely to the plane (x1x2) and therefore out-of-plane stresses
may be small but large enough to cause failure of the composite material.
With assuming
σ
3is zero does not follow that the associated strain
ε
3is also zero
and ignorable, for the stresses in the (x1x2)-plane can cause a significant strain
response in the x3-direction.
4.1.1 On-Axis Stiffness and Compliances of UD-Laminae
A thin lamina is assumed to be in a plane stress state (Sect. 2.1.5). Three cases of
material behavior of laminae are of special interest for engineering applications:
4.1 Elastic Behavior of Laminae 105
1. Short fibres or particle reinforced components with random orientation in the
matrix
The elastic behavior has no preferred direction and is macroscopically quasi-
homogenous and isotropic. The effective elastic moduli are Eand
ν
and the
relations of the in-plane stress components with the in-plane strain components
are described (Tables 2.6 and 2.7) by
σ
1
σ
2
σ
6
=
Q11 Q12 0
Q12 Q11 0
0 0 Q66
ε
1
ε
2
ε
6
,
ε
1
ε
2
ε
6
=
S11 S12 0
S12 S11 0
0 0 S66
σ
1
σ
2
σ
6
(4.1.1)
with
Q11 =E
1
ν
2,S11 =1
E,
Q12 =E
ν
1
ν
2,S12 =
ν
E,
Q66 =G=E
2(1+
ν
),S66 =1
G=2(1+
ν
)
E
2. Long fibres with one unidirectional fibre orientation, so-called unidirectional
laminae or UD-laminae, with loading along the material axis (on-axis case)
This type of material forms the basic configuration of fibre composites and is
the main topic of this textbook. The elastic behavior of UD-laminae depends on
the loading reference coordinate systems. In the on-axis case the reference axes
(1,2) are identical to the material or principal axes of the lamina parallel and
transverse to the fibre direction (Fig. 4.1). The 1-axis is also denoted as L-axis
and the 2-axis as T-axis (on-axis case). The elastic behavior is macroscopically
quasi-homogeneous and orthotropic with four independent material moduli (Ta-
ble 2.6)
E
1EL,E
2ET,E
6G
12 GLT,
ν
12
ν
LT (4.1.2)
and the in-plane stress-strain relations are
Fig. 4.1 Unidirectional lam-
ina with principal material
axis Land T(on-axis)
1L
2T
r
r
r
r
r
r
r
r
r
r
r
r
r
r
h
106 4 Elastic Behavior of Laminate and Sandwich Composites
σ
1
σ
2
σ
6
=
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q
66
ε
1
ε
2
ε
6
,
ε
1
ε
2
ε
6
=
S
11 S
12 0
S
12 S
22 0
0 0 S
66
σ
1
σ
2
σ
6
(4.1.3)
with
Q
11 =E
1/(1
ν
12
ν
21),S
11 =1/E
1,
Q
22 =E
2/(1
ν
12
ν
21),S
22 =1/E
2,
Q
66 =G
12 =E
6,S
66 =1/G
12 =1/E
6,
Q
12 =E
2
ν
12/(1
ν
12
ν
21),S
12 =
ν
12/E
1=
ν
21/E
2
3. UD-laminae with loading along arbitrary axis (x1,x2) different from the material
axis (off-axis case). The elastic behavior is macroscopically quasi-homogeneous
and anisotropic. The in-plane stress-strain relations are formulated by fully pop-
ulated matrices with all Qi j and Si j different from zero but the number of inde-
pendent material parameters is still four as in case 2. The transformation rules
are given in detail in Sect. 4.1.2.
A UD-lamina has different stiffness in the direction of the material axes. With
EfEmthe stiffness in the L-direction is fibre dominated and for the effective
moduli (Sect. 3.1) ELET. Figure 4.2 illustrates qualitatively the on-axis elastic
behavior of the UD-lamina. In thickness direction x3Torthogonal to the (L-T)-
plane a UD-lamina is macro-mechanically quasi-isotropic. The elastic behavior in
the thickness direction is determined by the matrix material and a three-dimensional
model of a single UD-layer yields a transversely-isotropic response with five inde-
pendent material engineering parameters:
E
1EL,E
2ET=E
3ET,E
5=E
6GLT,
ν
12
ν
LT =
ν
13
ν
LT,
E
4GTT=E
2/[2(1+
ν
23)] ET/[2(1+
ν
TT)]
(4.1.4)
Fig. 4.2 On-axis stress-strain
equations for UD-lamina
(qualitative)
σ
1,
σ
2,
σ
6
ε
1,
ε
2,
ε
6
σ
1(
ε
1)
σ
2(
ε
2)
σ
6(
ε
6)
σ
1
σ
1
σ
2
σ
2
σ
6
σ
6
4.1 Elastic Behavior of Laminae 107
The material behavior in the 2T and 3 =Tdirections is equivalent. Therefore,
the notation of the engineering parameters is given by E
2=E
3ET,E
5=E
6GLT,
ν
12 =
ν
13
ν
LT,E
4=E
2/[2(1+
ν
23)] GTT =ET/[2(1+
ν
TT)]. Summarizing, the
stress-strain relations for on-axis loading of UD-laminae in a contracted vector-
matrix notation leads the equations
σ
σ
σ
=Q
Q
Q
ε
ε
ε
or
σ
i=Q
i j
ε
j,Q
i j =Qji
ε
ε
ε
=S
S
S
σ
σ
σ
or
ε
i=S
i j
σ
j,S
i j =Sji
i,j=1,2,6 (4.1.5)
The values Q
i j of the reduced stiffness matrix Q
Q
Qand the S
i j of the compliance ma-
trix S
S
Sdepend on the effective moduli of the UD-lamina. The term reduced stiffness
is used in relations given by Eqs. (2.1.76) and (4.1.3). These relations simplify the
problem from a three-dimensional to a two-dimensional or plane stress state. Also
the numerical values of the stiffness Q
i j are actually less than the numerical values
of their respective counterparts C
i j, see Eq. (2.1.79), of the three-dimensional prob-
lem and therefore the stiffness are reduced in that sense also. For on-axis loading
the elastic behavior is orthotropic and with Q
16 =Q
26 =0 and S
16 =S
26 =0, there
is as in isotropic materials no coupling of normal stresses and shear strains and also
shear stresses applied in the (L-T)-plane do not result in any normal strains in the
L and T direction. The UD-lamina is therefore also called a specially orthotropic
lamina.
Composite materials are generally processed at high temperature and then cooled
down to room temperature. For polymeric fibre reinforced composites the tempera-
ture difference is in the range of 20003000C and due to the different thermal ex-
pansion of the fibres and the matrix, residual stresses result in a UD-lamina and ex-
pansion strains are induced. In addition, polymeric matrix composites can generally
absorb moisture and the moisture change leads to swelling strains and stresses sim-
ilar to these due to thermal expansion. Therefore we speak of hygrothermal stresses
and strains in a lamina. The hygrothermal strains in the longitudinal direction and
transverse the fibre direction of a lamina are not equal since the effective elastic
moduli ELand ETand also the thermal and moisture expansion coefficients
α
th
L,
α
th
T
and
α
mo
L,
α
mo
Trespectively, are different.
The stress-strain relations of a UD-lamina, including temperature and moisture
differences are given by
ε
1
ε
2
ε
6
=
S
11 S
12 0
S
12 S
22 0
0 0 S
66
σ
1
σ
2
σ
6
+
ε
th
1
ε
th
2
0
+
ε
mo
1
ε
mo
2
0
(4.1.6)
with
ε
th
1
ε
th
2
0
=
α
th
1
α
th
2
0
T,
ε
mo
1
ε
mo
2
0
=
α
mo
1
α
mo
2
0
M(4.1.7)
Tis the temperature change and Mis weight of moisture absorption per unit weight
of the lamina.
α
mo
L,
α
mo
Tare also called longitudinal and transverse swelling coeffi-
108 4 Elastic Behavior of Laminate and Sandwich Composites
cients. Equation (4.1.6) can be inverted to give
σ
1
σ
2
σ
6
=
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q
66
ε
1
ε
th
1
ε
mo
1
ε
2
ε
th
2
ε
mo
2
ε
6
(4.1.8)
Note that the temperature and moisture changes do not have any shear strain terms
since no shearing is induced in the material axes. One can see that the hygrothermal
behavior of an unidirectional lamina is characterized by two principal coefficients
of thermal expansion,
α
th
1,
α
th
2, and two of moisture expansion,
α
mo
1,
α
mo
2. These
coefficients are related to the material properties of fibres and matrix and of the fibre
volume fraction.
Approximate micro-mechanical modelling of the effective hygrothermal coef-
ficients were given by Schapery1and analogous to the micro-mechanical mod-
elling of elastic parameters, Chap. 3, for a fibre reinforced lamina and isotropic
constituents the effective thermal expansion coefficients are
α
th
L=
α
th
fvfEf+
α
th
m(1vf)Em
vfEf+ (1vf)Em
,
α
th
T=
α
th
fvf(1+
ν
f) +
α
th
m(1vf)(1+
ν
m)[vf
ν
f+ (1vf)
ν
m]
α
th
L
(4.1.9)
If the fibres are not isotropic but have different material response in axial and trans-
verse directions, e.g. in the case of carbon or aramid fibres, the relations for
α
th
Land
α
th
Thave to be changed to
α
th
L=
α
th
LfvfELf +
α
th
m(1vf)Em
vfELf + (1vf)Em
,
α
th
T= (
α
th
Tf +
ν
Tf
α
th
Lf)vf+ (1+
ν
m)
α
th
m(1vf)
[vf
ν
Tf + (1vf)
ν
m]
α
th
L
(4.1.10)
In most cases the matrix material can be considered isotropic and therefore the ori-
entation designation L,T of the matrix material parameters can be dropped.
Discussion and conclusions concerning effective moduli presented in Chap. 3 are
valid for effective thermal expansion coefficients too. The simple micro-mechanical
approximations of effective moduli yield proper results for
α
th
Lbut fails to predict
α
th
Twith the required accuracy. For practical applications
α
th
Land
α
th
Tshould be
normally determined by experimental methods.
Micro-mechanical relations for effective coefficients of moisture expansion can
be modelled analogously. However, some simplification can be taken into consider-
ation. Usually the fibres, e.g. glass, carbon, boron, etc., do not absorb moisture that
means
α
mo
f=0. For isotropic constituents the formulae for
α
mo
Land
α
mo
Tare
1Richard Allan Schapery (3 March 1935 Duluth, Minnesota, United States) - engineering educa-
tor, contributions in the field of mechanics of composite materials
4.1 Elastic Behavior of Laminae 109
α
mo
L=
α
mo
mEm(1vf)
vfEf+ (1vf)Em
=
α
mo
mEm(1vf)
E1
,
α
mo
T=
α
mo
m(1vf)[(1+
ν
m)Efvf+ (1vf)Em
ν
fvfEm]
vfEf+ (1vf)Em
=
α
mo
m(1vf)
E1{(1+
ν
m)Efvf+ [(1vf)
ν
fvf]Em}
(4.1.11)
and for a composite with isotropic matrix but orthotropic fibres the effective moduli
are given by
α
mo
L=
α
mo
m
Em(1vf)
EL
,
α
mo
T=
α
mo
m
1vf
EL{(1+
ν
m)ELfvf+ [(1vf)
ν
LTfvfEm}
(4.1.12)
The formulae (4.1.9) - (4.1.12) completes the discussion about micro-mechanics
in Chap. 3. Note that for a great fibre volume fraction negative
α
th
Lvalues can be
predicted reflecting the dominance of negative values of fibre expansion
α
f
L, e.g. for
graphite-reinforced material.
Summarizing one has to keep in mind that with the plane stress assumption re-
ferred to the principal material axis L,T the mechanical shear strains and the total
shear strain are identical, i.e.
ε
th
6=
ε
mo
60, Eqs. (4.1.6) - (4.1.8). Also the through-
the-thickness total strains
ε
4,
ε
5are zero and there are no mechanical, thermal or
moisture strains. The conclusion regarding the normal strains
ε
3is not the same.
Using the condition
σ
3=0 follows
ε
3=S
13
σ
1+S
23
σ
2+
α
th
3T+
α
mo
3M
This equation is the basis for determining the out-of-plane or through-the-thickness
thermal and moisture effects of a laminate.
4.1.2 Off-Axis Stiffness and Compliances of UD-Laminae
A unidirectional lamina has very low stiffness and strength properties in the trans-
verse direction compared with these properties in longitudinal direction. Laminates
are constituted generally of different layers at different orientations. To study the
elastic behavior of laminates, it is necessary to take a global coordinate system for
the whole laminate and to refer the elastic behavior of each layer to this reference
system. This is necessary to develop the stress-strain relationship for an angle lam-
ina, i.e. an off-axis loaded UD-lamina.
The global and the local material reference systems are given in Fig. 4.3. We
consider the ply material axes to be rotated away from the global axes by an angle
θ
, positive in the counterclockwise direction. This means that the (x1,x2)-axes are
at an angle
θ
clockwise from the material axes. Thus transformation relations are
110 4 Elastic Behavior of Laminate and Sandwich Composites
Fig. 4.3 UD-lamina with the
local material principal axis
(1,2)(L,T)and the global
reference system (x1,x2)
yx2
1x
1
xx1
2x
2
θ
needed for the stresses, the strains, the stress-strain equations, the stiffness and the
compliance matrices.
The transformation equations for a rotating the reference system (x
1,x
2) or
(x1,x2) counterclockwise or clockwise by an angle
θ
follow from Sect. 2.1.2 and
are given in Table 4.1. Note the relations for the transformation matrices derived in
Sect. 2.1.2
Table 4.1 Transformation rules of the coordinates, displacements, strains and stresses of a lamina
a) Rotation of the reference systems
x
1
x
2=c s
s c x1
x2,x1
x2=cs
s c x
1
x
2
x
x
x=R
R
Rx
x
x,x
x
x=R
R
RTx
x
x
b) Transformation of displacements
u
1
u
2=c s
s c u1
u2,u1
u2=cs
s c u
1
u
2
u
u
u=R
R
Ru
u
u,u
u
u=R
R
RTu
u
u
c) Transformation of strains
ε
1
ε
2
ε
6
=
c2s2sc
s2c2sc
2sc 2sc c2s2
ε
1
ε
2
ε
6
,
ε
1
ε
2
ε
6
=
c2s2sc
s2c2sc
2sc 2sc c2s2
ε
1
ε
2
ε
6
ε
ε
ε
=T
T
T
εε
ε
ε
=T
T
T
σ
T
ε
ε
ε
,
ε
ε
ε
=T
T
T
ε
ε
ε
ε
= (T
T
T
σ
)T
ε
ε
ε
d) Transformation of stresses
σ
1
σ
2
σ
6
=
c2s22sc
s2c22sc
sc sc c2s2
σ
1
σ
2
σ
6
,
σ
1
σ
2
σ
6
=
c2s22sc
s2c22sc
sc sc c2s2
σ
1
σ
2
σ
6
σ
σ
σ
=T
T
T
σσ
σ
σ
=T
T
T
ε
T
σ
σ
σ
,
σ
σ
σ
=T
T
T
σ
σ
σ
σ
= (T
T
T
ε
)T
σ
σ
σ
with ssin
θ
,ccos
θ
4.1 Elastic Behavior of Laminae 111
T
T
T
ε
=T
T
T
ε
1=T
T
T
σ
T,T
T
T
ε
= (T
T
T
ε
)1= (T
T
T
σ
)T,
T
T
T
σ
=T
T
T
σ
1=T
T
T
ε
T,T
T
T
σ
= (T
T
T
σ
)1= (T
T
T
ε
)T(4.1.13)
Table 4.2 summarizes the transformation rules for the stress-strain relations and for
the values of the stiffness and the compliance matrices. The transformation matrices
of Table 4.2 follow from Sect. 2.1.2. Using the relations (4.1.13) the transformation
rules can be formulated in matrix notation
Table 4.2 Transformation of the reduced stiffness matrix Q
i j and compliance matrix S
i j in the ref-
erence system (x
1,x
2) to the reduced stiffness matrix Qi j and compliance matrix Si j in the (x1,x2)-
system
a) Constitutive equations in the (x1,x2)-reference system
σ
1
σ
2
σ
6
=
Q11 Q12 Q16
Q12 Q22 Q26
Q16 Q26 Q66
ε
1
ε
2
ε
6
,
ε
1
ε
2
ε
6
=
S11 S12 S16
S12 S22 S26
S16 S26 S66
σ
1
σ
2
σ
6
σ
σ
σ
=Q
Q
Q
ε
ε
ε
,
ε
ε
ε
=S
S
S
σ
σ
σ
,
Q
Q
Q= (T
T
T
ε
)TQ
Q
QT
T
T
ε
,S
S
S= (T
T
T
σ
)TS
S
ST
T
T
σ
,
Qi j =Qji ,Q
i j =Qji,Si j =Sj i,S
i j =Sji
b) Transformation of the reduced stiffnesses
Q11
Q12
Q16
Q22
Q26
Q66
=
c42c2s2s44c2s2
c2s2c4+s4c2s24c2s2
c3scs(c2s2)cs32cs(c2s2)
s42c2s2c44c2s2
cs3cs(c2s2)c3s2cs(c2s2)
c2s22c2s2c2s2(c2s2)2
Q
11
Q
12
Q
22
Q
66
c) Transformation of the compliances
S11
S12
S16
S22
S26
S66
=
c42c2s2s4c2s2
c2s2c4+s4c2s2c2s2
2c3s2cs(c2s2)2cs3cs(c2s2)
s42c2s2c4c2s2
2cs32cs(c2s2)2c3s cs(c2s2)
4c2s28c2s24c2s2(c2s2)2
S
11
S
12
S
22
S
66
with ssin
θ
,ccos
θ
112 4 Elastic Behavior of Laminate and Sandwich Composites
Q
Q
Q=T
T
T
ε
T
Q
Q
QT
T
T
ε
=T
T
T
σ
Q
Q
Q(T
T
T
σ
)T,
S
S
S=T
T
T
σ
T
S
S
ST
T
T
σ
=T
T
T
ε
S
S
S(T
T
T
ε
)T,
Q
Q
Q= (T
T
T
ε
)TQ
Q
QT
T
T
ε
,
S
S
S= (T
T
T
σ
)TS
S
ST
T
T
σ
(4.1.14)
Starting with the stiffness equation
σ
σ
σ
=Q
Q
Q
ε
ε
ε
and introducing
σ
σ
σ
=Q
Q
Q
ε
ε
ε
in the
transformation
σ
σ
σ
=T
T
T
σ
σ
σ
σ
=T
T
T
ε
T
σ
σ
σ
it follows that
σ
σ
σ
=T
T
T
ε
T
Q
Q
Q
ε
ε
ε
and with
ε
ε
ε
=T
T
T
εε
ε
ε
this gives
σ
σ
σ
= (T
T
T
ε
)TQ
Q
QT
T
T
εε
ε
ε
. Comparison of equations
σ
σ
σ
=Q
Q
Q
ε
ε
ε
and
σ
σ
σ
= (T
T
T
ε
)TQ
Q
QT
T
T
εε
ε
ε
yields
Q11 Q12 Q16
Q12 Q22 Q26
Q16 Q26 Q66
= (T
T
T
ε
)T
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q66
T
T
T
ε
(4.1.15)
In an analogous way
S11 S12 S16
S12 S22 S26
S16 S26 S66
= (T
T
T
σ
)T
S
11 S
12 0
S
12 S
22 0
0 0 Q66
T
T
T
σ
(4.1.16)
can be derived. Note that in (4.1.15) and (4.1.16) the matrices [Qi j ]and [Si j ]have
six different elements but the matrices [Q
i j]and [S
i j]have only four independent
elements. The elements in Qi j or Si j are functions of the four independent material
characteristics Q
i j or S
i j and the angle
θ
. The experimental testing is therefore more
simple than for a real anisotropic material with 6 independent material values, if the
material axes of the lamina are known.
From the transformation c) in Table 4.2 follows the transformation of the engi-
neering parameters EL,ET,GLT,
ν
LT of the UD-lamina in the on-axis-system to the
engineering parameters in the global system (x1,x2). From equation a) in Table 4.2
for an angle lamina it can be seen that there is a coupling of all normal and shear
terms of stresses and strains. In Fig. 4.4 these coupling effects in an off-axes loaded
UD-lamina are described.
The coupling coefficients
ν
12 =
ε
2
ε
1
=S21
S11
,
ν
21 =
ε
1
ε
2
=S12
S22
(4.1.17)
are the known Poisson’s ratios and the ratios
ν
16 =
ε
6
ε
1
=S61
S11
,
ν
26 =
ε
6
ε
2
=S62
S22
(4.1.18)
are so called shear coupling values. They are non-dimensional parameters like Pois-
son’s ratio and relate normal stresses to shear strains or shear stresses to normal
strains.
4.1 Elastic Behavior of Laminae 113
σ
1
σ
1
σ
2
σ
2
σ
6
σ
6
ε
1
σ
1
=S11 =1
E1
ε
2
σ
1
=S21 =
ν
12
E1
ε
6
σ
1
=S61 =
ν
16
E1
ε
2
σ
2
=S22 =1
E2
ε
1
σ
2
=S12 =
ν
21
E2
ε
6
σ
2
=S62 =
ν
26
E2
ε
6
σ
6
=S66 =1
E6
ε
1
σ
6
=S16 =
ν
61
E6
ε
2
σ
6
=S26 =
ν
62
E6





Fig. 4.4 Off-axis loaded UD-lamina with one stress component in each case
Hence the strain-stress equation of an angle lamina can be written in terms of
engineering parameters of the off-axis case as
ε
1
ε
2
ε
6
=
1/E1
ν
21/E2
ν
61/E6
ν
12/E11/E2
ν
62/E6
ν
16/E1
ν
26/E21/E6
σ
1
σ
2
σ
6
(4.1.19)
With the compliance engineering parameters
S11 =1
E1
,S12 =
ν
21
E2
,S21 =
ν
12
E1
,
S66 =1
E6
,S16 =
ν
61
E6
,S61 =
ν
16
E1
,
S26 =
ν
62
E6
,S62 =
ν
26
E2
,S22 =1
E2
(4.1.20)
it follows from the symmetry considerations of the compliance matrix that
Si j =Sji,i,j=1,2,6; i.e.
ν
12
E2
=
ν
21
E1
,
ν
16
E1
=
ν
61
E6
,
ν
26
E2
=
ν
62
E6
(4.1.21)
but the anisotropic coupling coefficients are
ν
i j 6=
ν
ji,i,j=1,2,6 (4.1.22)
Equation (4.1.19) can be inverted to yield the stress-strain equations in terms of
engineering parameters but these relations would be more complex than (4.1.19).
Using the relationships between engineering parameters and compliances (4.1.20)
in the compliance transformation rule (Table 4.2) we obtain the following transfor-
mations for the engineering parameters of the angle lamina including shear coupling
114 4 Elastic Behavior of Laminate and Sandwich Composites
ratios
ν
16 and
ν
26
1
E1
=1
E
1
c4+1
G
12 2
ν
12
E
1s2c2+1
E
2
s4,
1
E2
=1
E
1
s4+1
G
12 2
ν
12
E
1s2c2+1
E
2
c4,
1
G12
=22
E
1
+2
E
2
+4
ν
12
E
11
G
12 s2c2+1
G
12
(c4+s4),(4.1.23)
ν
12
E1
=
ν
21
E2
=
ν
12
E
1
(c4+s4)1
E
1
+1
E
21
G
12 c2s2,
ν
16 =E
1 2
E
1
+2
ν
12
E
11
G
12 sc32
E
2
+2
ν
12
E
11
G
12 s3c,
ν
26 =E
2 2
E
1
+2
ν
12
E
11
G
12 s3c2
E
2
+2
ν
12
E
11
G
12 sc3
Equation (4.1.23) can be also written in the following form
E1=EL
c4+EL
GLT 2
ν
LTs2c2+EL
ET
s4
,
E2=ET
c4+ET
GLT 2
ν
TLs2c2+ET
EL
s4
,
G12 =GLT
c4+s4+22GLT
EL
(1+2
ν
LT) + 2GLT
ET1s2c2
,
ν
12 =
ν
LT(c4+s4)1+EL
ETEL
GLT c2s2
c4+EL
GLT 2
ν
LTc2s2+EL
ET
s4
,
ν
21 =
ν
TL(c4+s4)1+ET
ELET
GLT c2s2
c4+ET
GLT 2
ν
TLc2s2+ET
EL
s4
,
ν
16 =EL 2
EL
+2
ν
LT
EL1
GLT sc32
ET
+2
ν
LT
EL1
GLT s3c,
ν
26 =ET 2
EL
+2
ν
LT
EL1
GLT s3c2
ET
+2
ν
LT
EL1
GLT sc3
(4.1.24)
The engineering parameters can change rapidly with angel
θ
. This can be inter-
preted as if the fibres are not oriented exactly as intended the values of engineering
parameters are very less or more than expected.
4.1 Elastic Behavior of Laminae 115
A computational procedure for calculating the elastic parameters of a UD-lamina
in off-axis loading can be illustrated by the following steps:
1. Input the basic engineering parameters EL,ET,GLT ,
ν
LT referred to the material
axes of the lamina and obtained by material tests or mathematical modelling.
2. Calculate the compliances S
i j and the reduced stiffness Q
i j.
3. Application of transformations to obtain the lamina stiffness Qi j and compliances
Si j.
4. Finally calculate the engineering parameters E1,E2,G12 ,
ν
12,
ν
21,
ν
16,
ν
26 re-
ferred to the (x1,x2)-system.
Otherwise the engineering parameters referred to the (x1,x2)-coordinate can be cal-
culated directly by Eqs. (4.1.24).
Analogous to the on-axis loaded UD-lamina also the off-axis lamina is in thick-
ness direction x3Torthogonal to the (x1,x2)-plane macro-mechanically quasi-
isotropic and the three-dimensional material behavior is transversely-isotropic. The
mechanical properties transverse to the fibre direction are provided by weaker ma-
trix material and the effects of transverse shear deformation may be significant. For
such cases, the stress and the strain vector should include all six components
σ
σ
σ
T= [
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6],
ε
ε
ε
T= [
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6]
For a rotation about the direction e
e
e3(Fig. 2.6) the transformation matrices (2.1.39)
and (2.1.40) are valid and relations for stress and strain vectors in the on-axis and
the off-axis reference system are given by
σ
σ
σ
=
3
T
T
T
σσ
σ
σ
,
ε
ε
ε
=
3
T
T
T
εε
ε
ε
,
3
T
T
T
ε
=3
T
T
T
σ
T
=3
T
T
T
ε
1
,
σ
σ
σ
=3
T
T
T
σ
1
σ
σ
σ
,
ε
ε
ε
=3
T
T
T
ε
1
ε
ε
ε
,
3
T
T
T
σ
=3
T
T
T
ε
T
=3
T
T
T
σ
1(4.1.25)
When the stiffness matr ix C
C
Ccorresponding to an orthotropic material behavior, see
Eq. (2.1.46) , the transformed stiffness matrix C
C
Cmay be written in detail as for mon-
oclinic material behavior, see Eq. (2.1.42)
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
=
C11 C12 C13 0 0 C16
C22 C23 0 0 C26
C33 0 0 C36
C44 C45 0
S Y M C55 0
C66
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
(4.1.26)
The Ci j are the transformed stiffness, i.e. in vector-matrix notation
σ
σ
σ
=C
C
C
ε
ε
ε
σ
σ
σ
=C
C
C
ε
ε
ε
,
σ
σ
σ
=
3
T
T
T
σ
σ
σ
σ
= 3
T
T
T
ε
!T
C
C
C
ε
ε
ε
= 3
T
T
T
ε
!T
C
C
C3
T
T
T
εε
ε
ε
=C
C
C
ε
ε
ε
116 4 Elastic Behavior of Laminate and Sandwich Composites
we finally obtain
C
C
C= 3
T
T
T
ε
!T
C
C
C3
T
T
T
ε
,(4.1.27)
in which the Ci j
C11 =C
11c4+2C
12c2s2+C
22s4+4C
66c2s2,
C12 =C
11c2s2+C
12(c4+s4) + C
22c2s24C
66c2s2,
C13 =C
13c2+C
23s2,C14 =0,C15 =0,
C16 =C
11c3sC
12cs(c2s2)C
22cs32C
66cs(c2s2),
C22 =C
11s4+2C
12c2s2+C
22c4+4C
66c2s2,
C23 =C
13s2+C
23c2,C24 =0,C25 =0,
C26 =C
11cs3+C
12cs(c2s2)C
22c3s+2C
66cs(c2s2),
C33 =C
33,C34 =0,C35 =0,
C36 =C
13cs C
23cs,
C44 =C
44c2+C
55s2,
C45 =C
44cs +C
55cs,C46 =0,
C55 =C
44s2+C
55c2,C56 =0,
C66 =C
11c2s22C
12c2s2+C
22c2s2+C
66(c2s2)2(4.1.28)
The 13 non-zero stiffness of Cij are not independent material values. They are func-
tions of 9 C
i j for a three-dimensional orthotropic material, i.e. of
C
11,C
12,C
13,C
22,C
23,C
33,C
44,C
55,C66
and of 5 C
i j for a transverse-isotropic behavior, i.e. of
C
11,C
12,C
22,C
23,C
55
because
C
13 =C
12,C
33 =C
22,C
44 =1
2(C
22 C
23),C
66 =C
55
With
ε
ε
ε
=S
S
S
σ
σ
σ
,
ε
ε
ε
=S
S
S
σ
σ
σ
follow analogously the transformed compliances
S
S
S=3
T
T
T
σ
T
S
S
S3
T
T
T
σ
,(4.1.29)
in which the Si j
4.1 Elastic Behavior of Laminae 117
S11 =S
11c4+2S
12c2s2+S
22s4+S
66c2s2,
S12 =S
11c2s2+S
12(c4+s4) + S
22c2s2S
66c2s2,
S13 =S
13c2+S
23s2,S14 =0,S15 =0,
S16 =2S
11c3s2S
12cs(c2s2)2S
22cs3S
66cs(c2s2),
S22 =S
11s4+2S
12c2s2+S
22c4+S
66c2s2,
S23 =S
13s2+S
23c2,S24 =0,S25 =0,
S26 =2S
11cs32S
12cs(c2s2)2S
22c3sS
66cs(c2s2),
S33 =S
33,S34 =0,S35 =0,
S36 =2S
13cs 2S
23cs,
S44 =S
44c2+S
55s2,
S45 =S
44cs +S
55cs,S46 =0,
S55 =S
44s2+S
55c2,S56 =0,
S66 =4S
11c2s24S
12c2s2+2S
22c2s2+S
66(s2c2)2
(4.1.30)
There are again 13 non-zero compliances, but only 9 independent material values
for the orthotropic and 5 independent material values for the transversal-isotropic
case.
The stress-strain relationship for an angle lamina, i.e. an off-axis loaded UD-
lamina, including hygrothermal effects takes the following form
ε
1
ε
2
ε
6
=
S11 S12 S16
S12 S22 S26
S16 S26 S66
σ
1
σ
2
σ
6
+
ε
th
1
ε
th
2
ε
th
6
+
ε
mo
1
ε
mo
2
ε
mo
6
,(4.1.31)
where
ε
th
1
ε
th
2
ε
th
6
=
α
th
1
α
th
2
α
th
6
T,
ε
mo
1
ε
mo
2
ε
mo
6
=
α
mo
1
α
mo
2
α
mo
6
M(4.1.32)
with the thermal and moisture expansion coefficients
α
th
i,
α
mo
i,i=1,2,6 and the
temperature change Tor the weight of moisture absorption per unit weight M,
respectively. It should be remembered that although the coefficients of both ther-
mal and moisture expansions are pure dilatational in the material coordinate system
(L,T), rotation into the global (x1,x2) system results in coefficients
α
th
6,
α
mo
6. Fur-
thermore if there are no constraints placed on a UD-lamina, no mechanical strains
will be included in it and therefore no mechanical stresses are induced. But in lami-
nates, even if there are no constraints on the laminate, the difference in thermal and
moisture expansion coefficients of the various laminae of a laminate induces differ-
ent expansions in each layer and results in residual stresses. This will be explained
fully in Sects. 4.2.4 and 4.2.5. With
ε
th
1=
α
th
1T,
ε
th
2=
α
th
2T,
ε
th
6=0,
(4.1.33)
118 4 Elastic Behavior of Laminate and Sandwich Composites
and
ε
th
1=
ε
th
1c2+
ε
th
2s2,
ε
th
2=
ε
th
1s2+
ε
th
2c2,
ε
th
6=2(
ε
th
1
ε
th
2)cs,
(4.1.34)
follow
α
th
1=
α
th
1c2+
α
th
2s2,
α
th
2=
α
th
1s2+
α
th
2c2,
α
th
6=2(
α
th
1
α
th
2)cs
(4.1.35)
In an anisotropic layer, uniform heating induces not only normal strains, but also
shear thermal strains. For a transversal isotropic material behavior there is additional
ε
th
3=
α
th
3T,
α
th
3=
α
th
3,
α
th
4=
α
th
5=0. Because
ε
3=
ε
3the strain
ε
3can be obtained
directly from
ε
3=
α
th
3T+
α
mo
3M+S
13
σ
1+S
23
σ
2
However, the stresses
σ
1,
σ
2can be written in terms
σ
1,
σ
2,
σ
6referred to the off-
axis coordinate system to obtain an expression for
ε
3that represents the normal
strain in the x3-direction in terms of the global coordinate system
ε
3=
ε
th
3+
ε
mo
3+ (S13c2+S23s2)
σ
1+ (S13s2+S23c2)
σ
2
+2(S13 S23s2)sc
σ
1
(4.1.36)
4.1.3 Stress Resultants and Stress Analysis
Sections 4.1.1 and 4.1.2 describe the constitutive equations for UD-laminae in an
on-axis and an off-axis reference system as a relation between stresses and strains.
For each lamina, the stress components can be integrated across their thickness hand
yield stress resultants. Stress resultants can be in-plane forces, transverse forces and
resultant moments. The constitutive equations may then be formulated as relations
between mid-plane strain and in-plane forces, transverse shear strains and transverse
forces and mid-plane curvatures and resultant moments, respectively.
The in-plane stress resultant force vector, denoted by
N
N
N= [N1N2N6]T,(4.1.37)
is defined by
N
N
N=
h/2
Z
h/2
σ
σ
σ
dx3(4.1.38)
The Niare forces per unit length, N1,N2are normal in-plane resultants and N6is
a shear in-plane resultant, respectively. They are illustrated in Fig. 4.5 for constant
in-plane stresses
σ
1,
σ
2,
σ
6across the thickness. In this case we have
4.1 Elastic Behavior of Laminae 119
x1
x2
x3
N1
N6
N6
N2
h
Fig. 4.5 In-plane force resultants per unit length N
N
NT= [N1N2N6]
N
N
N=
σ
σ
σ
h(4.1.39)
The reduced stiffness matrix Q
Q
Qof the lamina has also constant components across
h. The strains of the midplane x3=0 of the lamina are given by
ε
ε
ε
(x1,x2,x3=0) =
ε
ε
ε
(x1,x2)
and Eq. (4.1.39) yields
N
N
N=Q
Q
Q
ε
ε
ε
h=A
A
A
ε
ε
ε
,A
A
A=Q
Q
Qh,
ε
ε
ε
T= [
ε
1
ε
2
ε
6](4.1.40)
Q
Q
Qis the reduced stiffness matrix (Table 4.2 a) and A
A
Ais the off-axis stretching or
extensional stiffness matrix of the lamina. From (4.1.40) it follows that
ε
ε
ε
=A
A
A1N
N
N=a
a
aN
N
N,a
a
a=A
A
A1=S
S
Sh1(4.1.41)
a
a
ais the off-axis in-plane compliance matrix. A
A
Aand a
a
aare, like Q
Q
Qand S
S
Ssymmet-
ric matrices, which have in the general case only non-zero elements. In the special
cases of on-axis reference systems or isotropic stiffness and compliances, respec-
tively, the structure of the matrices is simplified. A
A
Ais the extensional stiffness and a
a
a
the extensional compliance matrix expressing the relationship between the in-plane
stress resultant N
N
Nand the mid-plane strain
ε
ε
ε
:
120 4 Elastic Behavior of Laminate and Sandwich Composites
Off-axis extensional stiffness and compliance matrices
A
A
A=
A11 A12 A16
A12 A22 A26
A16 A26 A66
,a
a
a=
a11 a12 a16
a12 a22 a26
a16 a26 a66
(4.1.42)
On-axis extensional stiffness and compliance matrices
A
A
A=
A11 A12 0
A12 A22 0
0 0 A66
,a
a
a=
a11 a12 0
a12 a22 0
0 0 a66
(4.1.43)
If the stresses are not constant across h, resultant moments can be defined
M
M
M=
h/2
Z
h/2
σ
σ
σ
x3dx3(4.1.44)
The resultant moment vector is denoted by
M
M
M= [M1M2M6]T(4.1.45)
The Miare moments per unit length, M1,M2are bending moments and M6is a
torsional or twisting moment. Figure 4.6 illustrates these moments and a linear stress
distribution across h. The resultant moments yield flexural strains, e.g. bending and
x1
x2
x3
M6
M1
M2
M6
h
Qs
1
Qs
2
Fig. 4.6 Resultant moment vector M
M
MT= [M1M2M6]and transverse shear resultants Q
Q
QsT= [Qs
1Qs
2]
4.1 Elastic Behavior of Laminae 121
twisting strains, which are usually expressed by the relationship
ε
ε
ε
(x1,x2,x3) = x3
κ
κ
κ
,
κ
κ
κ
T= [
κ
1
κ
2
κ
6](4.1.46)
κ
κ
κ
is the vector of curvature,
κ
1,
κ
2correspond to the bending moments M1,M2and
κ
6to the torsion moment M6, respectively. The flexural strains are assumed linear
across h. With Qi j =const across h,i,j=1,2,6 follow
M
M
M=Q
Q
Q
κ
κ
κ
h/2
Z
h/2
x2
3dx3=Q
Q
Q
κ
κ
κ
h3
12 =D
D
D
κ
κ
κ
,
κ
κ
κ
=D
D
D1M
M
M=d
d
dM
M
M(4.1.47)
D
D
D=Q
Q
Qh3/12 is the flexural stiffness matrix and d
d
dthe flexural compliance matrix
expressing the relations between stress couples M
M
Mand the curvatures. For off-axis
and on-axis reference systems the matrices are given by:
Off-axis flexural stiffness and compliance matrices
D
D
D=
D11 D12 D16
D12 D22 D26
D16 D26 D66
,d
d
d=
d11 d12 d16
d12 d22 d26
d16 d26 d66
(4.1.48)
On-axis flexural stiffness and compliance matrices
D
D
D=
D11 D12 0
D12 D22 0
0 0 D66
,d
d
d=
d11 d12 0
d12 d22 0
0 0 d66
(4.1.49)
The transverse shear resultants can be defined in the same way by
Q
Q
Qs=Qs
1
Qs
2=
h/2
Z
h/2
σ
s
5
σ
s
4dx3(4.1.50)
Q
Q
Qsis (like the in-plane resultants) a load vector per unit length in the cross section
of the lamina x1=const or x2=const, respectively. The transverse shear resultant
vector Q
Q
Qsis written with a superscript sto distinguish it from the reduced stiffness
matrix Q
Q
Q. When modelling a plane stress state, there are no constitutive equations
for
σ
4,
σ
5and the shearing stresses are calculated with the help of the equilibrium
equations, Eq. (2.2.1), or with help of equilibrium conditions of stress resultant,
e.g. Chap. 8. In three-dimensional modelling, including transverse shear strains,
however, constitutive equations for transverse shear resultant can be formulated.
For a lamina with resultant forces N
N
Nand moments M
M
Mthe in-plane strain
ε
ε
ε
and
the curvature term
κ
κ
κ
have to be combined
ε
ε
ε
(x1,x2,x3) =
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2)(4.1.51)
122 4 Elastic Behavior of Laminate and Sandwich Composites
For the stress vector is
σ
σ
σ
(x1,x2,x3) = Q
Q
Q[
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2)]
=Q
Q
Q
ε
ε
ε
(x1,x2) + Q
Q
Qx3
κ
κ
κ
(x1,x2)(4.1.52)
and by integration through the lamina thickness hfollow
N
N
N=
h/2
Z
h/2
σ
σ
σ
dx3=Q
Q
Q
ε
ε
ε
h+Q
Q
Q
κ
κ
κ
h/2
Z
h/2
x3dx3=A
A
A
ε
ε
ε
+B
B
B
κ
κ
κ
,
M
M
M=
h/2
Z
h/2
σ
σ
σ
x3dx3=Q
Q
Q
ε
ε
ε
h/2
Z
h/2
x3dx3+Q
Q
Q
κ
κ
κ
h3
12 =B
B
B
ε
ε
ε
+D
D
D
κ
κ
κ
(4.1.53)
The coupling stiffness matrix B
B
Bis zero for a lamina, which is symmetric to the
midplane x3=0, i.e. there are no coupling effects between the N
N
Nand
κ
κ
κ
or M
M
Mand
ε
ε
ε
, respectively. In Table 4.3 the constitutive equations of the lamina resultants are
summarized for a symmetric general angle lamina, for a UD-orthotropic lamina and
for an isotropic layer. In a contracted vector-matrix notation, we can formulate the
constitutive equation of a lamina by
N
N
N
M
M
M=A
A
A0
0
0
0
0
0D
D
D
ε
ε
ε
κ
κ
κ
,(4.1.54)
where the in-plane stiffness submatrix A
A
A=Q
Q
Qh and the plate stiffness submatrix
D
D
D=Q
Q
Qh3/12. 0
0
0 are zero submatrices. The inverted form of (4.1.54) is important for
stress analysis
ε
ε
ε
κ
κ
κ
=a
a
a0
0
0
0
0
0d
d
dN
N
N
M
M
M,a
a
a=A
A
A1,
d
d
d=D
D
D1,
A
A
A=Q
Q
Qh,
D
D
D=Q
Q
Q(h3/12)(4.1.55)
Equation (4.1.52) yields the stress components
σ
i,i=1,2,6
σ
1=Q11(
ε
1+x3
κ
1) + Q12(
ε
2+x3
κ
2) + Q16(
ε
6+x3
κ
6) =
σ
1M +
σ
1B,
σ
2=Q21(
ε
1+x3
κ
1) + Q22(
ε
2+x3
κ
2) + Q26(
ε
6+x3
κ
6) =
σ
2M +
σ
2B,
σ
6=Q16(
ε
1+x3
κ
1) + Q62(
ε
2+x3
κ
2) + Q66(
ε
6+x3
κ
6) =
σ
6M +
σ
6B
σ
iMare the membrane or in-plane stresses coupled with Niand
σ
iBthe curvature or
plate stresses coupled with Mi. The membrane stresses are constant and the bending
stresses linear through the lamina thickness (Fig. 4.7).
The transverse shear stresses
σ
4,
σ
5for plane stress state condition are ob-
tained by integration of the equilibrium equations (2.2.1). If the volume forces
p1=p2=0, Eqs. (2.2.1) yield
σ
5(x3) =
x3
Z
h/2
σ
1
x1
+
σ
6
x2dx3,
σ
4(x3) =
x3
Z
h/2
σ
6
x1
+
σ
2
x2dx3(4.1.56)
4.1 Elastic Behavior of Laminae 123
Table 4.3 Stiffness matrices of laminae
Anisotropic single layer or UD-lamina, off-axis
N1
N2
N6
M1
M2
M6
=
A11 A12 A16 000
A12 A22 A26 000
A16 A26 A66 000
0 0 0 D11 D12 D16
0 0 0 D12 D22 D26
0 0 0 D16 D26 D66
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6
,
Ai j =Qi j h,
Di j =Qi j
h3
12
Orthotropic single layer or UD-laminae, on-axis
N1
N2
N6
M1
M2
M6
=
A11 A12 0 0 0 0
A12 A22 0 0 0 0
0 0 A66 000
0 0 0 D11 D12 0
0 0 0 D12 D22 0
0 0 0 0 0 D66
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6
,
Q11 =E1
1
ν
12
ν
21
,
Q22 =E2
1
ν
12
ν
21
,
Q12 =
ν
12E2
1
ν
12
ν
21
=
ν
21E1
1
ν
12
ν
21
,
Q66 =G12
Isotropic single layer
N1
N2
N6
M1
M2
M6
=
A11 A12 0 0 0 0
A12 A11 0 0 0 0
0 0 A66 000
0 0 0 D11 D12 0
0 0 0 D12 D11 0
0 0 0 0 0 D66
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6
,
Q11 =E
1
ν
2,
Q12 =
ν
E
1
ν
2,
Q66 =E
2(1+
ν
)
and with Eq. (4.1.52) follows
σ
5(x3) =
x3
Z
h/2
x1
[Q11(
ε
1+x3
κ
1) + Q12(
ε
2+x3
κ
2) + Q16(
ε
6+x3
κ
6)]
+
x2
[Q61(
ε
1+x3
κ
1) + Q62(
ε
2+x3
κ
2) + Q66(
ε
6+x3
κ
6)]dx3,
(4.1.57)
124 4 Elastic Behavior of Laminate and Sandwich Composites
h
0 0 0
σ
iM
σ
iB
σ
i=
σ
iM+
σ
iB
N6
h=
σ
6M
N2
h=
σ
2M
N1
h=
σ
1M
M6
h3/12 x3=
σ
6B
M2
h3/12 x3=
σ
2B
M1
h3/12 x3=
σ
1B
σ
6=
σ
6M +
σ
6B
σ
2=
σ
2M +
σ
2B
σ
1=
σ
1M +
σ
1B
Fig. 4.7 In-plane membrane stresses
σ
iM, bending stresses
σ
iBand total stresses
σ
iacross h(qual-
itative)
σ
4(x3) =
x3
Z
h/2
x1
[Q61(
ε
1+x3
κ
1) + Q62(
ε
2+x3
κ
2) + Q66(
ε
6+x3
κ
6)]
+
x2
[Q21(
ε
1+x3
κ
1) + Q22(
ε
2+x3
κ
2) + Q26(
ε
6+x3
κ
6)]dx3
Substituting the midplane strain
ε
ε
ε
and the curvature
κ
κ
κ
by the resultants N
N
Nand M
M
M
Eqs. (4.1.57) takes the form
σ
5(x3) =
x3
Z
h/2nQ11
x1
[a11N1+a12N2+a16N6+x3(d11M1+d12M2+d16M6)]
+Q12
x1
[a21N1+a22N2+a26N6+x3(d21M1+d22M2+d26M6)]
+Q16
x1
[a61N1+a62N2+a66N6+x3(d61M1+d62M2+d66M6)]
+Q61
x2
[a11N1+a12N2+a16N6+x3(d11M1+d12M2+d16M6)]
+Q62
x2
[a21N1+a22N2+a26N6+x3(d21M1+d22M2+d26M6)]
4.1 Elastic Behavior of Laminae 125
+Q66
x2
[a16N1+a26N2+a66N6+x3(d16M1+d26M2+d66M6)]odx3,
(4.1.58)
σ
4(x3) =
x3
Z
h/2nQ61
x1
[a11N1+a12N2+a16N6+x3(d11M1+d12M2+d16M6)]
+Q62
x1
[a21N1+a22N2+a26N6+x3(d21M1+d22M2+d26M6)]
+Q66
x1
[a61N1+a62N2+a66N6+x3(d61M1+d62M2+d66M6)]
+Q21
x2
[a11N1+a12N2+a16N6+x3(d11M1+d12M2+d16M6)]
+Q22
x2
[a21N1+a22N2+a26N6+x3(d21M1+d22M2+d26M6)]
+Q26
x2
[a61N1+a62N2+a66N6+x3(d61M1+d62M2+d66M6)]odx3
The distribution of the transverse shear stresses through the thickness his obtained
by integration from the bottom surface x3=h/2 of the lamina to x3
x3
Z
h/2
Q
Q
Qdx3=˜
A
A
A(x3) = Q
Q
Q(x3+h/2),
x3
Z
h/2
Q
Q
Qx3dx3=˜
B
B
B(x3) = Q
Q
Q1
2x2
3h2
4
(4.1.59)
and we have
˜
A
A
A(h/2) = ˜
B
B
B(h/2) = 0
0
0,˜
A
A
A(h/2)A
A
A,˜
B
B
B(h/2)B
B
B
Finally, the transverse shear stress equations (4.1.57) take the form
σ
5(x3) =˜
A11(x3)
ε
1
x1
+˜
A12(x3)
ε
2
x1
+˜
A16(x3)
ε
6
x1
+˜
B11(x3)
κ
1
x1
+˜
B12(x3)
κ
2
x1
+˜
B16(x3)
κ
6
x1
+˜
A61(x3)
ε
1
x2
+˜
A62(x3)
ε
2
x2
+˜
A66(x3)
ε
6
x2
+˜
B61(x3)
κ
1
x2
+˜
B62(x3)
κ
2
x2
+˜
B66(x3)
κ
6
x2,
126 4 Elastic Behavior of Laminate and Sandwich Composites
σ
4(x3) =˜
A61(x3)
ε
1
x1
+˜
A62(x3)
ε
2
x1
+˜
A66(x3)
ε
6
x1
+˜
B61(x3)
κ
1
x1
+˜
B62(x3)
κ
2
x1
+˜
B66(x3)
κ
6
x1
+˜
A21(x3)
ε
1
x2
+˜
A22(x3)
ε
2
x2
+˜
A26(x3)
ε
6
x2
+˜
B21(x3)
κ
1
x2
+˜
B22(x3)
κ
2
x2
+˜
B26(x3)
κ
6
x2
or an analogous equation to (4.1.59) by substituting the in-plane strains
ε
ε
ε
and the
mid-plane curvatures by the stress resultants N
N
Nand M
M
M.
The transverse shear stresses
σ
4,
σ
5are parabolic functions across the lamina
thickness. If there are no surface edge shear stresses, the conditions
σ
5(h/2) =
σ
4(h/2) = 0 are controlling the performance of the equilibrium equations and the
accuracy of the stress analysis.
4.1.4 Problems
Exercise 4.1. For a single layer unidirectional composite, the on-axis elastic behav-
ior is given by EL=140 GPa, ET=9 GPa,
ν
LT =0.3. Calculate the reduced stiffness
matrix Q
Q
Qand the reduced compliance matrix S
S
S.
Solution 4.1. ELE
1,ETE
2,
ν
LT =
ν
12. Equation (4.1.3) yields
σ
1
σ
2
σ
6
=
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q
66
ε
1
ε
2
ε
6
,
ε
1
ε
2
ε
6
=
S
11 S
12 0
S
12 S
22 0
0 0 S
66
σ
1
σ
2
σ
6
,
Q
11 =E
1/(1
ν
12
ν
21),S
11 =1/E
1,
Q
22 =E
2/(1
ν
12
ν
21),S
22 =1/E
2,
Q
12 =E
2
ν
12/(1
ν
12
ν
21),S
12 =
ν
12/E
2,
Q
66 =G
12 =E
2/2(1+
ν
12),S
66 =1/G
12,
ν
21 =
ν
12E
2/E
1=0,0192,G
12 =E
1/2(1+
ν
12) = 53,846 GPa,
Q
11 =140,811 GPa, Q
22 =9,052 GPa,
Q
12 =2,716 GPa, Q
66 =53,846 GPa,
S
11 =7,143 103GPa1,S
22 =111,111 103GPa1,
S
12 =2,143 103GPa1,S
66 =18,571 103GPa1,
4.1 Elastic Behavior of Laminae 127
Q
Q
Q=
140,811 2,716 0
2,716 9,052 0
0 0 53,846
GPa,
S
S
S=
7,143 2,143 0
2,143 111,111 0
0 0 18,576
103GPa1
Exercise 4.2. A composite panel is designed as a single layer lamina with the fol-
lowing properties EL=140 GPa, ET=10 GPa, GLT =6,9 GPa,
ν
LT =0,3 and
θ
=450. Calculate the strains
ε
1,
ε
2and
ε
6when the panel is loaded by a shear stress
σ
6=±
τ
=±10 MPa.
Solution 4.2. From Fig. 4.4 follows
ε
1=±S16
σ
6,
ε
2=±S26
σ
6,
ε
6=±S66
σ
6
With ELE
1,ETE
2,GLT G
12 and
ν
LT
ν
12 and Eq. (4.1.17) is
S
11 =1/E
1=7,143 103GPa1,
S
22 =1/E
2=100,000 103GPa1,
S
12 =
ν
12/E
1=2,143 103GPa1,
S
66 =1/G
12 =144,928 103GPa1
The transformation rule, Table 4.2, yields
S16 =S26 =0,46 101GPa1,S66 =1,11 101GPa1
The strains are
ε
1=0,46 ·103,
ε
2=
ε
1,
ε
6=±1,11 ·103.
Conclusion 4.1. A positive shear load
σ
6= +
τ
shortens the composite panel in both
directions, a negative shear load
σ
6=
τ
would enlarge the panel in both directions.
Exercise 4.3. In a unidirectional single layer is a strain state
ε
11 =1% =102,
ε
22 =0.5% =0.5 102,
γ
12 =2% =2 102. In the principal directions, the
following engineering parameters of the composite material are E
1=EL=40 GPa,
E
2=ET=10 GPa, G12 =GLT =5 GPa,
ν
LT =0.3. Determine the plane stress state
for the axis (x1,x2)and (x
1,x
2)(L,T)and
θ
=450.
Solution 4.3. The stress states
σ
σ
σ
and
σ
σ
σ
are to calculate for a given strain state
ε
1=102,
ε
2=0,5 102,
ε
6=2 102for a UD lamina with E
1=40 GPa,
E
2=10 GPa, G
12 =5 GPa,
ν
12 =0,3 and a fibre angle
θ
=450. With Table
4.1 the strains for the off-axis system x1,x2are transferred to the strains for the on-
axis system x
1,x
2. Taking into account cos450=sin 450=2/2=0,707107 we
obtain
ε
1
ε
2
ε
6
=
c2s2sc
s2c2sc
2sc 2sc c2s2
ε
1
ε
2
ε
6
=
0,5 0,5 0,5
0,5 0,50,5
1 1 0
10
5
20
103=
12,5
7,5
15
103,
128 4 Elastic Behavior of Laminate and Sandwich Composites
ε
1=12,5 103,
ε
2=7,5 103,
ε
6=15 103
and
ν
21 =
ν
12E
2
E
1
=0,075
The reduced stiffness Q
i j and the stresses
σ
iin the on-axis system follow from
(4.1.3)
Q
11 =E
1/(1
ν
12
ν
21) = 40,9207 GPa, Q
22 =E
2/(1
ν
12
ν
21) = 10,2302 GPa,
Q
12 =E
2
ν
12/(1
ν
12
ν
21) = 3,0691 GPa, Q
66 =G
12 =5,0 GPa,
σ
1
σ
2
σ
6
=
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q
66
ε
1
ε
2
ε
6
=
488,5
38,36
75,0
103GPa
The stresses
σ
iin the off-axis system are calculated with the help of the transforma-
tion rules Table 4.1
σ
1
σ
2
σ
6
=
c2s22sc
s2c22sc
sc sc c2s2
σ
1
σ
2
σ
6
=
0,5 0,51
0,5 0,5 1
0,50,5 0
488,5
38,36
75,0
103GPa =
0,300
0,150
0,263
GPa
Exercise 4.4. Sketch the variation curves E1/E
2and G12/E
2against the fibre ori-
entation
θ
for a carbon-epoxy and glass-epoxy lamina using the following material
data:
carbon-epoxy E
1=140 GPa, E
2=10 GPa, G
12 =7 GPa,
ν
12 =0,3
glass-epoxy E
1=43 GPa, E
2=9 GPa, G
12 =4,5 GPa,
ν
12 =0,27
Discuss the curves.
Solution 4.4. From (4.1.23) follows
(E1)1=c4
E
1
+1
G
12 2
ν
12
E
2s2c2+s4
E
2
,
(G12)1=22
E
1
+2
E
2
+4
ν
12
E
11
G
12 s2c2+1
G
12
(c4+s4)
Now the functions f1(
θ
) = E1(
θ
)/E
2and f2(
θ
) = G12(
θ
)/E
2can be sketched. The
results are shown on Figs. 4.8 and 4.9. Discussion of the functions f1(
θ
)and f2(
θ
):
1. The anisotropic ratio E1/E
2is higher for carbon- than for glass-epoxy.
2. The longitudinal effective modulus E1of the lamina drops sharply as the loading
direction deviates from the fibre direction, especially for-carbon-epoxy.
3. The effective shear modulus of the lamina attains a maximum value at
θ
=450.
4.1 Elastic Behavior of Laminae 129
Fig. 4.8 Variation of
E1(
θ
)/E
2against the fibre
orientation for two compos-
ites
carbon-epoxy glass-epoxy
0
010 20 30 40 50 60 70 80 90
2
4
6
8
10
12
14
E1(
θ
)/E
2
θ
Fig. 4.9 Variation of
G12(
θ
)/E
2against the fibre
orientation for two compos-
ites
carbon-epoxy glass-epoxy
010 20 30 40 50 60 70 80 90
0.5
0.6
0.7
0.8
0.9
G12(
θ
)/E
2
θ
Exercise 4.5. For a UD-lamina with the elastic properties E
1=180 GPa,
E
2=10 GPa, G
12 =7 GPa,
ν
12 =0,3 calculate
1. the on-axis compliances S
i j and the on-axis strains, if the applied on-axis stresses
are
σ
1=2 MPa,
σ
2=3 MPa,
σ
6=4 MPa,
2. the off-axis compliances Si j and the off-axis and on-axis strains
ε
ε
ε
,
ε
ε
ε
if
θ
=450
(Fig. 4.3) and the applied off-axis stresses are
σ
1=2 MPa,
σ
2=3 MPa,
σ
6=4 MPa,
3. the coefficients of thermal expansion in the off-axis system if
α
th
1=9 106/0K,
α
th
2=22 106/0K.
Solution 4.5. 1. Using (4.1.3) follows
130 4 Elastic Behavior of Laminate and Sandwich Composites
S
11 = (E
1)1=5,556 1012 Pa1,
S
12 =
ν
12/E
1=1,667 1012 Pa1,
S
22 = (E
2)1=100 1012 Pa1,
S
66 = (G
12)1=142,86 1012 Pa1,
ε
1
ε
2
ε
6
=
5,556 1,667 0
1,667 100 0
0 0 142,86
1012Pa1
2
3
4
106Pa
=
16,113
303,334
571,440
106
2. With Table 4.2 the transformed compliances Si j can be calculated (note that c=
cos450=0.7071,s=sin 450=0.7071)
S11 =61,270 1012 Pa1,S22 =61,271 1012 Pa1,
S12 =10,160 1012 Pa1,S66 =108,89 1012 Pa1,
S16 =S26 =47,222 1012 Pa1
Equations (4.1.19) and (4.1.20) yield the strains
ε
i
ε
1
ε
2
ε
6
=
S11 S12 S16
S12 S22 S26
S16 S26 S66
σ
1
σ
2
σ
6
=
61,270 10,160 47,222
10,160 61,271 47,222
47,222 47,222 108,89
1012Pa1
2
3
4
106Pa
=
35,868
393,01
482,782
106
Using Table 4.1 the strains
ε
ε
ε
can be transformed to the strains
ε
ε
ε
ε
1
ε
2
ε
6
=
0.5 0.5 0.5
0.5 0.50.5
1 1 0
ε
1
ε
2
ε
6
=
0.5 0.5 0.5
0.5 0.50.5
1 1 0
35,868
393,01
482,782
106
=
26,95
455,83
357,14
106
3. The transformed thermal expansion coefficients
α
th
ifollow like the strains with
Table 4.1 to
4.2 Elastic Behavior of Laminates 131
α
th
1
α
th
2
α
th
6
=
0.5 0.50.5
0.5 0.5 0.5
11 0
α
th
1
α
th
2
α
th
6
=
15,5
15,5
13,0
106/K
Note that in the off-axis system
α
th
66=0.
Exercise 4.6. The micro-mechanical material parameters of a carbon-epoxy com-
posite are
ELf =411 GPa, ETf =6.6 GPa,
ν
TLf =0.06,
ν
LTf =0.35,
α
th
Lf =1.2 1061/K,
α
th
Tf =27.3 1061/K,
Em=5.7 GPa,
ν
m=0.316,
α
m=45 1061/K, vf=0.5
The experimental tested values of the lamina are
EL=208.6 GPa, ET=6.3 GPa,
ν
LT =0.33,
α
th
L=0.5 1061/0C,
α
th
T=29.3 1061/0C,
Predict the lamina values using the micro-mechanical modelling and compare the
calculated and the experimental measured values.
Solution 4.6. Using Eqs. (3.1.27) and (4.1.9)
EL=vfELf + (1vf)Em=208.35 GPa,
ν
LT =vf
ν
LTf + (1vf)
ν
m=0.33,
ET=ETfEm/[vfEm+ (1vf)ETf ] = 6.12 GPa,
α
th
L= [
α
th
LfvfELf +
α
th
m(1vf)Em]/EL=0.57 1061/K,
α
th
T= (
α
th
Tf +
ν
Tf
α
th
Lf)vf+ (1+
ν
m)
α
th
m(1vf)
ν
LT
α
th
L=4.43 1061/K
It can be concluded that the simple rules of mixture providing proper results for lon-
gitudinal material characteristics EL,
ν
LT and
α
th
L. In this case also ETis predicted
quite well but the formula for
α
Tfails to predict the transverse thermal expansion
coefficient with required accuracy. For engineering applications the thermal expan-
sion coefficients should be normally determined by experimental methods.
4.2 Elastic Behavior of Laminates
In Sect. 4.1, stress-strain equations were developed for a single lamina. Mostly im-
portant in engineering applications are isotropic, quasi-isotropic (stochastic distri-
bution of short fibres or particles) and quasi-orthotropic (unidirectional fibre re-
inforced) laminae. Reduced stiffness Qi j , compliances Si j, membrane or in-plane
stiffness Ai j and plate or out-of-plane stiffness Di j were defined. Assuming symme-
try about the midplane of a lamina in-plane and plate responses are uncoupled in the
form of a first order theory (linear force-displacement relations).
The mechanics of laminated composite materials is generally studied at two dis-
tinct levels, commonly called micromechanics and macromechanics. In Chap. 3 the
micromechanics was used to study the interaction between the fibres and matrix in a
132 4 Elastic Behavior of Laminate and Sandwich Composites
lamina such that the mechanical behavior of the lamina could be predicted from the
known behavior of the constituents. Micromechanics establishes the relationship be-
tween the properties of the constituents and those of the lamina. All micromechanics
approaches suffer from the problem of measuring the material properties of the con-
stituents and generally require correction factors to correlate with measured lamina
properties. For most engineering design applications an analysis that addressed to
the micro-mechanical level is unrealistic.
At the macro-mechanical level the properties of the individual layers of a lami-
nate are assumed to be known a priori. Macromechanics involves investigation of the
interaction of the individual layers of a laminate with one another and their effects
on the overall response quantities, e.g. elastic stiffness, influence of temperature and
moisture on the response of laminated composites, etc. Such global response quan-
tities can be predicted well on this level. Thus, the use of macromechanical formu-
lations in designing composite laminates for desired material characteristics is well
established. Macromechanics is based on continuum mechanics, which models each
lamina as homogeneous and orthotropic and ignoring the fibre/matrix interface.
The lamination theory is the mathematical modelling to predict the macro-
mechanical behavior of a laminate based on an arbitrary assembly of homoge-
neous orthotropic laminae. A two-dimensional modelling is most common, a three-
dimensional theory is very complex and should be limited to selected problems, e.g.
the analysis of laminates near free edges.
A real structure generally will not consist of a single lamina. A laminate consist-
ing of more than one lamina bonded together through their thickness, for a single
lamina is very thin and several laminae will be required to take realistic structural
loads. Furthermore the mechanical characteristics of a UD-lamina are very limited
in the transverse direction and by stacking a number of UD-laminae it may be an
optimal laminate for unidirectional loading only. One can overcome this restriction
by making laminates with layers stacked at different fibre angles corresponding to
complex loading and stiffness requirements. To minimize the increasing costs and
weights for such approach one have to optimize the laminae angles. It may be also
useful to stack layers of different composite materials.
4.2.1 General Laminates
In the following section the macro-mechanical modelling and analysis of laminates
will be considered. The behavior of a multidirectional laminate is a function of the
laminae properties, i.e. their elastic moduli, thickness, angle orientations, and the
stacking sequence of the individual layers. The macro-mechanical modelling may
be in the framework of the following assumptions:
There is a monolithic bonding of all laminae i.e. there is no slip between laminae
at their interface and no special interface layers are arranged between the angle
plies.
4.2 Elastic Behavior of Laminates 133
Each layer is quasi-ho mogeneous and orthotropic, but the an gle orientations may
be different.
The strains and displacements are continuous throughout the laminate. The in-
plane displacements and strains vary linearly through the laminate thickness.
We will see that the stacking codes of laminates have a great influence on the global
mechanical laminate response (Sect. 4.2.3), but there are some rules to guarantee an
optimal global laminate behavior:
Symmetric laminate stacking yields an uncoupled modelling and analysis of in-
plane and bending/torsion stress-strain relations and avoids distorsions in the pro-
cessing.
Laminates should be made up of at least three UD-laminae with different fibre
angle orientation.
The differences of the mechanical properties and the fibre orientations between
two laminae following in the stacking sequence should not be so large that the
so-called interlaminar stresses are small.
Although it is possible to determine an optimum orientation sequence of lami-
nates for any given load condition, it is more practical from a fabrication stand-
point and from effective experimental lamina testing to limit the number of fibre
orientations to a few specific laminae types, e.g. fibre orientations of 00,±450
and 900, etc.
Consider a laminate made of nplies shown in Fig. 4.10. Each lamina has a thickness
of h(k),k=1,2,...,n, and we have
x3
x1
x2
Mid-plane x3=0
1
2
n
h
2
h
2
x(2)
3
x(1)
3
x(0)
3=h
2
x(n1)
3
x(n)
3=h
2
.
.
.
.
.
.
Fig. 4.10 Laminate made of nsingle layers, coordinate locations
134 4 Elastic Behavior of Laminate and Sandwich Composites
h(k)=x(k)
3x(k1)
3,k=1,2,...,nthickness of a lamina,
h=
n
k=1
h(k)thickness of the laminate,
x(k)
3=h
2+
k
i=1
h(i)distance from the mid-plane,
(4.2.1)
Then
x(n)
3=h
2and x(0)
3=h
2
are the coordinates of the top and the bottom surface of the laminate,
x(k)
3and x(k1)
3,k=1,2,...,n
are the location coordinates of the top and the bottom surface of lamina k. Each
layer of a laminate can be identified by its location in the laminate, its material and
its fibre orientation. The layers of the laminate may be symmetric, antisymmetric
or asymmetric to the midplane x3=0. h(k)and the reduced stiffness Q
Q
Q(k)may be
different for each lamina, but Q
Q
Q(k)is constant for the kth lamina. The following
examples illustrate the laminate code. In Fig. 4.11 the laminate codes for an un-
symmetric laminate with four layers and a symmetric angle-ply laminate with eight
layers are illustrated. A slash sign separates each lamina. The codes in Fig. 4.11
imply that each lamina is made of the same material and is of the same thickness.
Regular symmetric are those laminates which have an odd number of UD-laminae
of equal thicknesses and alternating angle orientations (Fig. 4.12). Since the number
of laminae is odd and symmetry exists at the mid-surface, the 900lamina is denoted
with a bar on the top. The subscript Soutside the code brackets, e.g., in Fig. 4.11 b),
represents that the four plies are repeated in the reverse order.
x3=0x3=0
1
2
3
4
a
θ
=00
θ
=300
θ
=300
θ
=900
b
θ
=00
θ
=450
θ
=900
θ
=450
θ
=450
θ
=900
θ
=450
θ
=00
8
7
6
5
4
3
2
1
[90/30/30/0][0/45/90/45/45/90/45/0]
[0/45/90/45]S
Fig. 4.11 Angle-ply laminates. aunsymmetric 4-layer laminate, bsymmetric 8-layer-laminate
4.2 Elastic Behavior of Laminates 135
x3=0
5
4
3
2
1
θ
=450
θ
=450
θ
=900
θ
=450
θ
=450
[45/45/90/45/45][45/45/¯
90]S
Fig. 4.12 Regular symmetric angle-ply laminate: orientation of the midplane
θ
=900
A general laminate has layers of different orientations
θ
with 900
θ
900.
An angle-ply laminate has ply orientations of
θ
and
θ
with 00
θ
900and at
least one lamina has an orientation other than 00or 900. Cross-ply laminates are
those which have only ply orientations of 00and 900.
A laminate is balanced when it consists of pairs of layers with identical thickness
and elastic properties but have +
θ
and
θ
orientations of their principal material
axes with respect to the laminate reference axes. A balanced laminate can be sym-
metric, antisymmetric or asymmetric
[+
θ
1/
θ
1/+
θ
2/
θ
2]Ssymmetric lay-up,
[
θ
1/
θ
2/
θ
2/
θ
1]antisymmetric lay-up,
[
θ
1/
θ
2/
θ
1/
θ
2]asymmetric lay-up
(4.2.2)
Antisymmetric laminates are a special case of balanced laminates, having the bal-
anced +
θ
and
θ
pairs of layers symmetrically situated about the middle surface.
Generally each layer of a laminate can have different fibre angles, different thick-
nesses and different composite materials. The influence of the laminate codes, i.e
the properties and the stacking sequences, on the elastic behavior of laminates will
be considered in Sect. 4.2.3.
4.2.2 Stress-Strain Relations and Stress Resultants
The stiffness matrix of a single lamina referred to the reference system xi,i=1,2,3,
has been formulated in Sect. 4.1.2, Eq. (4.1.26). Extending the assumption of a
plane stress state to laminates with in-plane and out-of-plane loading, the stress-
strain relation (4.1.26) can be rewritten by separating the transverse shear stresses
and strains. The stresses in the kth layer are expressed by means of the reduced
stiffness coefficients Qi j
136 4 Elastic Behavior of Laminate and Sandwich Composites
σ
1
σ
2
σ
6
σ
4
σ
5
(k)
=
Q11 Q12 Q16 0 0
Q21 Q22 Q26 0 0
Q61 Q62 Q66 0 0
0 0 0 Q44 Q45
0 0 0 Q54 Q55
(k)
ε
1
ε
2
ε
6
ε
4
ε
5
(k)
(4.2.3)
or in contracted notation
σ
(k)
i=Q(k)
i j
ε
(k)
j,i,j=1,2,6,
σ
(k)
i=Q(k)
i j
ε
(k)
j,i,j=4,5
(4.2.4)
with (see also 2.1.78)
σ
(k)
3=0,
ε
(k)
3=1
C33
(C13
ε
1+C23
ε
2+C36
ε
6)(4.2.5)
Q(k)
i j ,i,j=1,2,6 are the reduced stiffness of the kth layer and functions of Q
i j
(k)and
the fibre orientation angle, the Q(k)
44 ,Q(k)
45 =Q(k)
54 ,Q(k)
55 are identical to the material co-
efficients C(k)
44 ,C(k)
45 =C(k)
54 ,C(k)
55 , which are not reduced by the assumption of a plane
stress state. The discontinuity of Q(k)
i j from layer to layer implies the discontinuity
of the stresses when passing from one lamina to another.
From the assumption of macro-mechanical modelling of laminates it follows that
ε
ε
ε
(x1,x2,x3) =
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2)(4.2.6)
i.e the strains
ε
1,
ε
2,
ε
6vary linearly through the laminate thickness.
ε
ε
ε
(x1,x2)is
the vector of the in-plane or membrane strains and x3
κ
κ
κ
(x1,x2)the vector of flex-
ural strains (bending and twisting).
κ
κ
κ
(x1,x2)is the vector of curvature subjected to
bending and twisting. We shall see later (Sect. 5.4) that there are different curvature
components in the classical and the shear deformation theory of laminates.
The in-plane stress resultant force vector N
N
Nof a laminate follows by summarizing
the adequate vectors of all laminae
N
N
N=
n
k=1
N
N
N(k),N
N
NT= [N1N2N6],N
N
N(k)T= [N(k)
1N(k)
2N(k)
6](4.2.7)
By analogy it follows that the resultant moment vector is
M
M
M=
n
k=1
M
M
M(k),M
M
MT= [M1M2M6],M
M
M(k)T= [M(k)
1M(k)
2M(k)
6](4.2.8)
The positive directions are corresponding to Figs. 4.5 and 4.6 for a single layer. The
transverse shear resultants given in (4.1.50)
4.2 Elastic Behavior of Laminates 137
Q
Q
Qs(x1,x2) =
n
k=1
Q
Q
Qs(k),Q
Q
QsT= [Qs
1Qs
2],Q
Q
Qs(k)T= [Qs(k)
1Qs(k)
2](4.2.9)
Equations (4.2.4) and (4.2.6) yield
σ
σ
σ
(k)=Q
Q
Q(k)
ε
ε
ε
=Q
Q
Q(k)(
ε
ε
ε
+x3
κ
κ
κ
)(4.2.10)
and the resultants N
N
Nand M
M
Mfor the laminate are (k=1,2,...,n)
N
N
N(k)=Z
(h(k))
σ
σ
σ
(k)dx3=
σ
σ
σ
(k)h(k),h(k)=x(k)
3x(k1)
3,
N
N
N=
n
k=1
σ
σ
σ
(k)h(k)
(4.2.11)
and
M
M
M(k)=Z
(h(k))
σ
σ
σ
(k)x3dx3=
σ
σ
σ
(k)1
2x(k)2
3x(k1)2
3=
σ
σ
σ
(k)h(k)x(k),
M
M
M=
n
k=1
σ
σ
σ
(k)h(k)x(k)
(4.2.12)
with
x(k)=1
2x(k)
3+x(k1)
3
For each layer the membrane strains
ε
1,
ε
2,
ε
6, the curvatures
κ
1,
κ
2,
κ
6and the
reduced stiffness Q(k)
11 ,Q(k)
12 ,Q(k)
16 ,Q(k)
22 ,Q(k)
26 ,Q(k)
66 are constant through each thickness
h(k)and (4.2.11) and (4.2.12) reduces to:
N
N
N=
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
dx3
ε
ε
ε
+
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1
3)
x3dx3
κ
κ
κ
=A
A
A
ε
ε
ε
+B
B
B
κ
κ
κ
,
M
M
M=
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
x3dx3
ε
ε
ε
+
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
x2
3dx3
κ
κ
κ
=B
B
B
ε
ε
ε
+D
D
D
κ
κ
κ
(4.2.13)
A
A
A,B
B
B,D
D
Dare the extensional, coupling and bending stiffness matrices, respectively.
From (4.2.4) and (4.2.9), the relations for the transverse shear resultants are
138 4 Elastic Behavior of Laminate and Sandwich Composites
Q
Q
Qs=
n
k=1
C
C
Cs(k)
x(k)
3
Z
x(k1)
3
σ
σ
σ
s(k)dx3=
n
k=1
C
C
Cs(k)
x(k)
3
Z
x(k1)
3
dx3
γ
γ
γ
s=A
A
As
γ
γ
γ
s(4.2.14)
with
C
C
Cs=C44 C45
C54 C55 ,
σ
σ
σ
s=
σ
4
σ
5,
γ
γ
γ
s=
ε
4
ε
5,A
A
As=A44 A45
A54 A55
Equation (4.2.14) is a first approach and consists of taking the transverse shear strain
independent of the coordinate x3.A
A
Asis the transverse shear stiffness matrix. An
improvement is possible by replacing the transverse shear stiffness As
i j by (kA)s
i j.
ks
i j are so called shear correction factors (Sect. 5.4). The elements of the matrices
A
A
A,B
B
B,D
D
D,A
A
Asare
Ai j =
n
k=1
Q(k)
i j x(k)
3x(k1)
3=
n
k=1
Q(k)
i j h(k),i,j=1,2,6,
Bi j =1
2
n
k=1
Q(k)
i j x(k)
3
2x(k1)
3
2=
n
k=1
Q(k)
i j x(k)
3h(k),
Di j =1
3
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3=
n
k=1
Q(k)
i j x(k)
3
2+h(k)2
12 !h(k),
As
i j =
n
k=1
C(k)
i j x(k)
3x(k1)
3=
n
k=1
C(k)
i j h(k),i,j=4,5
(4.2.15)
The constitutive equation for laminates including extensional, bending/torsion and
transverse shear strains is the superposition from the so-called classical equations
for N
N
Nand M
M
Mand the equation that involves the transverse shear resultant Q
Q
Qs. The
constitutive equation can be written in the following contracted hypermatrix form
N
N
N
M
M
M
Q
Q
Qs
=
A
A
A B
B
B0
0
0
B
B
B D
D
D0
0
0
0
0
0 0
0
0A
A
As
ε
ε
ε
κ
κ
κ
γ
γ
γ
s
(4.2.16)
The stiffness Q(k)
i j and C(k)
i j in (4.2.15) referred to the laminate’s global reference
coordinate system xi,i=1,2,3, are given in Table 4.2 as functions of the Q
i j
(k)and
in (4.2.17) as functions of the C
i j
(k)referred to the material principal directions of
each lamina (k)
C44 =C
44c2+C
55s2,
C45 = (C
55 C
44)sc,
C55 =C
44s2+C
55c2
(4.2.17)
Equation (4.2.16) illustrates the coupling between stretching and bending/twisting
of a laminate, i.e. in-plane strains result in in-plane resultants but also bending
and/or torsion moments and vice versa. Since there are no coupling effects with
4.2 Elastic Behavior of Laminates 139
the transverse shear strains or shear resultants we consider the in-plane and flexural
simultaneous equations, (4.2.18), separately
N
N
N
···
M
M
M
=
A
A
A.
.
.B
B
B
. . . .
B
B
B.
.
.D
D
D
ε
ε
ε
···
κ
κ
κ
,
or
N1
N2
N6
···
M1
M2
M6
=
A11 A12 A16
.
.
.B11 B12 B16
A12 A22 A26
.
.
.B12 B22 B26
A16 A26 A66
.
.
.B16 B26 B66
. . . . . . . . . . . . . .. . . . . . . . . .
B11 B12 B16
.
.
.D11 D12 D16
B12 B22 B26
.
.
.D12 D22 D26
B16 B26 B66
.
.
.D16 D26 D66
ε
1
ε
2
ε
6
···
κ
1
κ
2
κ
6
(4.2.18)
The following steps are necessary for analyzing a laminated composite subjected to
forces and moments:
Calculate the values of the reduced stiffness Q
i j for each lamina kusing the four
elastic moduli, EL,ET,
ν
LT,GLT (4.1.2) and (4.1.3).
Calculate the values of the transformed reduced stiffness Qi j for each lamina k
(Table 4.2).
Knowing the thickness h(k)of each lamina kcalculate the coordinates x(k)
3,x(k1)
3
to the top and the bottom surface of each ply.
Calculate all Ai j ,Bi j and Di j from (4.2.15).
Substitute the calculated stiffness and the applied resultant forces and moments
in (4.2.18) and calculate the midplane strains
ε
iand curvatures
κ
i.
Calculate the global strains
ε
ε
ε
(k)in each lamina using (4.2.6) and then the global
stresses
σ
σ
σ
(k)for each lamina kusing (4.2.10).
Calculate the local strains
ε
ε
ε
,(k)and the local stresses
σ
σ
σ
,(k)for each lamina kusing
Table 4.1
The inverted relation (4.2.18) leads to the compliance hypermatrix for the in-plane
and flexural resultants
ε
ε
ε
κ
κ
κ
=a
a
a b
b
b
c
c
c d
d
dN
N
N
M
M
M,a
a
a b
b
b
c
c
c d
d
d=A
A
A B
B
B
C
C
C D
D
D1
(4.2.19)
The compliance submatrices a
a
a,b
b
b,c
c
c,d
d
dfollow from the stiffness submatrices A
A
A,B
B
B,D
D
D.
With
N
N
N=A
A
A
ε
ε
ε
+B
B
B
κ
κ
κ
(4.2.20)
it follows that
140 4 Elastic Behavior of Laminate and Sandwich Composites
ε
ε
ε
=A
A
A1(N
N
NB
B
B
κ
κ
κ
)(4.2.21)
and using (4.2.13)
M
M
M=B
B
BA
A
A1N
N
N(B
B
BA
A
A1B
B
BD
D
D)
κ
κ
κ
(4.2.22)
The first result is a mixed-type constitutive equation
ε
ε
ε
···
M
M
M
=
A
A
A.
.
.B
B
B
. . . . . .
C
C
C.
.
.D
D
D
N
N
N
···
κ
κ
κ
,
A
A
A=A
A
A1,B
B
B=A
A
A1B
B
B,
C
C
C=B
B
BA
A
A1=B
B
BT,D
D
D=D
D
DB
B
BA
A
A1B
B
B
(4.2.23)
With
κ
κ
κ
=D
D
D1M
M
MD
D
D1C
C
CN
N
N(4.2.24)
it follows that
ε
ε
ε
= (A
A
AB
B
BD
D
D1C
C
C)N
N
N+B
B
BD
D
D1M
M
M(4.2.25)
and the compliance relation has in contracted notation the form
ε
ε
ε
···
κ
κ
κ
=
a
a
a.
.
.b
b
b
. . . .
c
c
c.
.
.d
d
d
N
N
N
···
M
M
M
,
a
a
a=A
A
AB
B
BD
D
D1C
C
C=A
A
A+B
B
BD
D
D1B
B
BT,
b
b
b=B
B
BD
D
D1,
c
c
c=D
D
D1C
C
C=D
D
D1B
B
BT=b
b
bT,
d
d
d=D
D
D1
(4.2.26)
Equations (4.2.18) and (4.2.26) are inverse relations of the constitutive equation for
the resultants and the strains of a laminate. The elements of the submatrices A
A
A,B
B
B,D
D
D
and a
a
a,b
b
b,c
c
c,d
d
dare functions of the geometry, the material properties and the structure
of a laminate and therefore averaged effective elastic laminate moduli. The subma-
trices A
A
A,B
B
B,D
D
D,a
a
a,d
d
dare symmetric submatrices. That is not the case for the submatri-
ces b
b
band c
c
cbut with c
c
c=b
b
bTthe compliance hypermatrix is symmetric. The coupling
of different deformation states is a very important quality of the constitutive equa-
tions of laminates. In the general case, considered in this section, all coupling effects
are present. Figure 4.13 illustrates for example the coupling states for the resultant
force N1and the resultant moment M1.
In the next Sect. 4.2.3 we shall see that the stacking sequence of a laminate in-
fluences the coupling behavior of loaded laminates. In engineering applications it is
desired to specify the stacking sequence such that a number of coefficients of the
stiffness matrix will be zero and undesirable couplings between stretching, bending
4.2 Elastic Behavior of Laminates 141
N1
N2
N6
M1
M2
M6
=
A11 A12 A16 B11 B12 B16
A22 A26 ·B22 B26
A66 · · B66
D11 D12 D16
S Y M ·D22 D26
· · D66
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6
N1
M1
N1
M1
A11 A12 A16
D11 D12 D16
B11 B12 B16
B11 B12 B16
strain
ε
1strain
ε
2shear
ε
6
curvature
κ
1curvature
κ
2twisting
κ
6
curvature
κ
1curvature
κ
2twisting
κ
6
strain
ε
1strain
ε
2shear
ε
6
Fig. 4.13 Coupling of strain states: Influence of the stiffness A1j,D1jand B1j(j=1,2,6)on the
strains
ε
jand the curvature
κ
jof the middle surface of a general laminate loading with N1or M1.
In each case 6 deformation states of N1and M1have to be superposed.
142 4 Elastic Behavior of Laminate and Sandwich Composites
and/or twisting will be avoided. But it is rather difficult to specify an optimum stack-
ing sequence without detailed information about the performance requirements.
Engineering composite design has continued to evolve over many years. Most
early applications of composite materials were aimed particularly at weight reduc-
tion. Metals were replaced by composites with little or no emphasis placed on tailor-
ing the composite properties. Engineering design created quasi-isotropic laminates
that largely suppressed the directional material properties of unidirectional lami-
nae and made the laminate material response similar to that of isotropic materials,
e.g. of metals. We shall see in the following discussion that one of such quasi-
isotropic laminate is given if it has equal percentages of 00,+450,450and 900
layers placed symmetrically with respect to the laminate mid-plane. Quasi-isotropic
laminates have elastic properties that are independent of the direction in the plane
of the laminate, like traditional isotropic engineering materials. Therefore, quasi-
isotropic laminates were in the first applications of composites a convenient replace-
ment for steel or alloys in weight critical applications, e.g. in aerospace industries.
Weight saving could be achieved by simple replacing the isotropic metal with a
similar stiffness laminate that was lighter and probably stronger. If we compare a
graphite/epoxy laminate with an quasi-isotropic stacking sequence of laminae and
aluminium we find nearly the same elastic moduli, e.g. E70 GPa, but the density
values
ρ
and the specific stiffness E/
ρ
differ significantly. The specific stiffness
of graphite/epoxy laminate can be twice that of aluminium. Such applications of
quasi-isotropic laminates required a minimal amount of redesign effort and there-
fore minimal changes in structural configuration.
By the time the number of design engineers which are trained in composite ma-
terials increased and the tailoring of material properties gained more acceptance.
To maximize the utility of the non-isotropic nature of laminates, the influence of
the stacking sequence on the structural behavior must be investigated in detail and
optimized. Particularly the coupling effects of in-plane and out-of-plane responses
affect the effort of laminate structural analysis.
4.2.3 Laminates with Special Laminae Stacking Sequences
Now special cases of laminates which are important in the engineering design of
laminated structures will be introduced. Quite often the design of laminates is done
by using laminae that have the same constituents, the same thicknesses, etc. but
have different orientations of their fibre reinforcement direction with respect to the
global reference system of the laminate and a different stacking sequence of these
layers. In other cases layers with different materials or thicknesses are bonded to a
laminate. The stacking sequence of the layers may result in reducing the coupling
of normal and shear forces, of bending and twisting moments etc. It can simplify
the mechanical analysis but also gives desired mechanical performance. In the fol-
lowing, the mechanical behavior of special symmetric and unsymmetric laminates
are considered.
4.2 Elastic Behavior of Laminates 143
4.2.3.1 Symmetric Laminates
A laminate is called symmetric if the material, angle and thickness of laminae are
the same above and below the midplane, i.e. two symmetric arranged layers to the
midplane have the same reduced stiffness matrix Q
Q
Q(k)Q
Q
Q(k)and the same thickness
h(k)h(k)for opposite coordinates ¯x(k)and ¯x(k)=¯x(k)(Fig. 4.14). It follows that
the coefficients Bi j of the coupling submatrix B
B
Bare zero and there are no coupling
relations of stretching and bending
Bi j =1
2
n
k=1
Q(k)
i j x(k)
3
2x(k1)
3
2
=1
2
n
k=1
Q(k)
i j x(k)
3+x(k1)
3x(k)
3x(k1)
3
=
n
k=1
Q(k)
i j x(k)
3h(k)=0,i,j=1,2,6
(4.2.27)
With x(k)
3=x(k)
3the sum above have two pairs of equal absolute values but oppo-
site signs. Thus the ABD
ABD
ABD-matrix of symmetric laminates is uncoupled, i.e. all terms
of the coupling submatrix [Bi j]are zero, see following equation
x3=0
n1
1
k
k
x(k)
3
x(k)
3
x3
Fig. 4.14 Symmetric laminate with identical layers kand kopposite to the middle surface (h(k)=
h(k),Q
Q
Q(k)=Q
Q
Q(k))
144 4 Elastic Behavior of Laminate and Sandwich Composites
N1
N2
N6
. .
M1
M2
M6
=
A11 A12 A16
.
.
. 0 0 0
A12 A22 A26
.
.
. 0 0 0
A16 A26 A66
.
.
. 0 0 0
........................
0 0 0 .
.
.D11 D12 D16
0 0 0 .
.
.D12 D22 D26
0 0 0 .
.
.D16 D26 D66
ε
1
ε
2
ε
6
. .
κ
1
κ
2
κ
6
(4.2.28)
The extensional submatrix A
A
Aand the bending submatrix D
D
Dare in the case of sym-
metric angle ply laminates fully populated and we have in-plane normal and shear
strain and out-of-plane bending and torsion couplings. Since the coupling submatrix
B
B
Bis zero the elastic behavior of symmetric laminates is simpler to analyze than that
of general laminates and symmetric laminates have no tendency to warp as a result
of thermal contractions induced during the composite processing. Some important
special cases of symmetric laminates are:
Symmetric laminate with isotropic layers
Q(k)
11 =Q(k)
22 =Q(k)
11 =Q(k)
22 =E(k)
1
ν
(k),
Q(k)
12 =Q(k)
12 =
ν
(k)E(k)
1
ν
(k),
Q(k)
16 =Q(k)
26 =Q(k)
16 =Q(k)
26 =0,
Q(k)
66 =Q(k)
66 =E(k)
2(1+
ν
(k))=G(k),
Ai j =
n
k=1
Q(k)
i j h(k)
=A11 =A22,A16 =A26 =0,
Di j =
n
k=1
Q(k)
i j h(k) x(k)
3
2+h(k)2
12 !
=D11 =D22,D16 =D26 =0,
(4.2.29)
4.2 Elastic Behavior of Laminates 145
N1
N2
N6
. .
M1
M2
M6
=
A11 A12 0.
.
. 0 0 0
A12 A11 0.
.
. 0 0 0
0 0 A66
.
.
. 0 0 0
. . . . . . . . . . . . . . .. . . . . . . . .
0 0 0 .
.
.D11 D12 0
0 0 0 .
.
.D12 D11 0
0 0 0 .
.
. 0 0 D66
ε
1
ε
2
ε
6
. .
κ
1
κ
2
κ
6
(4.2.30)
This type of symmetric laminates has no stretching-shearing or bending-torsion
coupling.
Symmetric cross-ply laminate
A laminate is called a cross-ply laminate or a laminate with specially orthotropic
layers if only 00and 900plies were used. The material principal axes and the
global reference axes are identical. If for example for the kth layer the fibre ori-
entation and the x1-direction of the global reference system coincide, we have
Q(k)
11 Q(k)
11 =E(k)
1
1
ν
(k)
12
ν
(k)
21
,Q(k)
22 Q(k)
22 =E(k)
2
1
ν
(k)
12
ν
(k)
21
,
Q(k)
12 Q(k)
12 =
ν
(k)
21 E(k)
1
1
ν
(k)
12
ν
(k)
21
,Q(k)
16 Q(k)
16 =0,
Q(k)
66 Q(k)
66 =G(k)
12 ,Q(k)
26 Q(k)
26 =0
(4.2.31)
and with (4.2.15) A16 =A26 =0,D16 =D26 =0. The stiffness matrix of the
constitutive equation has an adequate structure as for isotropic layers, but now
A11 6=A22 and D11 6=D22, i.e. the laminate has an orthotropic structure
N1
N2
N6
···
M1
M2
M6
=
A11 A12 0.
.
. 0 0 0
A12 A22 0.
.
. 0 0 0
0 0 A66
.
.
. 0 0 0
. . . . . . . . . . . . . .. . . . . . . . . .
000.
.
.D11 D12 0
000.
.
.D12 D22 0
000.
.
. 0 0 D66
ε
1
ε
2
ε
6
···
κ
1
κ
2
κ
6
(4.2.32)
Figure 4.15 illustrates examples of symmetric cross-ply laminates. With
A16 =A26 =0,D16 =D26 =0 there is uncoupling between the normal and
shear in-plane forces and also between the bending and the twisting moments.
146 4 Elastic Behavior of Laminate and Sandwich Composites
S S S
2
1
3
900
00
00
h
h
h
a
1
2
3
4
5
6
900
00
900
900
00
900
h(1)
h(2)
h(3)
h(4)=h(3)
h(5)=h(2)
h(6)=h(1)
1
2
3
400
900
900
00h
h
h
h
b
c
Fig. 4.15 Symmetric cross-ply laminate. a3-layer laminate with equal layer thickness, b6-layer
laminate with equal layer thickness in pairs, c4-layer laminate with equal layer thickness
Symmetric balanced laminate
A laminate is balanced when it consists of pairs of layers of the same thick-
ness and material where the angles of plies are +
θ
and
θ
. An example is the
8 - layer-laminate [±
θ
1/±
θ
2/]s. The stiffness coefficients Ai j and Di j will be
calculated from
Ai j =n
k=1
Q(k)
i j h(k),h(k)=h(k),
θ
(k)=
θ
(k)
=A16 =A26 =0,
Di j =1
3
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3
=
n
k=1
Q(k)
i j x(k)
3
2+h(k)2
12 !h(k)
(4.2.33)
with
h(k)=h(k),
θ
(k)=
θ
(k),x(k)
3=x(k)
3
and the constitutive equation yields
N1
N2
N6
. .
M1
M2
M6
=
A11 A12 0.
.
. 0 0 0
A12 A22 0.
.
. 0 0 0
0 0 A66
.
.
. 0 0 0
. . . . . . . . . . . . . . .. . . . . . . . .
0 0 0 .
.
.D11 D12 D16
0 0 0 .
.
.D12 D22 D26
0 0 0 .
.
.D16 D26 D66
ε
1
ε
2
ε
6
. .
κ
1
κ
2
κ
6
(4.2.34)
The fact that the in-plane shear coupling stiffness A16 and A26 are zero is a defin-
ing characteristic of a balanced laminate. In general the bending/twisting cou-
pling stiffness D16 and D26 are not zero unless the laminate is antisymmetric.
4.2 Elastic Behavior of Laminates 147
Summarizing the results on symmetric laminates above, it is most important that all
components of the B
B
B-matrix are identical to zero and the full (6×6)ABD
ABD
ABD - matrix
decouples into two (3×3)matrices, namely
N
N
N=A
A
A
ε
ε
ε
,M
M
M=D
D
D
κ
κ
κ
(4.2.35)
Therefore also the inverse relations degenerates from (6×6)into two (3×3)rela-
tions
ε
ε
ε
=a
a
aN
N
N,
κ
κ
κ
=d
d
dM
M
M(4.2.36)
In these matrix equations a
a
ais the inverse of A
A
Aand d
d
dthe inverse of D
D
D
a11 =A22A66 A2
26
Det[A
A
A],d11 =D22D66 D2
26
Det[D
D
D],
a12 =A26A16 A12 A66
Det[A
A
A],d12 =D26D16 D12 D66
Det[D
D
D],
a16 =A12A26 A22 A16
Det[A
A
A],d16 =D12D26 D22 D26
Det[D
D
D],
a22 =A11A66 A2
16
Det[A
A
A],d22 =D11D66 D2
16
Det[D
D
D],
a26 =A12A16 A11 A26
Det[A
A
A],d26 =D12D16 D11 D26
Det[D
D
D],
a66 =A11A22 A2
12
Det[A
A
A],d66 =D11D22 D2
12
Det[D
D
D]
(4.2.37)
with
Det[X
X
X] = X11(X22X66 X2
26)X12(X12X66 X26X16)
+X16(X12X26 X22X16);Xi j =Ai j ,Di j
For the special symmetric stacking cases one can find
1. Isotropic layers
a11 =a22,a16 =a26 =0,A11 =A22 ,A16 =A26 =0,
d11 =d22,d16 =d26 =0,D11 =D22 ,D16 =D26 =0,
A11 =Eh
1
ν
,A12 =
ν
A11,A66 =E h
2(1+
ν
),
D11 =Eh3
12(1
ν
2),D12 =
ν
D11,D66 =E h3
24(1+
ν
)
with
h=
n
k=1
h(k)
148 4 Elastic Behavior of Laminate and Sandwich Composites
2. Cross-play layers
a11 =A22
A11A22 A2
12
,d11 =D22
D11D22 D2
12
,
a12 =A12
A11A22 A2
12
,d12 =D12
D11D22 D2
12
,
a22 =A11
A11A22 A2
12
,d22 =D11
D11D22 D2
12
,
a66 =1
A66
,d66 =1
D66
3. Balanced layers
The ai j are identical to cross-ply layers. The di j are identical to the general sym-
metric case.
4.2.3.2 Antisymmetric Laminates
A laminate is called antisymmetric if the material and thickness of the laminae are
the same above and below the midplane but the angle orientations at the same dis-
tance above and below of the midplane are of opposite sign, i.e two symmetric ar-
ranged layers to the midplane with the coordinates ¯x(k)and x(k)=x(k)having the
same thickness h(k)=h(k)and antisymmetric orientations
θ
(k)and
θ
(k)=
θ
(k)
(Fig. 4.14).
Antisymmetric cross-ply laminate
Antisymmetric cross-ply laminates consist of 00and 900laminae arranged in
such a way that for all 00-laminae (k)at a distance ¯x(k)from the midplane there
are 900-laminae (k)at a distance ¯x(k)=¯x(k)and vice versa. By definition
these laminates have an even number of plies. The reduced stiffness fulfills the
conditions
Q(k)
11 =Q(k)
22 ,Q(k)
22 =Q(k)
11 ,Q(k)
12 =Q(k)
12 ,
Q(k)
16 =Q(k)
16 =Q(k)
26 =Q(k)
26 =0,
which yield considering (4.2.15) and the 00and 900layers have the same thick-
ness
A11 =A22,A16 =A26 =0,
B11 =B22,B12 =B16 =B26 =B66 =0,
D11 =D22,D16 =D26 =0
(4.2.38)
and the constitutive equation has the form
4.2 Elastic Behavior of Laminates 149
N1
N2
N6
. .
M1
M2
M6
=
A11 A12 0.
.
.B11 0 0
A12 A11 0.
.
. 0 B11 0
0 0 A66
.
.
. 0 0 0
...........................
B11 0 0 .
.
.D11 D12 0
0B11 0.
.
.D12 D11 0
0 0 0 .
.
. 0 0 D66
ε
1
ε
2
ε
6
. .
κ
1
κ
2
κ
6
(4.2.39)
The constitutive equation (4.2.39) shows that antisymmetric cross-ply laminates
only have a tension/bending coupling. It is important to note that the coupling
coefficient B11 approaches zero as the number of plies increases for a constant
laminate thickness since it is inversely proportional to the total number of layers.
Antisymmetric balanced laminate
Antisymmetric balanced laminates consist of pairs of laminae (k)and (k)at a
distance ¯x(k)and ¯x(k)=¯x(k)with the same material and thickness but orien-
tations
θ
(k)and
θ
(k)=
θ
(k). Examples of these laminates are [
θ
1/
θ
1],
[
θ
1/
θ
2/
θ
2/
θ
1], etc. As for all balanced laminates A16 =A26 =0 and with
Di j =1
3
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3(4.2.40)
and x(k)
3
3x(k1)
3
3=x(k)
3
3x(k1)
3
3,
Q(k)
16 =Q(k)
16 ,Q(k)
26 =Q(k)
26
it follows that
D16 =D26 =0
Note that x(k)
3=x(k)
3,h(k)=h(k)and
Q(k)
11 =Q(k)
11 ,Q(k)
22 =Q(k)
22 ,Q(k)
12 =Q(k)
12 ,
Q(k)
66 =Q(k)
66 ,Q(k)
16 =Q(k)
16 ,Q(k)
26 =Q(k)
26
Equation (4.2.15) yields B11 =B22 =B12 =B66 =0. Balanced antisymmetric
laminates have no in-plane shear coupling and also no bending/twisting cou-
pling but a co upling of stretching/twisting and bending/shearing. T he constitutive
equation has the following structure
150 4 Elastic Behavior of Laminate and Sandwich Composites
N1
N2
N6
···
M1
M2
M6
=
A11 A12 0.
.
. 0 0 B16
A12 A22 0.
.
. 0 0 B26
0 0 A66
.
.
.B16 B26 0
. . . . . . . . . . . . . .. . . . . . . . . .
0 0 B16
.
.
.D11 D12 0
0 0 B26
.
.
.D12 D22 0
B16 B26 0.
.
. 0 0 D66
ε
1
ε
2
ε
6
···
κ
1
κ
2
κ
6
(4.2.41)
4.2.3.3 Stiffness Matrices for Symmetric and Unsymmetric Laminates in
Engineering Applications
Table 4.4 summarizes the stiffness matrices for symmetric and unsymmetric lam-
inates which are used in engineering applications. Symmetric laminates avoid the
stretching/bending coupling. But certain applications require the use of nonsym-
metric laminates. If possible symmetric balanced laminates should be used. The
bending and shearing couplings are eliminated and one can show that for symmet-
ric laminates with a constant total thickness hthe values of the bending or flexural
stiffness D16 and D26 decrease with an increasing number of layers and approach
zero for k . If the stiffness Ai j ,Bi j and Di j are calculated, the compliances
ai j,bi j ,ci j =bT
i j,di j follow from (4.2.26) or for symmetric laminates from (4.2.37).
The experimental identification of the compliances is simpler than for the stiffness
parameters.
The coupling stiffness Bi j and A16,A26,D16 and D26 complicate the analysis of
laminates. To minimize coupling effects symmetric balanced laminates should be
created with a fine lamina distribution. Then all Bi j and the A16 ,A26 are identical
to zero and the D16 ,D26 couplings are relatively low because of the fine lamina
distribution. Whenever possible it is recommended to limit the number of fibre ori-
entations to a few specific one, that are 00,±450,900to minimize the processing and
experimental testing effort and to select a symmetric and balanced lay-up with a fine
lamina interdispersion in order to eliminate in-plane and out-of-plane coupling and
the in-plane tension/shearing coupling and to minimize torsion coupling.
There is furthermore a special class of quasi-isotropic laminates. The layers of
the laminate can be arranged in such a way that the laminate will behave as an
isotropic layer under in-plane loading. Actually, such laminates are not isotropic,
because under transverse loading normal to the laminate plane and under interlami-
nar shear their behavior is different from real isotropic layer. That is why one use the
notation quasi-isotropic layer. Because all quasi-isotropic laminates are symmetric
and balanced the shear coupling coefficients A16,A26 are zero. It can be checked in
general any laminate with a lay-up of
4.2 Elastic Behavior of Laminates 151
Table 4.4 Stiffness matrices for symmetric and unsymmetric laminates
Symmetric laminate Unsymmetric laminate
Isotropic layers Balanced laminate
A11 A12 0 0 0 0
A12 A11 0 0 0 0
0 0 A66 000
0 0 0 D11 D12 0
0 0 0 D12 D11 0
0 0 0 0 0 D66
A11 A12 0B11 B12 B16
A12 A22 0B12 B22 B26
0 0 A66 B16 B26 B66
B11 B12 B16 D11 D12 D16
B12 B22 B26 D12 D22 D26
B16 B26 B66 D16 D26 D66
Eq. (4.2.30)
Cross-ply laminate Antimetric balanced laminate
A11 A12 0 0 0 0
A12 A22 0 0 0 0
0 0 A66 000
0 0 0 D11 D12 0
0 0 0 D12 D22 0
0 0 0 0 0 D66
A11 A12 0 0 0 B16
A12 A22 0 0 0 B26
0 0 A66 B16 B26 0
0 0 B16 D11 D12 0
0 0 B26 D12 D22 0
B16 B26 0 0 0 D66
Eq. (4.2.32) Eq. (4.2.41)
Balanced laminate Cross-ply
A11 A12 0 0 0 0
A12 A22 0 0 0 0
0 0 A66 000
0 0 0 D11 D12 D16
0 0 0 D12 D22 D26
0 0 0 D16 D26 D66
A11 A12 0B11 0 0
A12 A11 0 0 B11 0
0 0 A66 000
B11 0 0 D11 D12 0
0B11 0D12 D11 0
0 0 0 0 0 D66
Eq. (4.2.34) Eq. (4.2.39)
Angle-ply laminate Cross-ply
(approximate solution k)
A11 A12 A16 000
A12 A22 A26 000
A16 A26 A66 000
0 0 0 D11 D12 D16
0 0 0 D12 D22 D26
0 0 0 D16 D26 D66
A11 A12 0 0 0 0
A12 A22 0 0 0 0
0 0 A66 0 0 0
0 0 0 D11 D12 0
0 0 0 D12 D11 0
0 0 0 0 0 D66
Eq. (4.2.28)
0/
π
n/2
π
n/ . . ./ (n1)
π
nS
or
π
n/2
π
n/.../
π
S
152 4 Elastic Behavior of Laminate and Sandwich Composites
is quasi-isotropic for any integer ngreater than 2. The simplest types are laminates
with the following lay-up
[0/60/60]S,n=3,1200 600
and
[0/+45/45/90]S,n=4,1350 450
Summarizing the mechanical performance of laminates with special laminae
stacking sequences which are used in laminate design, we have considered the fol-
lowing classification:
1. General laminates
The stacking sequence, the thickness, the material and the fibre orientations of
all laminae is quite general. All extensional stiffness Bi j are not zero.
2. Symmetric laminates
For every layer to one side of the laminate reference surface there is a corre-
sponding layer to the other side of the reference surface at an equal distance and
with identical thickness, material and fibre orientation. All coupling stiffness Bi j
are zero.
3. Antisymmetric laminates
For every layer to one side of the laminate reference surface there is a corre-
sponding layer to the other side of the reference surface at an equal distance,
with identical thickness and material, but opposite fibre orientation. The stiffness
A16,A26 ,D16 and D26 are zero.
4. Balanced laminates
For every layer with a specified thickness, specific material properties and spe-
cific fibre orientation there is another layer with identical thickness and material
properties, but opposite fibre orientation anywhere in the laminate, i.e. the corre-
sponding layer with opposite fibre orientation does not have to be on the opposite
side of the reference surface, nor immediately adjacent to the other layer nor any-
where particular. A balanced laminate can be
General or unsymmetric: A16 =A26 =0
Symmetric A16 =A26 =0, Bi j =0
Antisymmetric A16 =A26 =0,D16 =D26 =0
An antisymmetric laminate is a special case of a balanced laminate, having its
balanced ±pairs of layers symmetrically situated to the middle surface.
5. Cross-ply laminates
Every layer of the laminate has its fibers oriented at either 00or 900. Cross-ply
laminates can be
General or unsymmetric: A16 =A26 =0, B16 =B26 =0, D16 =D26 =0
Symmetric A16 =A26 =0, D16 =D26 =0, Bi j =0
Antisymmetric A16 =A26 =0,D16 =D26 =0, B12 =B16 =B26 =B66 =0,
A11 =A22,B11 =B22 ,D11 =D22
4.2 Elastic Behavior of Laminates 153
Symmetric cross-ply laminates are orthotropic with respect to both in-plane and
bending behavior and all coupling stiffness are zero.
6. Quasi-isotropic laminates
For every laminate with a symmetric lay-up of
[0/
π
n/ . . ./ (n1)
π
n]Sor [0/
π
n/2
π
n/.../
π
]S
the in-plane stiffness are identical in all directions. Because all these quasi-
isotropic laminates are also balanced we have A11 =A22 =const,A12 =const,
A16 =A26 =0, Bi j 0,Di j 6=0
7. Laminates with isotropic layers
If isotropic layers of possible different materials properties and thicknesses are
arranged symmetrically to the middle surface the laminate is symmetric isotropic
and we have A11 =A22,D11 =D22 ,A16 =A26 =0,D16 =D26 =0, Bi j =0, i.e.
the mechanical performance is isotropic.
For symmetric laminates the in-plane and flexural moduli can be defined with
help of effective engineering parameters. We start with (4.2.26). a
a
a,b
b
b,c
c
c=b
b
bT,d
d
dare
the extensional compliance matrix, coupling compliance matrix and bending com-
pliance matrix, respectively. For a symmetric laminate B
B
B=0
0
0 and it can be shown
that a
a
a=A
A
A1and d
d
d=D
D
D1. The in-plane and the flexural compliance matrices a
a
aand
d
d
dare uncoupled but generally fully populated
ε
ε
ε
=a
a
aN
N
N,
κ
κ
κ
=d
d
dM
M
M(4.2.42)
Equations (4.2.42) lead to effective engineering moduli for symmetric laminates.
1. Effective in-plane engineering moduli EN
1,EN
2,GN
12,
ν
N
12:
Substitute N16=0,N2=N6=0 in
ε
ε
ε
=a
a
aN
N
Nas
ε
1
ε
2
ε
6
=
a11 a12 a16
a12 a22 a26
a16 a26 a66
N1
0
0
(4.2.43)
which gives
ε
1=a11N1
and the effective longitudinal modulus EN
1is
EN
1
σ
1
ε
1
=N1/h
a11N1
=1
ha11
(4.2.44)
In an analogous manner with N1=0,N26=0,N6=0 or N1=N2=0,N66=0,
the effective transverse modulus EN
2or the effective shear modulus GN
12 are
EN
2
σ
2
ε
2
=N2/h
a22N2
=1
ha22
,(4.2.45)
154 4 Elastic Behavior of Laminate and Sandwich Composites
GN
12
σ
6
ε
6
=N6/h
a66N6
=1
ha66
(4.2.46)
The effective in-plane Poisson’s ratio
ν
N
12 can be derived in the following way.
With N16=0,N2=N6=0 (4.2.43) yields
ε
2=a12N1,
ε
1=a11N1and
ν
12 is de-
fined as
ν
N
12 =
ε
2
ε
1
=a12N1
a11N1
=a12
a11
(4.2.47)
The Poisson’s ratio
ν
N
21 can be derived directly by substituting N1=N6=0,
N26=0 in (4.2.42) and define
ν
N
21 =
ε
1/
ε
2or by using the reciprocal relationship
ν
N
12/EN
1=
ν
N
21/EN
2. In both cases
ν
N
21 is given as
ν
N
21 =a12
a22
(4.2.48)
The effective in-plane engineering moduli can be also formulated in terms of the
elements of the A
A
A-matrix
EN
1=A11A22 A2
12
A22h,EN
2=A11A22 A2
12
A11h,GN
12 =A66
h,
ν
N
12 =A12
A22
,
ν
N
21 =A12
A11
(4.2.49)
2. Effective flexural engineering moduli EM
1,EM
2,GM
12,
ν
M
12,
ν
M
21:
To define the effective flexural moduli we start with
κ
κ
κ
=d
d
dM
M
M. Apply M16=0,
M2=0,M6=0 and substitute in the flexural compliance relation to give
κ
1=d11M1=M1
EM
1I,I=h3
12
and the effective flexural longitudinal modulus EM
1is
EM
1=12M1
κ
1h3=12
h3d11
(4.2.50)
Similarly, one can show that the other flexural elastic moduli are given by
EM
2=12
h3d22
,
ν
M
12 =d12
d11
,GM
12 =12
h3d66
,
ν
M
21 =d12
d22
(4.2.51)
Flexural Poisson’s ratios also have a reciprocal relationship
ν
M
12
EM
1
=
ν
M
21
EM
2
(4.2.52)
In terms of the elements of the D
D
Dmatrix we find
4.2 Elastic Behavior of Laminates 155
EM
1=12(D11D22 D2
12)
D22h3,EM
2=12(D11D22 D2
12)
D11h3,
GM
12 =12D66
h3,
ν
M
12 =D12
D22
,
ν
M
21 =D12
D11
(4.2.53)
Consider unsymmetric laminates, the laminate stiffness or compliance matrices
are not uncoupled and therefore it is not meaningful to use effective engineering
laminate moduli.
4.2.4 Stress Analysis
Laminate stresses may be subdivided into in-plane stresses, which are calculated
below with the classical assumption of linear strain functions of x3, and the through-
the-thickness stresses, which are calculated approximately by integration of the
equilibrium conditions. Taking into account the assumptions of macro-mechanical
modelling of laminates the strains
ε
1,
ε
2,
ε
6vary linearly across the thickness of the
laminate
ε
ε
ε
(x1,x2,x3) =
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2),h=
n
k=1
h(k)(4.2.54)
These global strains can be transformed to the local strains in the principal material
directions of the kth layer through the transformation equations (Table 4.1)
ε
ε
ε
(k)=T
T
T
εε
ε
ε
(k),x(k1)
3x3x(k)(4.2.55)
If the strains are known at any point along the thickness of the laminate, the stress-
strain relation (Table 4.2) calculates the global stress in each lamina
σ
σ
σ
(k)=Q
Q
Q(k)
ε
ε
ε
(k)=Q
Q
Q(k)(
ε
ε
ε
+x3
κ
κ
κ
),x(k1)
3x3x(k)(4.2.56)
By applying the transformation equation for the stress vector (Table 4.1) the stresses
expressed in the principal material axes can be calculated
σ
σ
σ
(k)=T
T
T
σσ
σ
σ
(k)(4.2.57)
Starting from the strains
ε
ε
ε
(k), the stresses in the kth layer are expressed as follows
σ
σ
σ
(k)=Q
Q
Q(k)
ε
ε
ε
(k)(4.2.58)
From (4.2.56), the stresses vary linearly through the thickness of each lamina and
may jump from lamina to lamina since the reduced stiffness matrix Q
Q
Q(k)changes
from ply to ply since Q
Q
Q(k)depends on the material and orientation of the lamina (k).
Figure 4.16 illustrates qualitatively the stress jumps of the membrane stresses
σ
σ
σ
(k)
M
156 4 Elastic Behavior of Laminate and Sandwich Composites
h
Q(3)
Q(2)
Q(1)
h(1)
h(2)
h(3)
σ
(k)
6M =Q(k)
66
ε
6
σ
(k)
2M =Q(k)
22
ε
2
σ
(k)
1M =Q(k)
11
ε
1
σ
(k)
6B =Q(k)
66 x3
κ
6
σ
(k)
2B =Q(k)
22 x3
κ
2
σ
(k)
1B =Q(k)
11 x3
κ
1
σ
(k)
6=
σ
(k)
6M +
σ
(k)
6B
σ
(k)
2=
σ
(k)
2M +
σ
(k)
2B
σ
(k)
1=
σ
(k)
1M +
σ
(k)
1B
Fig. 4.16 Qualitatively variation of the in-plane membrane stresses
σ
iM, the bending stresses
σ
iBand the total stress
σ
ithrough the thickness of the laminate. Assumptions h(1)=h(3),
Q(1)=Q(3)<Q(2),i=1,2,6
which follow from the in-plane resultants N
N
Nand are constant through each lamina
and the bending/torsion stresses
σ
σ
σ
(k)
Bfollowing from the moment resultants M
M
Mand
vary linearly through each ply thickness. The transverse shear stresses
σ
4,
σ
5follow
for a plane stress state assumptions that is in the framework of the classical laminate
theory, Sect. 5.1, not from a constitutive equation but as for the single layer, (4.1.56),
by integration of the equilibrium equations. For any lamina mof the laminate by
analogy to (4.1.57) can be established
σ
(m)
5(x3) =
m1
k=1(x(k)
3
Z
x(k1)
3
Q(k)
11
x1
(
ε
1+x3
κ
1) + Q(k)
12
x1
(
ε
2+x3
κ
2)
+Q(k)
16
x1
(
ε
6+x3
κ
6) + Q(k)
61
x2
(
ε
1+x3
κ
1)
+Q(k)
62
x2
(
ε
2+x3
κ
2) + Q(k)
66
x2
(
ε
6+x3
κ
6)dx3)
x3
Z
x(m1)
3
Q(m)
11
x1
(
ε
1+x3
κ
1) + Q(m)
12
x1
(
ε
2+x3
κ
2)
+Q(m)
16
x1
(
ε
6+x3
κ
6) + Q(m)
61
x2
(
ε
1+x3
κ
1)
+Q(m)
62
x2
(
ε
2+x3
κ
2) + Q(m)
66
x2
(
ε
6+x3
κ
6)dx3,(4.2.59)
σ
(m)
4(x3) =
m1
k=1(x(k)
3
Z
x(k1)
3
Q(k)
61
x1
(
ε
1+x3
κ
1) + Q(k)
62
x1
(
ε
2+x3
κ
2)
4.2 Elastic Behavior of Laminates 157
+Q(k)
66
x1
(
ε
6+x3
κ
6) + Q(k)
21
x2
(
ε
1+x3
κ
1)
+Q(k)
22
x2
(
ε
2+x3
κ
2)dx3+Q(k)
26
x2
(
ε
6+x3
κ
6)dx3)(4.2.60)
x3
Z
x(m1)
3
Q(m
61
x1
(
ε
1+x3
κ
1) + Q(m)
62
x1
(
ε
2+x3
κ
2)
+Q(m)
66
x1
(
ε
6+x3
κ
6) + Q(m)
21
x2
(
ε
1+x3
κ
1)
+Q(m)
22
x2
(
ε
2+x3
κ
2) + Q(m)
26
x2
(
ε
6+x3
κ
6)dx3,
σ
(m)
i(x3=x(m)
3) =
σ
(m+1)
i(x3=x(m)
3),i=4,5;m=1,2,...,n,
x(m+1)
3x3x(m)
3
With the relationships
x(k)
3
Z
x(k1)
3
Q
Q
Q(k)dx3=Q
Q
Q(k)x(k)
3x(k1)
3=Q
Q
Q(k)h(k),
x(k)
3
Z
x(k1)
3
Q
Q
Q(k)x3dx3=Q
Q
Q(k)1
2x(k)
3
2x(k1)
3
2=Q
Q
Q(k)h(k)x(k)
3=Q
Q
Q(k)s(k)
(4.2.61)
¯x(k)
3is the distance of lamina kfrom the midplane. The shear stresses
σ
(m)
i(x3=x(m)
3),i=4,5, at the top surface of the mth lamina can be formulated by
σ
(m)
5x3=x(m)
3
σ
(m)
4x3=x(m)
3
=
m
k=1
F
F
F(k)
1
TF
F
F(k)
6
T
F
F
F(k)
6
TF
F
F(k)
2
T
"
η
η
η
1
η
η
η
2#(4.2.62)
with
F
F
F(k)
1
T= [h(k)Q(k)
(11)h(k)Q(k)
(12)h(k)Q(k)
(16)s(k)Q(k)
(11)s(k)Q(k)
(12)s(k)Q(k)
(16)],
F
F
F(k)
2
T= [h(k)Q(k)
(21)h(k)Q(k)
(22)h(k)Q(k)
(26)s(k)Q(k)
(21)s(k)Q(k)
(22)s(k)Q(k)
(26)],
F
F
F(k)
6
T= [h(k)Q(k)
(61)h(k)Q(k)
(62)h(k)Q(k)
(66)s(k)Q(k)
(61)s(k)Q(k)
(62)s(k)Q(k)
(66)]
and
158 4 Elastic Behavior of Laminate and Sandwich Composites
η
η
η
T
1=
ε
1
x1
ε
2
x1
ε
6
x1
κ
1
x1
κ
2
x1
κ
6
x1,
η
η
η
T
2=
ε
1
x2
ε
2
x2
ε
6
x2
κ
1
x2
κ
2
x2
κ
6
x2
The transverse shear stresses only satisfy the equilibrium conditions but violate the
other fundamental equations of anisotropic elasticity. They vary in a parabolic way
through the thicknesses h(k)of the laminate layers and there is no stress jump if one
crosses the interface between two layers.
4.2.5 Thermal and Hygroscopic Effects
In Sect. 4.1.2 the hygrothermal strains were calculated for unidirectional and angle
ply laminae. As mentioned above, no residual mechanical stresses would develop
in the lamina at the macro-mechanical level, if the lamina is free to expand. Free
thermal strain, e.g., refers to the fact that fibres and matrix of an UD-lamina are
smeared into a single equivalent homogeneous material and that the smeared ele-
ments are free of any stresses if temperature is changed. When one considers an
unsmeared material and deals with the individual fibres and the surrounding matrix,
a temperature change can create significant stresses in the fibre and matrix. When
such selfbalanced stresses are smeared over a volume element, the net result is zero.
However, in a laminate with various laminae of different materials and orientations
each individual lamina is not free to deform. This results in residual stresses in the
laminate. As in Eqs. (4.1.31) and (4.1.32)
α
th,
α
mo are the thermal and moisture
expansion coefficients, Tis the temperature change and Mthe weight of moisture
absorption per unit weight. In the following equations, Tand Mare independent of
the x3-coordinate, i.e. they are constant not only through the thickness h(k)of a sin-
gle layer but through the thickness hof the laminate. Heat transfer in thin laminates,
e.g., is generally quite rapid and, hence, thermal gradients in x3-direction are seldom
taken into account and the temperature change Tis then approximately independent
of x3. Analogous considerations are valid for changes in moisture.
For a single layer in off-axis coordinates, the hygrothermal strains and stresses
are given by
ε
ε
ε
=S
S
S
σ
σ
σ
+
α
α
α
thT+
α
α
α
moM,
σ
σ
σ
=Q
Q
Q(
ε
ε
ε
α
α
α
thT
α
α
α
moM)(4.2.63)
or substituting
ε
ε
ε
=
ε
ε
ε
+x3
κ
κ
κ
σ
σ
σ
=Q
Q
Q(
ε
ε
ε
+x3
κ
κ
κ
α
α
α
thT
α
α
α
moM)(4.2.64)
The definitions of the force and moment resultants N
N
Nand M
M
M
4.2 Elastic Behavior of Laminates 159
N
N
N=Z
(h)
σ
σ
σ
dx3=Z
(h)
Q
Q
Q(
ε
ε
ε
+x3
κ
κ
κ
α
α
α
thT
α
α
α
moM)dx3,
M
M
M=Z
(h)
σ
σ
σ
x3dx3=Z
(h)
Q
Q
Q(
ε
ε
ε
+x3
κ
κ
κ
α
α
α
thT
α
α
α
moM)x3dx3
(4.2.65)
yield the equations
N
N
N=A
A
A
ε
ε
ε
+B
B
B
κ
κ
κ
N
N
Nth N
N
Nmo,
M
M
M=B
B
B
ε
ε
ε
+D
D
D
κ
κ
κ
M
M
Mth M
M
Mmo,(4.2.66)
A
A
A=Q
Q
Qh,B
B
B=0
0
0,D
D
D=Q
Q
Qh3
12
B
B
B=0
0
0 follows from the symmetry of a single layer to its midplane. N
N
Nth,N
N
Nmo,
M
M
Mth,M
M
Mmo are fictitious hygrothermal resultants which are defined in (4.2.67). If T
and Mare independent of x3one can introduce unit thermal and unit moisture
stress resultants ˆ
Nth,ˆ
Mth,ˆ
Nmo,ˆ
Mmo, i.e. resultants per unit temperature or moisture
change
N
N
Nth =Z
(h)
Q
Q
Q
α
α
α
thTdx3=Q
Q
Q
α
α
α
thT h =ˆ
NthT,
M
M
Mth =Z
(h)
Q
Q
Q
α
α
α
thT x3dx3=1
2Q
Q
Q
α
α
α
thT h2=ˆ
MthT,
N
N
Nmo =Z
(h)
Q
Q
Q
α
α
α
moMdx3=Q
Q
Q
α
α
α
moMh=ˆ
NmoM,
M
M
Mmo =Z
(h)
Q
Q
Q
α
α
α
moMx3dx3=1
2Q
Q
Q
α
α
α
moMh2=ˆ
MmoM.
(4.2.67)
Nth and Nmo have the units of the force resultant, namely N/m, and Mth and Mmo
the units of the moment resultants, namely Nm/m. The integral form of the resultant
definitions makes these definitions quite general, i.e. if Tor Mare known func-
tions of x3, the integration can be carried out. But for the temperature changes with
x3and if the material properties change with temperature, the integration can be
complicated, but in general the simple integrated form, Eqs. (4.2.67) can be used.
With the total force and moment resultants ˜
N
N
N,˜
M
M
M, equal to the respective sums of
their mechanical and hygrothermal components
˜
N
N
N=N
N
N+N
N
Nth +N
N
Nmo,˜
M
M
M=M
M
M+M
M
Mth +M
M
Mmo,(4.2.68)
the extended hygrothermal constitutive equation for a lamina can be written
˜
N
N
N
···
˜
M
M
M
=
A
A
A.
.
. 0
0
0
. . . .
0
0
0.
.
.D
D
D
ε
ε
ε
···
κ
κ
κ
(4.2.69)
160 4 Elastic Behavior of Laminate and Sandwich Composites
This constitutive equation is identical to that derived for mechanical loading only,
(4.1.54), except for the fact that here the hygrothermal forces and moments are
added to the mechanically applied forces and moments. The inversion of (4.2.69)
yields the compliance relation
ε
ε
ε
···
κ
κ
κ
=
a
a
a.
.
. 0
0
0
. . . .
0
0
0.
.
.d
d
d
˜
N
N
N
···
˜
M
M
M
,a
a
a=A
A
A1,d
d
d=D
D
D1(4.2.70)
The values of the stiffness Ai j ,Di j and compliances ai j ,di j are the same as for pure
mechanical loading (Table 4.3) and unit stress resultants ˆ
Nth,ˆ
Mth,ˆ
Nmo,ˆ
Mmo are
ˆ
Nth
1= (Q11
α
th
1+Q12
α
th
2+Q16
α
th
6)h,
ˆ
Nth
2= (Q12
α
th
1+Q22
α
th
2+Q26
α
th
6)h,
ˆ
Nth
6= (Q16
α
th
1+Q26
α
th
2+Q66
α
th
6)h,
ˆ
Mth
1= (Q11
α
th
1+Q12
α
th
2+Q16
α
th
6)1
2h2,
ˆ
Mth
2= (Q12
α
th
1+Q22
α
th
2+Q26
α
th
6)1
2h2,
ˆ
Mth
6= (Q16
α
th
1+Q26
α
th
2+Q66
α
th
6)1
2h2
(4.2.71)
and analogous for unit moisture stress resultant with
α
mo
i,i=1,2,6. In the more
general case the integral definitions have to used.
When a laminate is subjected to mech anical and hygrothermal load ing, a lamina k
within the laminate is under a state of stress
σ
σ
σ
(k)and strain
ε
ε
ε
(k). The hygrothermoe-
lastic superposition principle shows that the strains
ε
ε
ε
(k)in the lamina kare equal
to the sum of the strains produced by the existing stresses and the free, i.e. unre-
strained, hygrothermal strains and the stresses
σ
σ
σ
(k)follow by inversion
ε
ε
ε
(k)=S
S
S(k)
σ
σ
σ
(k)+
α
α
α
th(k)T+
α
α
α
mo(k)M,
σ
σ
σ
(k)=Q
Q
Q(k)(
ε
ε
ε
(k)
α
α
α
th(k)T
α
α
α
mo(k)M)
=Q
Q
Q(k)(
ε
ε
ε
(k)+x3
κ
κ
κ
(k)
α
α
α
th(k)T
α
α
α
mo(k)M)
(4.2.72)
When in the lamina kall strains are restrained, then
ε
ε
ε
(k)=0
0
0 and the hygrothermal
stresses are
σ
σ
σ
(k)=Q
Q
Q(k)
α
α
α
th(k)T
α
α
α
mo(k)M(4.2.73)
Integration of the stresses
σ
σ
σ
(k)and the stresses
σ
σ
σ
(k)multiplied by the x3-coordinate
across the thickness h(k)and summation for all laminae gives the force and moment
resultants of the laminate
4.2 Elastic Behavior of Laminates 161
N
N
N=
n
k=1
σ
σ
σ
(k)h(k)
=
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
dx3
ε
ε
ε
+
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
x3dx3
κ
κ
κ
+
n
k=1
Q
Q
Q(k)
α
α
α
th(k)T
x(k)
3
Z
x(k1)
3
dx3
+
n
k=1
Q
Q
Q(k)
α
α
α
mo(k)M
x(k)
3
Z
x(k1)
3
dx3
,
M
M
M=
n
k=1
σ
σ
σ
(k)h(k)x(k)
3
=
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
x3dx3
ε
ε
ε
+
n
k=1
Q
Q
Q(k)
x(k)
3
Z
x(k1)
3
x2
3dx3
κ
κ
κ
+
n
k=1
Q
Q
Q(k)
α
α
α
th(k)T
x(k)
3
Z
x(k1)
3
x3dx3
+
n
k=1
Q
Q
Q(k)
α
α
α
mo(k)M
x(k)
3
Z
x(k1)
3
x3dx3
(4.2.74)
With the stiffness matrices A
A
A,B
B
Band D
D
D, Eqs. (4.2.74), can be rewritten in a brief
matrix form
N
N
N=A
A
A
ε
ε
ε
+B
B
B
κ
κ
κ
N
N
Nth N
N
Nmo,
M
M
M=B
B
B
ε
ε
ε
+D
D
D
κ
κ
κ
M
M
Mth M
M
Mmo (4.2.75)
The fictitious hygrothermal resultants are given by
N
N
Nth =T
n
k=1
Q
Q
Q(k)
α
α
α
th(k)h(k)=T
n
k=1
ˆ
N(k)th,
N
N
Nmo =Mn
k=1
Q
Q
Q(k)
α
α
α
mo(k)h(k)=Mn
k=1
ˆ
N(k)mo,
M
M
Mth =1
2T
n
k=1
Q
Q
Q(k)
α
α
α
th(k)x(k)
3
2x(k1)
3
2(4.2.76)
=T
n
k=1
Q
Q
Q(k)
α
α
α
th(k)x(k)
3h(k)=T
n
k=1
ˆ
M(k)th,
M
M
Mmo =1
2Mn
k=1
Q
Q
Q(k)
α
α
α
mo(k)x(k)
3
2x(k1)
3
2
=Mn
k=1
Q
Q
Q(k)
α
α
α
mo(k)x(k)
3h(k)=Mn
k=1
ˆ
M(k)mo
162 4 Elastic Behavior of Laminate and Sandwich Composites
By analogy to the single layer one can introduce total force and moment resultants
˜
N
N
Nand ˜
M
M
Mthe stiffness and compliance equations expanded to the hygrothermal com-
ponents
N
N
N+N
N
Nth +N
N
Nmo =˜
N
N
N,M
M
M+M
M
Mth +M
M
Mmo =˜
M
M
M,
˜
N
N
N
···
˜
M
M
M
=
A
A
A.
.
.B
B
B
. . . .
B
B
B.
.
.D
D
D
ε
ε
ε
···
κ
κ
κ
,
ε
ε
ε
···
κ
κ
κ
=
a
a
a.
.
.b
b
b
.....
b
b
bT.
.
.d
d
d
˜
N
N
N
···
˜
M
M
M
,
a
a
a=A
A
AB
B
BD
D
D1C
C
C,A
A
A=A
A
A1,
b
b
b=B
B
BD
D
D1,B
B
B=A
A
A1B
B
B,
d
d
d=D
D
D1,D
D
D=D
D
DB
B
BA
A
A1B
B
B
(4.2.77)
The coupling effects discussed in Sects. 4.2.2 and 4.2.3 stay unchanged and the
stiffness matrices in Table 4.4 can be transferred.
If we can classify a laminate as symmetric, balanced, cross-plied or some combi-
nations of these three laminate stacking types, some of the thermal or moisture force
or moment resultant coefficients may be zero. For temperature or moisture change
that depends on x3, no general statements can be given. However for changes in-
dependent of x3, the following simplifications for the unit stress resultants can be
considered, J=th,mo
Symmetric laminates
ˆ
NJ
16=0,ˆ
MJ
1=0,
ˆ
NJ
26=0,ˆ
MJ
2=0,
ˆ
NJ
66=0,ˆ
MJ
6=0
Balanced laminates ˆ
NJ
16=0,ˆ
MJ
16=0,
ˆ
NJ
26=0,ˆ
MJ
26=0,
ˆ
NJ
6=0,ˆ
MJ
66=0
Symmetric balanced laminates
ˆ
NJ
16=0,ˆ
MJ
1=0,
ˆ
NJ
26=0,ˆ
MJ
2=0,
ˆ
NJ
6=0,ˆ
MJ
6=0
Cross-ply laminates
ˆ
NJ
16=0,ˆ
MJ
16=0,
ˆ
NJ
26=0,ˆ
MJ
26=0,
ˆ
NJ
6=0,ˆ
MJ
6=0
4.2 Elastic Behavior of Laminates 163
Symmetric cross-ply laminates
ˆ
NJ
16=0,ˆ
MJ
1=0,
ˆ
NJ
26=0,ˆ
MJ
2=0,
ˆ
NJ
6=0,ˆ
MJ
6=0
Summarizing the hygrothermal effects one can see that if both mechanical and
hygrothermal loads are applied the mechanical and fictitious hygrothermal loads can
be added to find ply by ply stresses and strains in the laminate, or the mechanical
and hygrothermal loads can be applied separately and then the resulting stresses and
strains of the two problems are added.
4.2.6 Problems
Exercise 4.7. A symmetric laminate under in-plane loading can be considered as
an equivalent homogeneous anisotropic plate in plane stress state by introducing
average stress
σ
σ
σ
=N
N
N/hand N
N
N=A
A
A
ε
ε
ε
. Calculate the effective moduli for general
symmetric laminates and for symmetric cross-ply laminates.
Solution 4.7. Equations (4.2.37) yields
ε
ε
ε
=a
a
aN
N
N,a
a
a=A
A
A1. The components of the
inverse matrix a
a
aare
a11 = (A22A66 A2
26)/
,a12 = (A16A26 A12 A66)/
,
a22 = (A11A66 A2
16)/
,a16 = (A12A26 A22 A16)/
,
a26 = (A12A16 A11A26)/
,a66 = (A11A22 A2
12)/
,
=Det(Ai j) = A11
A22 A26
A26 A66 A12
A12 A26
A16 A66
+A16
A12 A22
A16 A26
The comparison
ε
ε
ε
=a
a
aN
N
N=ha
a
a
σ
σ
σ
with (4.1.19) leads to
E1=1/ha11,E2=1/ha22,E6=1/ha66,
ν
12 =a12/a11,
ν
21 =a12/a22,
ν
16 =a16/a11,
ν
61 =a16/a66 ,
ν
26 =a26/a22 ,
ν
62 =a26/a6
These are the effective moduli in the general case. For cross-ply laminates is
A16 =A26 =0 (Eqs. 4.2.31). The effective moduli can be explicitly expressed in
terms of the in-plane stiffness Aij . With Det(Ai j) = A11 A22A66 A2
12A66 follow the
effective moduli
E1= (A11A22 A2
12)/hA22,
ν
12 =A12/A22 ,
E2= (A11A22 A2
12)/hA11,
ν
21 =A12/A11 ,
G12 E6=A66/h,
ν
16 =
ν
61 =
ν
26 =
ν
62 =0
Note that these simplified formulae are not only valid for symmetric cross-ply lam-
inates (0/90)Sbut also for laminates (±45)S.
164 4 Elastic Behavior of Laminate and Sandwich Composites
Exercise 4.8. Show that a symmetric laminate [±450/00/900]Swith E
1=140 GPa,
E
2=10 GPa, E
6=7 GPa,
ν
12 =0,3 has a quasi-isotropic material behavior. The
thicknesses of all plies are onstant h(k)=0,1 mm.
Solution 4.8. The solution will obtained in four steps:
1. Calculation of the on-axis reduced stiffness Q
i j
With respect to (2.1.56) we obtain
ν
12/E
1=
ν
21/E
2=
ν
21 = (
ν
12E
2)/E
1=0,0214
and finally from Eqs. (4.1.3) follow
Q
11 =E
1(1
ν
12
ν
21) = 140,905 GPa,
Q
22 =E
2(1
ν
12
ν
21) = 10,065 GPa,
Q
12 =E
2
ν
12/(1
ν
12
ν
21) = 3,019 GPa,
Q
66 =E
6=7 GPa
2. Calculation of the reduced stiffness in the laminae (Table 4.2)
Qi j[00]Q
i j,
Q11[900]=Q
22 =10,065 GPa,
Q12[900]=Q
12 =3,019 GPa,
Q22[900]=Q
11 =140,905 GPa,
Q66[900]=Q
66 =7 GPa,
Q16[900]=Q26[900]=0,
Q11[±450]=46,252 GPa,
Q12[±450]=32,252 GPa,
Q22[±450]=46,252 GPa,
Q66[±450]=36,233 GPa,
Q16[±450]=±32,71 GPa,
Q26[±450]=±32,71 GPa
3. Calculation of the axial stiffness Ai j (4.2.15)
Ai j =
8
n=1
Q(k)
i j h(k)=2
4
n=1
Q(k)
i j h(k),
A11 =48,695 106Nm1=A22 ,A12 =14,108 106Nm1,
A66 =17,293106Nm1,A16 =A26 =0
4. Calculation of the effective moduli (example 1)
E1=E2= (A2
11 A2
12)/hA22 =446,1 GPa, E6=A66 /h=172,9 GPa,
ν
12 =
ν
21 =A12/A22 =0,29
Note that E=2(1+
ν
)G=446,1 GPa, i.e. the isotropy condition is satisfied.
Exercise 4.9. Calculate the laminar stresses
σ
and
σ
in the laminate of previous
example loaded by uniaxial tension N1.
4.2 Elastic Behavior of Laminates 165
Solution 4.9. The following reduced stiffness matrices are calculated
Q
Q
Q[00]=
140,9 3,02 0
3,02 10,06 0
0 0 7,0
109Pa,
Q
Q
Q[900]=
10,06 3,02 0
3,02 140,9 0
0 0 7,0
109Pa,
Q
Q
Q[±450]=
46,25 32,25 ±32,71
32,25 46,25 ±32,71
±32,71 ±32,71 36,23
109Pa
The axial stiffness matrix A
A
Ais also calculated
A
A
A=
48,70 14,11 0
14,11 48,70 0
0 0 17,2 9
106N/m
With
a11 =A22/(A2
11 A2
12) = 0,02241 106m/N,
a22 =A11/(A2
11 A2
12) = a11 ,
a66 =1/A66 =0,05784 106m/N,
a12 =A12/(A11 A22 A2
12) = 0,00645 106m/N
follows the inverse matrix a
a
a=A
A
A1
a
a
a=
22,41 6,49 0
6,49 22,41 0
0 0 57,84
109m/N
The strains are with (4.2.42)
ε
ε
ε
=a
a
aN
N
N=
ε
1
ε
2
ε
6
=a
a
a
N1
0
0
=
22,41
6,49
0
109(m/N)N1
N1is given in N/m, i.e.
ε
iare dimensionless.
Now the laminar stresses are (Table 4.2)
σ
1
σ
2
σ
6
[00]
=Q
Q
Q[00]
ε
ε
ε
=
3138
2,4
0
N1[N/m2],
σ
1
σ
2
σ
6
[900]
=Q
Q
Q[900]
ε
ε
ε
=
205,8
846,8
0
N1[N/m2],
166 4 Elastic Behavior of Laminate and Sandwich Composites
σ
1
σ
2
σ
6
[±450]
=Q
Q
Q[±450]
ε
ε
ε
=
827,2
422,6
±520,7
N1[N/m2]
The stresses jump from lamina to lamina. Verify that the resultant force N2=0.
The stress components in reference to the principal material axes follow with the
transformation rule (Table 4.1)
σ
1
σ
2
σ
6
=
c2s22sc
s2c22sc
cs cs c2s2
σ
1
σ
2
σ
6
,
σ
1
σ
2
σ
6
[00]
=
3138
2,4
0
N1[N/m2],
σ
1
σ
2
σ
6
[900]
=
846,8
205,8
0
N1[N/m2],
σ
1
σ
2
σ
6
[±450]
=
1146
104,2
202,3
N1[N/m2]
Note 4.1. These stresses are used in failure analysis of a laminate.
Exercise 4.10. A laminate with an unsymmetric layer stacking [450/300/00]has
three layers of equal thickness h(1)=h(2)=h(3)=5 mm. The mechanical properties
of all UD-laminae are E
1=181 GPa, E
2=10,30 GPa, G
12 =7,17 GPa,
ν
12 =0,28
GPa. Determine the laminate stiffness Ai j ,Bi j ,Di j.
Solution 4.10. Using (4.1.3) the elements S
i j of the compliance matrix S
S
Sare
S
11 =1/E
1=0,0055 GPa1,
S
12 =
ν
12/E
1=0,0015 GPa1,
S
22 =1/E
2=0,0971 GPa1,
S
66 =1/G
12 =0,1395 GPa1
The minor Poisson’s ratio follows with
ν
21 =
ν
12E
2/E
1=0,01593
Using (4.1.3) the elements Q
i j of the reduced stiffness matrix Q
Q
Qare
Q
11 =E
1/(1
ν
12
ν
21) = 181,8 109Pa,
Q
12 =
ν
12E
2/(1
ν
12
ν
21) = 2,897 109Pa,
Q
22 =E
2/(1
ν
12
ν
21) = 10,35 109Pa,
Q
66 =G
12 =7,17 109Pa
4.2 Elastic Behavior of Laminates 167
Verify that the reduced stiffness matrix could also be obtained by inverting the
compliance matrix, i.e. Q
Q
Q=S
S
S′−1. Now the transformed reduced stiffness ma-
trices Q
Q
Q[00],Q
Q
Q[300],Q
Q
Q[450]have to be calculated with the help of Table 4.2 tak-
ing into account that c=cos 00=1,cos 300=0,8660,cos(450) = 0,7071 and
s=sin00=0,sin300=0,5,sin(450) = 0,7071
Q
Q
Q[00]=
181,8 2,897 0
2,897 10,35 0
0 0 7,17
109Pa,
Q
Q
Q[300]=
109,4 32,46 54,19
32,46 23,65 20,05
54,19 20,05 36,74
109Pa,
Q
Q
Q[450]=
56,66 42,32 42,87
42,32 56,66 42,87
42,87 42,87 46,59
109Pa
The location of the lamina surfaces are x(0)
3=7,5 mm, x(1)
3=2,5 mm, x(2)
3=2,5
mm, x(3)
3=7,5 mm. The total thickness of the laminate is 15 mm. From (4.2.15) the
extensional stiffness matrix A
A
Afollows with
Ai j =
3
k=1
Q(k)
i j h(k)=5
3
k=1
Q(k)
i j mm,
the coupling matrix B
B
Bfollows with
Bi j =
3
k=1
Q(k)
i j h(k)¯x(k)
3=5
3
k=1
Q(k)
i j ¯x(k)
3mm
and the bending stiffness matrix Di j follows with
Di j =1
3
3
k=1
Q(k)
i j h(k) (¯x(k)
3)2+h(k)2
12 !=5
3
3
k=1
Q(k)
i j 25
12 + ( ¯x(k)
3)2mm3
with
¯x(k)
3=1
2(x(k)
3+x(k1)
3),
i.e. ¯x(1)
3=5 mm, ¯x(2)
3=0 mm, ¯x(3)
3=5 mm. Summarizing the formulas for Ai j ,Bi j
and Di j we have the equations
Ai j =5[Q(1)
i j +Q(2)
i j +Q(3)
i j ]mm,
Bi j =5[5Q(1)
i j +0Q(2)
i j 5Q(3)
i j mm2],
Di j =5[(2 5 +25/12)Q(1)
i j + (25/12)Q(2)
i j + (25 +25/12)Q(3)
i j ]mm3
168 4 Elastic Behavior of Laminate and Sandwich Composites
and the stiffness matrices follow to
A
A
A=
17,39 3,884 0,566
3,884 4,533 1,141
0,566 1,141 4,525
108Pa m,
B
B
B=
3,129 0,986 1,072
0,986 1,158 1,072
1,072 1,072 0,986
106Pa m2,
D
D
D=
33,43 6,461 5,240
6,461 9,320 5,596
5,240 5,596 7,663
103Pa m3
4.3 Elastic Behavior of Sandwiches
One special group of laminated composites used extensively in engineering appli-
cations is sandwich composites. Sandwich panels consist of thin facings, also called
skins or sheets, sandwiching a core. The facings are made of high strength material
while the core is made of thick and lightweight materials, Sect. 1.3. The motiva-
tion for sandwich structure elements is twofold. First for beam or plate bending
the maximum normal stresses occur at the top and the bottom surface. So it makes
sense using high-strength materials at the top and the bottom and using low and
lightweight strength materials in the middle. The strong and stiff facings also sup-
port axial forces. Second, the bending resistance for a rectangular cross-sectional
beam or plate is proportional to the cube of the thickness. Increasing the thickness
by adding a core in the middle increases the resistance. The maximum shear stress
is generally in the middle of the sandwich requiring a core to support shear. The
advantages in weight and bending stiffness make sandwich composites attractive in
many applications.
The most commonly used facing materials are aluminium alloys and fibre re-
inforced plastics. Aluminium has a high specific modulus, but it corrodes without
treatment and can be prone to denting. Therefore fibre reinforced laminates, such
as graphite/epoxy or glass/epoxy are becoming more popular as facing materials.
They have high specific modulus and strength and corrosion resistance. The fibre
reinforced facing can be unidirectional or woven laminae.
The most commonly used core materials are balsa wood, foam, resins with spe-
cial fillers and honeycombs (Fig. 1.3). These materials must have high compressive
and shear strength. Honeycombs can be made of plastics, paper, card-boards, etc.
The strength and stiffness of honeycomb sandwiches depend on the material and
the cell size and thickness. The following sections consider the modelling and anal-
ysis of sandwiches with thin and thick cover sheets.
4.3 Elastic Behavior of Sandwiches 169
4.3.1 General Assumptions
A sandwich can be defined as a special laminate with three layers. The thin cover
sheets, i.e. the layers 1 and 3, are laminates of the thicknesses h(1)for the lower skin
and h(3)for the upper skin. The thickness of the core is h(2)hc. In a general case
h(1)does not have to be equal to h(3), but in the most important practical case of
symmetric sandwiches h(1)=h(3)hf.
The assumptions for macro-mechanical modelling of sandwiches are:
1. The thickness of the core is much greater than that of the skins, h(2)h(1),h(3)
or hchf
2. The strains
ε
1,
ε
2,
ε
6vary linearly through the core thickness hc
ε
ε
ε
=
ε
ε
ε
+x3
κ
κ
κ
3. The sheets only transmit stresses
σ
1,
σ
2,
σ
6and the in-plane strains are uniform
through the thickness of the skins. The transverse shear stresses
σ
4,
σ
5are ne-
glected within the skin.
4. The core only transmits transverse shear stresses
σ
4and
σ
5, the stresses
σ
1,
σ
2
and
σ
6are neglected.
5. The strain
ε
3is neglected in the sheets and the core.
With these additional assumptions in the frame of linear anisotropic elasticity, the
stresses and strains can be formulated.
Strains in the lower and upper sheets:
ε
ε
ε
(l)=
ε
ε
ε
1
2hc
κ
κ
κ
,
ε
(l)
i=
ε
i1
2hc
κ
i,l=1,3,i=1,2,6 (4.3.1)
The transverse shear strains
ε
4,
ε
5are neglected.
Strains in the sandwich core:
ε
ε
ε
(2)=
ε
ε
ε
c=
ε
ε
ε
+x3
κ
κ
κ
,hc
2x3+hc
2(4.3.2)
The transverse shear strains are, in a first approach, independent of the coordinate
x3(4.2.14)
γ
γ
γ
sc = [
ε
c
5
ε
c
4]T(4.3.3)
We shall see in Chap. 5 that in the classical laminate theory and the laminate theory
including transverse shear deformations the strain vector
ε
ε
ε
is written in an analogous
form, and only the expressions for the curvatures are modified.
Stresses in the lower and upper sheets:
In the sheets a plane stress state exists and with assumption 3. the transverse shear
stresses
σ
4and
σ
5are neglected. These assumptions imply that for laminated sheets
in all layers of the lower and the upper skins
σ
(k)
4=
σ
(k)
5=0
170 4 Elastic Behavior of Laminate and Sandwich Composites
The other stresses are deduced from the constant strains
ε
(l)
1,
ε
(l)
2,
ε
(l)
6,l=1,3
by the relationships
σ
(k)
i=Q(k)
i j
ε
(l)
j,i,j=1,2,6,l=1,3 (4.3.4)
for the kth layer of the lower (l=1) or the upper (l=3) skin.
Stresses in the sandwich core:
From assumption 4. it follows
σ
c
1=
σ
c
2=
σ
c
6=0
and the core transmits only the transverse shear stresses
σ
c
5
σ
c
4=C55 C45
C45 C44 
ε
5
ε
4(4.3.5)
or in matrix notation (Eq. 4.2.14)
σ
σ
σ
s=C
C
Cs
γ
γ
γ
s(4.3.6)
The coefficients Cc
i j of C
C
Csare expressed as functions of the coefficients Cc
i j referred
to the principal directions by the transformation equation (4.2.17). The coefficients
Cc
i j in the principal directions are themselves written as functions of the shear moduli
of the core (Sect. 2.1, Table 2.5), measured in principal directions as follows
Cc
44 =Gc
23,Cc
55 =Gc
13 (4.3.7)
For an isotropic core material a transformation is not required.
4.3.2 Stress Resultants and Stress Analysis
The in-plane resultants N
N
Nfor sandwiches are defined by
N
N
N=
1
2hc
Z
(1
2hc+h(1))
σ
σ
σ
dx3+
1
2hc+h(3)
Z
1
2hc
σ
σ
σ
dx3,(4.3.8)
the moment resultants by
4.3 Elastic Behavior of Sandwiches 171
M
M
M=
1
2hc
Z
(1
2hc+h(1))
σ
σ
σ
x3dx3+
1
2hc+h(3)
Z
1
2hc
σ
σ
σ
x3dx3(4.3.9)
and the transverse shear force by
Q
Q
Qs=
1
2hc
Z
1
2hc
σ
σ
σ
sdx3(4.3.10)
For the resultants N
N
Nand M
M
Mthe integration is carried out over the sheets only and for
the transverse shear force over the core.
By substituting Eqs. (4.3.4) - (4.3.7) for the stresses into the preceding expres-
sions for the force and moment resultants, we obtain analogous to (4.2.16) the con-
stitutive equation
N
N
N
M
M
M
Q
Q
Qs
=
A
A
A B
B
B0
0
0
C
C
C D
D
D0
0
0
0
0
0 0
0
0A
A
As
ε
ε
ε
κ
κ
κ
γ
γ
γ
s
(4.3.11)
with the stiffness coefficients
Ai j =A(1)
i j +A(3)
i j ,Bi j =1
2hcA(3)
i j A(1)
i j ,
Ci j =C(1)
i j +C(3)
i j ,Di j =1
2hcC(3)
i j C(1)
i j
(4.3.12)
and
A(1)
i j =
1
2hc
Z
(1
2hc+h1)
Q(k)
i j dx3=
n1
k=1Z
h(k)
Q(k)
i j dx3=
n1
k=1
Q(k)
i j h(k),
A(3)
i j =
1
2hc+h3
Z
1
2hc
Q(k)
i j dx3=
n2
k=1Z
h(k)
Q(k)
i j dx3=
n2
k=1
Q(k)
i j h(k),
C(1)
i j =
1
2hc
Z
(1
2hc+h1)
Q(k)
i j x3dx3=
n1
k=1Z
h(k)
Q(k)
i j x3dx3=
n1
k=1
Q(k)
i j h(k)¯x(k)
3,
C(3)
i j =
1
2hc+h3
Z
1
2hc
Q(k)
i j x3dx3=
n2
k=1Z
h(k)
Q(k)
i j x3dx3=
n2
k=1
Q(k)
i j h(k)¯x(k)
3
(4.3.13)
with x(k)
3=1
2(x(k)
3+x(k1)
3and
172 4 Elastic Behavior of Laminate and Sandwich Composites
As
i j =hcCs
i j ,i,j=4,5 (4.3.14)
n1and n2are the number of layers in the lower and the upper sheet respectively and
Cs
i j are the transverse shear moduli of the core. The constitutive equations (4.3.11)
for a sandwich composite has a form similar to the constitutive equation for lami-
nates including transverse shear. It differs only by the terms Ci j instead of Bi j which
induce an unsymmetry in the stiffness matrix.
In the case of symmetric sandwiches with identical sheets h(1)=h(3)=hf,
A(1)
i j =A(3)
i j =Af
i j,C(1)
i j =C(3)
i j =Cf
i j and from this it results that the stiffness
coefficients Eq. (4.3.12) are
Ai j =2Af
i j,Di j =hcCf
i j,Bi j =0,Ci j =0 (4.3.15)
As developed for laminates including shear deformations, the coefficients As
i j can
be corrected by shear correction factors ks
i j and replaced by shear constants (ksAs)i j
to improve the modelling.
In the case of symmetric sandwiches there is no coupling between stretching
and bending and the form of the constitutive equation is identical to the constitutive
equation for symmetric laminates including transverse shear.
4.3.3 Sandwich Materials with Thick Cover Sheets
In the case of thick cover sheets it is possible to carry out the modelling and analysis
with the help of the theory of laminates including transverse shear. Considering the
elastic behavior of sandwich composites we have:
The stretching behavior is determined by the skins.
The transverse shear is imposed by the core.
The modelling assumption 1. of Sect. 4.3.1 is not valid. Restricting the modelling
to the case of symmetric sandwich composites and to the case where the core’s
principal direction is in coincidence with the directions of the reference coordinate
system. The elastic behavior of the composite material is characterized by
the reduced stiffness parameters Qf
i j for the face sheets,
the reduced stiffness parameters Qc
i j and the transverse shear moduli Cc
i j for the
core
Application of the sandwich theory, Sect. 4.3.2, leads to the following expressions
for the stiffness coefficients of the constitutive equation (upper index Sa), one lamina
ASa
i j =2hfQf
i j,BSa
i j =0,CSa
i j =0,DSa
i j =1
2Qf
i j(hf+hc)hfhc,i,j=1,2,6 (4.3.16)
The shear stiffness coefficients As
i j are in the sandwich theory
AsSa
i j =hcCi j,i,j=4,5,C44 =Gc
23,C55 =Gc
13,C45 =0 (4.3.17)
4.3 Elastic Behavior of Sandwiches 173
Application of the laminate theory including transverse shear, Sect. 4.2.2, (4.2.15)
leads (upper index La) to
ALa
i j =2hfQf
i j +hcQc
i j,BLa
i j =0,
DLa
i j =1
2Qf
i jhf(hf+hc)2+1
3(hf)2+1
12 Qc
i j(hc)3,i,j=1,2,6
(4.3.18)
The shear stiffness coefficients As
i j are now
AsLa
i j =2hfCf
i j +hcCc
i j,i,j=4,5,Cf/c
44 =Gf/c
23 ,Cf/c
55 =Gf/c
13 ,Cf/c
45 =0 (4.3.19)
For symmetric faces with nlaminae Eqs. (4.3.16) - (4.3.19) yield
ASa
i j =2
n
k=1
Qf(k)
i j hf(k),
DSa
i j =hc
n
k=1
Qf(k)
i j hf(k)¯x(k)
3,
ALa
i j =2
n
k=1
Qf(k)
i j hf(k)+hcQc
i j,
DLa
i j =
n
k=1
Qf(k)
i j hf(k)(¯x(k)
3)2+hf(k)3
12 !+Qc
i j
hc3
12
The comparison of the analysis based on the sandwich or the laminate theory yields
ALa
i j =ASa
i j 1+hcQc
i j
2hfQf
i j !,
DLa
i j =DSa
i j 1+hf
hc
hc+ (4/3)hf
hc+hf+Qc
i j
6Qf
i j
(hc)2
hf(hc+hf)!,i,j=1,2,6
(4.3.20)
AsLa
i j =AsSa
i j 1+hfCf
i j
2hcCc
i j !,i,j=4,5 (4.3.21)
Generally the core of the sandwich is less stiff than the cover sheets
Qc
i j Qf
i j
and the relations (4.3.20) can be simplified
ALa
i j ASa
i j ,DLa
i j DSa
i j 1+hf
hc
hc+ (4/3)hf
hc+hf(4.3.22)
Equation (4.3.21) stays unchanged.
The bending stiffness Di j are modified with respect to the theory of sand-
wiches and can be evaluated by the influence of the sheet thickness. If for ex-
174 4 Elastic Behavior of Laminate and Sandwich Composites
ample hc=10 mm and the sheet thickness hf=1 mm/ 3 mm or 5 mm we find
DLa
i j =1,103DSa
i j /1,326DSa
i j or 1,555DSa
i j , a difference of more than 10%, 30% or
50 %.
4.4 Problems
Exercise 4.11. The reduced stiffness Qc
i j and Qf
i j of a symmetric sandwich satisfies
the relation Qc
i j Qf
i j. Evaluate the influence of the sheet thickness on the bending
stiffness ratio DLa
i j /DSa
i j if the core thickness is constant (hc=10 mm) and the sheet
thickness vary: hf=0.5/1.0/3.0/5.0/8.0/10.0 mm.
Solution 4.11. Using the simplified formula (4.3.22) the ratio values are
1.051/1.103/1.323/1.555/1.918/2.167 i.e. the difference
DLa
i j DSa
i j
DSa
i j
100%
of the stiffness values for DLa
i j and DSa
i j are more than 5%/10%/32%/55%/91% or
116%.
Conclusion 4.2. The sandwich formulas of Sect. 4.3.2 should be used for thin cover
sheets only, i.e. hfhc.
Exercise 4.12. A sandwich beam has faces of aluminium alloy and a core of
polyurethane foam. The geometry of the cross-section is given in Fig. 4.17. Cal-
PPPPPP
P
hf
hf
x3
x2
hch
b
abc
x3x3
Fig. 4.17 Sandwich beam. aGeometry of the cross-section of a sandwich beam, bDistribution
of the bending stress, if the local stiffness of the faces and the bending stiffness of the core are
dropped, cDistribution of the shear stress, if only the core transmit shear stresses
4.4 Problems 175
culate the bending stiffness Dand the distributions of the bending and the shear
stress across the faces and the core, if the stress resultants Mand Qare given.
Solution 4.12. The bending stiffness Dof the sandwich beam is the sum of the
flexural rigidities of the faces and the core
D=2Efbhf3
12 +2Efbhfhc+hf
22
+Ecbhc3
12
Efand Ecare the effective Young’s moduli of the face and the core materials. The
first term presents the local bending stiffness of the faces about their own axes, the
third term represents the bending stiffness of the core. Both terms are generally very
small in comparison to the second term. Provided that
(hc+hf)/hf>5,77,[(Efhf)/(Echc)][(hc+hf)/hc]2>100/6
i.e.
Efbhf3/6
Ef[bhf(hc+hf)2]/2<1
100
Ecbhc3/12
Ef[bhf(hc+hf)2]/2<1
100
the first and the third term are less than 1% of the second term and the bending
stiffness is approximately
DEfbhf(hc+hf)2/2
The bending stress distributions through the faces and the core are
σ
f=MEf
Dx3 ±MEf
D
hc+hf
2,
σ
c=MEc
Dx30
The assumptions of the classical beam theory yield the shear stress equation for the
core
τ
=QS(x3)
bI =Q
bD [EfSf(x3) + EcSc(x3)]
Q
D"Efhf(hc+hf)
2+Ec
2 hc2
4x2
3!#
The maximum core shear stress will occur at x3=0. If
Efhf(hc+hf)
Echc2/4>100
the ratio of the maximum core shear stress to the minimum core shear stress is <1%
and the shear stress distribution across the core can be considered constant
τ
Q
D
Efhf(hc+hf)
2
and with
176 4 Elastic Behavior of Laminate and Sandwich Composites
DEfbhf(hc+hf)2/2
follow
τ
Q/b(hc+hf)Q/bh. In Fig. 4.17 the distributions of the bending
and shear stresses for sandwich beams with thin faces are illustrated. Note that for
thicker faces the approximate flexural bending rigidity is
DEfbhf(hc+hf)/+Efbhf3/12
Chapter 5
Classical and Improved Theories
In this chapter, the theoretical background for two commonly used structural the-
ories for the modelling and analysis of laminates and sandwiches is considered,
namely the classical laminate theory and the first-order shear deformation theory.
The classical laminate theory (CLT) and the first-order shear deformation theory
(FSDT) are the most commonly used theories for analyzing laminated or sand-
wiched beams, plates and shells in engineering applications. The CLT is an ex-
tension of Kirchhoffs1classical plate theory for homogeneous isotropic plates to
laminated composite plates with a reasonable high width-to-thickness ratio. For ho-
mogeneous isotropic plates the Kirchhoff s theory is limited to thin plates with ratios
of maximum plate deflection wto plate thickness h<0.2 and plate thickness/ mini-
mum in-plane dimensions <0.1. Unlike homogeneous isotropic structure elements,
laminated plates or sandwich structures have a higher ratio of in-plane Young’s mod-
uli to the interlaminar shear moduli, i.e. such composite structure elements have a
lower transverse shear stiffness and often have significant transverse shear deforma-
tions at lower thickness-to span ratios <0.05. Otherwise the maximum deflections
can be considerable larger than predicted by CLT. Furthermore, the CLT cannot
yield adequate correct through-the-thickness stresses and failure estimations. As a
result of these considerations it is appropriate to develop higher-order laminated and
sandwich theories which can be applied to moderate thick structure elements, e.g.
the FSDT. CLT and FSDT are so-called equivalent single-layer theories (ESLT).
Moreover a short overview of so-called discrete-layer or layerwise theories is given,
which shall overcome the drawbacks of equivalent single layer theories.
1Gustav Robert Kirchhoff (12 March 1824 onigsberg - †17 October 1887 Berlin) - physi-
cist who contributed to the fundamental understanding of electrical circuits, spectroscopy, and the
emission of black-body radiation by heated objects, in addition, he formulated a plate theory which
was an extension of the Euler-Bernoulli beam theory
177
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_5
178 5 Classical and Improved Theories
5.1 General Remarks
A classification of the structural theories in composite mechanics illustrates that the
following approaches for the modelling and analysis of beams and plates composed
of composite materials can be used:
1. So called equivalent single-layer theories: These theories are derived from the
three-dimensional elasticity theory by making assumptions concerning the kine-
matics of deformation and/or the stress distribution through the thickness of a
laminate or a sandwich. With the help of these assumptions the modelling can be
reduced from a 3D-problem to a 2D-problem. In engineering applications equiv-
alent single-layer theories are mostly used in the form of the classical laminate
theory, for very thin laminates, and the first order shear deformation theory, for
thicker laminates and sandwiches.
An equivalent single layer model is developed by assuming continuous displace-
ment and strain functions through the thickness. The stresses jump from ply to
ply and therefore the governing equations are derived in terms of thickness aver-
aged resultants. Also second and higher order equivalent single layer theories by
using higher order polynomials in the expansion of the displacement components
through the thickness of the laminate are developed. Such higher order theories
introduce additional unknowns that are often difficult to interpret in mechanical
terms. The CLT requires C1-continuity of the transverse displacement, i.e. the
displacement and the derivatives must be continuous, unlike the FSDT requires
C0-continuity only. Higher order theories generally req uire at least C1-continuity.
2. Three-dimensional elasticity theories such as the traditional 3D-formulations of
anisotropic elasticity or the so-called layerwise theories: In contrast to the equiv-
alent single-layer theories only the displacement components have to be contin-
uous through the thickness of a laminate or a sandwich but the derivatives of
the displacements with respect to the thickness coordinate x3may be discon-
tinuous at the layer interfaces. We say that the displacement field exhibits only
C0-continuity through the thickness directions.
The basic assumption of modelling structural elements in the framework of the
anisotropic elasticity is an approximate expression of the displacement components
in the form of polynomials for the thickness coordinate x3. Usually the polynomials
are limited to degree three and can be written in the form
u1(x1,x2,x3) = u(x1,x2) +
α
x3
w(x1,x2)
x1
+
β
x3
ψ
1(x1,x2)
+
γ
x2
3
φ
1(x1,x2) +
δ
x3
3
χ
1(x1,x2),
u2(x1,x2,x3) = v(x1,x2) +
α
x3
w(x1,x2)
x2
+
β
x3
ψ
2(x1,x2)
+
γ
x2
3
φ
2(x1,x2) +
δ
x3
3
χ
2(x1,x2),
u3(x1,x2,x3) = w(x1,x2) +
β
x3
ψ
3(x1,x2) +
γ
x2
3
φ
3(x1,x2)
(5.1.1)
5.1 General Remarks 179
A displacement field in the form of (5.1.1) satisfies the compatibility conditions for
strains, Sect. 2.2.1, and allows possible cross-sectional warping, transverse shear
deformations and transverse normal deformations to be taken into account. The dis-
placement components of the middle surface are u(x1,x2),v(x1,x2),w(x1,x2). In the
case of dynamic problems the time tmust be introduced in all displacement func-
tions.
The polynomial approach (5.1.1) of the real displacement field yields the follow-
ing equivalent single-layer theories
Classical laminate theories
α
=1,
β
=
γ
=
δ
=
β
=
γ
=0
First-order shear deformation theory
α
=0,
β
=1,
γ
=
δ
=
β
=
γ
=0
Second order laminate theory
α
=0,
β
=1,
γ
=1,
δ
=
β
=
γ
=0
Third order laminate theory
α
=0,
β
=1,
γ
=1,
δ
=1,
β
=
γ
=0
Theories higher than third order are not used because the accuracy gain is so little
that the effort required to solve the governing equations is not justified. A third order
theory based on the displacement field u1,u2,u3has 11 unknown functions of the
in-plane coordinates x1,x2.u,v,wdenote displacements and
ψ
1,
ψ
2rotations of the
transverse normals referred to the plane x3=0.
ψ
3has the meaning of extension of
a transverse normal and the remaining functions can be interpreted as warping func-
tions that specify the deformed shape of a straight line perpendicular to the reference
plane of the undeformed structure. In addition, any plate theory should fulfill some
consistency requirements which was first time discussed for the simplest case of a
homogeneous isotropic plate in Kienzler (2002) and later extended to other cases
by Schneider and Kienzler (2015); Schneider et al (2014). Also implementations of
higher order theories into finite element approximations cannot be recommended. If
a laminated plate is thick or the 3D stress field must be calculated in local regions,
a full 3D analysis should be carried out.
The most widely used approach reduces the polynomial function of degree three
to a linear or first order approximation, which includes the classical and the first-
order shear deformation theory
u1(x1,x2,x3) = u(x1,x2) + x3
ψ
1(x1,x2),
u2(x1,x2,x3) = v(x1,x2) + x3
ψ
2(x1,x2),
u3(x1,x2,x3) = w(x1,x2)
(5.1.2)
180 5 Classical and Improved Theories
The classical approximation can be obtained if
ψ
1(x1,x2) =
w
x1
,
ψ
2(x1,x2) =
w
x2
The number of unknown functions reduces to three, that are u,v,w. On the other
hand there are five independent unknown functions u,v,w,
ψ
1,
ψ
2.
The strain-displacement equations (2.2.3) give for the first order displacement
approximation a first order strain field model with transverse shear
ε
1=
u
x1
+x3
ψ
1
x1
,
ε
2=
v
x2
+x3
ψ
2
x2
,
ε
3=0,
ε
4=
w
x2
+
ψ
2,
ε
5=
w
x1
+
ψ
1,
ε
6=
u
x2
+
v
x1
+x3
ψ
2
x1
+
ψ
1
x2
(5.1.3)
For the in-plane strains one can write in contracted form
ε
i(x1,x2,x3) =
ε
i(x1,x2) + x3
κ
i(x1,x2),i=1,2,6,
i.e. the in-plane strains
ε
1,
ε
2and
ε
6vary linearly through the thickness h.
The stress-strain relations in on-axis coordinates are
σ
i=C
i j
ε
j,i,j=1,2,...,6
From the transformation rule (4.1.27) follow the stiffness coefficients in the off-
axis-coordinates
C
C
C=3
T
T
T
ε
T
C
C
C3
T
T
T
ε
and with (4.1.26) the constitutive equation is
σ
1
σ
2
σ
3
σ
4
σ
5
σ
6
=
C11 C12 C13 0 0 C16
C12 C22 C23 0 0 C26
C13 C23 C33 0 0 C36
0 0 0 C44 C45 0
0 0 0 C45 C55 0
C16 C26 C36 0 0 C66
ε
1
ε
2
ε
3
ε
4
ε
5
ε
6
(5.1.4)
Assuming
σ
30, the stiffness matrix can be rewritten by separating the transverse
shear stresses and strains in analogy to (4.2.3) - (4.2.5)
σ
1
σ
2
σ
6
σ
4
σ
5
=
Q11 Q12 Q16 0 0
Q12 Q22 Q26 0 0
Q16 Q26 Q66 0 0
0 0 0 C44 C45
0 0 0 C45 C55
ε
1
ε
2
ε
6
ε
4
ε
5
(5.1.5)
5.1 General Remarks 181
and from
σ
3=C13
ε
1+C23
ε
2+C33
ε
3+C36
ε
6=0
it follows
ε
3=1
C33
(C13
ε
1+C23
ε
2+C36
ε
6)
The Qi j are the reduced stiffness in the off-axis reference system
Qi j =Ci j Ci3Cj3
C33
,i,j=1,2,6,Qi j =Cij ,i,j=4,5
Summarizing, one can say that the first order displacement approach (5.1.2) includes
the classical and the shear deformation theory for laminates and sandwiches. In both
cases the in-plane displacements and strains vary linearly through the thickness, but
the explicit expressions for the curvature vector
κ
κ
κ
differ. The force and moment re-
sultants can be defined for both theories in the usual way, e.g. (4.2.13), (4.2.14), but
in the classical theory there are only constitutive equations for the in-plane force
and the moment resultants N
N
N,M
M
M. It can be proved that a CLT approach is sufficient
for very thin laminates and it has been used particularly to determine the global
response of thin composite structure elements, i.e. deflections, overall buckling, vi-
bration frequencies, etc. The FSDT approach is sufficient for determining in-plane
stresses even if the structure slenderness is not very high.
The CLT neglects all transverse shear and normal effects, i.e. structural deforma-
tion is due entirely to bending and in-plane stretching. The FSDT relaxes the kine-
matic restrictions of CLT by including a constant transverse shear strain. Both first
order theories yield a complete understanding of the through-the-thickness laminate
response. Transverse normal and shear stresses, however, play an important role in
the analysis of beams, plates and shells since they significantly affect characteris-
tic failure modes like, e.g., delamination. The influence of interlaminar transverse
stresses are therefore taken into account by several failure criteria. Simple but suf-
ficient accurate methods for determination of the complete state of stress in com-
posite structures are needed to overcome the limitations of the simple first order 2D
modelling in the frame of an extended 2D modelling. In Sects. 5.2 and 5.3 a short
description of CLT and FSDT is given including some remarks to calculate trans-
verse stress components. In Chap. 11 will be seen that both the CLT and the FSDT
yield finite elements with an economical number of degrees of freedom, both have
some drawbacks. CLT-models require C1-continuity which complicates the imple-
mentation in commonly used FEM programs. FSDT-models have the advantage of
requiring only C0-continuity but they can exhibit so-called locking effects if lami-
nates becomes thin. Further details are given in Chap. 11.
182 5 Classical and Improved Theories
5.2 Classical Laminate Theory
The classical laminate theory uses the first-order model equations (5.1.2) but makes
additional assumptions:
1. All layers are in a state of plane stress, i.e.
σ
3=
σ
4=
σ
5=0
2. Normal distances from the middle surface remain constant, i.e. the transverse
normal strain
ε
3is negligible compared with the in-plane strains
ε
1,
ε
2.
3. The transverse shear strains
ε
4,
ε
5are negligible. This assumption implies that
straight lines normal to the middle surface remain straight and normal to that
surface after deformation (Bernoulli/Kirchhoff/Love2hypotheses in the theory
of beams, plates and shells).
Further we recall the general assumption of linear laminate theory that each layer
is quasi-homogeneous, the displacements are continuous through the total thickness
h, the displacements are small compared with the thickness hand the constitutive
equations are linear.
From assumptions 2. and 3. it follows from (5.1.3) that
ψ
1(x1,x2) =
w
x1
,
ψ
2(x1,x2) =
w
x2
,(5.2.1)
and the displacement approach (5.1.2) and the strain components (5.1.3) are written
by
u1(x1,x2,x3) = u(x1,x2)x3
w(x1,x2)
x1
,
u2(x1,x2,x3) = v(x1,x2)x3
w(x1,x2)
x2
,
u3(x1,x2,x3) = w(x1,x2),
(5.2.2)
ε
1=
u
x1x3
w2
x2
1
,
ε
2=
v
x2x3
w2
x2
2
,
ε
3=0,
ε
4=0,
ε
5=0,
ε
6=
u
x2
+
v
x12x3
w2
x1
x2
(5.2.3)
The condensed form for the in-plane strains can be noted as
ε
i(x1,x2,x3) =
ε
i(x1,x2) + x3
κ
i,i=1,2,6
with
2Augustus Edward Hough Love (17 April 1863, Weston-super-Mare †5 June 1940, Oxford) -
mathematician, mathematical theory of elasticity
5.2 Classical Laminate Theory 183
ε
1=
u
x1
,
ε
2=
v
x2
,
ε
6=
u
x2
+
v
x1
,
κ
1=
w2
x2
1
,
κ
2=
w2
x2
2
,
κ
6=2
w2
x1
x2
ε
ε
ε
T= [
ε
1
ε
2
ε
6]is the vector of midplane strains (stretching and shearing) and
κ
κ
κ
T= [
κ
1
κ
2
κ
6]the vector of curvature (bending and twisting). For all klayers the
stresses are given in condensed form by
σ
(k)
i=Q(k)
i j
ε
i+x3Q(k)
i j
κ
i,i,j=1,2,6 (5.2.4)
and the stiffness equations for the stress resultants follow from (4.2.13) - (4.2.18).
The classical laminate theory is also called shear rigid theory, the material equa-
tions yield zero shear stresses
σ
4,
σ
5for zero strains
ε
4,
ε
5, in the case that the shear
stiffness has finite values. But the equilibrium conditions yield non-zero stresses
σ
4,
σ
5, if the stresses
σ
1,
σ
2and
σ
6are not all constant. This physical contradiction
will be accepted in the classical theory and the transverse shear stresses are approx-
imately calculated with the given stresses
σ
1,
σ
2,
σ
6by the equilibrium equations
(4.1.56).
The approximate calculation of transverse shear stresses can be simplified if one
assumes the case of cylindrical bending, i.e. N1=N2=N60,M60. The consti-
tutive equation (4.2.18) or the inverted Eq. (4.2.19) with N
N
N0
0
0 gives
0
0
0
···
M
M
M
=
A
A
A.
.
.B
B
B
. . . .
B
B
B.
.
.D
D
D
ε
ε
ε
···
κ
κ
κ
,
ε
ε
ε
···
κ
κ
κ
=
a
a
a.
.
.b
b
b
. . . . .
b
b
bT.
.
.d
d
d
0
0
0
···
M
M
M
(5.2.5)
that is with Eqs. (4.2.20) - (4.2.26)
ε
ε
ε
=A
A
A1B
B
B
κ
κ
κ
,M
M
M= (D
D
DB
B
BA
A
A1B
B
B)
κ
κ
κ
=D
D
D
κ
κ
κ
ε
ε
ε
=b
b
bM
M
M=B
B
BD
D
D∗−1M
M
M,
κ
κ
κ
=d
d
dM
M
M=D
D
D∗−1M
M
M(5.2.6)
For symmetric laminates are B
B
B=0
0
0,B
B
B=0
0
0,D
D
D=D
D
Dand Eqs. (5.2.6) can be replaced
by
ε
ε
ε
=0
0
0,
κ
κ
κ
=D
D
D1M
M
M(5.2.7)
The partial extensional and coupling stiffness ˜
A
A
A(x3),˜
B
B
B(x3), Fig. 5.1, become
184 5 Classical and Improved Theories
x(m1)
3x(m2)
3x(0)
3
x3
N
(m1)
Fig. 5.1 Derivation of partial stiffness ˜
A
A
A(x3)and ˜
B
B
B(x3)for the shaded part of the cross-section
˜
A
A
A(x3) =
x3
Z
x(0)
3
Q
Q
Q(x3)dx3
=
m1
k=1
Q
Q
Q(k)h(k)+Q
Q
Q(m)x3x(m1)
3,
˜
B
B
B(x3) =
x3
Z
x(0)
3
Q
Q
Q(x3)x3dx3
=
m1
k=1
Q
Q
Q(k)s(k)+1
2Q
Q
Q(m)x2
3x(m1)
3
2,
(5.2.8)
h(k)=x(k)
3x(k1)
3,s(k)=h(k)x(k)
3=1
2x(k)
3+x(k1)
3x(k)
3x(k1)
3
Outgoing from the equilibrium equations (2.2.1) the shear stress equations can be
written
5.2 Classical Laminate Theory 185
σ
5(x3) =
x3
Z
x(0)
3
σ
1
x1
+
σ
6
x2dx3
=
x3
Z
x(0)
3
x1Q(k)
1j
ε
j+x3Q(k)
1j
κ
j+
x2Q(k)
6j
ε
j+x3Q(k)
6j
κ
jdx3,
σ
4(x3) =
x3
Z
x(0)
3
σ
6
x1
+
σ
2
x2dx3
=
x3
Z
x(0)
3
x1Q(k)
6j
ε
j+x3Q(k)
6j
κ
j+
x2Q(k)
2j
ε
j+x3Q(k)
2j
κ
jdx3
(5.2.9)
or in vector-matrix notation
σ
5(x3)
σ
4(x3)=
x3
Z
x(0)
3
1 0 0
0 0 1
Q(k)
1j(
ε
j,x1+x3
κ
j,x1)
Q(k)
2j(
ε
j,x1+x3
κ
j,x1)
Q(k)
6j(
ε
j,x1+x3
κ
j,x1)
dx3
x3
Z
x(0)
3
1 0 0
0 0 1
Q(k)
1j(
ε
j,x2+x3
κ
j,x2)
Q(k)
2j(
ε
j,x2+x3
κ
j,x2)
Q(k)
6j(
ε
j,x2+x3
κ
j,x2)
dx3
(5.2.10)
with (...),x
α
=
.../
x
α
,
α
=1,2,j=1,2,6. Using Eqs. (5.2.6) - (5.2.8)
σ
σ
σ
s(x3) = B
B
B1F
F
F(x3)M
M
M,x1B
B
B2F
F
F(x3)M
M
M,x2(5.2.11)
with
σ
σ
σ
s(x3) = [
σ
5
σ
4]T,M
M
M,xi= [M1,xiM2,xiM6,xi]T,
F
F
F(x3) = [ ˜
A
A
A(x3)A
A
A1B
B
B˜
B
B
B(x3)]D
D
D∗−1=
F11 F12 F16
F21 F22 F26
F61 F62 F66
(5.2.12)
in the general case if non-symmetrical laminate and
F
F
F(x3) = ˜
B
B
B(x3)D
D
D1(5.2.13)
for symmetrical laminates,
B
B
B1=1 0 0
0 0 1 ,B
B
B2=0 0 1
0 1 0 (5.2.14)
186 5 Classical and Improved Theories
are so called Boolean3matrices. Equation (5.2.11) can also be written in component
notation.
Equations (5.2.11) and (5.2.16) constitute the straight forward equilibrium ap-
proach for transverse shear stresses which only neglects the influence of the in-plane
force derivatives N
N
N,xi, but this is a very minor restriction, since, in most engineer-
ing applications, the dominating source for transverse shear stresses are transverse
force resultants. To express the bending moment derivatives by transverse shear
stress resultants it is necessary to assume special selected displacements modes.
If one selects the cylindrical bending around the x1- and the x2-axis one obtains
M6=0,M1,x2=0,M2,x1=0
M1,x1(x1) = Qs
1(x1),M2,x2(x2) = Qs
2(x2)(5.2.15)
with the transverse forces
Qs
1(x1) = Z
(h)
σ
5(x3)dx3=
n
k=1Z
(h)
σ
(k)
5(x3)dx3,
Qs
2(x2) = Z
(h)
σ
4(x3)dx3=
n
k=1Z
(h)
σ
(k)
4(x3)dx3
(5.2.16)
Equation (5.2.11) becomes in matrix notation
σ
σ
σ
s(x3) = F
F
F(x3)Q
Q
Qs,
σ
σ
σ
s= [
σ
5(x3)
σ
4(x3)]T,Q
Q
Qs= [Qs
1(x1)Qs
2(x2)]T,
F
F
F=F11 (x3)F62 (x3)
F61 (x3)F22 (x3)(5.2.17)
Summarizing the derivations of transverse shear stresses we have considered two
cases
1. N
N
N0
0
0,M
M
M= [M1M2M6]T,
2. N
N
N0
0
0,M
M
M= [M1(x1)M2(x2)]T
In case 1. follow Eqs. (5.2.18) and in case 2. Eqs. (5.2.19)
σ
5(x3)
σ
4(x3)=F11 (x3)F12 (x3)F16(x3)
F61 (x3)F62 (x3)F66(x3)
x1
M1
M2
M6
+F61 (x3)F62 (x3)F66(x3)
F21 (x3)F22 (x3)F26(x3)
x2
M1
M2
M6
,
(5.2.18)
3George Boole (2 November 1815 Lincoln - †8 December 1864 Ballintemp) - mathematician,
educator, philosopher and logician
5.2 Classical Laminate Theory 187
σ
5(x3)
σ
4(x3)=F11 (x3)F62 (x3)
F61 (x3)F22 (x3)Q1
Q2,Qs
1=Qs
1(x1)
Qs
2=Qs
2(x2),
M1(x1)
x1
=Q1,
M2(x2)
x2
=Q2
(5.2.19)
Symmetric laminates are preferred in engineering applications. In this case
D
D
D=D
D
D,B
B
B0
0
0 and F
F
F(x3) = ˜
B
B
B(x3)D
D
D1. The calculation of the transverse shear
stresses is more simple. The approximate solution for transverse shear stresses in
the classical laminate theory satisfies the equilibrium condition. The shear stresses
are layerwise parabolic functions and there is no stress jump at the layer interfaces.
Also in the frame of the classical laminate theory an approximate constitutive
equation can be formulated
Q
Q
Qs=A
A
As
ε
ε
ε
sor Qs
1
Qs
2=A55 A45
A45 A44
ε
5
ε
4(5.2.20)
Regarding the complementary transverse shear theory formulated in shear stresses
Ws
1=1
2Z
(h)
σ
σ
σ
sT(C
C
C)1
σ
σ
σ
sdx3(5.2.21)
and in shear forces
Ws
2=1
2Q
Q
QsT(A
A
As)1Q
Q
Qs(5.2.22)
The stress vector
σ
σ
σ
sis a function of x3only, and therefore the integration is carried
out over x3. In
C
C
Cs=C55 C45
C45 C44
the Ci j,i,j=4,5 are the elastic parameters of the Hooke’s law. In Eq. (5.2.21) the
stress can be replaced by the transverse force resultants, Eq. (5.2.19). The Qs
ido not
depend on x3and Eq. (5.2.21) yields
Ws
1=1
2Q
Q
QsT
Z
(h)
F
F
FT(x3)(C
C
Cs)1F
F
F(x3)dx3
Q
Q
Qs(5.2.23)
F
F
F(x3)is the reduced elasticity matrix Eq. (5.2.18) and Eq. (5.2.23) leads to
Ws
1=1
2[Qs
1Qs
2]
Z
(h)F11 F62
F61 F22 TC55 C45
C45 C44 1F11 F62
F61 F22 dx3
Qs
1
Qs
2(5.2.24)
With Ws
1=Ws
2follows the approximate shear stiffness
188 5 Classical and Improved Theories
A
A
As=
Z
(h)
F
F
FT(C
C
Cs)1F
F
Fdx3
1
(5.2.25)
The Ci j are layerwise constant. The calculation of A
A
Asdemands an integration over
layerwise defined polynomials of 4th order and can be just simple carried out by
programming. For unsymmetrical laminates F
F
F(x3)is defined by Eq. (5.2.12).
Hygrothermal effects have no influence on the transverse shear stresses. In the
classical laminate theory for mechanical and hygrothermal loading as demonstrated
in Sect. 4.2.5, the resultants N
N
Nand M
M
Mmust be substituted by the effective resultants
˜
N
N
Nand ˜
M
M
M.
5.3 Shear Deformation Theory for Laminates and Sandwiches
The classical laminate theory allows us to calculate the stresses and strains with
high precision for very thin laminates except in a little extended region near the
free edges. The validity of the classical theory has been established by comparing
theoretical results with experimental tests and with more exact solutions based on
the general equations of the linear anisotropic elasticity theory.
If the width-to-thickness ratio is less about 20, the results derived from the clas-
sical theory show significant differences with the actual mechanical behavior and
the modelling must be improved.
A first improvement is to include approximately the effect of shear deformation
in the framework of a first-order displacement approach. A further improvement is
possible by introducing correction factors for the transverse shear moduli.
The model used now has the same general form, as (5.1.2), for the displacements,
but contrary to the classical theory,
ψ
1and
ψ
2are independent functions and a nor-
mal line to the middle surface of the composite remains straight under deformation,
however it is not normal to the deformed middle plane. In the shear deformation the-
ory the actual deformation state is approximated by 5 independent two-dimensional
functions u,v,w,
ψ
1,
ψ
2, in the classical theory by 3 functions u,v,w, respectively.
The strains are deduced from the displacements, (5.1.3). The components of the
strains
ε
ε
ε
(x1,x2,x3) =
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2),i=1,2,6
again vary linearly through the thickness hand are given by
ε
1=
u
x1
,
ε
2=
v
x2
,
ε
6=
u
x2
+
v
x1
,
κ
1=
ψ
1
x1
,
κ
2=
ψ
2
x2
,
κ
6=
ψ
2
x1
+
ψ
1
x2
(5.3.1)
5.3 Shear Deformation Theory for Laminates and Sandwiches 189
The components of the vector
ε
ε
ε
T= [
ε
1
ε
2
ε
6]are not changed, however the compo-
nents of the curvature vector
κ
κ
κ
T= [
κ
1
κ
2
κ
6]are now expressed by the derivatives
of the functions
ψ
1,
ψ
2. The stresses in the kth layer can be expressed by
σ
1
σ
2
σ
6
σ
4
σ
5
(k)
=Q
Q
Q(k)
ε
1
ε
2
ε
6
ε
4
ε
5
(k)
=Q
Q
Q(k)
u
x1
+x3
ψ
1
x1
v
x2
+x3
ψ
2
x2
u
x2
+
v
x1
+x3
ψ
2
x1
+
ψ
1
x2
w
x2
+
ψ
2
w
x1
+
ψ
1
(5.3.2)
The stresses
σ
1,
σ
2and
σ
6are superimposed on the extensional and the flexural
stresses and vary linearly through a layer thickness, the stresses
σ
4,
σ
5are, in con-
tradiction to the equilibrium equations, constant through h(k). The strains
ε
1,
ε
2,
ε
6
vary linearly and the strains
ε
4,
ε
5constant through the laminate thickness h, i.e. they
vary continuously through the total thickness. Unlike, the corresponding stresses
σ
1,
σ
2,
σ
6and
σ
4,
σ
5vary linearly or remain constant, respectively, through each
layer thickness h(k)only. Therefore is no stress continuity through the laminate
thickness but stress jumps from ply to ply at their interfaces depending on the re-
duced stiffness Q
Q
Qand Q
Q
Qs.
With the definition equations for the stress resultants N
N
N,M
M
M,Q
Q
Qsand the stiffness
coefficients Ai j,Bi j ,Di j ,As
i j for laminates (4.2.13) - (4.2.15) or sandwich (4.3.8) -
(4.3.10), (4.3.12) - (4.3.14), respectively, the constitutive equation can be written in
the condensed hypermatrix form, Eqs. (4.2.16)
N
N
N
M
M
M
Q
Q
Qs
=
A
A
A B
B
B0
0
0
B
B
B D
D
D0
0
0
0
0
0 0
0
0A
A
As
ε
ε
ε
κ
κ
κ
γ
γ
γ
s
(5.3.3)
The stretching, coupling and bending stiffness Ai j ,Bi j ,Di j stay unchanged in com-
parison to the classical laminate theory. The shear stiffness are approximately given
by
As
i j =
n
k=1
C(k)
i j h(k),i,j=4,5 (5.3.4)
The C(k)
i j are the constant shear moduli of the kth lamina. These approximated shear
stiffness overestimate the shear stiffness since they are based on the assumption of
constant transverse shear strains and also do not satisfy the transverse shear stresses
vanishing at the top and bottom boundary layers.
The stiffness values can be improved with help of shear correction factors (Vla-
choutsis, 1992; Altenbach, 2000; Gruttmann and Wagner, 2017). In this case the
part of the constitutive equation relating to the resultants N
N
N,M
M
Mis not modified. The
190 5 Classical and Improved Theories
other part relating to transverse shear resultants Q
Q
Qsis modified by replacing the
stiffness As
i j by (kA)s
i j . The constants ks
i j are the shear correction factors. A very
simple approach is to introduce a weighting function f(x3)for the distribution of
the transverse shear stresses through the thickness h.
Assume a parabolic function f(x3)
f(x3) = 5
4"1x3
h/22#(5.3.5)
and considering that for the kth layer
σ
(k)
4=Q(k)
44
ε
4+Q(k)
45
ε
5,
σ
(k)
5=Q(k)
45
ε
4+Q(k)
55
ε
5(5.3.6)
the transverse resultants are:
Q2=
n
k=1
x(k)
3
Z
x(k1)
3
σ
(k)
4f(x3)dx3
=5
4
n
k=1
Q(k)
44
ε
4
x(k)
3
Z
x(k1)
3
"1x3
h/22#dx3+
n
k=1
Q(k)
45
ε
5
x(k)
3
Z
x(k1)
3
"1x3
h/22#dx3
,
Q1=
n
k=1
x(k)
3
Z
x(k1)
3
σ
(k)
5f(x3)dx3
=5
4
n
k=1
Q(k)
45
ε
4
x(k)
3
Z
x(k1)
3
"1x3
h/22#dx3+
n
k=1
Q(k)
55
ε
5
x(k)
3
Z
x(k1)
3
"1x3
h/22#dx3
The shear stiffness coefficients As
i j of the constitutive equations
Q2=As
44
ε
4+As
45
ε
5,Q1=As
45
ε
4+As
55
ε
5(5.3.7)
are calculated by
As
i j =5
4
n
k=1
Q(k)
i j x(k)
3x(k1)
34
3h2x(k)
3
3x(k1)
3
3
=5
4
n
k=1
Q(k)
i j "h(k)4
h2h(k) h(k)2
12 +x(k)
3
2!#,i,j=4,5
(5.3.8)
5.3 Shear Deformation Theory for Laminates and Sandwiches 191
This approach yields for the case of single layer with Q44 =Q55 =G,Q45 =0 a
shear correction factor ks=5/6 for the shear stiffness Gh
As=5
4Gh4h
h2h2
12 +0=5
6Gh (5.3.9)
The weighting function (5.3.5) resulting in a shear correction factor ksis consistent
with the Reissner theory of shear deformable single layer plates (Reissner, 1944)
and slightly differ from Mindlin’s value (Mindlin, 1951).
A second method to determine shear correction factors consists of considering
the strain energy per unit area of the composite. Some remarks on this method are
given in Chaps. 7 and 8. However shear correction factors depend on the special
loading and stacking conditions of a laminate and not the only factors is generally
applicable.
A particularly physical foundation to improve the shear stiffness values A
A
Asis the
equilibrium approach, Eq. (5.2.25). The sequence of calculation steps for determin-
ing improved transverse shear stresses in the frame of the FSDT are analogous to
the CLT and shall be shortly repeated
firstly, calculate the improved shear stiffness
A
A
As=
Z
(h)
F
F
FTC
C
Cs1F
F
Fdx3
1
(5.3.10)
secondly, calculate the resultant transverse shear forces
Q
Q
Qs=A
A
As
ε
ε
ε
s(5.3.11)
thirdly, calculate the improved transverse shear stresses
σ
σ
σ
s=F
F
FQ
Q
Qs
A
A
As= [Ai j],i,j=5,4,C
C
Cs= [Ci j],i,j=5,4
F
F
F=F11 F62
F61 F22 ,Q
Q
Qs= [Qs
1Qs
2]T,
σ
σ
σ
= [
σ
5
σ
4]T,
ε
ε
ε
s= [
ε
5
ε
4]T
(5.3.12)
Relying on the results of calculation improved transverse shear stresses
σ
σ
σ
s, the
transverse normal stress can be evaluated. The following equations explain the prin-
cipal way. One starts with solving the equilibrium condition for
σ
3, Eq. (2.2.1)
σ
3(x3) =
x3
Z
x3=0
σ
5
x1
+
σ
4
x2dx3+p0(5.3.13)
p0denotes the transverse load at the starting point of integration.
With
192 5 Classical and Improved Theories
F
F
F(x3) = F11 F62
F61 F22 =f
f
fT
1
f
f
fT
2(5.3.14)
we are able to replace the transverse shear stresses in Eq. (5.3.13) by Eq. (5.3.12)
σ
3(x3) =
x3
Z
x3=0
f
f
fT
1dx3Q
Q
Qs
,x1+
x3
Z
x3=0
f
f
fT
2dx3Q
Q
Qs
,x2
+p0(5.3.15)
Only the components of f
f
f1and f
f
f2depend on x3and therefore the derivatives of Q
Q
Qs
remain unchanged by the integration process. Moreover, Eq. (5.2.12) demonstrates
that only the partial stiffness ˜
A
A
A(x3)and ˜
B
B
B(x3)depend on x3, but not the matrices A
A
A,B
B
B
and D
D
D. Therefore the integration of F
F
F(x3)yields
x3
Z
x3=0
F
F
F(x3)dx3=
x3
Z
x3=0
˜
A
A
A(x3)dx3A
A
A1B
B
B
x3
Z
x3=0
˜
B
B
B(x3)dx3
D
D
D∗−1=ˆ
F
F
F(x3)(5.3.16)
For symmetrical laminates is the coupling matrix B
B
B0
0
0 and ˆ
F
F
F(x3)can be simplified
to x3
Z
x3=0
F
F
F(x3)dx3=
x3
Z
x3=0
˜
B
B
B(x3)dx3D
D
D1=ˆ
F
F
F(x3)(5.3.17)
Now, Eq. (5.3.15) can be transformed into
σ
3(x3) = hˆ
f
f
fT
1Q
Q
Qs
,x1+ˆ
f
f
fT
2dx3Q
Q
Qs
,x2i+p0,(5.3.18)
where
ˆ
f
f
fT
1= [ ˆ
F11 ˆ
F62 ],ˆ
f
f
fT
2= [ ˆ
F61 ˆ
F22 ]
and
Qs
2,x2= (A
A
As
ε
ε
ε
s),x2
The boundary conditions of vanishing transverse shear stresses at both surfaces are
fulfilled automatically. The boundary conditions for the transverse normal stresses
must be regarded and are taken into account in the integration process.
Summarizing the considerations on single layers or smeared modelling of lam-
inated structures it can be seen that an increasing number of higher order theories
particularly for the analysis of laminated plates has been published. The vast ma-
jority falls into the class of plate theories known as displacement based ones. All
consideration in this textbook are restricted to such theories. The term ”higher order
theories” indicates that the displacement distribution over the thickness is repre-
sented by polynomials of higher than first order. In general, a higher approximation
will lead to better results but also requires more expensive computational effort and
the accuracy improvement is often so little that the effort required to solve the more
complicated equations is not justified. In addition, the mechanical interpretation of
the boundary conditions for higher order terms is very difficult. The most used ESLT
5.4 Layerwise Theories 193
in engineering applications of composite structure elements is the FSDT. The CLT
applications are limited to very thin laminates only, for in comparison to homoge-
neous isotropic plates, the values of the ratio thickness to minimum in-plane dimen-
sion to regard a plate as ”thin” or as ”moderate thick” must be considerably reduced.
Generally, fibre-reinforced material is more susceptible to transverse shear than its
homogeneous isotropic counterpart and reduces the range of applicability of CLT.
Increasing in-plane stiffness may alternatively be regarded as relevant reduction of
its transverse shear strength.
The FSDT yields mostly sufficient accurate results for the displacements and
for the in-plane stresses. However, it may be recalled, as an example, that transverse
shear and transverse normal stresses are main factors that cause delamination failure
of laminates and therefore an accurate determination of the transverse stresses is
needed.
In Sect. 5.3 it was demonstrated that one way to calculate the transverse stresses
is an equilibrium approach in the frame of an extended 2D-modelling. Another rel-
ative simple method is to expand the FSDT from five to six unknown functions or
degrees of freedom, respectively, by including an x3-dependent term into the poly-
nomial representation of the out-of-plane displacement u3(x1,x2,x3). Several other
possibilities can be found in the literature.
5.4 Layerwise Theories
Layerwise theories are developed for laminates or sandwiches with thick single lay-
ers. Layerwise displacement approximations provide a more kinematically correct
representation of the displacement functions through the thickness including cross-
sectional warping associated with the deformation of thick composite structures.
So-called partial layerwise theories are mostly used which assume layerwise expan-
sions for the in-plane displacement components only. Otherwise so-called full lay-
erwise theories use expansions for all three displacement components. Compared
with equivalent single layer models the partial layerwise model provides a more re-
alistic description of the kinematics of composite laminates and the discrete-layer
behavior of the in-plane components.
Assume a linear displacement approximation (5.1.2) for each layer
u(k)
1(x1,x2,x3) = u(k)(x1,x2) + x3
ψ
(k)
1(x1,x2),
u(k)
2(x1,x2,x3) = v(k)(x1,x2) + x3
ψ
(k)
2(x1,x2),
u(k)
3(x1,x2,x3) = w(x1,x2)
(5.4.1)
with x(k1)
3x3x(k)
3;k=1,2,...n. A laminate with nlayers is determined by
(4n+1)unknown functions u(k),v(k),
ψ
(k)
1,
ψ
(k)
2,w;k=1,2,...,n. The continuity
conditions of the displacements at the layer interfaces yield 2(n1)equations and
194 5 Classical and Improved Theories
the equilibrium for the transverse shear stresses yield additional 2(n1)equations.
With these 2 ·2(n1)equations the maximum number of the unknown functions
can be eliminated and we have independent of the number of layers in all cases
(4n+1)(4n4) = 5 unknown trial functions. An equivalent single layer model
in the first-order shear deformation theory and the partial layerwise model have
the same number of functional degrees of freedom, which are 5. The modelling
of laminates or sandwiches on the assumption of the partial layerwise theory is
often used in the finite element method. A comparison of equivalent single layer
and layerwise theories one can find in Reddy (1993).
Summarizing one can say for the class of partial or discrete layer-wise models
that all analytical or numerical equations are two-dimensional and in comparison to
a real three-dimensional modelling, their modelling and solution effort, respectively,
is less time and cost consuming. The transverse normal displacement does not have
a layerwise representation, but compared to the equivalent single layer modelling,
the partial layerwise modelling provides more realistic description of the kinematics
of composite laminates or sandwiches by introducing discrete layerwise transverse
shear effects into the assumed displacement field.
Discrete layerwise theories that neglect transverse normal strain are not capa-
ble of accurately determining interlaminar stresses and modelling localized effects
such as cutouts, free edges, delamination etc. Full or generalized layerwise theories
include in contrast to the partial layerwise transverse shear and transverse normal
stress effects.
Displacement based finite element models of partial and full layerwise theories
have been developed and can be found in the literature. In Chap. 11 the exemplary
consideration of finite beam and plate elements have been restricted to CLT and
FSDT.
5.5 Problems
Exercise 5.1. The displacement field of a third order laminate (5.1.1) may defined
by
α
=c0,
β
=1,
γ
=0,
δ
=c1,¯
β
=¯
γ
=0.
1. Formulate the displacement equations and recover the displacement equations
for the classical and the shear deformation laminate theory.
2. Introduce new variables
φ
1=
ψ
1c0
w/
x1,
φ
2=
ψ
2c0
w/
x2and express
the displacement field in terms of
φ
1and
φ
2.
3. Substitute the displacements into the linear strain-displacement relations.
4. Formulate the equations for the transverse shear stresses
σ
4,
σ
5. Find the equa-
tions for c1so that the transverse shear stresses vanish at the top and the bottom
of the laminate if c0=1.
Solution 5.1. In the case of a third order displacement field one obtains the follow-
ing answers:
1. The starting point is the displacement field
5.5 Problems 195
u1(x1,x2,x3) = u(x1,x2) + x3
ψ
1(x1,x2)c0
w(x1,x2)
x1x3
3c1
χ
1(x1,x2),
u2(x1,x2,x3) = v(x1,x2) + x3
ψ
2(x1,x2)c0
w(x1,x2)
x2x3
3c1
χ
2(x1,x2),
u3(x1,x2,x3) = w(x1,x2)
Classical laminate theory: c1=0,
ψ
1=0,
ψ
2=0,c0=1
First shear deformation theory: c0=c1=0
2. The starting point is now another displacement field
u1(x1,x2,x3) = u(x1,x2) + x3
φ
1(x1,x2)c1x3
3
χ
1(x1,x2),
u2(x1,x2,x3) = v(x1,x2) + x3
φ
2(x1,x2)c1x3
3
χ
2(x1,x2),
u3(x1,x2,x3) = w(x1,x2)
3. Using the strain-displacement equation (2.2.3) and substitute equations b) we
find
ε
1=
u1
x1
=
u
x1
+x3
φ
1
x1x3
3c1
χ
1
x1
=
ε
0
1+x3
ε
I
1+x3
3
ε
II
1,
ε
2=
u2
x2
=
v
x2
+x3
φ
2
x2x3
3c1
χ
2
x2
=
ε
0
2+x3
ε
I
2+x3
3
ε
II
2,
ε
6=
u2
x1
+
u1
x2
=
v
x1
+
u
x2
+x3
φ
2
x1
+
φ
1
x2x3
3c1
χ
2
x1
+
χ
1
x2
=
ε
0
6+x3
ε
I
6+x3
3
ε
II
6,
ε
4=
u3
x2
+
u2
x3
=
w
x2
+
φ
23c1x2
3
χ
2
=
ε
0
4+x2
3
ε
II
4,
ε
5=
u3
x1
+
u1
x3
=
w
x1
+
φ
23c1x2
3
χ
1
=
ε
0
5+x2
3
ε
II
5
Note
ε
0
i
ε
i.
4. The transverse shear stress in the kth layer of a laminate follow with (5.3.5) to
σ
(k)
4=Q(k)
44
ε
4+Q(k)
45
ε
5=Q(k)
44 (
ε
0
4+x2
3
ε
II
4) + Q(k)
45 (
ε
0
5+x2
3
ε
II
5),
σ
(k)
5=Q(k)
45
ε
4+Q(k)
55
ε
5=Q(k)
55 (
ε
0
5+x2
3
ε
II
5) + Q(k)
45 (
ε
0
4+x2
3
ε
II
4)
The transverse shear stresses shall vanish at the bottom and the top of the lami-
nate, i.e.
σ
(k)
4(±h/2) =
σ
(k)
5(±h/2) = 0 if k=1 or n.
196 5 Classical and Improved Theories
Q(1)
44
ε
0
4+h2
4
ε
II
4+Q(1)
45
ε
0
5+h2
4
ε
II
5=0,
Q(n)
44
ε
0
4+h2
4
ε
II
4+Q(n)
45
ε
0
5+h2
4
ε
II
5=0,
Q(1)
55
ε
0
5+h2
4
ε
II
5+Q(1)
45
ε
0
4+h2
4
ε
II
4=0,
Q(1)
55
ε
0
5+h2
4
ε
II
5+Q(1)
45
ε
0
4+h2
4
ε
II
4=0
=
ε
0
4+h2
4
ε
II
4=0,
ε
0
5+h2
4
ε
II
5=0
In view of the fact that for c0=1 follows
φ
2=
ψ
2
w
x2
ε
4=
ψ
2x2
33c1
χ
2,
ε
0
4+h2
4
ε
II
4=0
ψ
2=h2
43c1
χ
2
If 3c1=4/h2
χ
2=
ψ
2. Analogously follow with 3c1=4/h2that
χ
1=
ψ
1, i.e
ε
0
4+1
3c1
ε
II
4=
ψ
2
ψ
2=0,
ε
0
5+1
3c1
ε
II
5=
ψ
1
ψ
1=0
The condition 1/3c1=h2/4, i.e. c1=4/3h2is sufficient to make the transverse
shear stresses
σ
4and
σ
5zero at the top and the bottom of the laminate.
Exercise 5.2. A symmetric cross-ply laminate [00/900]Shas the properties h=1
mm, E
1=141 GPa, E
2=9,4 GPa, E
4G
23 =3,2 GPa, E
5G
13 =E
6G
12 =4,3
GPa,
ν
12 =0,3.
1. Using the simplified equations (5.2.8) to calculate the shear stresses
σ
5(x3),
σ
4(x3)and sketch their distribution across the laminate thickness hfor
given transverse force resultants Q1=
M1/
x1,Q2=
M2/
x2and M60.
2. Compare the average shear stiffness with the improved corrected stiffness values.
Solution 5.2. The solution can be obtained as follows.
1. The reduced stiffness matrix Q
Q
Qand the shear stiffness matrix C
C
CG
G
Gmust be
calculated for the four layers
00-layers,
ν
21 =
ν
12E
2/E
1:
5.5 Problems 197
Q
Q
Q[00]Q
Q
Q=
E
1
(1
ν
12
ν
21)
ν
12E
2
(1
ν
12
ν
21)0
ν
12E
2
(1
ν
12
ν
21)
E
2
(1
ν
12
ν
21)0
0 0 E
6
=
141,85 2,84 0
2,84 9,46 0
0 0 4,3
GPa,
G
G
G[00]G
G
G=G
13 0
0G
23 =4,3 0
0 3,2GPa
900-layers:
Q
Q
Q[900]=
9,46 2,84 0
2,84 141,85 0
0 0 4,3
GPa,
G
G
G[900]=3,2 0
0 4,3GPa
The bending stiffness matrix follows with (4.2.15)
Di j =
4
k=1
Q(k)
i j h(¯x(k)
3)2+ (h(k))2/12ih(k),
D
D
D=
9,654 0,207 0
0,207 1,379 0
0 0 0,314
GPamm3
The corrected flexural stiffness matrix D
D
D(5.2.6) is identical D
D
Dfor symmetric
laminates, i.e. D=D, and the F
F
F(x3)-matrix in (5.2.12) can be simplified
F
F
F(x3) = ˜
B
B
B(x3)D
D
D1
The inversion of the matrix D
D
Dyields with
=22,572 the elements D1
11 of the
inverse matrix D
D
D1
D1
11 =D22/
,D1
22 =D11/
,D1
12 =D12/
,D1
66 = (D66)1
D
D
D1=
0,104 0,016 0
0,016 0,727 0
0 0 3,185
[GPamm3]1
Using (5.2.6) the shearing coupling stiffness ˜
B
B
B(x3)for the layers of the laminate
can be calculated
198 5 Classical and Improved Theories
˜
B
B
B[00](x3) =
70,93 1,42 0
1,42 4,73 0
0 0 2,30
GPax2
3
17,73 0,36 0
0,36 1,18 0
0 0 0,58
kN,
˜
B
B
B[900](x3) =
4,73 1,42 0
1,42 70,93 0
0 0 2,30
GPax2
3
13,60 0,36 0
0,36 5,32 0
0 0 0,57
kN
and with F
F
F(x3) = ˜
B
B
B(x3)D
D
D1
σ
5(x3)
σ
4(x3)=F11 F62
F61 F22 Q1
Q2,
F[00]=6,79 0
0 2,17 x2
3+1,70 0
0 0,54 mm1,
F[900]=0,44 0
0 32,80 x2
3+1,30 0
0 2,46 mm1
σ
5[00](x3) = F11[00](x3)Q1= (6,79x2
3+1,70)Q1,
σ
5[900](x3) = F11[900](x3)Q1= (0,44x2
3+1,30)Q1,
σ
4[00](x3) = F22[00](x3)Q2= (2,17x2
3+0,54)Q2,
σ
4[900](x3) = F22[900](x3)Q2= (32,80x2
3+2,46)Q2
The distribution of the shear stresses through the laminate thickness his sketched
in Fig. 5.2.
x3
x3
x1x2
0
0
0.25
0.25
0.5
0.5
-0.25
-0.25
-0.5
-0.5
1.30
1.28
1.28
2.46
0.4
0.4
σ
5(x3)/Q1
σ
4(x3)/Q2
Fig. 5.2 Distribution of the shear stresses
σ
5(x3)/Q1and
σ
4(x3)/Q2across the laminate thickness
5.5 Problems 199
2. A simplified calculation of the average shear stiffness ¯
As
i j yields (4.2.15)
¯
As
i j =
4
k=1
G(k)
i j h(k)=¯
A
A
As=3,75 0
0 3,75 GPamm
An improved shear stiffness matrix which include the transverse shear stress dis-
tribution follows with the help of the complementary strain energy W
W=1
2Z
(h)
σ
σ
σ
sTG
G
G1
σ
σ
σ
sdx3
=1
2Q
Q
QT
Z
(h)
F
F
FTG
G
G1F
F
Fdx3
Q
Q
Q=1
2Q
Q
QTA
A
As1Q
Q
Q
With 0,5
Z
0,25
F
F
FT
[00]G
G
G1F
F
F[00]dx3=2,000 0
0 0,295 x5
3
1,667 0
0 0,246 x3
3
+0,625 0
0 0,092 x3GPa1,
0,25
Z
0
F
F
FT
[900]G
G
G1F
F
F[900]dx3=0,012 0
0 46,69 x5
3
0,119 0
0 11,68 x3
3
+0,529 0
0 1,314 x3GPa1
follows by the sum up over the four layers the improved matrix A
A
As
A
A
As=3,01 0
0 2,54 GPa
The comparison of ¯
A
A
Asand A
A
Ascan be carried out in the form
k
k
ks¯
A
A
As=A
A
As
which yields the shear correction vector
k
k
ks=0,7718
0,6513
200 5 Classical and Improved Theories
References
Altenbach H (2000) On the determination of transverse shear stiffnesses of or-
thotropic plates. Zeitschrift f¨ur angewandte Mathematik und Physik ZAMP
51(4):629–649
Gruttmann F, Wagner W (2017) Shear correction factors for layered plates and
shells. Computational Mechanics 59(1):129–146
Kienzler R (2002) On consistent plate theories. Archive of Applied Mechanics
72(4):229–247
Mindlin RD (1951) Influence of rotatory inertia and shear on flexural motions of
isotropic elastic plates. Trans ASME J Appl Mech 18:31–38
Reddy JN (1993) An evaluation of equivalent-single-layer and layerwise theories of
composite laminates. Composite Structures 25(1):21 35
Reissner E (1944) On the theory of bending of elastic plates. J Math and Phys
23:184–191
Schneider P, Kienzler R (2015) Comparison of various linear plate theories in the
light of a consistent second-order approximation. Mathematics and Mechanics of
Solids 20(7):871–882
Schneider P, Kienzler R, B ¨ohm M (2014) Modeling of consistent second-order plate
theories for anisotropic materials. ZAMM - Journal of Applied Mathematics and
Mechanics / Zeitschrift f¨ur Angewandte Mathematik und Mechanik 94(1-2):21–
42
Vlachoutsis S (1992) Shear correction factors for plates and shells. International
Journal for Numerical Methods in Engineering 33(7):1537–1552
Chapter 6
Failure Mechanisms and Criteria
Failure of structural elements can be defined in a different manner. As in the case of
buckling, a structural element may be considered failure though the material is still
intact, but there are excessive deformations. In Chap. 6 failure will be considered to
be the loss of integrity of the composite material itself.
The failure analysis procedures for metallic structures were well established a
long time ago. In the case of monolithic materials stress concentrations, e.g. around
notches and holes, cause localized failures. For brittle materials local failures may
lead to fracture and therefore to a total loss of load-carrying capability. For ductile
materials local failure may be in the form of yielding and remains localized, i.e., it
is tolerated better than brittle failure. The fail-safe philosophy has been employed
in the design of metallic structures and is standard in engineering applications. Sim-
ilar procedures for composite materials are not well defined and are the object of
intensive scientific research up to now. Failure of fibre-reinforced materials is a very
complex topic. While it is important to understand the principal mechanisms of fail-
ure, for many applications it is impossible to detail each step of the failure process.
Main causes of failure are design errors, fabrication and processing errors or unex-
pected service conditions. Design errors can be made in both material and structure.
The stress level carried by each lamina in a laminate depends on the elastic mod-
uli. This may cause large stress gradients between laminae which are oriented at
considerably large angles to each other (e.g. 900). If the stress gradients are close
to a limit value, fracture may occur. Such high levels of internal stresses in adja-
cent laminae may develop a result of external applied loads but also by temperature
and moisture changes. Though manufacturing control and material inspection tests
are carried out, structural composites with abnormalities can be produced. The me-
chanical properties of composites may be significantly reduced by high temperature
variations, impact damage, etc. Service anomalies can include improper operation,
faulty maintenance, overloads or environmental incurred damage.
If structural loadings produce local discontinuities inside the material we speak
of a crack. Micro-cracking is considered as the nucleation of micro-cracks at the
microscopic level starting from defects and may cause the initiation of material
fracture. Macro-cracking is the propagation of a fracture by the creation of new
201
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_6
202 6 Failure Mechanisms and Criteria
fracture surfaces at the macroscopic level. For composite materials the fraction ini-
tiation is generally well developed before a change in the macroscopic behavior can
be observed.
If in a laminate macro-cracks occur, it may not be catastrophic, for it is possible
that some layers fail first and the composite continues to take more loads until all
laminae fail. Failed laminae may still contribute to the stiffness and strength of the
laminate. Laminate failure estimations are based on procedures for finding the suc-
cessive loads between the first and the last ply failure of the laminate. The failure of
a single layer plays a central function in failure analysis of laminates.
In this section the elastic behavior of laminae is primarily discussed from a
macroscopic point of view. But in the case of failure estimations and strength anal-
ysis of a lamina it is important to understand the underlying failure mechanisms
within the constituents of the composites and their effect to the ultimate macro-
scopic behavior. For this reason some considerations on micro-mechanic failure
mechanisms are made first and then failure criteria are discussed more in detail.
Summarizing one can say that the ability of failure prediction is a key aspect in
design of engineering structures. The first step is to consider what is meant by fail-
ure. Material failure of metallic structures is mostly related with material yielding
or rupture, but with composites it is more complex. Therefore research is ongoing in
developing failure mechanisms and failure criteria for unidirectional fibre laminae
and their laminates and in evaluating the accuracy of the failure criteria.
6.1 Fracture Modes of Laminae
Composite fracture mechanisms are rather complex because of their anisotropic na-
ture. The failure modes depend on the applied loads and on the distribution of rein-
forcements in the composites. In continuous fibre reinforced composites the types
of fracture may be classified by these basic forms:
Intralaminar fracture,
interlaminar fracture,
translaminar fracture.
Intralaminar fracture is located inside a lamina, interlaminar fracture shows the fail-
ure developed between laminae and translaminar fracture is oriented transverse to
the laminate plane. Inter- and intralaminar fractures occur in a plane parallel to that
of the fibre reinforcement.
Composite failure is a gradual process. The degradation of a layer results in a
redistribution of stresses in the laminate. It is characterized by different local failure
modes
The failure is dominated by fiber degradation, e.g. rupture, microbuckling, etc.
The failure is dominated by matrix degradation, e.g. crazing.
The failure is dominated by singularities at the fiber-matrix interface, e.g. crack
propagation, delamination, etc.
6.1 Fracture Modes of Laminae 203
Failure modes of sandwich material may be characterized by
Tensile failure of the sandwich faces
Wrinkling failure of the faces due to compressive stresses. Wrinkling is charac-
terized by the eigenmodes of buckling faces.
Shear failure of core or adhesive failure between core and face.
Crushing failure of the face and core at a support or tensile respectively shear
failure at fasteners.
The following considerations are restricted to the strength of an unidirectional layer
and to the development of reliable criteria for the predicting of the failure of lam-
inae and laminates. The failure criteria in engineering applications are mainly of a
phenomenological character, i.e. analytical approximations of experimental results,
e.g. by curve fitting.
The fracture of a UD-lamina is the result of the accumulation of various elemen-
tary fracture mechanisms:
Fibre fracture,
transverse matrix fracture,
longitudinal matrix fracture
fracture of the fibre-matrix interface.
Figure 6.1 illustrates various fracture modes of a single layer. In the fibre direc-
tion, as a tensile load is applied, Fig. 6.1a, failure is due to fibre tensile fracture.
2=T
1=L
3
σ
L
σ
L
σ
L>0
σ
L<0
σ
T
σ
T>0
σ
LT
σ
LT
σ
T
σ
T<0
a
b
Fig. 6.1 Fracture modes of a single layer in the case of elementary load states. aFibre fracture by
pure tension
σ
L>0 or compression
σ
L<0 (micro-buckling), bMatrix fracture by pure tension
σ
T>0, pure shearing
σ
LT and pure compression
σ
T<0
204 6 Failure Mechanisms and Criteria
One fibre breaks and the load is transferred through the matrix to the neighboring
fibres which are overloaded and fail too. The failure propagates rapidly with small
increasing load. Otherwise a tensile fracture perpendicular to the fibres, Fig. 6.1b,
due a combination of different micromechanical failure mechanisms: tensile failure
of matrix material, tensile failure of fibres across the diameters, failure of the inter-
face between fibre and matrix. The shear strength, Fig. 6.1b, is limited by the shear
strength of the matrix material, the shear strength between the fibre and the ma-
trix, etc. Figure 6.2 shows the basic strength parameters of a unidirectional lamina
referred to the principal material axes. For in-plane loading of a lamina 5 strength
parameters are necessary, but it is important to have in mind that for composite ma-
terials different strength parameters are measured for tensile and for compression
tests. If the shear stresses act parallel or transverse to the fibre orientation there is
σ
L>0
σ
L>0
σ
L<0
σ
L<0
σ
T>0
σ
T>0
σ
T<0
σ
T<0
σ
LT
σ
LT
material property:
σ
Lt material property:
σ
Lc
material property:
σ
Tc
material property:
σ
Tt
material property:
τ
S
ab
d
c
e
Fig. 6.2 Basic strength parameters. aLongitudinal tensile strength
σ
Lt,bLongitudinal compres-
sive strength
σ
Lc,cTransverse tensile strength
σ
Tt,dTransverse compressive strength
σ
Tc,eIn-
plane (intralaminar) shear strength
τ
S
6.1 Fracture Modes of Laminae 205
no influence of the load direction (Fig. 6.3a). Otherwise the positive shear stress
σ
6>0 causes tensile in L-direction and compression in T-direction and vice versa
for
σ
6<0 and other strength parameters are standard. The required experimental
characterization is relatively simple for the parameters
σ
Lt and
σ
Tt, but more com-
plicated for the strength parameters
σ
Lc,
σ
Tc and
τ
S.
In the case of laminates, besides the basic failure mechanisms for a single
layer, such as fibre fracture, longitudinal and transverse matrix fraction, fibre-matrix
debonding, etc. described above, another new fracture mode occurs. This mode is
called delamination and consists of separation of layers from one another. Through-
the-thickness variation of stresses may be caused even if a laminate is loaded by
uniform in-plane loads. Generally, the matrix material that holds the laminae of a
laminate together has substantially smaller strength than the in-plane strength of the
layers. Stresses perpendicular to the interface between laminae may cause breaking
of the bond between the layers in mostly localized, small regions. However, even if
σ
6<0
❅
σ
L=
σ
6
σ
T=
σ
6
σ
6>0
❅
σ
L=
σ
6
σ
T=
σ
6
σ
LT <0
❅
σ
2=
σ
LT
σ
1=
σ
LT
σ
LT >0
❅
σ
2=
σ
LT
σ
1=
σ
LT
b
a
1
1=L
2
2=T
2=T
2
1=L
1
Fig. 6.3 In-plane shear. aPositive and negative shear stresses along the principal material axes,
bPositive and negative shear stresses at 450with the principal material axis
206 6 Failure Mechanisms and Criteria
the size of such delaminations is small they may affect the integrity of a laminate
and can degrade their in-plane load-carrying capability. Therefore, in practical en-
gineering applications it is important to calculate the interlaminar normal and shear
stresses
σ
3,
σ
4and
σ
5and to check interlaminar failure too.
The definition of failure may change from case to case and depends on the com-
posite material and the kind of loads. For composite material, such as UD-laminates,
the end of the elastic domain is associated with the development of micro-cracking.
But in the first stage, the initiated cracks do not propagate and their development
changes the stiffness of the material very gradually but the degradation is irre-
versible. In the following section failure criteria for laminae will be discussed first
to allow the designer to have an evaluation of the mechanical strength of laminae.
Secondly, concepts for laminate failure are considered.
6.2 Failure Criteria
Failure criteria for composites are many and varied. In their simplest form they are
similar, in principle, to those used for isotropic materials, e.g. maximum stress/strain
and distortional energy theories. The major difference between isotropic materials
and unidirectional fibrous composite materials is the directional dependence of the
strength on a macrosopic scale. It is important to realize that failure criteria are
purely empirical. Their purpose is to define a failure envelope by using a minimum
number of test data. Generally, these experimental data are obtained from relatively
simple uniaxial and pure shear tests. Combined stress tests are more difficult to per-
form and should be, if possible, not included in the determination failure envelopes.
We shall start by considering a single lamina before moving on to discuss failure
of laminates. Longitudinal tension or pressure, transverse tension or pressure and
shear are the five basic modes of failure of a lamina. Generally the strength in the
principal material axes are regarded as the fundamental parameters defining failure.
When the lamina is loaded at an angle to the fibres one has to determine the stresses
in the principal directions and compare them with the fundamental strength param-
eters. Failure criteria usually grouped in literature into three different classes: limit
criteria, interactive criteria and hybrid criteria which combine selected aspects of
limit and interactive methods. In the following we only discuss selected criteria of
the first two classes.
Failure criteria for homogeneous isotropic materials are well established. Macro-
mechanical failure theories for composite materials have been developed by extend-
ing and adapting isotropic failure theories to account for anisotropy in stiffness and
strength of the composite. All theories can be expressed as functions of the basic
strength parameters referred to the principal material axes (Fig. 6.2). Some criteria
do not account for interaction of stress components while others do. Some inter-
action criteria require additional strength parameters obtained by more expended
biaxial experimental tests.
6.2 Failure Criteria 207
Laminate failure criteria are applied on a ply-by-ply basis and the load-carrying
capability of the entire composite is predicted by the laminate or sandwich theories
given in Chaps. 4 and 5. A laminate may be assumed to have failed when the strength
criterion of any one of its laminae is reached (first-ply failure). However, the failure
of a single layer not necessarily leads to a total fracture of the laminate structure.
Criteria of an on-axis lamina can be determined with relative easily. Off-axis criteria
can be obtained by coordinate transformations of stresses or strains. Based on the
ply-by-ply analysis first-ply failure and last-ply failure concepts can be developed.
Failure criteria have been established in the case of a layer. Of all failure criteria
available, the following four are considered representative and more widely used:
Maximum stress theory
Maximum strain theory
Deviatoric or distorsion strain energy criteria of Tsai-Hill1
Interactive tensor polynomial criterion of Tsai-Wu2
Maximum stress and maximum strain criteria assume no stress interaction while
the other both include full stress interaction. In the maximum stress theory, failure
occurs when at least one stress component along one of the principal material axes
exceeds the corresponding strength parameter in that direction
σ
L=
σ
Lt,
σ
L>0,
σ
T=
σ
Tt,
σ
T>0,
σ
L=
σ
Lc,
σ
L<0,
σ
T=
σ
Tc,
σ
T<0,
|
σ
LT|=
τ
S,
(6.2.1)
Note that failure can occur for more than one reason. A layer failure does not occur
if
σ
Lc <
σ
L<
σ
Lt,
σ
Tc <
σ
T<
σ
Tt,
τ
S<
σ
LT <
τ
S
(6.2.2)
For a two-dimensional state of normal stresses, i.e.
σ
L6=0,
σ
T6=0,
σ
LT =0, the fail-
ure envelope, Fig. 6.4, takes the form of a rectangle. In the case of off-axis tension
or compression of a UD-lamina, Fig. 6.5, the transformed stresses are
σ
L=
σ
1cos2
θ
=
σ
1c2
σ
T=
σ
1sin2
θ
=
σ
1s2
σ
LT =
σ
1sin
θ
cos
θ
=
σ
1sc
σ
1=
σ
L/c2
σ
1=
σ
T/s2
σ
1=
σ
LT/sc
(6.2.3)
and the maximum stress criteria is expressed as follows
1Rodney Hill (11 June 1921, Stourton, Leeds - 2 February 2011, Yorkshire) - applied mathe-
matician and a former Professor of Mechanics of Solids at Gonville and Caius College, Cambridge,
UK
2Edward Ming-Chi Wu (30 September 1938 - †3 June 2009) - US-American engineer
208 6 Failure Mechanisms and Criteria
σ
T
σ
L
σ
T
σ
L
σ
T
σ
L
σ
Tt
σ
Tc
σ
Lc
σ
Lt
Fig. 6.4 Failure envelope for UD-lamina under biaxial normal loading (max. stress criterion)
θ
σ
1
σ
1
x2
x1
x
1=xL
x
2=xT
θ
Fig. 6.5 Off-axis unidirectional loading
σ
Lc <
σ
1c2<
σ
Lt,
σ
Tc <
σ
1s2<
σ
Tt,
τ
S<
σ
1sc <
τ
S
(6.2.4)
The ultimate strength for
σ
1corresponds to the smallest of the following six values
σ
1t =
σ
Lt/c2,
σ
1t =
σ
Tt/s2,
σ
1t =
τ
S/sc,
σ
1>0,
σ
1c =
σ
Lc/c2,
σ
1c =
σ
Tc/s2,
σ
1c =
τ
S/sc,
σ
1<0(6.2.5)
The failure modes depend on the corresponding ultimate strength
σ
1u
σ
1u =
σ
Lt/c2fibre failure,
σ
1u =
σ
Tt/s2transverse normal stress failure,
σ
1u =
τ
S/sc in-plane shear failure
In the more general case of off-axis loading, the stress transformation rule, Table
4.1, is used
6.2 Failure Criteria 209
σ
1
σ
L
σ
2
σ
T
σ
6
σ
LT
=
c2s22sc
s2c22sc
sc sc c2s2
σ
1
σ
2
σ
6
(6.2.6)
Because of the orthotropic symmetry, shear strength is independent of the sign of
σ
LT (Fig. 6.3) and there are ve independent failure modes in the maximum stress
criterion. There is no interaction among the modes although in reality the failure
processes are highly interacting. The maximum stress theory may be applicable
for brittle modes of failure of material, e.g. follow from transverse or longitudinal
tension (
σ
L>0,
σ
T>0).
The maximum strain theory is quite similar to the maximum stress theory. Now
the strains are limited instead of the stresses. Failure of a lamina occurs when at
least one of the strain components along the principal material axes exceeds the
corresponding ultimate strain in that direction
ε
L=
ε
Lt
ε
L>0,
ε
T=
ε
Tt
ε
T>0,
ε
L=
ε
Lc
ε
L<0,
ε
T=
ε
Tc
ε
T<0,
|
ε
LT|=
ε
S
(6.2.7)
The lamina failure does not occur if
ε
Lc <
ε
L<
ε
Lt,
ε
Tc <
ε
T<
ε
Tt,
ε
S<
ε
LT <
ε
S
(6.2.8)
In the case of unidirectional off-axis tension or compression (Fig. 6.5), the stress
relations are given by (6.2.3). For the in-plane stress state strains in the principal
material axes are
ε
L
ε
T
ε
LT
=
S
11 S
12 0
S
12 S
22 0
0 0 S
66
σ
L
σ
T
σ
LT
(6.2.9)
By associating (6.2.3) and (6.2.6) and expressing the compliance parameters S
i j as
functions of the engineering moduli in the principal directions, EL,ET,GLT ,
ν
LT,
ν
T L, it follows that
ε
L=1
EL
(c2
ν
LTs2)
σ
1,
ε
T=1
ET
(s2
ν
TLc2)
σ
1,
ε
LT=1
GLT
sc
σ
1
(6.2.10)
The maximum strain and the maximum stress criteria must lead to identical values
in the cases of longitudinal loading and
θ
=00or transverse unidirectional loading
and
θ
=900. The identity of the shear equations is given in both cases. This implies
that
210 6 Failure Mechanisms and Criteria
ε
Lt =
σ
Lt
EL
,
ε
Lc =
σ
Lc
EL
,
ε
Tt =
σ
Tt
ET
,
ε
Tc =
σ
Tc
ET
,
ε
S=
τ
S
GLT
(6.2.11)
and the maximum strain criterion may be rewritten as follows
σ
Lc <
σ
1(c2
ν
LTs2)<
σ
Lt,
σ
Tc <
σ
1(s2
ν
TLc2)<
σ
Tt,
τ
S<
σ
1sc <
τ
S
(6.2.12)
By comparing Eqs. (6.2.4) and (6.2.12) we establish that the two criteria differ by
the introduction of the Poisson’s ratio
ν
LT in the strain criterion. In practice these
terms modify the numerical results slightly. In the special case of a two-dimensional
stress state
σ
L6=0,
σ
T6=0,
σ
LT =0, compare Fig. 6.4, the failure envelope takes the
form of a parallelogram for the maximum strain criterion, Fig. 6.6.
One of the first interactive criteria applied to anisotropic materials was introduced
by Hill. For a two-dimensional state of stress referred to the principal stress direc-
tions, von Mises3developed a deviatoric or distortional energy criterion for isotropic
ductile metals (von Mises, 1913) which can be presented for the two-dimensional
stress state as
σ
2
I+
σ
2
II
σ
I
σ
II =
σ
eq
or in a general reference system
σ
2
1+
σ
2
2
σ
1
σ
2+3
σ
2
6=
σ
eq
σ
I,
σ
II are principal stresses,
σ
eq the equivalent stress. This criterion was extended
in von Mises (1928) and modified for the case of orthotropic ductile materials by
Hill (1948)
A
σ
2
1+B
σ
2
2+C
σ
1
σ
2+D
σ
2
6=1 (6.2.13)
σ
L
σ
T
σ
Tt
σ
Tc
σ
Lt
σ
Lc
σ
T
ν
TL
σ
L=
σ
Tt
σ
L
ν
LT
σ
T=
σ
Lt
σ
T
ν
TL
σ
L=
σ
Tc
σ
L
ν
LT
σ
T=
σ
Lc
Fig. 6.6 Failure envelope for UD-lamina under biaxial normal loading (max. strain criterion)
3Richard Edler von Mises (19 April 1883 Lemberg, Austria-Hungary (now Lviv, Ukraine) -
14 July 1953 Boston, Massachusetts) - mathematician who worked on solid mechanics, fluid
mechanics, aerodynamics, aeronautics, statistics and probability theory, one of founders of the
journal Zeitschrift ur Angewandte Mathematik und Mechanik
6.2 Failure Criteria 211
A,B,C,Dare material parameters. Equation (6.2.13) cannot be defined as distorsion
energy, since in anisotropy distorsion and dilatation energies are not separated. The
criterion (6.2.13) was applied to UD-laminae by Tsai and Wu (1971)
A
σ
2
L+B
σ
2
T+C
σ
L
σ
T+D
σ
2
LT =1 (6.2.14)
The material parameters A,B,C,Dcan be identified by tests with acting basic load-
ings
σ
L=
σ
LU,
σ
T=0,
σ
LT =0A=1
σ
2
LU
,
σ
L=0,
σ
T=
σ
TU,
σ
LT =0B=1
σ
2
TU
,
σ
L=0,
σ
T=0,
σ
LT =
τ
UD=1
τ
2
U
(6.2.15)
In dependence on the failure mode, the superscript U must be substituted by t,c or s
and denotes the ultimate value of stress at failure.
The remaining parameter Cmust be determined by a biaxial test. The C-term
yields the interaction between the normal stresses. Under equal biaxial normal load-
ing
σ
L=
σ
T6=0,
σ
LT =0 it can be assumed that the failure follows the maximum
stress criterion, i.e failure will occur when the transverse stress reaches the trans-
verse strength
σ
TU which is much lower than the longitudinal strength
σ
LU. Equa-
tion (6.2.14) yields
σ
L
σ
LU 2
+
σ
T
σ
TU 2
+C
σ
2
T=1,
σ
T=
σ
TU =C=1
σ
2
LU
(6.2.16)
The Tsai-Hill criterion in the case of plane stress state and on-axis loading may be
written
σ
L
σ
LU 2
+
σ
T
σ
TU 2
σ
L
σ
T
σ
2
LU
+
σ
LT
τ
U2
=1 (6.2.17)
In the case of tension or compression off the principal material directions, Fig. 6.5,
the Tsai-Hill criterion becomes
σ
1c2
σ
LU 2
+
σ
1s2
σ
TU 2
σ
1cs
σ
LU 2
+
σ
1sc
τ
U2
=1 (6.2.18)
and the strength parameter
σ
1Uin x1-direction is
1
σ
1U 2
=c2
σ
LU 2
+s2
σ
TU 2
+1
τ
2
U1
σ
2
LU c2s2
c2
σ
LU 2
+s2
σ
TU 2
+cs
τ
2
U2(6.2.19)
The approximated form presumes
σ
LU
τ
U.
212 6 Failure Mechanisms and Criteria
The Tsai-Hill criterion is a single criterion instead of the three subcriteria re-
quired in maximum stress and strain theories. It allows considerable interaction
among the strain components and for ductile material the failure estimation agrees
well with experimental results.
Gol’denblat and Kopnov (1965) proposed a tensor polynomial criterion. Tsai and
Wu modified this criterion by assuming the existence of a failure surface in stress
space. They took into account only the first two terms of the polynomial criterion
and postulated that fracture of an anisotropic material occurs when the following
equation is satisfied
ai j
σ
i j +ai jkl
σ
i j
σ
kl =1 (6.2.20)
or in a contracted notation
ai
σ
i+ai j
σ
i
σ
j=1 (6.2.21)
We are interested in the case of an orthotropic composite material, i.e. a unidirec-
tional lamina, subjected to plane stress state, and the Tsai-Wu criterion may be ex-
pressed as
aL
σ
L+aT
σ
T+aS
σ
S+aLL
σ
2
L+aTT
σ
2
T+
aSS
σ
2
S+2aLT
σ
L
σ
T+2aLS
σ
L
σ
S+2aTS
σ
T
σ
S=1(6.2.22)
Equation (6.2.22) is written in the on-axis system and
ν
LT aS.
The linear terms take account the actual differences between composite material
behavior under tension and compression. The term aLT
σ
L
σ
Trepresents independent
interaction among the stresses
σ
Land
σ
Tand the remaining quadratic terms describe
an ellipsoid in stress space. Since the strength of a lamina loaded under pure shear
stress
τ
Sin the on-axis system is independent of the sign of the shear stress, all linear
terms in
σ
Smust vanish
aS=aLS =aTS =0 (6.2.23)
Then the Tsai-Wu criterion for a single layer in on-axis system has the form
aL
σ
L+aT
σ
T+aLL
σ
2
L+aTT
σ
2
T+aSS
σ
2
S+2aLT
σ
L
σ
T=1 (6.2.24)
The four quadratic terms in (6.2.24) correspond to the four independent elastic char-
acteristics of orthotropic materials, the linear terms allow the distinction between
tensile and compressive strength. The coefficients of the quadratic Tsai-Wu crite-
rion are obtained by applying elementary basic loading conditions to the lamina
σ
L=
σ
Lt,
σ
T=
σ
S=0
σ
L=
σ
Lc,
σ
T=
σ
S=0aL
σ
Lt +aLL
σ
2
Lt =1
aL
σ
Lc +aLL
σ
2
Lc =1
aL=1
σ
Lt 1
σ
Lc
aLL =1
σ
Lt
σ
Lc
σ
T=
σ
Tt,
σ
L=
σ
S=0
σ
T=
σ
Tc,
σ
L=
σ
S=0aT
σ
Tt +aTT
σ
2
Tt =1
aT
σ
Tc +aTT
σ
2
Tc =1
aT=1
σ
Tt 1
σ
Tc
aTT =1
σ
Tt
σ
Tc
6.2 Failure Criteria 213
σ
S=
τ
S,
σ
L=
σ
T=0aSS
τ
2
S=1aSS =1
τ
2
S
(6.2.25)
The remaining coefficient aLT must be obtained by biaxial testing
σ
L=
σ
T=
σ
U,
σ
S=0
(aL+aT)
σ
U+ (aLL +aTT +2aLT)
σ
2
U=1(6.2.26)
σ
Uis the experimentally measured strength under equal biaxial tensile loading
σ
L=
σ
T.
In many cases the interaction coefficient is not critical and is given approxi-
mately. A sufficient approximation is in this case
aLT 1
2aLLaTT (6.2.27)
The Tsai-Wu criterion may also be formulated in strain space.
Summarizing the considerations on interactive failure criteria lead: The Tsai-Hill
and the Tsai-Wu failure criteria are quadratic interaction criteria which have the
general form
Fi j
σ
i
σ
j+Fi
σ
i=1,i,j=L,T,S (6.2.28)
Fi j and Fiare strength parameters and
σ
i,
σ
jthe on axis stress components.
For plane stress state six strength parameters FLL,FTT ,FSS ,FLT,FL,FTare re-
quired for implementation of the failure criterion, FLS =FTS =FS=0, see Eq.
(6.2.23). Five of these strength parameters are conventional tensile, compressive
or shear strength terms which can be measured in a conventional experimental test
programme. The strength parameter FLT is more difficult to obtain, since a biaxial
test is necessary and such test is not easy to perform. The two-dimensional repre-
sentation of the general quadratic criterion (6.2.28) in the stress space can be given
in the equation below
σ
2
L
σ
Lt
σ
Lc
+
σ
2
T
σ
Tt
σ
Tc
+
σ
2
LT
τ
2
S
+2FLT
σ
L
σ
T+1
σ
Lt 1
σ
Lc
σ
L+1
σ
Tt 1
σ
Tc
σ
T=1
(6.2.29)
Equation (6.2.29) reduces, e.g., for
σ
Lt =
σ
Lc,
σ
Tt =
σ
Tc,FLT =1
2
σ
2
Lt
to the Tsai-Hill criterion, for
σ
Lt 6=
σ
Lc,
σ
Tt 6=
σ
Tc,FLT =1
2
σ
Lt
σ
Lc
to the Hoffman criterion, and for
σ
Lt 6=
σ
Lc,
σ
Tt 6=
σ
Tc,FLT =1
2
σ
Lt
σ
Lc
σ
Tt
σ
Tc
214 6 Failure Mechanisms and Criteria
to the Tsai-Wu criterion. Hoffman’s criterion is a simple generalization of the Hill
criterion that allows different tensile and compressive strength parameters (Hoff-
man, 1967).
If one defines dimensionless stresses as
σ
L=FLL
σ
L,
σ
T=FTT
σ
T,
σ
LT =pFSS
σ
LT
and normalized strength coefficients as
F
L=FL/FLL,F
T=FT/FT,F
LT =FLT/FLLFTT
Equation (6.2.29) can be rewritten as
σ
2
L+
σ
2
T+
σ
2
LT +2F
LT
σ
L
σ
2
T+F
L
σ
2
L+F
T
σ
2
T=1 (6.2.30)
Note that in the case of isotropic materials with
σ
Lt =
σ
Lc =
σ
Tt =
σ
Tc follow
F
L=F
T=0. There the principal stress state will have
σ
LT =
σ
LT =0. Equation
(6.2.30) reduces with F
LT =1
2to the known von Mises criterion.
Using the above failure criteria the possibility of a lamina failing can be deter-
mined, for example. In the maximum stress criterion, the lamina failes if any of the
inequalities (6.2.4) are violated. However, the criterion does not give information
about how much the load can be increased by if the lamina is safe or how much it
can be decreased if the lamina has failed. To overcome this problem, strength ratios
are defined as
R=maximum load which can be applied
load applied (6.2.31)
This definition is applicable to all failure criteria. If R>1, then the lamina is safe
and the applied load can be increased by a factor of R. If R<1 the lamina is unsafe
and the applied load needs to be reduced. A value of R=1 implies the failure load.
The stress ratio factor assumes that the material is linear elastic, for each state of
stress there is a corresponding state of strain and all components of stress and strain
increase by the same proportion.
Summarizing the discussion above, the strength ratio for the four criteria can be
formulated:
Maximum stress criterion
RLt
σ
=
σ
Lt/
σ
L,
σ
L>0 Strength factor fibre fracture,
RTt
σ
=
σ
Tt/
σ
T,
σ
T>0 Strength factor matrix fracture,
RLc
σ
=
σ
Lc/|
σ
L|,
σ
L<0 Strength factor micro-buckling,
RTc
σ
=
σ
Tc/|
σ
T|,
σ
T<0 Strength factor matrix fracture,
RS
σ
=
τ
S/|
σ
LT|,Strength factor matrix fracture
(6.2.32)
6.2 Failure Criteria 215
Maximum strain criterion
RLt
ε
=
ε
Lt/
ε
L,
ε
L>0,
RTt
ε
=
ε
Tt/
ε
T,
ε
T>0,
RLc
ε
=
ε
Lc/|
ε
L|,
ε
L<0,
RTc
ε
=
ε
Tc/|
ε
T|,
ε
T<0,
RS
ε
=
ε
S/|
ε
LT|
(6.2.33)
Tsai-Hill-criterion
Only one strength ratio can be introduced
RTH
σ
L
σ
LU 2
+RTH
σ
T
σ
TU 2
RTH
σ
LRTH
σ
T
σ
2
LU
+RTH
σ
LT
τ
U2
=1 (6.2.34)
With the ultimate strength
σ
LU,
σ
TU for tension and compression the strength ratio
RTH follows from
1
(RTH)2=
σ
L
σ
LU 2
+
σ
T
σ
TU 2
σ
L
σ
T
σ
2
LU
+
σ
LT
τ
U2
Tsai-Wu-criterion
The Tsai-Hill and the Tsai-Wu criterion define only one strength ratio RTW
(aL
σ
L+aT
σ
T)RTW + (aLL
σ
2
L+aTT
σ
2
T+aSS
σ
2
S+2aLT
σ
L
σ
T)RTW2=1
or in symbolic notation
ARTW +B(RTW)2=1(RTW )2+A
BRTW =1
B
with the solutions
RTW
1/2=1
2
A
B±r1
4
A2
B2+1
B=1
2BA±pA2+4B
RTW must be positive
RTW =pA2+4BA/2B(6.2.35)
The procedure for laminate failure estimation on the concept of first ply and last
ply failure is given as follows:
1. Use laminate analysis to find the midplane strains and curvatures depending on
the applied mechanical and hygrothermical loads.
2. Calculate the local stresses and strains in each lamina under the assumed load.
3. Use the ply-by-ply stresses and strains in lamina failure theory to find the strength
ratios. Multiplying the strength ratio to the applied load gives the load level of
the failure of the first lamina. This load may be called the first ply failure load.
Using the conservative first-ply-failure concept stop here, otherwise go to step 4.
216 6 Failure Mechanisms and Criteria
4. Degrade approximately fully the stiffness of damaged plies. Apply the actual
load level of previous failure.
5. Start again with step 3. to find the strength ratios in the undamaged laminae. If
R>1 multiply the applied load by the strength ratio to find the load level of the
next ply failure. If R<1, degrade the stiffness and strength characteristics of all
damaged lamina.
6. Repeat the steps above until all plies have failed. That is the last-ply-failure con-
cept.
The laminate failure analysis can be subdivided into the following four parts. The
first-ply-failure concept demands only one run through, the last-ply-failure requires
several iterations with degradation of lamina stiffness.
Failure analysis of laminates in stress space:
Step 1
Calculate the stiffnesses
Q
Q
Q(k)=
Q
11 Q
12 0
Q
12 Q
22 0
0 0 Q
66
QLL QLT 0
QLT QTT 0
0 0 QSS
of all ksingle layers in on-axis system with help of the layer moduli
E(k)
L,E(k)
T,
ν
(k)
TL ,G(k)
TL and the layer thicknesses h(k)
Transformation of the reduced stiffnesses Q
Q
Q(k)of single layers in on-axis
system to the reduced stiffnesses Q
Q
Q(k)of single layers in off-axis system
Q
Q
Q(k)= (T
T
T
ε
)TQ
Q
Q(k)T
T
T
ε
Calculate the laminate stiffnesses A
A
A,B
B
Band D
D
D
Ai j =
n
k=1
Q(k)
i j h(k),Bi j =1
2
n
k=1
Q(k)
i j x(k)
3
2x(k1)
3
2=
n
k=1
Q(k)
i j h(k)x(k)
3,
Di j =1
3
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3=
n
k=1
Q(k)
i j h(k)x(k)
3
2+1
12h(k)2,
x(k)
3=1
2x(k)
3+x(k1)
3
Inversion of the matrices A
A
A,B
B
Band D
D
D
Calculate the compliance matrices a
a
a,b
b
b,c
c
cand d
d
dof the laminate
a
a
a=A
A
AB
B
BD
D
D1C
C
C,b
b
b=B
B
BD
D
D1,c
c
c=D
D
D1C
C
C,d
d
d=D
D
D1,
A
A
A=A
A
A1,B
B
B=A
A
A1B
B
B,C
C
C=B
B
BA
A
A1,D
D
D=D
D
DB
B
BA
A
A1B
B
B
Step 2
Calculation of the laminate stress resultants N
N
Nand M
M
M
by structural analysis of beam or plate structures
6.2 Failure Criteria 217
Step 3.
Calculate the laminate strains
ε
ε
ε
=
ε
ε
ε
+x3
κ
κ
κ
ε
ε
ε
···
κ
κ
κ
=
a
a
a.
.
.b
b
b
. . . .
c
c
c.
.
.d
d
d
N
N
N
···
M
M
M
and the strains for all laminae at lamina interfaces
ε
ε
ε
(k)=
ε
ε
ε
0+x(k)
3
κ
κ
κ
,k=0,1,2,...,n
Calculate the stresses for all interface surfaces of single layers
σ
σ
σ
(k)=Q
Q
Q(k)
ε
ε
ε
(k1),bottom surface of lamina k
σ
σ
σ
(k)+ =Q
Q
Q(k)
ε
ε
ε
k,top surface of lamina k,k=0,1,2,...,n
Transformation of the interface stresses
σ
σ
σ
(k),
σ
σ
σ
(k)+
to the on-axis system of layer kk = 0, 1, 2, . . . , n
Step 4
Failure analysis based on a selected failure criterion in stress space
Summarizing the strength ratios concept to the general quadratic interaction criteria
Eq. (6.2.28) we formulate with the maximum values of stresses
Fi j
σ
max
i
σ
max
j+Fi
σ
max
i=1
Substituting R
σ
applied
ifor
σ
max
iyield the quadratic equation for the strength ratio R
(Fi j
σ
i
σ
j)R2+ (Fi
σ
i)R1=0
or
aR2+bR 1=0,a=Fi j
σ
i
σ
j,b=Fi
σ
i(6.2.36)
The strength ratio Ris equal to the positive quadratic root
R=b
2a+sb
2a2
+1
a
As considered above this approach is easy to use because the resulting ratio provides
a linear scaling factor, i.e.
if R1 failure occurs,
if R>1, e.g. R=2, the safety factor is 2 and the load can be doubled or the laminate
thickness reduced by 0.5 before failure occurs.
The same strength ratio can be determined from the equivalent quadratic criterion
in the strain space. With
σ
σ
σ
=Q
Q
Q
ε
ε
ε
follows, e.g. with Eqs. (6.2.24) - (6.2.27) the Tsai-
Wu criterion in the strain space as
bL
ε
L+bT
ε
T+bLL
ε
2
L+bTT
ε
2
T+bSS
ε
2
S+2bLT
ε
L
ε
T=1 (6.2.37)
218 6 Failure Mechanisms and Criteria
with
bL=aLQLL +aTQLT,
bT=aTQTT +aLQLT,
bLL =aLLQ2
LL +aTTQ2
LT +2aLTQLL QLT ,
bTT =aTTQ2
TT +aLLQ2
LT +2aLTQTT QLT ,
bLT =aLLQLLQLT +aTTQTTQLT +aLT(Q2
LT +QLLQTT)
(6.2.38)
In the more general form analogous to the strength ratio equation is
(Gi j
ε
i
ε
j)R2+ (Gi
ε
i)R1=0,
cR2+dR 1=0,
R=d
c+sd
2c+1
c
(6.2.39)
To determine Rfrom this equivalent quadratic criterion the strain space may be
preferred, because laminae strains are either uniform or vary linearly across each
lamina thickness.
As considered above, the most widely used interlaminar failure criteria are the
maximum stress criterion, the maximum strain criterion and the quadratic failure
criteria as a generalization of the von Mises yield criterion, in particular the Tsai-
Hill and the Tsai-Wu criterion. The interlaminar failure modes can be fibre breaking,
fibre buckling, fibre pullout, fibre-matrix debonding or matrix cracking. The predic-
tion of the First-Ply Failure with one of the above mentioned criteria is included in
nearly all available analysis tools for layered fibre reinforced composites.
Interlaminar failure, i.e. failure of the interface between adjacent plies, is a de-
lamination mode. Delamination failure can have different causes. Weakly bonded
areas impact initial delamination in the inner region of a laminate, whereas delam-
ination along free edges is a result of high interlaminar stresses. Free edges delam-
ination is one of the most important failure modes in layered composite structures.
Along a free edge a tri-axial stress state is present and must be considered. Free
edge delamination is subject of actual intensive research.
The strength analysis of laminate presupposes experimental measured ultimate
stresses or strains for the laminae and realistic or approximate assumptions for stiff-
ness degradation of damaged layers. Strength under longitudinal tensile and com-
pression stresses is usually determined with unidirectional plane specimen, strength
under transverse tension and compression is measured with plane specimen or cir-
cumferentially reinforced tubes and shear strength is determined in torsion test of
such tubes. Note that compression testing is much more difficult than tension testing
since there is a tendency of premature failure due to crushing or buckling.
Summarizing the discussion above on failure analysis one can say that for deter-
mination of safety factors of fibre reinforced laminated structural elements there is
a strong need for fracture criteria and degradation models which are simple enough
for engineering applications but being also in sufficient agreement with the physical
6.3 Problems 219
reality. In spite of many efforts were made during recent years strength analysis of
laminates is still underdeveloped in comparison to the stress and strain analysis.
Essential for recent success in failure analysis was to distinguish between fibre
failure and inter-fibre failure by separate failure criteria introduced by Puck4. The
theory and application of Puck’s criterion are detailed described in special literature
(Knops, 2008) and are not considered here. In addition, Christensen5has presented
some arguments concerning the best choice of failure criteria - stress or strain based
(Christensen, 2013). On some actual problems and the state of the art is reported in
Talreja (2016).
6.3 Problems
Exercise 6.1.
A UD lamina is loaded by biaxial tension
σ
L=13
σ
T,
σ
LT =0. The material is a
glass-fibre epoxy composite with EL=46 GPa, ET=10 GPa, GLT =4,6 GPa,
ν
LT =0,31. The basic strength parameters are
σ
Lt =1400 MPa,
σ
Tt =35 MPa,
τ
S=70 MPa. Compare the maximum stress and the maximum strain criteria.
Solution 6.1. Maximum stress criterion (
σ
L<
σ
Lt,
σ
T<
σ
Tt)
13
σ
T=
σ
L<
σ
Lt
σ
T=
σ
Tt <
σ
Lt =
σ
T<107,69 MPa
σ
T<35 MPa
The ultimate stress is determined by the smallest of the two values, i.e. failure occurs
by transverse fracture. The stress state is then
σ
T=35 MPa,
σ
L=13 ·35 =455 MPa <1400 MPa
Maximum strain criterion (
ε
L<
ε
Lt,
ε
T<
ε
Tt)
To determine the ultimate strains we assume approximately a linear stress-strain
relation up to fracture. Then follows the ultimate strains
ε
Lt =
σ
Lt/EL,
ε
Tt =
σ
Tt/ET
The strains caused by the biaxial tension state are
ε
L=SLL
σ
L+SLT
σ
T=1
EL
σ
L
ν
LT
EL
σ
T=1
EL
(
σ
L
ν
LT
σ
T)<
ε
Lt,
ε
T=SLT
σ
L+STT
σ
T=
ν
TL
ET
σ
L+1
ET
σ
T=1
ET
(
σ
T
ν
TL
σ
L)<
ε
Tt
The maximum strain criterion can be written
4Alfred Puck (1927) - engineer and professor, development of a physical-based strength criterion
for UD reinforced laminates
5Richard M. Christensen (3 July 1932 Idaho Falls, Idaho) - specialist in mechanics of materials
220 6 Failure Mechanisms and Criteria
σ
L
ν
LT
σ
T<
σ
Lt,
σ
T
ν
TL
σ
L<
σ
Tt,
ν
TL =ET
EL
ν
LT
Since
σ
L=13
σ
Tfollows
σ
T<
σ
Lt/(13
ν
LT) = 110,32 MPa,
σ
T<
σ
Tt/(113
ν
LTET/EL) = 282,72 MPa
The ultimate stress is given by the lowest of both values, i.e. failure occurs by
longitudinal fracture and the stress state is then
σ
L=13 ·110,32 =1434,16 MPa,
σ
T=110,32 MPa
The values of both criteria differ significantly and the fracture mode is reversed from
transverse to longitudinal fracture. Because linear elastic response is assumed to
fail, the criterion can predict strength also in terms of stresses. In reality the relation
between ultimate stress and strain is more complex.
Exercise 6.2. Consider an off-axis unidirectional tension of a glass fibre/polyster
resin laminate (Fig. 6.5),
σ
1=3,5 MPa,
θ
=600. Estimate the state of stress with
the help of the maximum stress, the maximum strain and the Tsai-Hill failure cri-
terion. The lamina properties are E
1=30 GPa, E
2=4 GPa, G
12 =1,2 GPa,
ν
12 =0,28,
ν
21 =0,037,
σ
Lt =1200 MPa,
σ
Tt =45 MPa,
τ
S=35 MPa,
ε
Lt =0,033,
ε
Tt =0,002,
ε
S=0,0078.
Solution 6.2. The solution is split with respect to different criteria.
1. Maximum stress criterion
Using (6.2.6) the stresses in the principal material axes can be calculated
σ
1=
σ
1cos2
θ
=0,875 MPa <
σ
Lt,
σ
2=
σ
1sin2
θ
=2,625 MPa <
σ
Tt,
σ
6=
σ
1sin
θ
cos
θ
=1,515 MPa <
τ
S
The off-axis ultimate tensile strength
σ
1t is the smallest of the following stresses
σ
1=
σ
Lt/cos2
θ
=4800 MPa,
σ
1=
σ
Tt/sin2
θ
=60 MPa,
σ
1=
τ
S/sin
θ
cos
θ
=80,8 MPa
i.e.
σ
1t =60 MPa. All stresses
σ
iare allowable, the lamina does not fail.
2. Maximum strain criterion
From the Hooke’s law for orthotropic materials follows
ε
1=
σ
1/E
1
ν
21
σ
2/E
2=
σ
1/E
1
ν
12
σ
2/E
1,
ε
2=
ν
12
σ
1/E
1+
σ
2/E
2,
ε
6=
σ
6/E
6
The transformation for
σ
iyields
6.3 Problems 221
ε
1=1
E
1
[cos2
θ
ν
12 sin2
θ
]
σ
1=0,0000047 <
ε
Lt,
ε
2=1
E
2
[sin2
θ
ν
12
E
2
E
1
cos2
θ
]
σ
1=0,0006 <
ε
Tt,
ε
6=1
G
12
sin
θ
cos
θσ
1=0,0013 <
ε
S
All strains are allowed. The composite does not fail.
3. Tsai-Hill criterion
Using (6.2.18) the criterion can be written
cos2
θ
σ
Lt 2
+sin2
θ
σ
Tt 2
sin
θ
cos
θ
σ
Lt 2
+sin
θ
cos
θ
τ
S2
<1
σ
2
1
,
"0,25
1200 2
+0,75
45 2
0,433
1200 2
+0,433
35 2#MPa2<0,00043MPa2,
1
σ
2
1
=0,0816MPa2
The left-hand side is smaller than the right-hand side, therefore the composite
does not fail.
Exercise 6.3. The plane stress state of a UD-lamina is defined by
σ
1=2
σ
,
σ
2=3
σ
,
σ
6=4
σ
,
σ
>0
The material properties are
E
1=181GPa,E
2=10,3GPa,
ν
12 =0,28,G
12 =7,17GPa,
ν
21 =0,01593,
σ
Lt =1500MPa,
σ
Lc =1500MPa,
σ
Tt =40MPa,
σ
Tc =246MPa,
τ
S=68MPa
The fibre angle is
θ
=600. Calculate the maximum value for
σ
by using the different
failure criteria.
Solution 6.3. The solution is given for the four cases separately.
1. Maximum stress criterion
Transformation of the stresses from the off-axis to on-axis reference system
yields (Table 4.1)
σ
1
σ
2
σ
6
=
0,250 0,750 0,866
0,750 0,250 0,866
0,433 0,433 0,500
2
σ
3
σ
4
σ
=
1,714
2,714
4,165
σ
Using (6.2.2) we find the inequalities
222 6 Failure Mechanisms and Criteria
1500 MPa <1,714
σ
<1500 MPa,
246 MPa <2,714
σ
<40 MPa,
68 MPa <4,165
σ
<68 MPa,
=875,1MPa <
σ
<875,1 MPa,
14,73 MPa <
σ
<90,64 MPa,
16,33 MPa <
σ
<16,33 MPa
The three inequalities are satisfied if 0 <
σ
<16.33 MPa. The maximum stress
state which can be applied before failure is
σ
1=32,66 MPa,
σ
2=48,99 MPa,
σ
6=65,32 MPa
The mode of failure is shear.
2. Maximum strain criterion
Using the transformation rule (4.1.5) for strains
ε
ifollows with S
11 =
1/E
1=0,5525 1011 Pa1,S
22 =1/E
2=9,709 1011 Pa1,S
66 =1/G
12 =
13,95 1011 Pa1,S
12 =
ν
12/E
1=0,1547 1011 Pa1
ε
1
ε
2
ε
6
=
S
11 S
12 0
S
12 S
22 0
0 0 S
66
σ
1
σ
2
σ
6
=
0,1367
2,662
5,809
1010
σ
MPa
MPa
Assuming a linear relationship between the stresses and the strains until failure,
we can calculate the ultimate strains in a simple way
ε
Lt =
σ
Lt/E
1=8,287 103,
ε
Lc =
σ
Lc/E
1=8,287 103,
ε
Tt =
σ
Tt/E
2=3,883 103,
ε
Tc =
σ
Tc/E
2=23,88 103,
ε
S=
τ
S/G
12 =9,483 103
and the inequalities (6.2.8) yield
8,287 103<0,1367 1010
σ
<8,287 103,
23,88 103<2,662 1010
σ
<3,883 103,
9,483 103<5,809 1010
σ
<9,483 103
or
606,2 106<
σ
<606,2 106,
14,58 106<
σ
<89,71 106,
16,33 106<
σ
<16,33 106
The inequalities are satisfied if 0 <
σ
<16,33 MPa, i.e. there is the same maxi-
mum value like using the maximum stress criterion, because the mode of failure
is shear. For other failure modes there can be significant differences, see Exercise
6.1.
3. Tsai-Hill criterion
Using (6.2.17) we have
6.3 Problems 223
"1,714
1500 2
+2,714
40 2
1,714
1500 2,714
1500 +4,165
68 2#
σ
2
1012 <1
i.e.
σ
<10,94.
The Tsai-Hill criterion is an interactive criterion which cannot distinguish the
failure modes. In the form used above it also does not distinguish between com-
pression and tensile strength which can result in an underestimation of the allow-
able loading in compression with other failure criteria. Generally the transverse
tensile strength of a UD-lamina is much less than the transverse compressive
strength. Therefore the criteria can be modified. In dependence of the sign of
the
σ
ithe corresponding tensile or compressive strength is substituted. For our
example follows
"1,714
1500 2
+2,714
246 2
1,714
1500 2,714
1500 +4,165
68 2#
σ
2
1012 <1
i.e.
σ
<16,06 MPa.
4. Tsai-Wu criterion
Now (6.2.24) must be applied. The coefficients can be calculated
aL=1
σ
Lt 1
σ
Lc =0,
aTT =1
σ
Tt
σ
Tc
=1,0162 1016Pa2,
aT=1
σ
Tt 1
σ
Tc =2,093 108Pa1,
aSS =1
τ
2
Tt
=2,1626 1016Pa2,
aLL =1
σ
Lt
σ
Lc =4,44441 1019Pa2,
aLT 1
2aLLaTT =3,360 1018Pa2
Substituting the values of the coefficients in the criterion it yields the following
equation
0·(1,714)
σ
+2,093(108)(2,714)
σ
+4,4444(1019)(1,714
σ
)2
+1,0162(1016)(2,714
σ
)2+2,1626(1016)(4,165
σ
)2
+2(3,360)(1018)(1,714)(2,714)
σ
2<1
The solution of the quadratic equation for
σ
yields
σ
<22,39 MPa.
224 6 Failure Mechanisms and Criteria
Summarizing the results of the four failure criteria we have
Max. stress criterion:
σ
=16,33 ¯
σ
(¯
σ
RS
σ
)
Max. strain criterion:
σ
=16,33 ¯
σ
(¯
σ
RS
ε
)
Tsai-Hill criterion:
σ
=10,94 ¯
σ
(¯
σ
RTH)
Mod. Tsai-Hill criterion:
σ
=16,06 ¯
σ
(¯
σ
RTHm)
Tsai-Wu criterion:
σ
=22,39 ¯
σ
(¯
σ
RTW)
The values ¯
σ
σ
are identical with the strength ratios (6.2.32) - (6.2.35).
Remark 6.1. A summary of the examples demonstrates that different failure criteria
can lead to different results. Unfortunately, there is no one universal criterion which
works well for all situations of loading and all materials. For each special class of
problems a careful proof of test data and predicted failure limits must be conducted
before generalizations can be made. In general, it may be recommended that more
than one criterion is used and the results are compared.
References
Christensen RM (2013) The Theory of Materials Failure. University Press, Oxford
Gol’denblat II, Kopnov VA (1965) Strength of glass-reinforced plastics in the com-
plex stress state. Polymer Mechanics 1(2):54–59
Hill R (1948) A theory of the yielding and plastic flow of anisotropic metals. Pro-
ceedings of the Royal Society of London A: Mathematical, Physical and Engi-
neering Sciences 193(1033):281–297
Hoffman O (1967) The brittle strength of orthotropic materials. Journal of Compos-
ite Materials 1(2):200–206
Knops M (2008) Analysis of Failure in Fiber Polymer Laminates: The Theory of
Alfred Puck. Springer, Heidelberg
von Mises R (1913) Mechanik des festen K ¨orpers im plastischen deformablen Zu-
stand. Nachrichten der K ¨oniglichen Gesellschaft der Wissenschaften G ¨ottingen,
Mathematisch-physikalische Klasse pp 589–592
von Mises R (1928) Mechanik der plastischen foranderung von kristallen. ZAMM
- Journal of Applied Mathematics and Mechanics / Zeitschrift ur Angewandte
Mathematik und Mechanik 8(3):161–185
Talreja R (2016) On failure theories for composite materials. In: Naumenko K,
Aßmus M (eds) Advanced Methods of Continuum Mechanics for Materials and
Structures, Springer Singapore, Singapore, pp 379–388
Tsai SW, Wu EM (1971) A general theory of strength for anisotropic materials.
Journal of Composite Materials 5(1):58–80
Part III
Analysis of Structural Elements
The third part (Chaps. 7–9) is devoted to the analysis of structural elements (beams,
plates and shells) composed of laminates and sandwiches. The modelling of lam-
inated and sandwich plates and shells is limited to rectangular plates and circular
cylindrical shells. The individual fiber reinforced laminae of laminated structured
elements are considered to be homogeneous and orthotropic, but the laminate is
heterogeneous through the thickness and generally anisotropic. An equivalent sin-
gle layer theory using the classical lamination theory, and the first order shear defor-
mation theory are considered. Multilayered theories or laminate theories of higher
order are not discussed in detail.
Chapter 7
Modelling and Analysis of Beams
In Chap. 1 the classification of composite materials, the significance, advantages
and limitations of composite materials and structures and the material characteris-
tics of the constituents of composite materials were considered. Chapter 2 gave a
short introduction to the governing equations of the linear theory of anisotropic ma-
terial behavior. Chapter 3 defined effective material moduli of composites including
elementary mixture rules and improved formulas. Chapter 4 developed in detail the
modelling of the mechanical behavior of laminates and sandwiches in the frame of
classical theories including thermal and hygroscopic effects. The constitutive equa-
tions, describing the relationships between stress resultants and in-plane strains and
mid-surface curvatures were developed for unidirectional laminae, laminates and
sandwiches with the assumptions of the classical laminate theory. Further the calcu-
lation of in-p lane and throug h-the-thickness stresses was considered. Chapter 5 gave
an introduction to classical and refined laminate theories. In Chap. 6 selected failure
mechanisms and criteria were briefly discussed. These parts of the book give the
basic knowledge, how the design engineer can tailor composite materials to obtain
the desired properties by the appropriate choice of the fibre and matrix constituents,
a laminate or a sandwich material, the stacking sequence of layers, etc. This ba-
sic knowledge can be utilized to develop the modelling and analysis of structural
elements and structures composed of composite materials.
7.1 Introduction
The analysis of structural elements can be performed by analytical and semi-
analytical approaches or by numerical methods. The advantage of analytical so-
lutions is their generality allowing the designer to take into account various design
parameters. Analytical solutions may be either closed form solutions or infinite se-
ries and may be exact solutions o f the governing equations or variational appro aches.
However, analytical solutions are restricted to the analysis of simple structural el-
ements such as beams, plates and shells with simple geometry. Otherwise numer-
227
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_7
228 7 Modelling and Analysis of Beams
ical methods have to be applied more general for structural analysis. Chapter 7 to
10 describe analytical solutions for one- and two-dimensional structural elements.
Chapter 11 gives an insight into numerical solutions based on the finite element
method.
In the following sections of Chap. 7 we consider rods, columns and beams. These
are one-dimensional structural elements with a thickness hand a width bwhich are
small relative to the element length l, i.e. h,bl. When this element is loaded by
an axial force only one speaks of a rod if the loading is tensile, and of columns if
the load is compressive. One calls this element a beam when it is acted upon by
lateral loads. In general a combination of lateral and axial loadings is possible and
so we shall speak of beams under lateral and axial loadings. The other type of one-
dimensional structural elements, so called plate strips under cylindrical bending,
are discussed in Chap. 8. The modelling and analysis of generalized beams based a
thin-walled folded structure are considered in Chap. 10 (generalized Vlasov beam
theory).
The elementary or classical beam theory assumes that the transverse shear strains
are negligible and plane cross-sections before bending remain plane and normal to
the axis of the beam after bending (Bernoulli-Euler beam theory, Sect. 7.2). The
assumption of neglecting shear strains is valid if the thickness his small relative to
the length l(h/l<1/20). In the Bernoulli-Euler beam theory the transverse deflec-
tion u3is assumed to be independent of coordinates x2,x3of the cross-section (Fig.
7.1), i.e. u3w=w(x1). In Sect. 7.2 the governing equations of the classical beam
theory for composite beams are considered. The differential equations and varia-
tional formulations will be developed in detail for bending only, the equations for
vibration and buckling are briefly summarized.
In the case of sandwich beams or moderately thick laminate beams, the results
derived from the Bernoulli-Euler theory can show significant differences with the
actual mechanical behavior, i.e. the deflection, stress distribution, etc. An improve-
ment is possible by introducing the effect of transverse shear deformation, i.e. we
apply Timoshenko beam theory (Sect. 7.3). The assumptions of the classical theory
axis
b
h
l
x1
x3
x2
Fig. 7.1 Rod/column/beam
7.2 Classical Beam Theory 229
have then to be relaxed in the following way: the transverse normals do not remain
perpendicular to the deformed axis of the beam after straining. Section 7.4 discuss
some special aspects of sandwich beams.
Laminate or sandwich beams with simple or double symmetric cross-sections are
most important in engineering applications. The derivations in Sects. 7.2 - 7.5 are
therefore limited to straight beams with simple or double symmetric constant cross-
sections which are predominantly rectangular. The bending moments act in a plane
of symmetry. Also cross-sections consisting of partition walls in and orthogonal to
the plane of bending, e.g. I- or box beams, are considered.
7.2 Classical Beam Theory
Frequently, as engineers try to optimize the use of materials, they design compos-
ite beams made from two or more materials. The design rationale is quite straight
forward. For bending loading, stiff, strong, heavy or expensive material must be far
away from the neutral axis at places where its effect will be greatest. The weaker,
lighter or less expensive material will be placed in the central part of the beam. At
one extreme is a steel-reinforced concrete beam, where weight is not a major con-
cern, but strength and cost are. At the other extreme is a sandwich beam used e.g. in
an aircraft with fibre-reinforced laminate cover sheets and a foam core. In that case,
stiffness and weight are essential but cost not.
First we consider elementary beam equations: The cross-section area Acan have
various geometries but must be symmetric to the x3-axis. The fibre reinforcement
of the beam is parallel to the x1-axis and the volume fraction is a function of the
cross-sectional coordinates x2,x3, i.e. vf=vf(x2,x3). The symmetry condition yields
vf(x2,x3) = vf(x2,x3)and E1(x2,x3) = E1(x2,x3).
With the known equations for the strain
ε
1and the stress
σ
1at x1=const
ε
1(x3) =
ε
1+x3
κ
1,
σ
1(x2,x3) =
ε
1E1(x2,x3) + x3
κ
1E1(x2,x3)(7.2.1)
follow the stress resultants N(x1),M(x1)of a beam
N=
ε
1Z
(A)
E1(x2,x3)dA+
κ
1Z
(A)
x3E1(x2,x3)dA,
M=
ε
1Z
(A)
x3E1(x2,x3)dA+
κ
1Z
(A)
x2
3E1(x2,x3)dA
(7.2.2)
The effective longitudinal modulus of elasticity is (3.1.8)
E1=Efvf+Emvm=Em+
φ
(x2,x3)(EfEm)(7.2.3)
and with Ef=const, Em=const,
φ
(x2,x3) =
φ
(x2,x3)it follows that
230 7 Modelling and Analysis of Beams
N=a
ε
1+b
κ
1,a=EmA+ (EfEm)Z
(A)
φ
(x2,x3)dA,
M=b
ε
1+d
κ
1,b= (EfEm)Z
(A)
φ
(x2,x3)x3dA,
I=Z
(A)
x2
3dA,d=EmI+ (EfEm)Z
(A)
φ
(x2,x3)x2
3dA
(7.2.4)
The inverse of the stress resultants, (7.2.4), are
ε
1=dN bM
ad b2,
κ
1=aM bN
ad b2(7.2.5)
and the stress equation (7.2.1) has the form
σ
1(x2,x3) = dN bM + (aM bN)x3
ad b2E1(x2,x3)(7.2.6)
Taking into consideration the different moduli Efand Em, the fibre and matrix
stresses are
σ
f(x3) = dN bM + (aM bN)x3
ad b2Ef,
σ
m(x3) = dN bM + (aM bN)x3
ad b2Em
(7.2.7)
In the case of a double symmetric geometry and fibre volume fraction function
φ
,
b=0 and the equations can be simplified
ε
1=N/a,
κ
1=M/d,
σ
f(x3) = (N/a+x3M/d)Ef,
σ
m(x3) = (N/a+x3M/d)Em(7.2.8)
For a uniform fibre distribution,
φ
(x2,x3) = const, (7.2.3) (7.2.4) give
a=EmA+ (EfEm)
φ
A=E1A,
b=0,
d=EmI+ (EfEm)
φ
I=E1I
(7.2.9)
and the stress relations for fibre and matrix, (7.2.8), are transformed to
σ
f(x3) = N
A+x3
M
IEf
E1,
σ
m(x3) = N
A+x3
M
IEm
E1(7.2.10)
If Ef=Emand E1=E, (7.2.10) becomes the classical stress formula for isotropic
beam with axial and lateral loadings
σ
(x3) = N
A+M
Ix3(7.2.11)
7.2 Classical Beam Theory 231
Now we consider laminate beams loaded by axial and lateral loading. For simplicity,
thermal and hygrothermal effects are ignored. The derivation of the beam equations
presume the classical laminate theory (Sects. 4.1 and 4.2). There are two different
cases of simple laminated beams with rectangular cross-section:
1. The beam is loaded orthogonally to the plane of lamination.
2. The beam is loaded in the plane of lamination.
In the first case, we start from the constitutive equations (4.2.18).
The beam theory makes the assumption that in the case of bending and stretch-
ing in the (x1x3)-plane of symmetry, i.e. no unsymmetrical or skew bending,
N2=N6=0, M2=M6=0 and that all Poisson’s effects are neglected.
With these assumptions (4.2.18) is reduced to
N1
M1=A11 B11
B11 D11
ε
1
κ
1(7.2.12)
and from (4.2.14)
Q1=A55
ε
5(7.2.13)
If the beam has a midplane symmetry, there is no bending-stretching coupling, so
that B11 =0 and (7.2.12) becomes
N1=A11
ε
1,M1=D11
κ
1(7.2.14)
Note that in the classical theory, the transverse shear strain will be ignored, i.e
ε
5=0, and there is no constitutive equation for resultant shear forces.
The starting point for derivation of structural equations for beams is the equilib-
rium equations for stress resultants N,Mand Qat the undeformed beam element,
Fig. 7.2. The in-plane and transverse stress resultants N1,Q1and the resultant mo-
x3
x1
MM+dM
QQ+dQ
NN+dN
q(x1)
n(x1)
m(x1)
dx1
Fig. 7.2 Stress resultants N,Qand Mof the infinite beam element, q(x1),n(x1)are line forces,
m(x1)is a line moment
232 7 Modelling and Analysis of Beams
Table 7.1 Differential relations for laminate beams based on the classical beam theory
(
ε
1=u(x1),
κ
1=w′′(x1))
Relations between stress resultants and loading
N(x1) = n(x1),Q(x1) = q(x1),
M(x1) = Q(x1)m(x1),M′′(x1) = q(x1)m(x1)
Relations between stress resultants and strains
N=bA11u(x1)bB11 w′′(x1)
M=bB11u(x1)bD11 w′′(x1)
N(x1)
M(x1)
=
bA11 bB11
bB11 bD11
u(x1)
w′′(x1)
Differential equations for the displacements
General case
(bA11u)′′ (bB11 w′′)′′ =n
(bB11u)′′ (bD11 w′′)′′ =qm(...)=d
dx1
Constant stiffness
bA11 bB11
bB11 bD11 u′′′
w′′′′ =n
qm
Midplane symmetric laminates (B11 =0)
(bA11u)=n
(bD11w′′)′′ =q+m
bA11u′′ =n
bD11w′′′′ =q+m
Special cases
m(x1) = 0 :
u′′(x1) = n(x1)
bA11
,w′′′′(x1) = q(x1)
bD11
N(x1) = n(x1),Q(x1)= q(x1),
M(x1) = Q(x1)m(x1)
m(x1) = 0,n(x1) = 0 :
u(x1) = N
bA11
=const,w′′′′(x1) = q(x1)
bD11
Q(x1) = q(x1),M(x1) = Q(x1)
ment M1in (7.2 .12)–(7.2.13) are load s per unit length and must be multiplied by th e
beam width b, i.e. the beam resultants are N=bN1,Q=bQ1,M=bM1.
The differential relations for laminate beams loaded orthogonally to the plane of
lamination are summarized in Table 7.1. Note that when Nis a compressive load,
we have to consider additional stability conditions.
The stresses
σ
(k)
1(x1,x3)in the kth layer are given by
7.2 Classical Beam Theory 233
σ
(k)
1=Q(k)
11
ε
(k)
1=Q(k)
11 (
ε
1+x3
κ
1) = Q(k)
11 du(x1)
dx1x3
d2w(x1)
dx2
1(7.2.15)
or with
ε
1=D11NB11M
A11D11 B2
11
,
κ
1=A11MB11N
A11D11 B2
11
,
one get
σ
(k)
1=Q(k)
11
1
bD11NB11M
A11D11 B2
11
+x3
A11MB11N
A11D11 B2
11 (7.2.16)
In the most usual case of midplane symmetric beams the stress equations (7.2.15),
(7.2.16) can be simplified to
σ
(k)
1M(x1) = Q(k)
11
du(x1)
dx1
=Q(k)
11
N(x1)
bA11
,
σ
(k)
1B(x1) = Q(k)
11 x3
d2w(x1)
dx2
1=Q(k)
11 x3
M(x1)
bD11
(7.2.17)
σ
(k)
1Mare the layerwise constant stretching or membrane stresses produced by N(x1)
and
σ
(k)
1Bthe layerwise linear distributed flexural or bending stresses produced by
M(x1). The strain
ε
1=
ε
1+x3
κ
1is continuous and linear through the total beam
thickness h. The stresses
σ
(k)
1are continuous and linear through each single layer
and have stress jumps at the layer interfaces (Fig. 7.3) With the help of effective
moduli EN
eff and EM
eff for stretching and flexural loading we can compare the stress
equations of a laminate beam with the stress equation of a single layer beam.
With N6=0,M=0 Eq. (7.2.12) becomes
N1
0=bA11 B11
B11 D11
ε
1
κ
1(7.2.18)
and with
0=bB11
ε
1+bD11
κ
1,
κ
1=B11
D11
ε
1(7.2.19)
1
2
3
4
5
6
σ
1M
ε
1
σ
1B
x3
κ
1
x1
S
h
Fig. 7.3 Qualitative distribution of the stresses and strains through the beam thickness hassuming
Q(1)
11 =Q(6)
11 >Q(3)
11 =Q(4)
11 >Q(2)
11 =Q(5)
11
234 7 Modelling and Analysis of Beams
follows the equations for N=bN1and
ε
1
N=bA11
ε
1+bB11
κ
1=
ε
1bA11D11 B2
11
D11
,
ε
1=D11
(A11D11 B2
11)bN=D11h
(A11D11 B2
11)
N
bh =N
EN
effA(7.2.20)
with
EN
eff =A11D11 B2
11
D11h,A=bh
The strain
ε
1on the beam axis of a single layer isotropic, homogeneous beam is
ε
1=N/EA. Replacing Eby EN
eff gives the strain equation for the laminate beam.
The stresses
σ
(k)
1in the klayers are then
σ
(k)
1=Q(k)
11 (
ε
1+x3
κ
1) = Q(k)
11
ε
11x3
B11
D11
=Q(k)
11 1x3
B11
D11 du
dx1
,
σ
(k)
1=Q(k)
11
EN
eff
N
bh 1x3
B11
D11 =E(k)
1
EN
eff
N
bh 1x3
B11
D11
(7.2.21)
or for midplane symmetric beams with B11 =0
σ
(k)
1=E(k)
1
EN
eff
N
bh =E(k)
1
EN
eff
N
A,EN
eff =A11
h(7.2.22)
In an analogous manner it follows from (7.2.12) with N=0,M6=0 that
M=bB11
ε
1+bD11
κ
1=
κ
1bA11D11 B2
11
A11
,
κ
1=A11
(A11D11 B2
11)bM=A11h3
12(A11D11 B2
11)
M
bh3
12
=M
EM
effI,(7.2.23)
with
EM
eff =12(A11D11 B2
11)
A11h3,I=bh3
12 (7.2.24)
For an isotropic homogeneous single layer beam of width band thickness hone get
κ
1=M/EI =12M/bh3,I=bh3/12. Replacing now Eby EM
eff, the stress equations
are
7.2 Classical Beam Theory 235
σ
(k)
1=Q(k)
11 (
ε
1+x3
κ
1) = Q(k)
11
κ
1B11
A11
+x3
=Q(k)
11 B11
A11 x3d2w
dx2
1
,
=Q(k)
11
Eeff
M
Ix3B11
A11 =E(k)
1
Eeff
M
Ix3B11
A11
(7.2.25)
or with B11 =0 for the symmetric case
σ
(k)
1=E(k)
1
EM
eff
M
Ix3,EM
eff =12D11
h3(7.2.26)
If both in-plane and lateral loads occur simultaneously, the stress in each lamina of
the beam is as for symmetric case
σ
(k)
1(x1,x3) = Q(k)
11 du
dx1x3
d2w
dx2
1=E(k)
1N
EN
effA+M
EM
effIx3,(7.2.27)
A=bh,I=bh3/12,EN
eff =A11/h,EM
eff =12D11/h3
Conclusion 7.1. Summarizing the equations for symmetric laminated beams, one
can say that the equations for u(x1)and w(x1)are identical in form to those of el-
ementary theory for homogeneous, isotropic beams. Hence all solutions available,
e.g. for deflections of isotropic beams under various boundary conditions, can be
used by replacing the modulus Ewith EN
eff or EM
eff, respectively. The calculation of
the stresses illustrates that constant in-plane layer stresses produced by Nare propor-
tional to the layer modulus E(k)
1(7.2.22). N/AE N
eff is for a cross-section x1=const
a constant value. Analogous are the flexural layer stresses proportional to E(k)
1x3
(7.2.26). In general, the maximum stress does not occur at the top or the bottom of
a laminated beam, but the maximum stress location through the thickness depends
on the lamination scheme.
From the bending moment-curvature relation (N=0)
M=bD11
κ
1(7.2.28)
it follows that
κ
1max =Mmax
bD11
=d2w
dx2max
,
and the maximum stress can be calculated for each lamina
σ
(k)
1max =Q(k)
11
κ
1maxx3=Q(k)
11 d2w
dx2max
x3=E(k)
1
EM
eff
Mmax
Ix3(7.2.29)
σ
(k)
1max must be compared with the allowable strength value.
236 7 Modelling and Analysis of Beams
The calculation of the transverse shear stress
σ
(k)
5(x1,x3)is analogous as in the
elementary beam theory. We restrict our calculation to midplane symmetric beams
and therefore all derivations can be given for the upper part of the beam element
(x30). The equilibrium equations in the x1-direction lead with
σ
(j)
1+d
σ
(j)
1
σ
(j)
1+d
σ
(j)
1
dx1
dx1
(Fig. 7.4) and no edge shear stresses on the upper and lower faces
σ
(k)
5dx1
N
j=k+1
x(j)
3
Z
x(j1)
3
"
σ
(j)
1+d
σ
(j)
1
dx1
dx1!
σ
(j)
1#dx3=0 (7.2.30)
or
σ
(k)
5=
N
j=k+1
x(j)
3
Z
x(j1)
3
d
σ
(j)
1
dx1
dx3=
N
j=k+1
x(j)
3
Z
x(j1)
3
E(j)
1
EM
effI
dM
dx1
x3dx3
With Q=dM/dx1it follows that
σ
(k)
5(x1,x3) = Q(x1)
EM
effI
N
j=k+1
x(j)
3
Z
x(j1)
3
E(j)
1x3dx3=Q(x1)
EM
effI
N
j=k+1
E(j)
1
1
2x(j)
3
2x(j1)
3
2
=Q(x1)
EM
effI
N
j=k+1
E(j)
1h(j)x(j)
3
(7.2.31)
x1
x3
σ
(k)
5
k+1
j
N
.
.
.
.
.
.
σ
(k+1)
1
σ
(j)
1
σ
(N)
1
σ
(k+1)
1+d
σ
(k+1)
1
σ
(j)
1+d
σ
(j)
1
σ
(N)
1+d
σ
(N)
1
dx1
x(k)
3
x(j1)
3
x(j)
3
x(j)
3
x(N)
3
Fig. 7.4 Beam element b(x(N)
3x(k)
3)dx1with flexural normal stresses
σ
(j)
1and the interlaminar
stress
σ
(k)
5
7.2 Classical Beam Theory 237
For a single layer homogeneous, isotropic beam (7.2.31) yields the known parabolic
shear stress distribution through h
σ
5(x1,x3) = Q
I
h/2
Z
x3
x3dx3=12Q
bh3
1
2h2
4x2
3=3Q
2bh 14x3
h2(7.2.32)
With an increasing number of equal thickness layers, the transverse shear stress
distribution (7.2.31) approaches the parabolic function of the single layer beam.
All stress equations presume that the Poisson’s effects can be completely ne-
glected, i.e. Q(k)
i j =Di j =0,i6=j,i,j=1,2,6. They are summarized for symmet-
ric laminated beams (N6=0,M6=0) in Table 7.2. For symmetric laminated beams
loaded orthogonally to the plane of lamination, the classical laminate theory yields
identical differential equations for u(x1)and w(x1)with to Bernoulli’s beam theory
of single layer homogeneous isotropic beams, if one substitutes bA11 by E A =Ebh
and bD11 by EI =Ebh3/12. An equal state is valid for beam vibration and beam
buckling.
The following equations are given without a special derivation (b,A,D11,
ρ
are
constant values):
Differential equation of flexure (N=0)
d2w(x1)
dx2
1
=M(x1)
bD11
,bD11
d4w(x1)
dx4
1
=q(x1)(7.2.33)
Table 7.2 Stress formulas for symmetric laminated beams, classical theory
σ
(k)
1(x1,x3) =
σ
(k)
1M(x1) +
σ
(k)
1B(x1,x3)
=Q(k)
11
du(x1)
dx1Q(k)
11
d2w(x1)
dx2
1
x3
=Q(k)
11
N(x1)
bA11
+Q(k)
11
M(x1)
bD11
x3
=E(k)
1
EN
eff
N(x1)
A11
+E(k)
1
EM
eff
Mx1)
Ix3,
σ
(k)
5(x1,x3) = Q(x1)
EM
effI
N
j=k+1
E(j)
1h(j)¯x(j)
3,
A=bh,I=bh3/12,
EN
eff =A11/h,EM
eff =12D11/h3,¯x(j)=1
2(x(j)
3+x(j1)
3)
238 7 Modelling and Analysis of Beams
Forced or free vibrations
bD11
d4w(x1,t)
dx4
1
+
ρ
Ad2w(x1,t)
dt2=q(x1,t),
ρ
=1
h
N
k=1
ρ
(k)h(k)(7.2.34)
Rotational inertia terms are neglected. For free vibration with q=0 the solution
is assumed periodic: w(x1,t) = W(x1)exp(i
ω
t).
Buckling equation
d2M(x1)
dx2
1
+N1(x1)d2w(x1)
dx2
1
=0,bD11
d4w(x1)
dx4
1N1(x1)d2w(x1)
dx2
1
=0
(7.2.35)
or with N(x1) = F
d2M(x1)
dx2
1Fd2w(x1)
dx2
1
=0,bD11
d4w(x1)
dx4
1
+Fd2w(x1)
dx2
1
=0
All solutions of the elementary beam theory for single layer isotropic beams can
transferred to laminate beams. Note that laminate composites are stronger shear
deformable than metallic materials and the classical beam theory is only acceptable
when the ratio l/h>20.
The equations for flexure, vibration and buckling can also be given in a vari-
ational formulation (Sect. 2.2.2). With the elastic potential for a flexural beam
(N=0,M6=0)
Π
(w) = 1
2
l
Z
0
bD11 d2w(x1)
dx2
12
dx1
l
Z
0
qdx1(7.2.36)
and the kinetic energy
T(w) = 1
2
l
Z
0
ρ
2w
t2
dx1(7.2.37)
the Lagrange function is given by L(w) = T(w)
Π
(w)(Sect. 2.2.2).
The variational formulation for a symmetric laminated beam without bending-
stretching coupling subjected to a lateral load qin x3-direction (N=0,M6=0,
ε
50,
ν
i j 0) based on the theorem of minimum of total potential energy is
given in the form
Π
[w(x1)] = 1
2
l
Z
0
bD11 d2w(x1)
dx2
12
dx1
l
Z
0
qdx1,
δΠ
[w(x1)] = 0
(7.2.38)
The variational formulation for the buckling of a symmetric laminate beam with
N(x1) = Fis
7.2 Classical Beam Theory 239
Π
[w(x1)] = 1
2
l
Z
0
bD11 d2w(x1)
dx2
12
dx11
2
l
Z
0
Fdw(x1)
dx12
dx1,
δΠ
[w(x1)] = 0
(7.2.39)
The variational formulation for free flexural beam vibration (additional to ap-
proaches noted above the rotatory inertia effects are neglected) can be given by
the Hamilton’s principle
H[w(x1,t)] =
t2
Z
t1
L[W(x1,t)]dt,
δ
H[w(x1,t)] = 0 (7.2.40)
The variational formulations can be used for approximate analytical solution with
the Rayleigh-Ritz procedure or numerical solutions.
In the second case of laminate beams, the loading is in the plane of lamination.
We restrict our considerations to symmetric layered beams and neglect all Poisson’s
ratio effects. The beam is illustrated in Fig. 7.5. For a symmetric layer stacking there
is no bending-stretching coupling and we have the constitutive equations
N1=A11
ε
1,M1=b3
12hA11
κ
1
or for the beam resultants N=hN1,M=hM1
Fig. 7.5 Laminate beam
loaded in the plane of lamina-
tion
q(x1)
x2
x3
x1
b
h
240 7 Modelling and Analysis of Beams
N=hA11
ε
1,M=b3
12 A11
κ
1,A11 =
N
k=1
Q(k)
11 h(k)(7.2.41)
The differential equations of flexure (N=0) are
d2w(x1)
dx2
1
=M(x1)
A11
b3
12 ,12A11
b3
d4w(x1)
dx4
1
=q(x1)(7.2.42)
and the additional equation for the case N6=0 is
hA11
du(x1)
dx1
=N(x1)(7.2.43)
The calculation of stresses is analogous to case 1 of layered beams.
When beam profiles consist of partition-walls in the plane of loading and orthog-
onal to the plane of loading, e.g. I-profiles or box-beams, the bending differential
equations can be written in the form, given above. The bending stiffness is obtained
by combining the results of orthogonal to plane loading and in-plane loading.
As an example for a box-beam, we consider the beam as shown in Fig. 7.6, which
may be subjected to axial loads in x1-direction, a bending moment with respect to
the x2-axis and a twisting moment with respect to the x1-axis. For an isotropic beam
the stiffness needed are the extensional stiffness, EA, the flexural stiffness, E I, and
the torsional stiffness, GIt.
The axial force resultant (per unit width) in x1-direction is N1=A11
ε
1and the
axial load carried by the whole section is then
N(x1) = 2NI
1b+2NII
1h=2[(A11)Ib+ (A11)I I h]
ε
1(7.2.44)
The extensional stiffness for the box cross-section is given by
Fig. 7.6 Laminated box-beam
with identical top and bottom
panels I and vertical walls II
x2
x3
x1
b
l
t
I
II
h
7.2 Classical Beam Theory 241
(EA)eff =2(A11)Ib+2(A11)I I h(7.2.45)
The box beam is bent in the (x1x3) plane, and the moment curvature relation is
M="2(D11)Ib+2(A11)Ibh
22
+2(A11)II
h3
12 #
κ
1
"2(A11)Ibh
22
+1
6(A11)II h3#
κ
1
(7.2.46)
Since the top and bottom panels are thin relative to the height of the box profile, i.e.
th,(D11)Ican be neglected and the bending stiffness of the box cross-section is
(EI )eff 2(A11)Ibh
22
+1
6(A11)II h3(7.2.47)
If the box-beam is acted by a torsional moment MTthis is equivalent to the moment
of the shear flows with respect to the x1-axis and we have
MT=2NI
6b(h/2) + 2NII
6h(b/2),NI
6=AI
66
ε
I
6,NII
6=AII
66
ε
II
6(7.2.48)
In the elementary theory of strength of materials the equation for the angle of twist
of a box-beam is given by
θ
=1
2AIq(s)
Gt ds(7.2.49)
q(s)is the shear flow. In our case, Fig. 7.6, the displacements of the contours of the
walls of the box beam are denoted by
δ
Iand
δ
II and the angle of twist becomes
θ
=
δ
I
(h/2)=
δ
II
(b/2)with
δ
I
l=
ε
I
6,
δ
II
l=
ε
II
6(7.2.50)
From (7.2.48) - (7.2.50) we have
MT=bh
l[(AI
66)h+ (AII
66)b]
θ
(7.2.51)
and the torsional stiffness of the cross-section is
(GIt)eff =bh
l[(AI
66)h+ (AII
66)b](7.2.52)
For the I-profile in Fig. 7.7 the calculation for bending is analogous. The bending
stiffness is
(EI )eff ="2(D11)b+2(A11)bh
22
+ (A11)h3
12 #
A11
6h2b+h3
12
(7.2.53)
242 7 Modelling and Analysis of Beams
Fig. 7.7 I-profile with uni-
form thickness t
b
t
h
if D11 0. Note that for a one-dimensional thin structural element which is sym-
metric with respect to all mid-planes and Poisson’s effect is neglected, we have the
simple relationships
Q11 =E1,A11 =E1t,
κ
1=M
(EI )eff
Summarizing the classical beam equations it must be noted that the effect of Pois-
son’s ratio is negligible only if the length-to-with ratio l/bis large (lb), otherwise
the structure behavior is more like a plate strip than a beam (Sect. 8.2). This is of
particular importance for angle-ply laminates, i.e. orthotropic axes of material sym-
metry in each ply are not parallel to the beam edges and anisotropic shear coupling
is displayed.
7.3 Shear Deformation Theory
The structural behavior of many usual beams may be satisfactorily approximated by
the classical Euler-Bernoulli theory. But short and moderately thick beams or lami-
nated composite beams which l/hratios are not rather large cannot be well treated
in the frame of the classical theory. To overcome this shortcoming Timoshenko ex-
tended the classical theory by including the effect of transverse shear deformation.
However, since Timoshenko’s beam theory assumed constant shear strains through
the thickness ha shear correction factor is required to correct the shear strain energy.
In this section we study the influence of transverse shear deformation upon the
bending of laminated beams. The similarity of elastic behavior of laminate and sand-
wich beams with transverse shear effects included allows us generally to transpose
the results from laminate to sandwich beams. When applied to beams, the first order
shear deformation theory is known as Timoshenko’s beam theory. Figure 7.8 illus-
trates the cross-section kinematics for the Bernoulli’s and the Timoshenko’sbending
beam.
7.3 Shear Deformation Theory 243
x1
x3
O
B
A
dx1
u(x1)
u1(x1)
O
B
AB′′
A′′
w(x1)
ψ
(x1)
w(x1)
Bernoulli kinematics
flexure curve
Timoshenko kinematics
w(x1)
x2
Fig. 7.8 Kinematics of a bent Timoshenko- and Bernoulli-beam in the (x1x3) plane
When all Poisson’s effects are neglected the constitutive equations are identical
with (7.2.12) - (7.2.13), but from Sect. 5.1 the strains of the Timoshenko’s beam are
ε
1=
u1
x1
=du
dx1
+x3
d
ψ
1
dx1
,
ε
20,
ε
30,
ε
5=
u1
x3
+
w
x1
=
ψ
1+dw
dx1
,
ε
40,
ε
60
(7.3.1)
i.e. we only have one longitudinal and one shear strain
ε
1(x1,x3) =
ε
1(x1) + x3
κ
1(x1),
ε
5(x1,x3) =
ψ
1(x1) + w(x1),
ε
1(x1) = du
dx1
,
κ
1(x1) = d
ψ
1(x1)
dx1
(7.3.2)
When the transverse shear strain are neglected it follows with
ε
50 that the rela-
tionship is
ψ
1(x1) = w(x1)and that is the Bernoulli’s kinematics.
244 7 Modelling and Analysis of Beams
In the general case of an unsymmetric laminated Timoshenko’s beam loaded
orthogonally to the laminated plane and N6=0,M6=0, the constitutive equations
(stress resultants - strain relations) are given by
N=A11
ε
1+B11
κ
1,M=B11
ε
1+D11
κ
1,Q=ksA55
γ
s(7.3.3)
with A11 =bA11,B11 =bB11 ,D11 =bD11,
γ
s=
ε
5and the stiffness equations are
from (4.2.15) and Q(k)
11 C(k)
11 =E(k)
1,Q(k)
55 C(k)
55 =G(k)
13
A11 =b
n
k=1
C(k)
11 x(k)
3x(k1
3)=b
n
k=1
C(k)
11 h(k),
A55 =b
n
k=1
C(k)
55 x(k)
3x(k1)
3=b
n
k=1
Ck
55h(k),
B11 =b1
2
n
k=1
C(k)
11 x(k)
3
2x(k1)
3
2,
D11 =b1
3
n
k=1
C(k)
11 x(k)
3
3x(k1)
3
3
ksis the shear correction factor (Sect. 5.3).
In the static case, the equilibrium equations for the undeformed beam element
(Fig. 7.2) yield again for lateral loading q6=0
dM
dx1Q=0,dQ
dx1
+q=0 (7.3.4)
When considerations are limited to symmetric laminated beams the coupling stiff-
ness B11 is zero and from (7.3.3) it follows that
M=D11
d
ψ
1
dx1
,Q=ksA55
ψ
1+dw
dx1(7.3.5)
Substituting the relations (7.3.5) into (7.3.4) leads to the differential equations of
flexure
[D11
ψ
1(x1)]ksA55[
ψ
1(x1) + w(x1)] = 0,
ksA55 [
ψ
1(x1) + w(x1)]+q(x1) = 0(7.3.6)
Derivation of the first equation of (7.3.6) and setting in the second equation yields a
differential equation of 3rd order for
ψ
1(x1)
[D11
ψ
(x)]′′ =q(x)(7.3.7)
and with
Q=dM
dx1
= [D11
ψ
1(x1)](7.3.8)
and
Q=ksA55[
ψ
1(x1) + w(x1)]
7.3 Shear Deformation Theory 245
follows an equation for w(x1)
w(x1) =
ψ
1(x1) + [D11
ψ
1(x1)]
ksA55
(7.3.9)
Summarizing the derivations above, the equations for a bent Timoshenko’s beam
are: [D11
ψ
1(x1)]′′ =q(x1),
M(x1) = D11
ψ
1(x1),
Q(x1) = [D11
ψ
1(x1)],
w(x1) =
ψ
1(x1) + [D11
ψ
1(x1)]
ksA55
(7.3.10)
When the laminated beam problem allows to write the bending moment Mand the
transverse force Qin terms of the known applied lateral loads q, like in statically
determined beam problems, (7.3.5) can be utilized to determine first
ψ
1(x1)and then
w(x1). Otherwise (7.3.6) or (7.3.10) are used to determine w(x1)and
ψ
1(x1).
Integrating the second Eq. (7.3.6) with respect to x1, we obtain
ksA55[w(x1) +
ψ
1(x1)] = Zq(x1)dx1+c1
Substituting the result into the first equation of (7.3.6) and integrating again with
respect to x1yields
D11
ψ
1(x1) = ZZ q(x1)dx1dx1+c1x1+c2,
D11
ψ
1(x1) = ZZZ q(x1)dx1dx1dx1+c1
x2
1
2+c2x1+c3
(7.3.11)
Substituting
ψ
1(x1)and
ψ
1(x1)in (7.3.9), considering (7.3.7) and integrating once
more with respect to x1we obtain
w(x1) = 1
D11 ZZZZ q(x1)dx1dx1dx1dx1+c1
x3
1
6+c2
x2
1
2+c3x1+c4
1
ksA55 ZZ q(x1)dx1dx1+c1x1
=wB(x1) + wS(x1)
(7.3.12)
The transverse deflection consists of two parts. The bending part wB(x1)is the same
as derived in the classical theory. When the transverse stiffness goes to infinity, the
shear deflection wS(x1)goes to zero,
ψ
1(x1)goes to w(x1)and the Timoshenko’s
beam theory reduces to the classical Bernoulli’s beam theory.
The relations for the stresses
σ
1are the same as in the classical theory. The trans-
verse shear stress can be computed via a constitutive equation in the Timoshenko
theory
246 7 Modelling and Analysis of Beams
σ
(k)
5(x1,x3) = Q(k)
55
Q(x1)
ksA55
(7.3.13)
The variational formulation for a lateral loaded symmetric laminated beam is given
by
Π
(w,
ψ
1) =
Π
i+
Π
a(7.3.14)
with
Π
i=1
2
l
Z
0"D11d
ψ
1
dx12
+ksA55
ψ
1+dw
dx12#dx1,
Π
a=
l
Z
0
q(x1)wdx1
(7.3.15)
In the more general case of unsymmetric laminated beams and axial and lateral
loadings we have
Π
(u,w,
ψ
1). The
Π
iexpression can be expanded to
Π
i=1
2
L
Z
0"A11 du
dx12
+2B11
du
dx1
d
ψ
1
dx1
+D11 d
ψ
1
dx12
+ksA55
ψ
+dw
dx12#dx1
(7.3.16)
and
Π
ahas to include axial and lateral loads.
Since the transverse shear strains are represented as constant through the lam-
inate thickness, it follows that the transverse stresses will also be constant. In the
elementary beam theory of homogeneous beams, the transverse shear stress varies
parabolically through the beam thickness and in the classical laminate theory the
transverse shear stresses vary quadratically through layer thickness. This discrep-
ancy between the stress state compatible with the equilibrium equations and the
constant stress state of the first order shear deformation theory can be overcome
approximately by introducing a shear correction factor (Sect. 5.3).
The shear correction factor kscan be computed such that the strain energy W1
due to the classical transverse shear stress equals the strain energyW2due to the first
order shear deformation theory. Consider, for example, a homogeneous beam with
a rectangular cross-section A=bh. The classical shear stress distribution following
from the course of elementary strength of materials is given by
σ
13 =
τ
1=3
2
Q
bh "12x3
h2#,h
2x3+h
2(7.3.17)
The transverse stress in the first order shear deformation theory is constant through
the thickness h
σ
13 =
τ
2=Q
bh ,
γ
2=Q
ksG(7.3.18)
With W1=W2it follows that
7.3 Shear Deformation Theory 247
1
2Z
(A)
τ
2
1
GdA=1
2Z
(A)
τ
2
2
ksGdA,
3
5
Q2
Gbh =1
ks
Q2
2Gbh =ks=5
6(7.3.19)
The shear correction factor for a general laminate depends on lamina properties and
lamina stacking and is given here without a special derivation by
1
ks=A55b
N
k=1
x(k)
3
Z
x(k1)
3
g(k)2(x3)
G(k)dx3,(7.3.20)
g(k)(z) = d11 (C(k)
11
z2
2+
k
j=1hC(j)
11 C(j1)
11 iz(j1)2
2),C(0)
11 0
d11 =1/D11 is the beam compliance, C(k)
11 =E(k)
1.
Summarizing the beam equations for the first order shear deformation theory
for symmetrically laminated cross-sections, including vibration and buckling, the
following relations are valid for constant values of h,b,A,D11 ,
ρ
:
Flexure equations (N=0,M6=0)
ksA55 [
ψ
1(x1) + w(x1)]+q(x1) = 0,
[D11
ψ
1(x1)]ksA55[
ψ
1(x1) + w(x1)] = 0(7.3.21)
or Eqs. (7.3.10).
Forced or free vibrations equations
ksA55 [
ψ
1(x1,t) + w(x1,t)]
ρ
0¨w(x1,t) + q(x1,t) = 0,
[D11
ψ
1(x1,t)]ksA55[
ψ
1(x1,t) + w(x1,t)]
ρ
2¨
ψ
1(x1,t) = 0,(7.3.22)
ρ
0=b
n
k=1
ρ
(k)x(k)
3x(k1)
3,
ρ
2=b
n
k=1
1
3
ρ
(k)x(k)3
3x(k1)3
3
The terms involving
ρ
0and
ρ
2are called translatory or rotatory inertia terms. For
free vibrations we assume that the transverse load qis zero and the motion is
periodic:
w(x1,t) = W(x1)exp(i
ω
t),
ψ
1(x1,t) =
Ψ
1(x1)exp(i
ω
t)
Buckling equations
ksA55 [
ψ
1(x1) + w(x1)]N(x1)w′′(x1) = 0,
[D11
ψ
1(x1)]ksA55[
ψ
1(x1) + w(x1)] = 0(7.3.23)
248 7 Modelling and Analysis of Beams
or with N(x1) = F
D11 1F
ksA55 w′′′′(x1) + F w′′(x1) = 0
The variational formulation for the symmetric bending beam is given by Eq.
(7.3.15).
For vibrations the Lagrange function L(w,
ψ
1) = T(w,
ψ
1)
Π
(w,
ψ
1)yields the
Hamilton’s principle
H[w(x1,t),
ψ
1(x1,t)] =
t2
Z
t1
L[w(x1,t),
ψ
1(x1,t)]dt,
δ
H[w(x1,t),
ψ
1(x1,t)] = 0
with
Π
[w(x1,t),
ψ
1(x1,t)] = 1
2
l
Z
0
[D11
ψ
1
2+ksA55(
ψ
1+w)2]dx1
l
Z
0
q(x1,t)wdx1,
T[w(x1,t),
ψ
1(x1,t)] = 1
2
l
Z
0
[
ρ
0˙w2+
ρ
2˙
ψ
2
1]dx1
(7.3.24)
For buckling problems with N(x1) = Fit follows that
Π
[w(x1,t),
ψ
1(x1,t)] = 1
2
l
Z
0
[D11
ψ
1
2+ksA55(
ψ
1+w)2]dx1
1
2
l
Z
0
Fw2dx1
(7.3.25)
Equations (7.3.21) to (7.3.25) summarize the bending, buckling and vibration dif-
ferential and variational statements for laminated beams based on the shear defor-
mation theory.
7.4 Sandwich Beams
The similarity of the elastic behavior between symmetric laminates and symmetric
sandwich beams in the first order shear deformation theory (Sects. 4.3 and 5.3) al-
lows us to transpose the results derived above to the bending of sandwich beams.
7.4 Sandwich Beams 249
In addition to the differences between the expressions for the flexural and trans-
verse shear stiffness D11 and A55 the essential difference is at the level of stress
distribution. The model assumptions for sandwich composites with thin and thick
cover sheets are considered in detail in Sects. 4.3.1 to 4.3.3. There one can find the
stiffness values A11,D11,A55 . With these values, all differential and variational for-
mulation of the theory of laminated beams including transverse shear deformation
can be transposed.
In the case of a symmetric sandwich beam with thin cover sheets we have, for
example, the stiffness values
A11 =2Af
11 =2
n
k=1
Q(k)
11 h(k),
D11 =2hcCf
11 =hcn
k=1
Q(k)
11 h(k)x(k),
As
55 =hcGc
13
(7.4.1)
nis the number of the face layers.
The coefficient As
55 can be corrected by a shear correction factor ks. In addition
to the calculation of ks, derived for a laminated beam an approximate formula was
developed by Reuss, for sandwich beams with thin cover sheets. With the inverse
effective shear stiffness G1
R, given by the Reuss model, and the effective shear
stiffness GV, given by the Voigt model, we have
GV=
n
k=1
G(k)h(k)
h=A55
bh ,G1
R=
n
k=1
1
G(k)
h(k)
h,ks=GR
GV
(7.4.2)
The use of sandwich structures is growing very rapidly. Sandwich beams has a
high ratio of flexural stiffness to weight and in comparison to other beam struc-
tures they have lower lateral deformations, higher buckling resistance and higher
natural frequencies. As a result sandwich constructions quite often provide a lower
structural weight than other structural elements for a given set of mechanical and
environmental loads.
The elastic behavior of sandwich beams was modelled by the laminate the-
ory, Sect. 4.3, but it is appropriately to distinguish thin and thick sandwich faces.
The differential equations or variational statements describing the structural behav-
ior of sandwich beams generally based in the first order shear deformation theory
Sect. 5.3, and only if very flexible cores are used a higher order theory may be
needed.
Because of the continuing popularity of sandwich structures Sect. 7.4 intends to
recall and summarizes the results of Sects. 4.3 and 5.3 to cover some of the most
important aspects of sandwich beam applications.
250 7 Modelling and Analysis of Beams
7.4.1 Stresses and Strains for Symmetrical Cross-Sections
Figure 7.9 shows a sandwich beam with a symmetrical lay up, i.e. the faces have the
same thickness hfand are of the same material. As derived in Sect. 4.3 and 4.4 the
flexural rigidy is (D11 DLa
11 ,Q11 E1=E)
bD11 =D=bEf(hf)3
6+Efhfd2
2+Ec(hc)3
12 =2Df+Do+Dc(7.4.3)
and both, 2Dfand Dcare less than 1% of Doif d/hf>5,77 and
(6Efhfd2)/Ec(hc)3>100. Thus, for a sandwich with thin faces hfhcand a weak
core, EcEf, the flexural rigidity is approximately
DDo=bEfhfd2
2(7.4.4)
It can be noted that in most engineering applications using structural sandwich
beam elements, the dominating term in flexural rigidity is that of the faces bending
about the neutral axes of the beam, i.e. the dominating part Doof the total rigidity
Doriginating from a direct tension-compression of the cover sheets. But is there
no monolithic bonding between the faces and the core the flexural rigidity will be
nearly lost.
The following derivations assume in-plane-, bending- and shear stiffness for all
layers, i.e. for the faces and the core. Therefore we use the laminate theory including
transverse shear, Sects. 4.3.3 and 5.3. All calculations are first restricted to midplane
symmetric beams.
The bending strains vary linearly with x3over the cross-section:
ε
M
1=M
Dx3(7.4.5)
Unlike the bending strains, which vary linearly with x3over the whole cross-section,
the bending stresses vary linearly within each material constituent, but there is a
jump in the stresses at the face/core interfaces:
x3
x1
3
2
1
Q
N
M
E3=Ef
E2=Ec
E3=Ef
d=hc+hf
b
hf
hc
hf
Fig. 7.9 Symmetrical sandwich beam: N=bN1,Q=bQ1,M=bM1are the beam stress resultants
7.4 Sandwich Beams 251
σ
(k)
1=ME(k)
Dx3(k=1,2,3)
σ
f=MEf
Dx3
σ
c=MEc
Dx3
(7.4.6)
With Eq. (7.2.12) follows
M=bD11
κ
1,
κ
1=M
bD11
=h3
12D11
M
bh3/12 =M
EM
effI,(7.4.7)
with
Eeff =12
h3D11,I=bh3
12
and the stress equations can be written as
σ
(k)
1=E(k)
EM
eff
M
Ix3(k=1,2,3)(7.4.8)
The strains due to in-plane loading are:
ε
N
1=N
3
k=1
Q(k)h(k)
=N
3
k=1
E(k)h(k)
=N
bh
h
A11
=N
EN
effA(7.4.9)
with
EN
eff =A11
h,A11
3
k=1
Q(k)
11 h(k)=
3
k=1
E(k)h(k),A=bh
ε
N
1is the strain of the neutral axis. The in-plane stresses follow to
σ
(k)
1=E(k)
EN
eff
N
A(k=1,2,3)
σ
f=E(1)
EN
eff
N
A=E(3)
EN
eff
N
A
σ
c=E(2)
EN
eff
N
A
(7.4.10)
The strains and stresses due to in-plane loads and bending can be superimposed.
In the same manner as outlined above a general definition can be found for shear
strains and shear stresses. Consider the beam element b(x(3)
3x3)dx1, Fig. 7.10. The
upper edge of the sandwich, i.e. x3= (d+hf)/2, Fig. 7.9, is stress free and we have
τ
[(d+hf)/2] = 0.
Since we restrict on calculations to midplane symmetric beams all derivation can
be given for the upper part of the beam element (x30). The equilibrium equation in
the x1-direction yield with
σ
1d
σ
1
σ
1+ (
σ
1/
x1)dx1and no edge shear stresses
on the upper face
252 7 Modelling and Analysis of Beams
x3
x1
dx1
τ
(x3)
(x(3)
3x3)
σ
f
σ
c
σ
f+d
σ
f
σ
c+d
σ
c
σ
5+d
σ
5
σ
5
σ
1
σ
1+d
σ
1
Fig. 7.10 Sandwich beam element b(x(3)
3x3)dx1:
σ
(3)(x3)
σ
f(x3),
σ
(2)(x3)
σ
c(x3),
σ
5(x3)
τ
(x3)
τ
(x3)bdx1
(d+hf)/2
Z
x3
σ
1+
σ
1
x1
dx1
σ
1bdx3=0
τ
(x3) =
(d+hf)/2
Z
x3
σ
1
x1
dx3=0 (7.4.11)
Using the relations dM(x1)/dx1=Q(x1)and
σ
1=M(E(x3)/D)x3we have
d
σ
1
dx1
=Q(x1)
DE(x3)x3,
τ
(x3) = Q
bD
(d+hf/2)
Z
x3
E(x3)x3bdx3=Q
bD S(x3)(7.4.12)
S(x3)is the first moment of the area (x(3)
3x3)b. For a single layer homogeneous
and isotropic beam we have the well-known formula
S(x3) = b
2h2
4x2
3;h= (d+hf)
For sandwich beams we have a more generalized definition for the first moment of
area:
(d+hf)/2
Z
x3
E(x3)x3bdx3=
bEfhfd
2+Ec
2hc
2x3hc
2+x3,
|x3| hc
2
bEf
2hc
2+hfx3hc
2+hf+x3,
hc
2|x3| hc
2+hf
(7.4.13)
7.4 Sandwich Beams 253
The shear stresses for the core and the faces are
τ
c(x3) = Q
DEfhfd
2+Ec
2(hc)2
4x2
3,
τ
f(x3) = Q
D
Ef
2(hc)2
4+hchf+ (hf)2x2
3(7.4.14)
The maximum shear stress appears at the neutral axes:
τ
max =
τ
c(x3=0) = Q
DEfhfd
2+Ec(hc)2
8(7.4.15)
The shear stress in the core/face interface is
τ
c
min
τ
f
max =
τ
hc
2=Q
DEfhfd
2(7.4.16)
There is no jump in the shear stresses at the interfaces and the shear stresses are zero
at the outer fibres of the faces. If we have
4Efhfd
Ec(hc)2>100 (7.4.17)
the shear stresses in the core are nearly constant. The difference between
τ
c
max and
τ
c
min is less than 1%. As it was outlined in Sect. 4.3, the stress equation in sandwich
beams very often can be simplified.
Summarizing the stress estimations due to bending and shear for symmetrical
sandwich beams we have the following equations:
1. The core is weak, EcEf, but the faces can be thick
σ
f(x3)MEf
Do+2Dfx3,
σ
c(x3)0,
τ
f(x3)Q
Do+2Df
Ef
2(hc)2
4+hchf+ (hf)2x2
3,
τ
c(x3)QEfhfd
2(Do+2Df)
(7.4.18)
2. The core is weak, EcEf, and the faces are thin, hfhc
σ
f(x3) ± M
bhfd,
σ
c(x3)0,
τ
f(x3)0,
τ
c(x3)Q
bd (7.4.19)
This approximation can be formulated as: The faces of the sandwich beam carry
bending moments as constant tensile and compressive stresses and the core car-
ries the transverse forces as constant shear stresses.
254 7 Modelling and Analysis of Beams
7.4.2 Stresses and Strains for Non-Symmetrical Cross-Sections
In engineering applications also sandwich beams with dissimilar faces are used,
Fig. 7.11. The first moment of area is zero when integrated over the entire cross-
section and x3is the coordinate from the neutral axes
ZE(x3)x3bdx3=0 (7.4.20)
The location of neutral axis is unknown. With the coordinate transformation
x
3=x3efrom a known axis of the cross-section the equation above becomes
S(x3) = ZE(x3)x3bdx3=ZE(x
3+e)bdx
3=0,eZEdx
3=ZEx
3dx
3
For the sandwich cross-section, Fig. 7.11, follows
eE(1)h(1)+E(2)h(2)+E(3)h(3)
=E(1)h(1)1
2h(1)+h(2)+1
2h(3)+1
2E(2)h(2)h(2)+h(3)
and we get an equation for the unknown value e
e=
E(1)h(1)h(1)+2h(2)+h(3)+E(2)h(2)h(2)+h(3)
2E(1)h(1)+E(2)h(2)+E(3)h(3)(7.4.21)
If the core is weak, E(2)(E(1),E(3))we have approximately
e=E(1)h(1)d
E(1)h(1)+E(3)h(3)or de=E(3)h(3)d
E(1)h(1)+E(3)h(3),(7.4.22)
where d=1
2h(1)+h(2)+1
2h(3).
The bending stiffness D=RE(x3)x2
3bdx3yields in the general case
Fig. 7.11 Definition of the
neutral axis (N.A.) of an
unsymmetrical sandwich:
x3=x
3+e
x2
x3
E(3)
E(2)
E(1)
d
h(3)
b
hc=h(2)
h(1)
e
x
3
N.A.
7.4 Sandwich Beams 255
D=1
12 hE(1)(h(1))3+E(2)(h(2))3+E(3)(h(3))3i
+E(1)h(1)(de)2+E(3)h(3)e2+E(2)h(2)1
2(h(2)+h(3))e2(7.4.23)
and can be simplified for E(2)(E(1),E(3))but thick faces as
DE(1)(h(1))3
12 +E(3)(h(3))3
12 +E(1)h(1)E(3)h(3)d2
E(1)h(1)+E(3)h(3)(7.4.24)
For thin faces the first two terms vanish
DDo=E(1)h(1)E(3)h(3)d2
E(1)h(1)+E(3)h(3)(7.4.25)
Now the bending and shearing stresses can be calculated in the usual way
σ
1(k)(x3) = ME(k)
Dx3,
τ
(k)(x3) = Q
bD S(x3)(7.4.26)
For sandwich beams with weak core and thin but dissimilar faces the stress formulas
are approximately
σ
(3)
1
σ
f1
1MEf1
De=M
bhf1d,
σ
(1)
1
σ
f2
1 MEf2
D(de) = M
bhf2d,
τ
(2)
τ
cQ
bd ,
τ
(3)=
τ
(1)0
(7.4.27)
7.4.3 Governing Sandwich Beam Equations
The following derivations assumed, as generally in Chap. 7, straight beams with at
least single symmetric constant cross-sections which are rectangular, i.e we consider
single core sandwich beams. The faces can be thin or thick and symmetrical or non-
symmetrical. The bending moments and axial forces act in the plane of symmetry
(x1x3). The influence of transverse shear deformation is included, because the
core of sandwich beams has a low transverse modulus of rigidity G13. The shear
correction factor ksis determined approximately for sandwich beams with thin cover
sheets with the Reuss formula (7.4.1), or more generally using equivalent shear
strain energy, i.e the potential energy of the applied load equals the strain energy
of the beam to account for the nonuniform shear distribution through the thickness.
The shear deformation theory (Sect. 7.3) is valid and we can adapt the equations of
this section to the special case of sandwich beams.
The strains
ε
1and
ε
5
γ
are given, (7.3.2), as
256 7 Modelling and Analysis of Beams
ε
1=du
dx1
+x3
d
ψ
dx1
,
γ
=dw
dx1
+
ψ
With
A11 =A,B11 =B,D11 =D,ksA55 =S(7.4.28)
the constitutive equations (7.2.12), (7.2.13) yield
N
M
Q
=
A B 0
B D 0
0 0 S
u
ψ
w+
ψ
(7.4.29)
For static loading q(x1)6=0,n(x1)0 the equilibrium equations are as in the clas-
sical beam theory, Table 7.1,
N=0,Q+q=0,MQ=0 (7.4.30)
If N0 the neutral axes position xN.A.
3is constant along the length of the beam and
is given by
ε
1(xN.A.
3) = u+xN.A.
3
ψ
=0,N=Au+B
ψ
=0
xN.A.
3=u
ψ
=B
A(7.4.31)
Thus, if the stiffness A,B,D,Sare constant, the substitution of Eq. (7.4.29) into
(7.4.30) yields the following two governing differential equations for sandwich
beams
DR
ψ
′′(x1)S[w(x1) +
ψ
(x1)] = 0,
S[w′′(x1) +
ψ
(x1)] = q(x1)(7.4.32)
with DR=D(B2/A). Derivation of the first equation and setting in the second
equation yield
DR
ψ
′′′(x1) = q(x1)(7.4.33)
and with
M=Q=S(w+
ψ
),M=Bu+DR
ψ
,M=Bu′′ +DR
ψ
′′
follow
w(x1) =
ψ
(x1) + 1
S(Bu′′ +DR
ψ
′′)
For symmetrical cross-sections is B=0,DR=D. In the general case, if all stiff-
ness are constant and unequal zero the substitution of (7.4.4) into (7.4.5) yields the
governing simultaneous differential equations for unsymmetrical sandwich beams
DR
ψ
′′(x1)S[w(x1) +
ψ
(x1)] = 0,DR=DB2
A
S(w′′(x1) +
ψ
(x1)) = q(x1)
u(x1) = B
A
ψ
(x1)
(7.4.34)
7.4 Sandwich Beams 257
Derivation of the first equation and setting in the second equation yield one uncou-
pled equation for
ψ
(x1)
DR
ψ
′′′(x1) = q(x1)(7.4.35)
The constitutive equations (7.4.4) give the relations
M(x1) = Bu(x1) + D
ψ
(x1),Q(x1) = S[w(x1) +
ψ
(x1)] (7.4.36)
and with
M(x1) = Q(x1) = Bu′′(x1) + D
ψ
′′(x1),u(x1) = B
A
ψ
(x1)
follow
w(x1) =
ψ
(x1) + 1
S[Bu′′(x1) + DR
ψ
′′(x1)] =
ψ
(x1) + DR
S
ψ
′′(x1)(7.4.37)
Thus we have three uncoupled differential equations:
DR
ψ
′′′(x1) = q(x1),
w(x1) =
ψ
(x1) + DR
S
ψ
′′(x1),
u(x1) = B
A
ψ
(x1)
(7.4.38)
For symmetrically cross-sections is B0, DRDand the differential equations
reduce to
D
ψ
′′′(x1) = q(x1),
w(x1) =
ψ
(x1) + D
S
ψ
′′(x1)),
M(x1) = D
ψ
(x1)or
ψ
(x1) = M(x1)
EM
effI
Q(x1) = D
ψ
′′(x1)or M(x1) = S(w(x1) +
ψ
(x1))
(7.4.39)
The equations (7.4.39) correspond to the equations (7.3.10) of the laminated beam
and the analytical solutions (7.3.11), (7.3.12) can be transposed with D11 =D,
ksA55 =S. In dependence of the calculation of the stiffness Dand Sthe equation
are valid for sandwich beams with thin or thick faces.
The stresses
σ
1and
τ
can be calculated with the help of the stress formulas
derived in Sects. 7.4.1 and 7.4.2. For statically determinate structures, M(x1)and
Q(x1)can be calculated with the equilibrium equations and the last two equations
(7.4.39) can directly used for static calculations.
We consider as an example the cantilever beam, Fig. 7.12, then:
258 7 Modelling and Analysis of Beams
Fig. 7.12 Symmetrical can-
tilever beam with thin faces
x3
x1
hf
hf
hcd
l
F
M(x1) = F(lx1),Q(x1) = F,
ψ
(x1) = M(x1)
D=1
DF(lx1),
ψ
(x1) = F
Dlx11
2x2
1+C1,
ψ
(0) = 0C1=0,
w(x1) = F
Dlx11
2x2
1+Q
S=F
Dlx11
2x2
1F
S,
w(x1) = F
D1
2lx2
11
6x3
1Fx1
S+C2,
w(0) = 0C2=0,
w(x1) = F
D1
2lx2
11
6x3
1Fx1
S=wB(x1) + wS(x1),
w(l) = Fl 3
3DFl
S
Assume thin faces and weak core, i.e
hfhc,EcEf
we have
D=Do=bEfhfd2
2,S=ksGcbd
and the stresses are
σ
f=±MEf
D
d
2=±M
bdhf=±F(lx1)
bdhf,
τ
c=Q
bd =F
bd ,
σ
c=
τ
f=0
Consider w(l) = wB(l) + wS(l)it can be seen that the shear deformation strongly
depends on land S. It is important for short and shear weak beams and negligible
for slender shear stiff beams.
Summarizing the aspects of sandwich beams it could be demonstrated in the
static case that the shear deformation theory for laminated beams is valid for sand-
wich beams, if the stiffness A11 ,B11 and D11 of a laminated beam are replaced by
the stiffness A,B,Dand S, Eq. (7.4.3), of the sandwich beam. The same conclusion
is valid not only for bending but also for buckling and vibration and for differential
7.5 Hygrothermo-Elastic Effects on Beams 259
and variational formulations. In this way all formulas (7.3.19) to (7.3.23) can easy
transposed to symmetrically sandwich beams.
In the considerations above we have assumed that the effect of core transverse
deformability is negligible on the bending, vibration and the overall buckling of
sandwich beams. But in a special case of buckling, called face wrinkling the trans-
verse normal stiffness of the core has an important influence. Wrinkling is a form
of local instability of thin faces associated with short buckling waves. This phe-
nomenon was not discussed here.
7.5 Hygrothermo-Elastic Effects on Beams
In Sects. 7.2 and 7.4 the effect of mechanical loads acting upon fibre reinforced
beams with E1(x2,x3) = E1(x2,x3)and laminated or sandwich beams was consid-
ered. The considerations for laminated beams as derived are valid in the framework
of the classical laminate theory, Sect. 7.2, and of the first order shear deformation
theory, Sect. 7.3. Section 7.4 considered some special aspects of sandwich beams
with thin or thick cover sheets and different stiffness of the core.
In the present section the effects of hygrothermally induced strains, stresses and
displacements are examined. We assume a moderate hygrothermal loading such that
the mechanical properties remain unchanged for the temperature and moisture dif-
ferences considered.
With Eqs. (4.2.63) to (4.2.68) the beam equations (7.2.1) have additional terms
ε
1(x3) =
ε
1+x3
κ
1=
σ
1(x2,x3)
E1(x2,x3)
+ [
α
th(x2,x3)T(x2,x3) +
α
mo(x2,x3)M(x2,x3)],
(7.5.1)
σ
1(x2,x3) = E1(x2,x3)[
ε
1+x3
κ
1
α
th(x2,x3)T(x2,x3)
α
mo(x2,x3)M(x2,x3)] (7.5.2)
α
th,
α
mo are the thermal and moisture expansion coefficients, Tthe temperature
change and Mthe weight of moisture absorption per unit weight. Equations (7.2.4)
have now additional terms Nth ,Nmo,Mth ,Mmo, the so-called fictitious hygrothermal
resultants, (4.2.67), and with
˜
N=N+Nth +Nmo,˜
M=M+Mth +Mmo (7.5.3)
the extended hygrothermal constitutive equation for the composite beam are
˜
N
˜
M=a b
b d
ε
1
κ
1,
ε
1
κ
1=a b
b d 1˜
N
˜
M(7.5.4)
The stress formula (7.2.6) yields with Eq. (7.2.5)
260 7 Modelling and Analysis of Beams
σ
1(x2,x3) = (dN bM) + (aM bN)x3
ad b2E1(x2,x3)
E1(x2,x3)[
α
thT+
α
moM]
(7.5.5)
For double symmetric cross-sectional geometry the coupling coefficient is zero and
the stress equation can be simplified.
For uniform fibre distribution, i.e.
φ
=const, (7.2.9) follow for a,b,dand the
stress relations for fibres and matrices material are
σ
f(x3) = ( ˜
N/A+x3˜
M/I)(Ef/E1)Ef(
α
thT+
α
moM),
σ
m(x3) = ( ˜
N/A+x3˜
M/I)(Em/E1)Em(
α
thT+
α
moM)(7.5.6)
With Ef=Em=E1=Ecomes the stress equation for isotropic beams with mechan-
ical and hygrothermal loadings
σ
(x3) = ˜
N
A+x3
˜
M
IE(
α
thT+
α
moM)(7.5.7)
For laminate or sandwich beams the developments are similar. All problems are
linear and the principle of superposition is valid and can be used to calculate the
hygrothermal effects. Consider for example a symmetric laminate beam in the frame
of the classical laminate theory and include hygrothermal loads, (7.2.27) yield
σ
(k)
1=E(k)
1
˜
N
EN
effA+x3
˜
M
EM
effIE(k)
1(
α
th(k)T(k)+
α
mo(k)M(k))(7.5.8)
The differential equations for deflection and midplane displacement of a symmetric
laminated beam are
[D11w′′(x1)]′′ =q(x1)Mth(x1)′′ Mmo(x1)′′,
[A11ux1)]=n(x1) + Nth (x1)Nmo(x1)(7.5.9)
In an analogous manner the differential equations including shear deformation can
be found. The differential equation for a symmetric Timoshenko’s beam with lateral
loading and hygrothermal effects follows with (7.3.10)
[D11
ψ
1(x1)]′′ =q(x1) + Mth(x1)′′ +Mmo(x1)′′ (7.5.10)
The relation for the layer stresses
σ
(k)
1are identical to the classical theory. The trans-
verse shear stresses
σ
(k)
5are not changed by hygrothermal effects.
7.6 Analytical Solutions
The differential equations for bending, vibration and buckling of symmetric lam-
inated beams loaded orthogonally to the plane of lamination are summarized by
7.6 Analytical Solutions 261
(7.2.33) to (7.2.35) for the classical Bernoulli’s beam theory and by (7.3.21) to
(7.3.23) for the Timoshenko’s beam theory including transverse shear deformation.
All stiffness and material parameters are constant values.
The simplest problem is the analysis of bending. The general solution of the
differential equation of 4th order (7.2.33) for any load q(x1)is given by
bD11w(x1)bD11 wB(x1)
=C1
x3
1
6+C2
x2
1
2+C3x1+C4+ZZZZ q(x1)dx1dx1dx1dx1
(7.6.1)
The general solution of the Timoshenko’s beam is given by (7.3.12) in the form
w(x1) = wB(x1) + wS(x1). The correction term wS(x1)describes the influence of the
shear deformation and it decreases with increasing shear stiffness ksA55.
The free vibration of Bernoulli’s beams is modelled by (7.2.34), rotatory inertia
terms are neglected. The partial differential equation
4w(x1,t)
x4
1
+
ρ
A
bD11
2w(x1,t)
t2=0 (7.6.2)
can be separated with w(x1,t) = W(x1)T(t)and yields
W′′′′(x1)T(t) =
ρ
A
bD11
W(x1)¨
T(t)(7.6.3)
or ¨
T(t)
T(t)=
ρ
A
bD11
W′′′′(x1)
W(x1)=
ω
2(7.6.4)
We get two differential equations
¨
T(t) +
ω
T(t) = 0,W′′′′(x1)
ρ
A
bD11
ω
2W(x1) = 0 (7.6.5)
with the solutions
T(t) = Acos
ω
t+Bsin
ω
t,
W(x1) = C1cos
λ
lx1+C2sin
λ
lx1+C3cosh
λ
lx1+C4sinh
λ
lx1,
λ
l4
=
ρ
A
bD11
ω
2
(7.6.6)
The vibration mode is periodic, and
ω
is called the natural circular frequency. The
mode shapes depend on the boundary conditions of the beam. Consider, for exam-
ple, a simply supported beam, we have W(0) = W′′(0) = W(l) = W′′(l) = 0 and
therefore C1=C3=C4=0 and C2sin(
λ
/l)l=C2sin
λ
=0, which implies that
262 7 Modelling and Analysis of Beams
λ
=n
π
,
ω
2
n=n4
π
4
ρ
A
bD11
l4,
ω
n=n
π
l2sbD11
ρ
A(7.6.7)
For each nthere is a different natural frequency and a different mode shape. The
lowest natural frequency, corresponding to n=1, is termed the fundamental fre-
quency. If the laminate beam is unsymmetric to the middle surface, i.e. B11 6=0,
then D11 in Eq. (7.6.7) can be approximately replaced by (A11D11 B2
11)/A11, the
so called reduced or apparent flexural stiffness.
Including shear deformation effects, i.e. using the Timoshenko vibration equa-
tion (7.3.22), involves considerable analytical complications. To prove whether
the transverse shear deformation can be important for the natural frequencies,
we compare the natural frequencies for a simply supported Bernoulli and Timo-
shenko beam. Using (7.3.22) the boundary conditions for the Timoshenko beam are
w(0,t) = w(l,t) =
ψ
1(0,t) =
ψ
1(l,t) = 0 and by introducing
w=Asin
ω
tsin n
π
x1
l,
ψ
1=Bsin
ω
tcos n
π
x1
l,
in Eqs. (7.3.22) we can calculate the natural frequencies
ω
2
n=n4
π
4
ρ
A
bD11
l4/1+
π
2D11n2
l2ksA55 ,
ρ
A
ρ
0(7.6.8)
i.e.
ω
n=
ω
Bernoulli
ns1
1+ (n2
π
2D11)/(l2ksA55 )
Transverse shear deformation reduces the values of vibration frequencies. As in
the case of static bending the influence of shear on the values of vibration frequen-
cies depends on the ratio E1/G13 E1/E5and the ratio l/h, i.e the span length
between the supports to the total thickness of the laminate. For more general bound-
ary conditions we can develop a mode shape function similar to (7.6.6).
In an analogous way, one can show that the buckling loads for a simply supported
Bernoulli and Timoshenko beam with a compression load Ffollow from (7.2.35)
and (7.3.23) and are
Fcr =
π
2bD11
l2(Minimum Euler load, Bernoulli beam)
Fcr =
π
2bD11
l2
1
1+
π
2D11
l2ksA55
(Minimum buckling load, Timoshenko beam)
(7.6.9)
The buckling loads for clamped beams or beams with more general boundary con-
ditions can be calculated analytically analogous to eigenfrequencies of vibration
problems.
7.7 Problems 263
For non constant cross-section, stiffness or material parameters there are no exact
analytical solutions. Approximate analytical solutions can be found with the help of
the Rayleigh-Ritz procedure. In Sect. 7.7 exact and approximate analytical solution
procedures are illustrated for selected beam problems.
7.7 Problems
Exercise 7.1. A reinforced concrete beam is loaded by a bending moment M
(Fig. 7.13). It is assumed that the concrete has zero strength in tension so that the
entire tensile load associated with the bending moment is carried by the steel rein-
forcement. Calculate the stresses
σ
m(x3)and
σ
f(x3)in the concrete part (Am,Em)
and the steel reinforcements (Af,Ef).
Solution 7.1. The neutral axis x1of the beam is in an unknown distance
α
hfrom the
top, AcAmis the effective area of the concrete above the x1-axis. The strains will
vary linearly from the x1-axis and the stresses will equal strain times the respective
moduli. The stress resultant N(x1)must be zero
σ
fAf1
2
σ
m(
α
h)b
α
h=0,
σ
m(
α
h) =
σ
max
m
With (7.2.1) follows
σ
f= (h
α
h)
κ
1Ef,
σ
m(
α
h) =
α
h
κ
1Em
i.e.
(h
α
h)EfAf1
2(
α
h)2bEm=0,
α
=EfAf
Embh 1+s1+2bh
Af
Em
Ef!
or
x3
x1
α
h
b
(h
α
h)
h
h
α
h
3
Af
σ
max
m
1
2
σ
max
mb
α
h
(
σ
fAf)
M
AcAm
Fig. 7.13 Reinforced concrete beam loaded by pure bending
264 7 Modelling and Analysis of Beams
α
=1
m(1+1+2m),m=Embh
EfAf
Now the bending moment is with
σ
fAf=1
2
σ
mbh2
α
2
M= (
σ
fAf)h
α
h
3=
σ
fAfh1
α
3
=
κ
1EfAf(h
α
h)h
α
3
The maximal stress in the concrete is
σ
m(
α
h) =
κ
1
α
hEm=MEm
α
h
EfAf(h
α
h)(h
α
h/3)
and the reinforcement stress is
σ
f=
κ
1(h
α
h)Ef=M
Af(h
α
h/3)
Exercise 7.2. A symmetric cross-ply laminate beam is shown in Fig. 7.14. The ma-
terial properties and the geometry are defined by
E
1=17,24 104MPa,E
2=0,6895 104MPa,
G
12 =G
13 =0,3448 104MPa,G
23 =0,1379 104MPa,
ν
12 =0,25,
L=240 mm,b=10 mm,h(1)=h(2)=h(3)=8 mm,h=24 mm,
q0=0,6895 N/mm
x1
x3
L
h
3
h
3
h
3
h
q(x1) = q0sin
π
x1
L
Fig. 7.14 Simply supported cross-ply laminated beam [0/90/0]
7.7 Problems 265
Calculate a approximative solution using the Timoshenko beam model and two one-
term Ritz procedures.
Solution 7.2. Let us introduce the cross-section geometry
x[0]
3=12 mm,x[1]
3=4 mm,x[2]
3=4 mm,x[3]
3=12 mm
The shear correction factor can be calculated with Eq. (7.3.20) to ks=0,569. The
bending stiffness ¯
D11 and the shear stiffness ks¯
A55 follow with Eq. (7.3.3)
¯
D11 =b
3[E1((4)3(12)3) + E2(43(4)3) + E1(12343)]
=1,92 109Nmm2,
ks¯
A55 =ksb[G128+G238+G128] = 3,76 105N
The variational formulation for a lateral loaded symmetric laminate beam is given
with (7.3.15)
Π
(w,
ψ
) = 1
2
L
Z
0"¯
D11 d
ψ
dx12
+ks¯
A55
ψ
+dw
dx12#dx1
L
Z
0
q0sin
π
x1
Lwdx1
The essential boundary conditions are
w(x1=0) = 0,w(x1=L) = 0,
ψ
(x1=0) = 0,
ψ
(x1=L) = 0
The approximate functions are
˜w(x1) = a1sin
π
x1
L,˜
ψ
(x1) = b1cos
π
x1
L
and it follows
˜
Π
(˜w,˜
ψ
) = 1
2
L
Z
0¯
D11 b1
π
L2
sin2
π
x1
L
+ks¯
A55 b1cos
π
x1
L+a1
π
Lcos
π
x1
L2dx1
L
Z
0
q0sin
π
x1
La1sin
π
x1
Ldx1=˜
Π
(a1,b1)
With (2.2.41) must be
˜
Π
/
a1=0,
˜
Π
/
b1=0 which yields the two equations
¯
D11
ks¯
A55
π
L+L
π
b1+a1=0,
b1+
π
La1=q0L
ks¯
A55
π
266 7 Modelling and Analysis of Beams
and the solution for the unknown constants a1,b1
a1=q0L4
¯
D11
π
41+¯
D11
ks¯
A55
π
2
L2,
b1=q0L3
¯
D11
π
3
The approximate solutions are now
˜w(x1) = q0L4
¯
D11
π
41+¯
D11
ks¯
A55
π
2
L2sin
π
x1
L,
˜
ψ
(x1) = q0L3
¯
D11
π
3cos
π
x1
L
Note that
˜wmax =˜wx1=L
2=q0L4
¯
D11
π
41+¯
D11
ks¯
A55
π
2
L2
The transverse deflection consists of two parts
˜wB(x1) = q0L4
¯
D11
π
4sin
π
x1
L(bending deflections),
˜wS(x1) = q0L4
¯
D11
π
4
¯
D11
ks¯
A55
π
2
L2sin
π
x1
L
For ks¯
A55 follows ˜wS0, i.e. ˜wB(x1)is the solution of the Bernoulli beam
model and we found
wTimoshenko =k1wBernoulli
with
k1=1+¯
D11
ks¯
A55
π
2
L2=1,875
For the laminate beam with h/L=1/10 the Bernoulli model cannot be accepted, the
relative error for the maximum value of the deflection is 46,7 %. Equations (7.3.10)
lead
˜
Mmax =˜
Mx1=L
2=q0L2
π
2=12,64 Nm,
˜
Qmax =˜
Q(x1=0) = q0L
π
=52,67 N
The strains
ε
1follow from Eqs. (7.3.1) or (7.3.2) and (7.3.10)
ε
1(x1) = x3
ψ
(x1) = x3
κ
1=x3M(x1)/¯
D11
ε
1(x1)is linear distributed across hand we calculate the following values for the
cross-section x1=L/2
7.7 Problems 267
˜
ε
(3)
1x1=L
2,x[3]
3=2,51 105,
ε
(3)
1x1=L
2,x[2]
3=0,84 105
The bending stresses
σ
1(x1,x3)in the 3 layers are for x1=L/2 and x3=x(k)
3
˜
σ
(3)
1(x(3)
3) = ˜
ε
1(x(3)
3)E
1=4,327 MPa,
˜
σ
(3)
1(x(2)
3) = ˜
ε
1(x(2)
3)E
1=1,448 MPa,
˜
σ
(2)
1(x(2)
3) = ˜
ε
1(x(2)
3)E
2=0,579 MPa,
˜
σ
(2)
1(0) = ˜
ε
1(0)E
2=0 MPa
Exercise 7.3. Find the analytical solution for the natural vibrations of a simply sup-
ported symmetric laminate or sandwich beam. Test the influence of the transverse
shear deformation and the rotatory inertia upon the natural frequencies.
Solution 7.3. Starting point are the (7.3.22) with q(x1,t)0 and the boundary con-
ditions
w(0,t) = w(l,t) = 0,
ψ
(0,t) =
ψ
(l,t) = 0
For a simply supported beam we can assume the periodic motion in the form
w(x1,t) = W(x1)sin
ω
t,
ψ
1(x1,t) =
Ψ
(x1)sin
ω
t
These functions are substituted in (7.3.22)
ks¯
A55[W′′(x1) +
Ψ
(x1)] +
ρ
0
ω
2W(x1) = 0,
¯
D11
Ψ
′′(x1)ks¯
A55[W(x1) +
Ψ
(x1)] +
ρ
2
ω
2
Ψ
(x1) = 0
Now we can substitute
ks¯
A55
Ψ
(x1) =
ρ
0
ω
2W(x1)ks¯
A55W′′(x1),
i.e.
Ψ
(x1) =
ρ
0
ω
2
ks¯
A55
W(x1)W′′(x1)
into the derivative of the second equation and we find
¯
D11W′′′′(x1) + ¯
D11
ρ
0
ks¯
A55
+
ρ
2
ω
2W′′(x1)
1
ω
2
ρ
2
ks¯
A55
ρ
0
ω
2W(x1) = 0
or
aW ′′′′(x1) + bW′′(x1)cW(x1) = 0
with the coefficients
268 7 Modelling and Analysis of Beams
a=¯
D11,b=¯
D11
ks¯
A55
+
ρ
2
ρ
0
ρ
0
ω
2,c=1
ω
2
ρ
2
ks¯
A55
ρ
0
ω
2
The linear differential equation of 4th order has constant coefficients and the general
solutions follow with the solutions
λ
iof the bi-quadratic characteristic equation
a
λ
4b
λ
2c=0
i.e. (2a
λ
2b)2=b2+4ac
λ
14=±r1
2a(b±pb2+4ac)
W(x1) = C1sin
λ
1x1+C2cos
λ
2x1+C3sinh
λ
3x1+C4cosh
λ
4x1
For a simply supported beam the boundary conditions are
W(0) = 0,W(L) = 0,
Ψ
(0) = 0,
Ψ
(L) = 0
or the equivalent equations
W(0) = 0,W(L) = 0,W′′(0) = 0,W′′(L) = 0
The boundary conditions lead to the result C2=C3=C4=0 and C1sin
λ
1L=0
which implies
λ
1n=n
π
L=
λ
n
The bi-quadratic equation can be written alternatively in terms of
ω
A
ω
4B
ω
2+C=0
with
A=
ρ
2
ks¯
A55
,B=1+¯
D11
ks¯
A55
+
ρ
2
ρ
0
λ
2,C=¯
D11
ρ
0
λ
4
i.e. the roots of the equation are
(
ω
2)1/2=1
2A(B±pB24AC)
It can be shown that B24AC >0. Therefore the frequency given by B24AC
is the smaller of the two roots.
When the rotatory inertia is neglected follows A0 and the frequency is given
by
ω
2=C
B
with
B=1+¯
D11
ks¯
A55
λ
2,C=¯
D11
ρ
0
λ
4
7.7 Problems 269
and for ks¯
A55 follow with ˜
B1 the natural frequency for the Bernoulli beam
model. Substitute
λ
n=n
π
/Lfor the simply supported beam we obtain the results:
General case
ω
2
n=ks¯
A55
2
ρ
21+¯
D11
ks¯
A55
+
ρ
2
ρ
0n
π
L2
s1+¯
D11
ks¯
A55
+
ρ
2
ρ
0n
π
L22
4
ρ
2
ks¯
A55
¯
D11
ρ
0n
π
L4
Rotatory inertia neglected (
ρ
20)
ω
2
n=n
π
L4¯
D11
ρ
0
1n
π
L2¯
D11
ks¯
A55 +¯
D11 n
π
L2
=n
π
L4¯
D11
ρ
0
1
1+¯
D11 n
π
L2/ks¯
A55
Classical beam theory (ks¯
A55 ,
ρ
2=0)
ω
2
n=n
π
L4¯
D11
ρ
0
Conclusion 7.2. (
ω
Timoshenko
n)2<(
ω
Bernoulli
n)2, i.e. the shear deformation de-
creases the frequencies of natural vibration. In the case of classical beam theory with
rotatory inertia (ks¯
A55 0,
ρ
26=0) we have A=0,B=1+
λ
2
ρ
2/
ρ
0,C=
λ
4¯
D11/
ρ
0,
i.e.
ω
2
n=n
π
L4¯
D11
ρ
0
1
1+n
π
L2
ρ
2
ρ
0
and we see that also the rotatory inertia decreases the eigenfrequencies. All formulas
can be used for computing natural frequencies for all symmetric laminate and sand-
wich beams. The values for L,
ρ
0,
ρ
2,ks,¯
D11,¯
A55 correspond to the special beam
model. Note that the classical laminate theory and the neglecting of rotatory inertia
lead to a overestimation of the natural frequencies.
Exercise 7.4. Calculate the buckling load of a simply supported and a clamped sym-
metric laminate or sandwich beam. Compare the results for the classical beam the-
ory and the beam theory including shear deformation.
Solution 7.4. Staring point are Eqs. (7.3.23) with N(x1) = F, i.e.
¯
D11 1F
ks¯
A55 w′′′′(x1) + F w′′(x1) = 0
270 7 Modelling and Analysis of Beams
or
w′′′′(x1) + k2w′′(x1) = 0,k2=F
¯
D11 1F
ks¯
A55 ,Fk2¯
D11
1+k2¯
D11/ks¯
A55
The linear differential equation with constant coefficients has the characteristic
equation
λ
4+k2
λ
2=0
λ
2(
λ
2+k2) = 0
with the four roots
λ
1/2=0,
λ
3/4=±ik
and the general solution is
w(x1) = C1sin kx1+C2coskx1+C3x1+C4
1. Simply supported beam
Boundary conditions are w(0) = w(L) = 0,w′′(0) = w′′(L) = 0, which leads the
constants C2=C3=C4=0 and for C16=0 follows sin kL =0 implies kL =
n
π
,k=n
π
/L. Substituting kinto the equation for Fwe obtain
F=n
π
L2¯
D11
ks¯
A55
ks¯
A55 +n
π
L2¯
D11
=n
π
L2¯
D11
1
¯
D11 n
π
L2/ks¯
A55
1+¯
D11 n
π
L2/ks¯
A55
The critical buckling load Fcr is given for the minimum (n=1)
F=
π
L2¯
D11
1
1+
π
L2¯
D11/ks¯
A55
For the classical beam model is ks¯
A55 and we obtain
F=
π
L2¯
D11
2. At both ends fixed beam (clamped beam)
Now we have the boundary conditions
w(0) = w(L) = 0,
ψ
(0) =
ψ
(L) = 0
From (7.3.23) follows
7.7 Problems 271
ks¯
A55[w′′(x1) +
ψ
(x1)] Fw′′(x1) = 0,
¯
D11
ψ
′′(x1)ks¯
A55[w(x1) +
ψ
(x1)] = 0
The first equation yields
ks¯
A55
ψ
(x1) = (ks¯
A55 F)w′′(x1),
ks¯
A55
ψ
(x1) = (ks¯
A55 F)w(x1) + K1
The boundary conditions lead the equations
C2+C4=0,C1sin k L +C2cos k L +C3L+C4=0,
1F
ks¯
A55 kC1C3=0,
1F
ks¯
A55 (kC1coskL kC2sinkL)C3=0
Note that
F
k2=¯
D11
1+k2¯
D11
ks¯
A55
=1F
ks¯
A55 ¯
D11,
i.e. 1F
ks¯
A55 =1
1+k2¯
D11
ks¯
A55
,
expressing C4and C3in terms of C1and C2and setting the determinant of the re-
maining homogeneous algebraic equations zero we obtain the buckling equation
2(coskL 1)1+k2¯
D11
ks¯
A55 +kL sin kL =0
With ks¯
A55 follows the buckling equation for the classical beam
kL sin kL +2 coskL 2=0
Conclusion 7.3. Transverse shear deformation has the effect of decreasing the
buckling loads, i.e. the classical laminate theory overestimates buckling loads.
The buckling equations can be applied to all symmetric laminate and sandwich
beams if the corresponded material and stiffness values are calculated and sub-
stituted.
Exercise 7.5. A sandwich beam is modelled by the laminated beam version and
the shear deformation theory. Consider the variational formulation for applied dis-
tributed transverse loading and calculate the Euler differential equation and the
boundary conditions.
272 7 Modelling and Analysis of Beams
Solution 7.5. The elastic potential
Π
(u,w,
ψ
)for unsymmetrical laminated beams
is given by Eq. (7.3.16). Using the notations for the stiffness of sandwich beams,
Sect. 7.4, we have
Π
(u,w,
ψ
) = 1
2
l
Z
0
[Au2+2Bu
ψ
+D
ψ
2+S(w+
ψ
)2]dx1
l
Z
0
qwdx1
Taking the variation
δΠ
=0 one can write the following equation
δΠ
=
l
Z
0{Au
δ
u+B(u
δψ
+
ψ
δ
u) + D
ψ
δψ
+S[(w+
ψ
)
δ
w+ (w+
ψ
)
δψ
]}dx1
l
Z
0
q
δ
wdx1=0
Integrating by parts, i.e
l
Z
0
fgdx= [ fg]l
0
l
Z
0
f′′gdx
yield
l
Z
0{(Au′′ +B
ψ
′′)
δ
u+ [Bu′′ +D
ψ
′′ S(w+
ψ
)]
δψ
+S[(w′′ +
ψ
) + q]
δ
w}dx1
[(Au+B
ψ
)
δ
u]l
0[(Bu+D
ψ
)
δψ
]l
0S[(w+
ψ
)
δ
w]l
0=0
and the associated differential equations and boundary conditions are
Au′′ +B
ψ
′′ =0,
Bu′′ +D
ψ
′′ S(w+
ψ
) = 0,
S(w′′ +
ψ
) + q=0
Putting in u′′ =(B/A)
ψ
′′ into the second equation yield
DB2
A
ψ
′′ S(w+
ψ
) = 0
and we obtain the differential equations (7.4.32)
DR
ψ
′′ S(w+
ψ
) = 0,
S(w′′ +
ψ
) = q
The boundary conditions for x1=0,lare
7.7 Problems 273
u=0 or Au+B
ψ
=N=0
ψ
=0 or Bu+D
ψ
=M=0
w=0 or S(w+
ψ
) = Q=0
u,
ψ
,wrepresent the essential and N,M,Qthe natural boundary conditions of the
problem. For symmetric sandwich beams the equations can be simplified: B=
0,DR=D.
Exercise 7.6. A sandwich beam is modelled by the differential equations and
boundary conditions of Excercise 7.5. Calculate the exact solution for a simply sup-
ported beam with q(x1) = q0,N(x1) = 0.
Solution 7.6. The boundary conditions are:
w(0) = w(l) = 0,M(0) = M(l) = 0
Using the equations (7.4.38)
ψ
′′′(x1) = q0
DR
,
ψ
′′(x1) = q0x1
DR
+C1,
ψ
(x1) = q0x2
1
2DR
+C1x1+C2,
ψ
(x1) = q0x3
1
6DR
+C1
x2
1
2+C2x1+C3,
ψ
(0) = 0C2=0,
ψ
(l) = 0C1=q0l
2DR
,
w(x1) =
ψ
(x1) + DR
S
ψ
′′ =q0x3
1
6DRq0lx2
1
4DR
+C3q0
Sx1l
2,
w(x1) = q0
24DR
(x4
12lx3
1) + C3x1q0
2Sx2
1lx1,
w(l) = 0C3=q0l3
24DR
,
u(x1) = B
A
ψ
(x1) = B
A
q0
2DR
(x2
1lx1),
u(x1) = B
A
q0
12DR
(2x3
13lx1) + C4,
u(0) = 0C4=0
Finally we get
274 7 Modelling and Analysis of Beams
ψ
(x1) = q0x3
1
6DR
+q0l
2DR
x2
1
2+q0l3
24DR
,
w(x1) = q0
24DR
(x4
12lx3
1+l3x1) + q0
2Slx1x2
1,
u(x1) = B
A
q0
12DR
(2x3
13lx1)
For symmetrical beams the solution simplified with B=0,DR=Dto
w(x1) = q0l4
24D(x4
12x3
1+x1) + q0l2
2S(x1x2
1)
=wB(x1) + wS(x1),
ψ
(x1) = q0l3
24D(1+6x2
14x3
1)
u(x1)0,x1=x/l
Chapter 8
Modelling and Analysis of Plates
The modelling and analysis of plates constituted of laminate or sandwich material is
a problem of more complexity than that of beams, considered in Chap. 7. Generally,
plates are two-dimensional thin structure elements with a plane middle surface. The
thickness his small relatively to the two other dimensions a,b(Fig. 8.1).
In Chap. 8 all derivatives are as a matter of priority restricted to rectangular plates
including the special case of a plate strip, i.e. a rectangular plate element which
is very long, for instance in the x2-direction and has finite dimension in the x1-
direction. When the transverse plate loading, the plate stiffness, and the boundary
conditions for the plate edges x1=const are independent of the coordinate x2, the
plate strip modelling can be reduced to a one-dimensional problem. The analysis is
x1,u1
x2,u2
x3,u3
a
b
h
x1
x1
x2
x2
x3
x3
N1
N2
Nn
Q1
Q2
Qn
N6
N6
Nnt
M1
M2
M6
M6
Mn
Mnt
ab
Fig. 8.1 Rectangular plate. aGeometry, bforce resultants N1,N2,N6,Q1,Q2and moment resul-
tants M1,M2,M6.Nn,Nnt ,Qnand Mn,Mnt are force and moment resultants for an oblique edge
275
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_8
276 8 Modelling and Analysis of Plates
nearly the same as in the beam theory.Chapter 8 gives a first introduction to the clas-
sical plate theory and the plate theory including transverse shear deformations. The
derivations of the principal equations for plates relies upon the basic considerations
of Chap. 5.
8.1 Introduction
In the theory of plate bending the most complex problem is the modelling and anal-
ysis of laminate plates with an arbitrary stacking of the layers. These plates present
couplings of stretching and bending, stretching and twisting and bending and twist-
ing and the design engineer has to look for simplifications.
The first and most important simplification is to design symmetric laminates
for which no coupling exists between in-plane forces and flexural moments. The
coupling terms Bi j of the constitutive equations vanish. An additional simplifica-
tion occurs when no bending-twisting coupling exist, i.e the terms D16 and D26 are
zero. As we discussed in Sect. 4.2, in some cases of layer stacking these coupling
terms decrease with an increasing number of layers. Symmetric laminates for which
no bending-twisting coupling exists are referred to as specially orthotropic lami-
nates. These laminates are considered in detail in this chapter, because analytical
solutions exist for various loadings and boundary conditions. Specially orthotropic
plates are obtained for single layer plates with orthotropic material behavior or sym-
metric cross-ply laminates. Symmetric balanced laminates with a great number of
layers have approximately a specially orthotropic behavior. This class of laminates
is greatly simplified and will be used to gain a basic understanding of laminate plate
response. Like in Chap. 7 for beams, we consider the plates in the framework of the
classical and the first order shear deformation theory. For a better understanding the
assumptions of both plate theories given in Sects. 5.1 and 5.2 are reviewed.
The first order shear deformation theory accounted for a constant state of trans-
verse shear stresses, but the transverse normal stress is often neglected. In the frame-
work of this plate theory, the computation of interlaminar shear stresses through
constitutive equations is possible, which is simpler than deriving them through equi-
librium equations.
The most significant difference between the classical and first-order shear defor-
mation theory is the effect of including transverse shear deformation in the predic-
tion of deflections, frequencies or buckling loads. It can be noted that the classical
laminate theory underestimates deflections and overestimates frequencies as well as
buckling loads when the plate side-to-thickness ratio is of the order 20 or less. For
this reason it is necessary to include shear deformation for moderately thick plates.
In general, moderately thick plates must be computed by numerical methods, ap-
plication of analytical methods are much more restricted than in the classical plate
theory.
8.2 Classical Laminate Theory 277
8.2 Classical Laminate Theory
In the classical laminate theory one presumes that the Kirchhoff hypotheses of the
classical plate theory remains valid:
Transverse normals before deformation remain straight after deformation and
rotate such that they remain normal to the middle surface.
Transverse normals are inextensible, i.e. they have no elongation.
These assumptions imply that the transverse displacement wis independent of the
thickness coordinate x3, the strains
ε
3,
ε
4and
ε
5are zero and the curvatures
κ
iare
given by
[
κ
1
κ
2
κ
6] =
2w
x2
1
2w
x2
22
2w
x1
x2(8.2.1)
Figure 8.1 shows the plate geometry and the plate stress resultants. The equilibrium
equations will be formulated for a plate element dx1dx2(Fig. 8.2) and yield three
force and two moments equations
ւ
N1
x1
+
N6
x2
=p1,
N6
x1
+
N2
x2
=p2,
Q1
x1
+
Q2
x2
=p3,
M1
x1
+
M6
x2
=Q1,
ւ
M6
x1
+
M2
x2
=Q2
(8.2.2)
x1
x2
x3
N1dx2
(N1+dN1)dx2
N2dx1(N2+dN2)dx1
N6dx1
(N6+dN6)dx1
dx2
dx2
dx1
dx1
h
h
M6dx1
(M6+dM6)dx1
M6dx2
(M6+dM6)dx2
M1dx2
(M1+dM1)dx2
M2dx1
(M2+dM2)dx1
Q1dx2
(Q1+dQ1)dx2
Q2dx1
(Q2+dQ2)dx1
p1dx1dx2
p2dx1dx2
p3dx1dx2
ab
Fig. 8.2 Stress resultants applied to a plate element
278 8 Modelling and Analysis of Plates
The transverse shear force resultants Q1,Q2can be eliminated and the five equations
(8.2.2) reduce to three equations. The in-plane force resultants N1,N2and N6are
uncoupled with the moment resultants M1,M2and M6
N1
x1
+
N6
x2
=p1,
N6
x1
+
N2
x2
=p2,
2M1
x2
1
+2
2M6
x1
x2
+
2M2
x2
2
=p3
(8.2.3)
The equations are independent of material laws and present the static equations for
the undeformed plate element. The further considerations neglect the in-plane plate
loads p1and p2, i.e. p1=p2=0, p36=0. In-plane reactions can be caused by
coupling effects of unsymmetric laminates or sandwich plates.
Putting the constitutive equations
N
N
N
···
M
M
M
=
A
A
A.
.
.B
B
B
. . . .
B
B
B.
.
.D
D
D
ε
ε
ε
···
κ
κ
κ
(8.2.4)
into the equilibrium (8.2.3) and replacing using Eqs. (5.2.3) the in-plane strains
ε
i
and the curvatures
κ
iby
ε
ε
ε
T= [
ε
1
ε
2
ε
6] =
u
x1
v
x2
u
x2
+
v
x1,
κ
κ
κ
T= [
κ
1
κ
2
κ
6] =
2w
x2
1
2w
x2
22
2w
x1
x2(8.2.5)
gives the differential equations for general laminated plates
A11
2u
x2
1
+2A16
2u
x1
x2
+A66
2u
x2
2
+A16
2v
x2
1
+ (A12 +A66)
2v
x1
x2
+A26
2v
x2
2B11
3w
x3
13B16
3w
x2
1
x2(B12 +2B66)
3w
x1
x2
2B26
3w
x3
2
=0,
A16
2u
x2
1
+ (A12 +A66)
2u
x1
x2
+A26
2u
x2
2
+A66
2v
x2
1
+2A26
2v
x1
x2
+A22
2v
x2
2B16
3w
x3
1(B12 +2B66)
3w
x2
1
x23B26
3w
x1
x2
2B22
3w
x3
2
=0,
D11
4w
x4
1
+4D16
4w
x3
1
x2
+2(D12 +2D66)
4w
x2
1
x2
2
+4D26
4w
x1
x3
2
+D22
4w
x4
2B11
3u
x3
13B16
3u
x2
1
x2(B12 +2B66)
3u
x1
x2
2B26
3u
x3
2
B16
3v
x3
1(B12 +2B66)
3v
x2
1
x23B26
3v
x1
x2
2B22
3v
x3
2
=p3
(8.2.6)
8.2 Classical Laminate Theory 279
Equations (8.2.6) are three coupled partial differential equations for the displace-
ments u(x1,x2),v(x1,x2),w(x1,x2). Equation (8.2.6) can be formulated in matrix
form as
L11 L12 L13
L21 L22 L31
L31 L32 L33
u
v
w
=
0
0
p
,Li j =Lji (8.2.7)
The differential operators are given in App. C.
The differential operators L11 ,L12 and L22 are of second order, L13 and L23 of
third order and L33 of fourth order. The homogeneous part of the coupled partial
differential equations (8.2.7) can be reduced to one partial equation of eight order
[(L11L22 L2
12)L33 (L11L2
23 2L12L13L23 +L2
13L22 )]w=0 (8.2.8)
Consistent with the eight order set of differential equations four boundary conditions
must be prescribed for each edge of the plate. The classical boundary conditions are:
Either
Nnor u,Nnt or v,Mnor
w
n,VnQn+
Mnt
tor w(8.2.9)
must be specified. The subscripts nand tin the boundary conditions above denote
the coordinates normal and tangential to the boundary. It is well known that in the
classical plate theory the boundary cannot responded separately to the shear force
resultant Qnand the twisting moment Mnt but only to the effective or Kirchhoff
shear force resultant
VnQn+
Mnt
t(8.2.10)
Equations (8.2.9) may be used to represent any form of simple edge conditions, e.g.
clamped, simply supported and free.
The boundary conditions (8.2.9) represent pairs of response variables. One com-
ponent of these pairs involve a force or a moment resultant, the other a displace-
ment or a rotation. Take into account that in addition to the edge conditions it can
be necessary to fulfil the point corner conditions, e.g. for a free corner. Sometimes
more general boundary conditions, which are applicable to edges having elastic con-
straints, are used, e.g. the transverse and/or rotatory plate conditions
Vn±cTw=0; Mn±cR
w
n=0 (8.2.11)
cTand cRdenote the spring stiffness of the constraints.
In applying the boundary conditions (8.2.9) it is useful to have explicit expres-
sions for the stress resultants in a displacement formulation. According to Eqs.
(8.2.5) and (8.2.4) the stress resultants can be written as
N1=A11
u
x1
+A12
v
x2
+A16
u
x2
+
v
x1
280 8 Modelling and Analysis of Plates
B11
2w
x2
1B12
2w
x2
22B16
2w
x1
x2
,
N2=A12
u
x1
+A22
v
x2
+A26
u
x2
+
v
x1
B12
2w
x2
1B22
2w
x2
22B26
2w
x1
x2
,
N6=A16
u
x1
+A26
v
x2
+A66
u
x2
+
v
x1
B16
2w
x2
1B26
2w
x2
22B66
2w
x1
x2
,
M1=B11
u
x1
+B12
v
x2
+B16
u
x2
+
v
x1
D11
2w
x2
1D12
2w
x2
22D16
2w
x1
x2
,
M2=B12
u
x1
+B22
v
x2
+B26
u
x2
+
v
x1(8.2.12)
D12
2w
x2
1D22
2w
x2
22D26
2w
x1
x2
,
M6=B16
u
x1
+B26
v
x2
+B66
u
x2
+
v
x1
D16
2w
x2
1D26
2w
x2
22D66
2w
x1
x2
,
Q1=B11
2u
x2
1
+2B16
2u
x1
x2
+B66
2u
x2
2
+B16
2v
x2
1
+ (B12 +B66)
2v
x1
x2
+B26
2v
2x2D11
3w
x3
13D16
3w
x2
1
x2(D12 +2D66)
3w
x1
x2
2D26
3w
x3
2
,
Q2=B16
2u
x2
1
+ (B12 +B66)
2u
x1
x2
+B26
2u
x2
2
+B66
2v
x2
1
+2B26
2v
x1
x2
+B22
2v
2x2D16
3w
x3
1(D12 +2D66)
3w
x2
1
x23D26
3w
x1
x2
2D22
3w
x3
2
,
V1=B11
2u
x2
1
+3B16
2u
x1
x2
+2B66
2u
x2
2
+B16
2v
x2
1
+ (B12 +2B66)
2v
x1
x2
+2B26
2v
2x2D11
3w
x3
14D16
3w
x2
1
x2(D12 +D66)
3w
x1
x2
22D26
3w
x3
2
,
8.2 Classical Laminate Theory 281
V2=2B16
2u
x2
1
+ (B12 +2B66)
2u
x1
x2
+B26
2u
x2
2
+2B66
2v
x2
1
+3B26
2v
x1
x2
+B22
2v
x2
22D16
3w
x3
1(D12 +4D66)
3w
x2
1
x24D26
3w
x1
x2
2D22
3w
x3
2
The coupled system of three partial differential equations (8.2.6) or (8.2.7), respec-
tively, can be simplified for special layer stacking, Sect. 4.2.3. The differential op-
erators Li j for some special cases are given in App. C.
1. Symmetric laminates
Because all coupling stiffness Bi j are zero the in-plane and the out-of-plane dis-
placement response are uncoupled. With L13 =L31 =0,L23 =L32 =0 Eq. (8.2.7)
simplifies to
L11 L12 0
L12 L22 0
0 0 L33
u
v
w
=
0
0
p3
(8.2.13)
The plate equation reduces to L33w=p3and corresponds to the plate equation
of an anisotropic homogeneous plate.
2. Antisymmetric laminates
The in-plane and the transverse part of Eq. (8.2.7) are coupled, but with
A16 =A26 =0,D16 =D26 =0 the differential operators L11 ,L22,L33 and L12
are reduced. It is no in-plane tension/shearing coupling and no bending/twisting
coupling.
3. Balanced laminates
For general balanced laminates with A16 =A26 only the in-plane ten-
sion/shearing coupling is zero, for an antisymmetric balanced laminate we have
A16 =A26 =0,D16 =D26 =0 and for symmetric balanced laminates follow
A16 =A26 =0, Bi j =0. The last case yields the equations
L11 L12 0
L12 L22 0
0 0 L33
u
v
w
=
0
0
p3
with simplified differential operators L11 and L22. Only the in-plane equations
correspond to an orthotropic stiffness behavior.
4. Cross-ply laminates
The stacking can be unsymmetrical, i.e. A16 =A26 =0,D16 =D26 =0,
B16 =B26 =0, antisymmetrical, i.e. A16 =A26 =0,D16 =D26 =0,
B12 =B16 =B26 =B66 =0, B22 =B11 or symmetrical with A16 =A26 =0,
D16 =D26 =0,Bi j =0. Cross-ply laminates have an orthotropic response to
both in-plane and bending and no in-plane/bending coupling. The plate equation
L33w=p3corresponds to the equation of an homogeneous orthotropic plate.
Summarizing the mathematical structures of the differential equations in depen-
dence on the layer stacking the following conclusions can be drawn:
282 8 Modelling and Analysis of Plates
The mathematical structure of a general balanced laminate is not much simpler
as for a general unsymmetric, unbalanced laminate
Compared to the general case the mathematical structure of the symmetric cross-
ply laminate is nearly trivial. A symmetric cross-ply is orthotropic with respect
to both in-plane and bending behavior, and both are uncoupled.
The most simple mathematical structure yields the laminate with symmetrical
arranged isotropic layers. With A11 =A22,A16 =A26 =0,Bi j =0, D11 =D22,
D16 =D26 =0, it corresponds to a single layer isotropic plate with in-plane and
transverse loading.
For special layer stacking also the force and moment resultant Eqs. (8.2.12) are
reduced to more simple equations.
The following developments are restricted to general symmetric plates and plates
with specially orthotropic behavior. The equations will be significant simplified, for
example in the general case all Bi j =0 and for specially orthotropic plates there are
additional D16 =D26 =0. The in-plane and the flexural equations are uncoupled. Ta-
ble 8.1 summarizes the most important plate equations. In Table 8.1 standard bound-
ary conditions are also expressed. The necessary and sufficient number of boundary
conditions for plates considered here are two at each of the boundaries. The stan-
dard conditions for the free edge reduce the three static conditions Mn=0, Qn=0
and Mnt =0 to two conditions Mn=0,Vn=0, where Vn=Qn+
Mnt/
t=0 is as
discussed above the Kirchhoff effective shear resultant. In order to avoid mistakes
in the application the equations of Table 8.1, a summary of plate stiffness is given.
Table 8.2 contains the plate stiffness for single layer plates. The plate stiffness for
symmetric laminates are given in Table 8.3. In all equations the hygrothermal effects
are neglected, but it is no problem to include thermal or moisture changes. In this
case (4.2.63), (4.2.64) must be used instead of (8.2.4) to put into the equilibrium
equations. This will be considered in Sect. 8.5.
The classical laminate theory can be used also for modelling and analysis of vi-
bration and buckling of laminated plates. We restrict the consideration to symmetric
plates. In the case of forced transversal vibration the momentum equilibrium equa-
tion (8.2.3) has an additional inertial term
2M1
x2
1
+2
2M6
x1
x2
+
2M2
x2
2
=p3+
ρ
h
2w
t2(8.2.14)
M1,M2,M6,wand p3are functions of x1,x2and the time t,his the total thickness of
the plate and
ρ
the mass density
h=
n
k=1
h(k),
ρ
=1
h
n
k=1
ρ
(k)x(k)
3x(k1)
3=1
h
n
k=1
ρ
(k)h(k)(8.2.15)
The rotatory inertia is neglected. The Eqs. (8.2.14), (8.2.4) and (8.2.5) yield the plate
equations for force vibration. For the both layer stacking discussed above we obtain:
8.2 Classical Laminate Theory 283
Table 8.1 Plate equation, boundary conditions and stress resultants of symmetric laminates
1. General case: Bi j =0,Di j 6=0i,j=1,2,6
D11
4w
x4
1
+4D16
4w
x3
1
x2
+2(D12 +2D66)
4w
x2
1
x2
2
+4D26
4w
x1
x3
2
+D22
4w
x4
2
=p3
2. Specially orthotropic laminates: Bi j =0,D16 =D26 =0
D11
4w
x4
1
+2(D12 +2D66)
4w
x2
1
x2
2
+D22
4w
x4
2
=p3
or with D11 =D1,D22 =D2,D12 +2D66 =D3
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D2
4w
x4
2
=p3
3. Laminates with isotropic layers: D11 =D22 =D1,(D12 +2D66 ) = D3
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D1
4w
x4
2
=p3
Typical boundary conditions: 1. Simply supported edge: w=0,Mn=0
2. Clamped edge: w=0,
w/
n=0
3. Free edge: Mn=0,Vn=Qn+
Mnt/
t=0
Stress resultants:
1. General case
M1
M2
M6
=
D11 D12 D16
D12 D22 D26
D16 D26 D66
2w/
x2
1
2w/
x2
2
2
2w/
x1
x2
,
Q1=
M1
x1
+
M6
x2
Q2=
M6
x1
+
M2
x2
2. Specially orthotropic
M1
M2
M6
=
D11 D12 0
D12 D22 0
0 0 D66
2w/
x2
1
2w/
x2
2
2
2w/
x1
x2
,
Q1=
M1
x1
+
M6
x2
Q2=
M6
x1
+
M2
x2
3. Isotropic layers (like 2. with D11 =D22)
1. General case of symmetric plates
L33 +
ρ
h
tw=p3
or explicitly
284 8 Modelling and Analysis of Plates
Table 8.2 Plate stiffness for single layer
Anisotropic single layer
Di j =Q(k)
i j
h3
12
Specially orthotropic single layer (on-axis)
D11 =Q11
h3
12 ,D12 =Q12
h3
12 ,D22 =Q22
h3
12 ,D66 =Q66
h3
12 ,
Q
11 =E
1
1
ν
12
ν
21 ,Q
22 =E
2
1
ν
12
ν
21 ,Q
12 =
ν
12E1
1
ν
12
ν
21 ,
Q
66 =G
12 =E
6
Isotropic single layer
D11 =D22 =Eh3
12(1
ν
2)=D,D12 =
ν
D=
ν
Eh3
1
ν
2,
D66 =1
ν
2D=Eh3
24(1+
ν
),
D11
4w
x4
1
+4D16
4w
x3
1
x2
+2(D12 +2D66)
4w
x2
1
x2
2
+4D26
4w
x1
x3
2
+D22
4w
x4
2
=p3
ρ
h
2w
t2
with w=w(x1,x2,t),p= (x1,x2,t).
2. Specially orthotropic plates
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D2
4w
x4
2
=p3
ρ
h
2w
t2(8.2.16)
The equation of symmetric laminate plates with isotropic layers follows from
(8.2.16) with D1=D2, the plate stiffness are taken from Table 8.2 (single layer
plates) or Table 8.3 (laminates). In the case of the computation of natural or eigen-
vibrations, the forcing function p3(x1,x2,t)is taken to be zero and the time depen-
dent motion is a harmonic oscillation. The differential equation is homogeneous,
leading an eigenvalue problem for the eigenvalues (natural frequencies) and the
eigenfunctions (mode shapes).
To predict the buckling for plates, in-plane force resultants must be included. For
a coupling of in-plane loads and lateral deflection, the equilibrium (8.2.2) will be
formulated for the deformed plate element with p1=p2=p3=0 and modified to
8.2 Classical Laminate Theory 285
Table 8.3 Plate stiffness for symmetric laminates
Symmetric angle ply laminate
Di j =
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3=
n
k=1
Q(k)
i j h(k) x(k)
3+h(k)2
12 !,
x(k)
3=1
2x(k)
3+x(k1)
3, the Q(k)
i j follow from Table 4.2.
Symmetric balanced laminates
Di j =
n
k=1
Q(k)
i j h(k) x(k)
3+h(k)2
12 !
The Q(k)
i j follow from Table 4.2.
Symmetric cross-ply laminate (specially orthotropic)
Di j =
n
k=1
Q(k)
i j h(k) x(k)
3+h(k)2
12 !,
D16 =D26 =0
Q(k)
11 =E1
1
ν
12
ν
21 (k)
,Q(k)
22 =E2
1
ν
12
ν
21 (k)
,
Q(k)
12 =
ν
12E1
1
ν
12
ν
21 (k)
,Q(k)
66 =G(k)
12
Symmetric laminate with isotropic layers (x1-direction equal fibre direction)
Di j =
n
k=1
Q(k)
i j h(k) x(k)
3+h(k)2
12 !,
D16 =D26 =0,D11 =D22,
Q(k)
11 =Q(k)
22 =E
1
ν
2(k)
,Q(k)
12 =
ν
E
1
ν
2(k)
,
Q(k)
66 =E
2(1+
ν
)(k)
2M1
x2
1
+2
2M6
x1
x2
+
2M2
x2
2
=N1
2w
x2
1
+N2
2w
x2
2
+2N6
2w
x1
x2
,
N1
x1
+
N6
x2
=0,
N6
x1
+
N2
x2
=0
(8.2.17)
In the general case of a symmetric laminate, the plate equation can be expressed by
D11
4w
x4
1
+4D16
4w
x3
1
x2
+2(D12 +2D66)
4w
x2
1
x2
2
+4D26
4w
x1
x3
2
+D22
4w
x4
2
=N1
2w
x2
1
+N2
2w
x2
2
+2N6
2w
x1
x2
(8.2.18)
286 8 Modelling and Analysis of Plates
and for specially orthotropic laminates
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D2
4w
x4
2
=N1
2w
x2
1
+N2
2w
x2
2
+2N6
2w
x1
x2
(8.2.19)
The special case of symmetric laminates with isotropic layers follows from (8.2.19)
with D1=D2. The buckling load is like the natural vibration independent of the
lateral load and p3is taken to be zero. The classical bifurcation buckling requires to
satisfy the governing differential equations derived above and the boundary equa-
tions. Both sets of equations are again homogeneous and represent an eigenvalue
problem for the buckling modes (eigenvalues) and the mode shapes (eigenfunc-
tions).
To calculate the in-plane stress resultants N1,N2,N6it is usually convenient to
represent they by the Airy stress function F(x1,x2)
N1=
2F
x2
2
,N2=
2F
x2
1
,N6=
2F
x1
x2
(8.2.20)
If Eqs. (8.2.19) are substituted into the first two equilibrium equations (8.2.3) it is
seen that these equations are identically satisfied. Using Eq. (4.2.22)
M
M
M=B
B
BA
A
A1N
N
N(B
B
BA
A
A1B
B
BD
D
D)
κ
κ
κ
and substitute N
N
Nwith help of the Airy‘s stress function and
κ
κ
κ
by the derivatives of
wthe third equilibrium equation (8.2.3) yields one coupled partial differential equa-
tion for Fand w. The necessary second equation yields the in-plane compatibility
condition (Sect. 2.2)
2
ε
1
x2
2
+
2
ε
2
x2
1
=
2
ε
6
x1
x2
together with Eq. (4.2.25) to substitute the strains by the stress resultants. Suppress-
ing the derivations and restricting to symmetric problems yield the following in-
plane equations which are summarized in Table 8.4. The stiffness A
A
A,B
B
B,C
C
C,D
D
Dfol-
low with Eq. (4.2.23) as A
A
A=A
A
A1,B
B
B=A
A
A1B
B
B,C
C
C=B
B
BA
A
A1,D
D
D=D
D
DB
B
BA
A
A1B
B
B.
One can see from Table 8.4 that in the general case the mathematical structure of
the partial differential equation corresponds to an anisotropic and in the special or-
thotropic case to an orthotropic in-plane behavior of a single layer homogeneous
anisotropic or orthotropic plate. A summary of the in-plane stiffness is given in Ta-
ble 8.5. The Q(k)
i j for angle-ply laminates are calculated in Table 4.2.
Similar to the beam theory the plate equations for flexure, vibration and buckling
can be given in a variational formulation (Sect. 2.2). This formulation provides the
basis for the development of approximate solutions. We restrict the variational for-
mulation to symmetric laminated plates and to the classical energy principles. From
(2.2.24) it follows with
ε
3=
ε
4=
ε
50 that the elastic potential
Π
is
8.2 Classical Laminate Theory 287
Table 8.4 In-plane equations, boundary conditions and stress resultants for symmetric laminates
1. Angle-ply laminates
B
i j =0,C
i j =0,i,j=1,2,6
A
22
4F
x4
12A
26
4F
x3
1
x2
+(2A
12 +A
66)
4F
x2
1
x2
22A
16
4F
x1
x3
2
+A
11
4F
x4
2
=0
2. Cross-ply laminates
B
i j =0,A
16 =A
26 =0,
A
22
4F
x4
1
+ (2A
12 +A
66)
4F
x2
1
x2
2
+A
11
4F
x4
2
=0
or with A
11 =A
1,A
22 =A
2,(2A
12 +A
66) = A
3,
A
2
4F
x4
1
+2A
3
4F
x2
1
x2
2
+A
1
4F
x4
2
=0
3. Laminates with isotropic layers
A
1=A
2=A
3=1,
4F
x4
1
+2
4F
x2
1
x2
2
+
4F
x4
2
=0
Typical boundary conditions
Edge x1=const
2F
x2
2
=N1(x1=const,x2),
2F
x1
x2
=N6(x1=const,x2),
For an unloaded edge follow N1=0,N6=0
Stress resultants
N1=
2F
x2
2
,N2=
2F
x2
1
,N6=
2F
x1
x2
Π
=1
2Z
V
(
σ
1
ε
1+
σ
2
ε
2+
σ
6
ε
6)dVZ
A
p3(x1,x2)w(x1,x2)dA(8.2.21)
=1
2
n
k=1Z
A
x(k)
3
Z
x(k1)
3
(
σ
(k)
1
ε
1+
σ
(k)
2
ε
2+
σ
(k)
6
ε
6)dx3dAZ
A
p3(x1,x2)w(x1,x2)dA
With
ε
ε
ε
(x1,x2,x3) =
ε
ε
ε
(x1,x2) + x3
κ
κ
κ
(x1,x2)(8.2.22)
288 8 Modelling and Analysis of Plates
Table 8.5 In-plane stiffness for symmetric laminates
1. Angle-ply laminates
Ai j =
n
k=1
Q(k)
i j x(k)
3x(k1)
3=
n
k=1
Q(k)
i j h(k),i,j=1,2,6
2. Cross-ply laminates
Ai j =
n
k=1
Q(k)
i j h(k),i,j=1,2,6,A16 =A26 =0,
Q(k)
11 =E1
1
ν
12
ν
21 (k)
,Q(k)
22 =E2
1
ν
12
ν
21 (k)
Q(k)
12 =
ν
12E1
1
ν
12
ν
21 (k)
,Q(k)
66 =G(k)
12
3. Laminates with isotropic layers
Ai j =
n
k=1
Q(k)
i j h(k),i,j=1,2,6,A16 =A26 =0,A11 =A22
Q(k)
11 =Q(k)
22 =E
1
ν
2(k)
,Q(k)
12 =
ν
E
1
ν
2(k)
,
Q(k)
66 =E
2(1+
ν
)(k)
=G(k)
4. Single layer
For anisotropic and orthotropic single layers the Ai j followed by 1. and 2.
For an isotropic single layer is
A11 =A22 =A=Eh
1
ν
2,A12 =
ν
A=
ν
Eh
1
ν
2,
A66 =(1
ν
)
2A=Eh
2(1+
ν
)=G
ε
ε
ε
T=
u
x1
v
x2
u
x2
+
v
x1,
κ
κ
κ
T=
2w
x2
1
2w
x2
22
2w
x1
x2
and the constitutive equations for the strains and the stress resultants
σ
σ
σ
(k)=Q
Q
Q(k)(
ε
ε
ε
+x3
κ
κ
κ
),
N
N
N
M
M
M
=
A
A
A.
.
. 0
0
0
. . . .
0
0
0.
.
.D
D
D
ε
ε
ε
κ
κ
κ
(8.2.23)
one obtains the elastic potential for the general case of symmetric plates and for the
special cases of orthotropic or isotropic structure behavior.
8.2 Classical Laminate Theory 289
Bending of plates, classical laminate theory
:
Angle-ply laminates
Π
(w) = 1
2Z
AD11
2w
x2
12+D22
2w
x2
22+2D12
2w
x2
1
2w
x2
2
+4D66
2w
x1
x22
+4D16
2w
x2
1
+D26
2w
x2
2
2w
x1x2dAZ
A
p3wdA
(8.2.24)
Cross-ply laminates
D16 =D26 =0
Laminates with isotropic layers
D16 =D26 =0,D11 =D22
The principle of minimum of the total potential yields
δΠ
[w(x1,x2)] = 0
as the basis to derive the differential equation and boundary conditions or to apply
the direct variational methods of Ritz, Galerkin or Kantorovich for approximate
solutions.
Vibration of plates, classical laminate theory:
The kinetic energy of a plate is (rotatory energy is neglected)
T=1
2Z
A
ρ
h
w
t2
dA,
ρ
=1
h
n
k=1
ρ
(k)h(k)(8.2.25)
The Hamilton principle for vibrations yields
δ
H(x1,x2,t) = 0
with
H=
t2
Z
t1
(T
Π
)dt =
t2
Z
t1
Ldt(8.2.26)
Buckling of plates, classical laminate theory:
To calculate buckling loads, the in-plane stress resultants N1,N2,N6must be in-
cluded into the potential
Π
. These in-plane stress resultants are computed in a first
step or are known a priori. With the known N1,N2,N6the potential
Π
can be formu-
lated for angle-ply laminates with bending in-plane forces
290 8 Modelling and Analysis of Plates
Π
(w) = 1
2Z
A("D11
2w
x2
12
+D22
2w
x2
22
+2D12
2w
x2
1
2w
x2
2
+4D66
2w
x1
x22
+4D16
2w
x2
1
+D26
2w
x2
2
2w
x1x2#
"N1
w
x12
+N2
w
x22
+2N6
w
x1
w
x2#2p3w
)dA
(8.2.27)
The buckling formulation one get with p30. With D16 =D26 =0 or
D16 =D26 =0, D11 =D22 follows the equations for cross-ply laminates plates
and for plates with isotropic layers. The plate stiffness can be taken from Tables 8.2
or 8.3.
The variational principle
δΠ
=0 applied to (8.2.24) and (8.2.27) yield solutions
for bending and bending with in-plane forces. Hamilton’s principle and p36=0 is
valid to calculate forced vibrations. With p3(x1,x2,t) = 0 in vibration equations
or p3(x1,x2) = 0 in (8.2.27), we have formulated eigenvalue problems to compute
natural frequencies or buckling loads.
Summarizing the derivations of governing plate equations in the frame of classi-
cal laminate theory there are varying degrees of complexity:
An important simplification of the classical two-dimensional plate equations is
the behavior of cylindrical bending. In this case one considers a laminated plate
strip with a very high length-to-width ratio. The transverse load and all displace-
ments are functions of only x1and all derivatives with respect to x2are zero.
The laminated beams, Chap. 7, and the laminated strips under cylindrical bend-
ing are the two cases of laminated plates that can be treated as one-dimensional
problems. In Sect. 8.6 we discuss some applications of cylindrical bending
In the case of two-dimensional plate equations the first degree of simplification
for plates is to be symmetric. Symmetric laminates can be broken into cross-
ply laminates (specially orthotropic plates) with uncoupling in-plane and bend-
ing response (Bi j =0) and vanishing bending-twisting terms (D16 =D26 =0)
and angle-ply laminates (only Bi j =0). The governing equations of symmetric
cross-ply laminates have the mathematical structure of homogeneous orthotropic
plates, symmetric angle-ply laminates of homogeneous anisotropic plates. For
special boundary conditions symmetric cross-ply laminated rectangular plates
can be solved analytically. The solutions were obtained in the same manner as
for homogeneous isotropic plates, Sect. 8.6.
Laminates with all coupling effects are more complicated to analyze. Generally,
approximate analytical or numerical methods are used.
8.3 Shear Deformation Theory 291
8.3 Shear Deformation Theory
In Sect. 8.2 we have neglected the transverse shear deformations effects. The anal-
ysis and results of the classical laminate theory are sufficiently accurate for thin
plates, i.e. a/h,b/h>20. Such plates are often used in civil engineering. For mod-
erately thick plates we have to take into account the shear deformation effects, at
least approximately. The theory of laminate or sandwich plates corresponds then
with the Reissner or Mindlin1plate theory. In the Reissner-Mindlin theory the as-
sumptions of the Kirchhoffs plate theory are relaxed only in one point. The trans-
verse normals do not remain perpendicular to the middle surface after deformation,
i.e. a linear element extending through the thickness of the plate and perpendicular
to the mid-surface prior to loading, upon the load application undergoes at most a
translation and a rotation. Plate theories based upon this assumption are called first
order shear deformation theories and are most used in the analysis of moderate thick
laminated plates and of sandwich plates. Higher order theories which do not require
normals to remain straight are considerably more complicated.
Based upon that kinematical assumption of the first order shear deformation the-
ory the displacements of the plate have the form (5.1.2)
u1(x1,x2,x3) = u(x1,x2) + x3
ψ
1(x1,x2),
u2(x1,x2,x3) = v(x1,x2) + x3
ψ
2(x1,x2),
u3(x1,x2,x3) = w(x1,x2)
(8.3.1)
and with (5.1.3) are the strains
ε
i(x1,x2,x3) =
ε
i(x1,x2) + x3
κ
i(x1,x2),i=1,2,6,
ε
ε
ε
T=
u
x1
v
x2
u
x2
+
v
x1,
κ
κ
κ
T=
ψ
1
x1
ψ
2
x2
ψ
1
x2
+
ψ
2
x1,
ε
4(x1,x2) =
w
x2
+
ψ
2,
ε
5(x1,x2) =
w
x1
+
ψ
1
(8.3.2)
One can see that a constant state of transverse shear stresses is accounted for. The
stresses for the kth layer are formulated in (5.3.2) to
σ
σ
σ
(k)=Q
Q
Q(k)
ε
ε
ε
(k),
σ
T= [
σ
1
σ
2
σ
6
σ
4
σ
5],
ε
ε
ε
T= [
ε
1
ε
2
ε
6
ε
4
ε
5](8.3.3)
σ
1,
σ
2,
σ
6vary linearly and
σ
4,
σ
5constant through the thickness hof the plate. With
the stress resultants N
N
N,M
M
M,Q
Q
Qsand stiffness coefficients Ai j,Bi j ,Di j ,As
i j for laminates
or sandwiches given in Eqs. (4.2.13) - (4.2.15) or (4.3.8) - (4.3.22), respectively, the
constitutive equation can be formulated in a hypermatrix form, (4.2.16). The stiff-
ness coefficients Ai j ,Bi j,Di j stay unchanged in comparison to the classical theory
and the As
i j are defined in (5.3.4) and can be improved with the help of shear correc-
1Raymond David Mindlin (17 September 1906 New York - †22 November 1987 Hanover, New
Hempshire) - mechanician, seminal contributions to many branches of applied mechanics, applied
physics, and engineering sciences
292 8 Modelling and Analysis of Plates
tion factors ks
i j of plates similar to beams (7.3.19) (7.3.20). The definition of the
positive rotations
ψ
1,
ψ
2is illustrated in Fig. 8.3. The equilibrium equations (8.2.2)
- (8.2.3) stay unchanged.
Substituting the kinematic relations (5.3.1) into the constitutive equations (5.3.3)
and then these equations into the five equilibrium equations (8.2.2) one obtains the
governing plate equations for the shear deformation theory in a matrix form as
˜
L11 ˜
L12 ˜
L13 ˜
L14 0
˜
L21 ˜
L22 ˜
L23 ˜
L24 0
˜
L31 ˜
L32 ˜
L33 ˜
L34 ˜
L35
˜
L41 ˜
L42 ˜
L43 ˜
L44 ˜
L45
0 0 ˜
L53 ˜
L54 ˜
L55
u
v
ψ
1
ψ
2
w
=
0
0
0
0
p
(8.3.4)
The differential operators ˜
Li j are given in App. C.2 for unsymmetric angle-ply, sym-
metric angle-ply and symmetric cross-ply laminates. Symmetric laminates leading,
additional to (8.3.4), the uncoupled plate equations
˜
L11 ˜
L12
˜
L21 ˜
L22 u
v=0
0,
˜
L33 ˜
L34 ˜
L35
˜
L43 ˜
L44 ˜
L45
˜
L53 ˜
L54 ˜
L55
ψ
1
ψ
2
w
=
0
0
p
(8.3.5)
Equation (8.3.4) can also formulated in a compact matrix form
˜
L
L
L˜
u
u
u=˜
p
p
p
˜
L
L
Lis a (5×5)matrix and ˜
u
u
u,˜
p
p
pare (5×1)matrices.
The governing plate equations including transverse shear deformations are a set
of three coupled partial equations of second order, i.e. the problem is of sixth order
an for each edge of the plate three boundary conditions must be prescribed. The
most usual boundary conditions are:
fixed boundary
w=0,
ψ
n=0,
ψ
t=0
free boundary
Mn=0,Mnt =0,Qn=0
Fig. 8.3 Positive definition of
rotations
ψ
i
x2,v
x3,w
x1,u
ψ
2
ψ
1
8.3 Shear Deformation Theory 293
free edge
Mn=0,
ψ
t=0,w=0
simply supported boundary
a) w=0, Mn=0,
ψ
t=0 or w=0,
ψ
n/
n=0,
ψ
t=0 (hard hinged support)
b) w=0, Mn=0, Mnt =0
Case b) is more complicated for analytical or semianalytical solutions. Generally,
boundary conditions require prescribing for each edge one value of each of the
following five pairs: (uor Nn), (vor Nnt), (
ψ
nor Mn), (
ψ
tor Mnt), (wor Qn).
With
ψ
1=
w/
x1and
ψ
2=
w/
x2Eq. (8.3.5) can be reduced to the classical
plate equation.
In the following we restrict our development to plates that are midplane symmet-
ric (Bi j =0), and additional all coupling coefficients (...)16,(...)26,(...)45 are zero.
The constitutive equations are then simplified to
N1=A11
ε
1+A12
ε
2,N2=A12
ε
1+A22
ε
1,N6=A66
ε
6,
M1=D11
κ
1+D12
κ
2,M2=D12
κ
1+D22
κ
2,M6=D66
κ
6,
Q1=ks
55A55
ε
5,Q2=ks
44A44
ε
4
(8.3.6)
or in a contracted notation
N
N
N
M
M
M=A
A
A0
0
0
0
0
0D
D
D
ε
ε
ε
κ
κ
κ
,Q
Q
Qs=A
A
As
ε
ε
ε
s,
N
N
NT= [N1N2N6],M
M
MT= [M1M2M6],Q
Q
QsT= [Q1Q2],
ε
ε
ε
T= [
ε
1
ε
2
ε
6],
κ
T= [
κ
1
κ
2
κ
6],
ε
ε
ε
sT= [
ε
5
ε
4],
A
A
A=
A11 A12 0
A12 A22 0
0 0 A66
,D
D
D=
D11 D12 0
D12 D22 0
0 0 D66
,A
A
As=ks
55A55 0
0ks
44A44
(8.3.7)
Substituting the constitutive equations for M1,M2,M6,Q1,Q2into the three equilib-
rium equations (8.2.2) of the moments and transverse force resultants results in the
following set of governing differential equations for a laminated composite plate
subjected to a lateral load p3(x1,x2)and including transverse shear deformation
D11
2
ψ
1
x2
1
+(D12 +D66)
2
ψ
2
x1
x2
+D66
2
ψ
1
x2
2ks
55A55
ψ
1+
w
x1=0,
D66
2
ψ
2
x2
1
+(D12 +D66)
2
ψ
1
x1
x2
+D22
2
ψ
2
x2
2ks
44A44
ψ
2+
w
x2=0,
ks
55A55
ψ
1
x1
+
2w
x2
1+ks
44A44
ψ
2
x2
+
2w
x2
2+p3(x1,x2) = 0
(8.3.8)
Analogous to the classical plate equations the shear deformation theory can be used
for modelling and analysis of forced vibrations and buckling of laminate plates. In
294 8 Modelling and Analysis of Plates
the general case of forced vibrations the displacements u,v,w, the rotations
ψ
1,
ψ
2
and the transverse load pin Eq. (8.3.4) are functions of x1,x2and t. In-plane loading
is not considered but in-plane displacements, rotary and coupling inertia terms have
to take into account. Therefore, generalized mass densities must be defined
ρ
0=
n
k=1
ρ
(k)x(k)
3x(k1)
3=
n
k=1
ρ
(k)h(k),
ρ
1=
n
k=1
ρ
(k)x(k)
3
2x(k1)
3
2,
ρ
2=
n
k=1
ρ
(k)x(k)
3
3x(k1)
3
3
(8.3.9)
Coupling inertia terms
ρ
1are only contained in unsymmetric plate problems.
If one wishes to determine the natural frequencies of the rectangular plate con-
sidered above, then in (8.3.8) p3(x1,x2)must be set zero but a term
ρ
0
2w/
t2
must be added on the right hand side. In addition, because
ψ
1and
ψ
2are both in-
dependent variables which are independent of the transverse displacement w, there
will be an oscillatory motion of a line element through the plate thickness which re-
sults in rotary inertia terms
ρ
2
2
ψ
1/
t2and
ρ
2
2
ψ
2/
t2, respectively, on the right
hand side of the first two equations of (8.3.8).
The governing equations for the calculation of natural frequencies of specially
orthotropic plates with A45 =0 are
D11
2
ψ
1
x2
1
+ (D12 +D66)
2
ψ
2
x1
x2
+D66
2
ψ
1
x2
2
ks
55A55
ψ
1+
w
x1=
ρ
2
2
ψ
1
t2,
D66
2
ψ
2
x2
1
+ (D12 +D66)
2
ψ
1
x1
x2
+D22
2
ψ
2
x2
2
ks
44A44
ψ
2+
w
x2=
ρ
2
2
ψ
2
t2,
ks
55A55
ψ
1
x2
+
2w
x2
1+ks
44A44
ψ
2
x2
+
2w
x2
2=
ρ
0
2w
t2,
(8.3.10)
ρ
0=
n
k=1
ρ
(k)(x(k)
3x(k1)
3) =
n
k=1
ρ
(k)h(k),
ρ
2=1
3
n
k=1
ρ
(k)(x(k)3
3x(k1)3
3)
w,
ψ
1and
ψ
2are functions of x1,x2and t.
In a similar way the governing equations for buckling problems can be derived.
In the matrix equations (8.3.4) and (8.3.5) only the differential operator ˜
L55 is sub-
stituted by
8.3 Shear Deformation Theory 295
˜
L55 N1
2
x2
1
+2N6
2
x1
x2
+N2
2
x2
2(8.3.11)
For a cross-ply symmetrically laminated plate is with Bi j =0,D16 =0, D26 =0,
A45 =0
D11
2
ψ
1
x2
1
+ (D12 +D66)
2
ψ
2
x1
x2
+D66
2
ψ
1
x2
2ks
55A55
ψ
1+
w
x1=0,
D66
2
ψ
2
x2
1
+ (D12 +D66)
2
ψ
1
x1
x2
+D22
2
ψ
2
x2
2ks
44A44
ψ
2+
w
x2=0,
ks
55A55
ψ
1
x2
+
2w
x2
1+ks
44A44
ψ
2
x2
+
2w
x2
2
=N1
2w
x2
1
+2N6
2w
x1
x2
+N2
2w
x2
2
(8.3.12)
The variational formulation of laminated plates including shear deformation may
be based for example upon the principle of minimum potential energy for static
problems and the Hamilton’s principle for dynamic problems. Formulating the elas-
tic potential
Π
we have to consider that in the general case of unsymmetric laminate
plates, including shear deformation,
Π
=
Π
(u,v,w,
ψ
1,
ψ
2)is a potential function of
five independent variables and that the strain energy
Π
ihas a membrane, a bending
and a transverse shearing term, i.e.
Π
i=1
2Z
V
(
σ
1
ε
1+
σ
2
ε
2+
σ
6
ε
6+
σ
5
ε
5+
σ
4
ε
4)dV
=
Π
m
i+
Π
b
i+
Π
s
i
(8.3.13)
with
Π
m
i=1
2Z
V
(N1
ε
1+N2
ε
2+N6
ε
6)dA,
Π
b
i=1
2Z
V
(M1
κ
1+M2
κ
2+M6
κ
6)dA,
Π
s
i=1
2Z
V
(Qs
1
ε
5+Qs
2
ε
4)dA
(8.3.14)
The stress resultants, stiffness and constitutive equations are formulated in Sect. 4.2,
e.g. (4.2.10) - (4.2.17). The elastic potential
Π
is then given by
Π
(u,v,w,
ψ
1,
ψ
2) = 1
2Z
A
(
ε
ε
ε
TA
A
A
ε
ε
ε
+
κ
κ
κ
TB
B
B
ε
ε
ε
+
ε
ε
ε
TB
B
B
κ
κ
κ
+
κ
κ
κ
TD
D
D
κ
κ
κ
+
ε
ε
ε
sTA
A
As
ε
ε
ε
s)dx1dx2Z
A
p3wdx1dx2
(8.3.15)
296 8 Modelling and Analysis of Plates
In (8.3.15) the in-plane loads p1,p2are not included and must be added in gen-
eral loading cases. Shear correction coefficients can be developed for plates quite
similar to beams. Approximately one considers a laminate strip of the width ”1”
orthogonal to the x1-direction and independently another laminate strip orthogonal
to the x2-direction and calculates the correction factors ks
55 and ks
44 like in Chap. 7
for beams. Sometimes the shear correction factors were used approximately equal
to homogeneous plates, i.e, ks
44 =ks
45 =ks
55 =ks=5/6.
Mostly we have symmetric laminates and the variational formulation for bending
Mindlin’s plates can be simplified
Π
(w,
ψ
1,
ψ
2) = 1
2Z
A
(
κ
TD
D
D
κ
κ
κ
+
ε
ε
ε
sTA
A
As
ε
s)dx1dx2Z
A
p3wdx1dx2(8.3.16)
If we restricted the Hamiltons principle to vibration of symmetric plates, the varia-
tional formulation yields
L(w,
ψ
1,
ψ
2) = T(w,
ψ
1,
ψ
2)
Π
(w,
ψ
1,
ψ
2),
T(w,
ψ
1,
ψ
2) = 1
2Z
A"
ρ
0
w
t2
+
ρ
2
ψ
1
t2
+
ρ
2
ψ
2
t2#dx1dx2
(8.3.17)
Π
is given by (8.3.16) and
ρ
0,
ρ
2by Eqs. (8.3.10), Tis the kinetic energy.
For a symmetric and specially orthotropic Mindlin’s plate assuming A45 =0 it
follows from (8.3.16) for bending problems that
Π
(w,
ψ
1,
ψ
2) = 1
2Z
A"D11
ψ
1
x12
+2D12
ψ
1
x1
ψ
2
x2+D22
ψ
2
x22
+D66
ψ
1
x2
+
ψ
2
x12
+ks
55A55
ψ
1+
w
x12
+ks
44A44
ψ
2+
w
x22#dx1dx2Z
A
p3wdx1dx2,
(8.3.18)
δΠ
(w,
ψ
1,
ψ
2) = 0
For natural vibration the variational formulation for that plate is
L(w,
ψ
1,
ψ
2) = T(w,
ψ
1,
ψ
2)
Π
(w,
ψ
1,
ψ
2),
δ
t2
Z
t1
L(w,
ψ
1,
ψ
2)dt=0 (8.3.19)
To calculate buckling loads the in-plane stress resultants must be, like in the Kirch-
hoff’s plate theory, included into the part
Π
aof
Π
. Consider a plate with a constant
in-plane force N1it follows
8.4 Sandwich Plates 297
Π
(w,
ψ
1,
ψ
2) =
Π
i(w,
ψ
1,
ψ
2)1
2ZN1
w
x12
dx1dx2(8.3.20)
The case of a more general in-plane loading can be transposed from (8.2.27). The
term
Π
istay unchanged.
8.4 Sandwich Plates
To formulate the governing differential equations or tht variational statement for
sandwich plates we draw the conclusion from the similarity of the elastic behavior
between laminates and sandwiches in the first order shear deformation theory that
all results derived above for laminates can be applied to sandwich plates. We restrict
our considerations to symmetric sandwich plates with thin or thick cover sheets.
Like in the beam theory, there are differences in the expressions for the flexural
stiffness D11,D12 ,D22,D66 and the transverse shear stiffness A55,A44 of laminates
and sandwiches (Sects. 4.3.2 and 4.3.3). Furthermore there are essential differences
in the stress distributions. The elastic behavior of sandwiches and the general model
assumptions are considered in detail in Sect. 4.3. The stiffness relations for sand-
wiches with thin and thick skins are also given there:
Symmetric sandwiches with thin cover sheets (4.3.12) - (4.3.14)
Ai j =2Af
i j =2
n
k=1
Q(k)
i j h(k),
Di j =hcCf
i j =hcn
k=1
Q(k)
i j h(k)¯x(k)
3,¯x(k)
3=1
2x(k)
3+x(k1)
3
(i j) = (11),(12),(22),(66)
As
i j =hcCs
i j,(i j ) = (44),(55),Cs
44 =Gs
23,Cs
55 =Gs
13
(8.4.1)
hcis the thickness of the core, nis the number of faces layers and G13 ,G23 are the
core shear stiffness moduli. Shear correction factors can be calculated similarly
to the beams approximately with the help of (7.4.2).
298 8 Modelling and Analysis of Plates
Symmetric sandwiches with thick cover sheets (4.3.16) - (4.3.17) of one lamina
Ai j =ASa
i j =2hfQf
i j or Ai j =ALa
i j =2hfQf
i j +hcQc
i j,i,j=1,2,6
Di j =DSa
i j =1
2Qf
i j hf+hchfhcor
Di j =DLa
i j =1
2hfQf
i j hf+hc2+1
3hf2+1
12 hc3Qc
i j,i,j=1,2,6
(i j) = (11),(12),(22),(66)
As
i j =As
i j
Sa =hcCc
i j,As
i j =As
i j
La =2hfCf
i j +hcCc
i j,(i j ) = (44),(55)
Cf
44 =Gf
23,Cc
44 =Gc
23,Cf
55 =Gf
13,Cc
55 =Gc
13
(8.4.2)
With these stiffness values for the two types of sandwich plates the differential equa-
tions (8.3.8), (8.3.10) or the variational formulations (8.3.18) - (8.3.20) of the theory
of laminate plates including transverse shear deformation can be transposed to sand-
wich plates.
Equation (4.3.22) demonstrated that for sandwich plates with thick faces the stiff-
ness ALa
i j ,DLa
i j ,(i j) = (11),(22),(66)and ALa
i j ,(i j) = (44),(55)should be used. Be-
cause generally Qc
i j Qf
i j usually the simplified stiffness
ALa
i j ASa
i j =2Af
i j,DLa
i j DSa
i j 1+hf
hc
hc+ (4/3)hf
hc+hf
yield satisfying results in engineering applications. Thus is valid for isotropic-facing
sandwich plates and for sandwich plates having orthotropic composite material fac-
ings (cross-ply laminates).
In Sect. 4.3 generally and in Sect. 7.4 for beams the continuing popularity of
sandwich structures was underlined. Sect. 7.4 also recalled and summarized the
main aspects of modelling and analysis of sandwich structures. Engineering ap-
plications to sandwich beams were discussed in detail. Keeping this in mind, the
derivations to sandwich plates can be restricted here to few conclusions:
Most sandwich structures can be modelled and analyzed using the shear defor-
mation theory for laminated plates.
Generally, the stiffness matrices A
A
A,B
B
Band D
D
Dof laminated plates are employed.
Consider the lower face as lamina 1, the core as lamina 2 and the upper face
as lamina 3 one can include or ignore the effect of the core on the response to
bending and in-plane loads and the effect of transverse shear deformation on the
response of the facings.
The shear deformation theory of laminated plates can be not only transposed to
sandwich plates for bending, vibration and buckling induced by mechanical loads
but include also other loading, e.g. hydrothermal effects.
With the special sandwich stiffness including or ignore in-plane, bending and trans-
verse shear deformation response all differential equations and variational formula-
tions of Sect. 8.3 stay valid. Some examples for sandwich plates are considered in
Sect. 8.7.
8.5 Hygrothermo-Elastic Effects on Plates 299
8.5 Hygrothermo-Elastic Effects on Plates
Elevated temperature and absorbed moisture can alter significantly the structural
response of fibre-reinforced laminated composites. In Sects. 8.2 to 8.4 the structural
response of laminated plates as result of mechanical loading was considered and
thermal or hygrosgopic loadings were neglected.
This section focuses on hygrothermally induced strains, stresses and displace-
ments of thin or moderate thick laminated plates. We assume as in Sect. 7.5 mod-
erate hygrothermal loadings such that the mechanical properties remain approx-
imately unchanged for the temperature and moisture differences considered. Be-
cause the mathematical formulations governing thermal and hygroscopic loadings
are analogous, a unified derivation is straightforward and will be considered in the
frame of the classical laminate theory and the shear deformation theory.
The following derivations use the basic equations, Sect. 4.2.5, on thermal and
hygroscopic effects in individual laminae and in general laminates. The matrix for-
mulations for force and moment resultants, Eq. (4.2.75), can be written explicitly
as
N1
N2
N6
M1
M2
M6
=
A11 A12 A16 B11 B12 B16
A22 A26 B12 B22 B26
A66 B16 B26 B66
SD11 D12 D16
YD22 D26
MD66
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6
Nth
1
Nth
2
Nth
6
Mth
1
Mth
2
Mth
6
Nmo
1
Nmo
2
Nmo
6
Mmo
1
Mmo
2
Mmo
6
(8.5.1)
with the known matrix elements, Eq. (4.2.15),
Ai j =
n
k=1
Q(k)
i j x(k)
3x(k1)
3,
Bi j =1
2
n
k=1
Q(k)
i j x(k)
3
2x(k1)
3
2,
Di j =1
3
n
k=1
Q(k)
i j x(k)
3
3x(k1)
3
3
(8.5.2)
The thermal and moisture stress resultants N
N
Nth,N
N
Nmo,M
M
Mth,M
M
Mmo are resultants per unit
temperature or moisture change, Eqs. (4.2.67).
Substituting the hygrothermal constitutive equation (8.5.1) into the equilibrium
equations (8.2.3) and replacing the in-plane strains
ε
iand the curvatures
κ
iby the
displacements u,v,w, Eq. (8.2.5), yield the following matrix differential equation
for the classical laminate theory
L
L
Lu
u
u=p
p
p
N
N
N
M
M
M(8.5.3)
300 8 Modelling and Analysis of Plates
L
L
Lu
u
u=p
p
pis identically to Eq. (8.2.7) with Li j given in App. C. N
N
N=N
N
Nth +N
N
Nmo,
M
M
M=M
M
Mth +M
M
Mmo are the hygrothermal stress results and
is a special (3 ×6)
differential matrix
=
x1
x2
0 0 0 0
0
x2
x1
0 0 0
0 0 0
2
x2
1
2
x2
22
2
x1
x2
(8.5.4)
For selected layer stacking Eq. (8.5.3) can be simplified. The matrix L
L
Land the dif-
ferential operators Li j are summarized for the most important special laminates in
App. C.
Hygrothermal induced buckling can be modelled as
Lu
Lu
Lu +
N
N
N
M
M
M=N1
2
x2
1
+2N6
2
x1
x2
+N2
2
x2
2u
u
u(8.5.5)
with u
u
uT= [0 0 w]. Prebuckling displacements and stress resultants are determined
by solving Eq. (8.5.5) with N
N
N0
0
0. For the corresponding buckling problem, N1,N2
and N6are taken to be the stress resultant functions corresponding to the prebuckling
state. The buckling loads are found by solving the eigenvalue problem associated
with (8.5.5), i.e. with N
N
N=0
0
0 and M
M
M=0
0
0.
Because energy methods are useful to obtain approximate analytical solutions
for hygrothermal problems the total potential energy
Π
is formulated. Restricting
to symmetrical problems with A16 =A26 =0 and D16 =D26 =0, i.e to cross-ply
laminates, we have
Π
(u,v,w) = 1
2Z
A(A11
u
x12
+2A12
u
x1
v
x2+A22
v
x22
+A66
u
x2
+
v
x12
+D11
2w
x2
12
+2D12
2w
x2
1
2w
x2
2
+D22
2w
x2
22
+4D66
2w
x1
x22
2N
1
u
x12N
2
v
x22N
6
u
x2
+
v
x1
+2M
1
2w
x2
1
+2M
2
2w
x2
24M
6
2w
x1
x2
"N1
2w
x2
12
+2N6
w
x1
w
x2+N2
2w
x2
22#)dA
(8.5.6)
The classical laminate theory which neglect transverse shear deformations can lead
to significant errors for moderately thick plates and hygrothermal loadings. Us-
8.5 Hygrothermo-Elastic Effects on Plates 301
ing the shear deformation theory, Sect. 8.3, we can formulate corresponding to Eq.
(8.3.4).
˜
L
L
L˜
u
u
u=˜
p
p
p˜
N
N
N
M
M
M(8.5.7)
The matrix ˜
is identically to Eq. (8.5.4). For hygrothermal induced buckling we
have analogous to Eq. (8.5.5)
˜
L
L
L˜
u
u
u+˜
N
N
N
M
M
M=N1
2
x2
1
+2N6
2
x1
x2
+N2
2
x2
2u
u
u(8.5.8)
For prebuckling analysis the terms involving N1,N2and N6are ignored. Then buck-
ling loads are calculated by substituting the values N1,N2,N6determined for the
prebuckling state into Eq. (8.5.8) dropping now the hygrothermal stress resultants
N
N
Nand M
M
M. For special laminate stacking the differential operators are summarized
in App. C.
The elastic potential
Π
is now a function of five independent functions
u,v,w,
ψ
1,
ψ
2. Restricting again to cross-ply laminates the elastic total potential
Π
can be formulated as
Π
=1
2Z
A(A11
u
x12
+2A12
u
x1
v
x2+A22
v
x22
+A66
u
x2
+
v
x12
+D11
ψ
1
x12
+2D12
ψ
1
x1
ψ
2
x2+D22
ψ
2
x22
+D66
ψ
1
x2
+
ψ
2
x1
+ks
44A44 dw
dx2
+
ψ
22
+ks
55A55 dw
dx1
+
ψ
12
(8.5.9)
2N
1
u
x12N
2
v
x22N
6
u
x2
+
v
x1
2M
1
ψ
1
x12M
2
ψ
2
x22M
6
ψ
1
x2
+
ψ
2
x1
"N1
2w
x2
12
+2N6
w
x1
w
x2+N2
2w
x2
22#)dA
Equations (8.5.6) and (8.5.9) are the starting point for solving hygrothermal induced
buckling problems e.g. with the Ritz- or Galerkin approximation or the finite ele-
ment method. Analytical solutions are in general not possible. As considered above,
the force resultants N1,N2and N6have to be calculated in the prebuckling state,
i.e. for N
N
N0
0
0 and the calculation force resultants are substituted into Eqs. (8.5.6) or
(8.5.9), respectively, with N
N
N=0
0
0,M
M
M=0
0
0 to calculate the buckling loads. If there are
transverse loads p, Eqs. (8.5.3) or (8.5.7) the bending problem follows from (8.5.6)
or (8.5.7) by setting N
N
N=0
0
0 and substitute an additional term
302 8 Modelling and Analysis of Plates
Z
A
pwdA
8.6 Analytical Solutions
The analysis of rectangular plates with selected layer stacking and boundary condi-
tions can be carried out analytically in a similar manner to homogeneous isotropic
and orthotropic plates. The analytical methods of homogeneous isotropic plates, e.g.
the double series solutions of Navier2or the single series solutions of adai3-L´evy4
can be applied to laminated plates with special layer stacking and analogous bound-
ary conditions. In Sect. 8.6 possibilities of analytical solutions in the frame of clas-
sical laminate theory and shear deformation theory are demonstrated for bending,
buckling and vibration problems.
8.6.1 Classical Laminate Theory
There are varying degrees of complexity in laminated plate analysis. The least com-
plicated problems are one-dimensional formulations of cylindrical plate bending.
For cylindrical bending both, symmetric and unsymmetric laminates, are handled in
a unique manner assuming all deformations are one-dimensional.
In the case of two-dimensional plate equations the most important degree of sim-
plification is for plates being midplane symmetric because of their uncoupling in-
plane and out-of-plane response. The mathematical structure of symmetric angle-ply
plate equations corresponds to homogeneous anisotropic plate equations and that of
symmetric cross-ply plate equations to homogeneous orthotropic plate equations.
To illustrate analytical solutions for rectangular plates in the framework of the clas-
sical laminate theory we restrict our developments to specially orthotropic, i.e. to
symmetric cross-ply plates. For this type of laminated plates the Navier solution
method can be applied to rectangular plates with all four edges simply supported.
The N´adai-L ´evy solution (N ´adai, 1925) method can be applied to rectangular plates
with two opposite edges have any possible kind of boundary conditions. For more
general boundary conditions of special orthotropic plates or other symmetric or un-
symmetric rectangular plates approximate analytical solutions are possible, e.g. us-
ing the Ritz-, the Galerkin- or the Kantorovich methods, Sect. 2.2.3, or numerical
methods are applied, Chap. 11.
2Claude Louis Marie Henri Navier (10 February 1785 Dijon - †21 August 1836 Paris) - engineer
and physicist
3´
Arp´ad N´adai (3 April 1883 Budapest - †18 July 1963 Pittsburgh) - professor of mechanics,
contributions to the plate theory and theory of plasticity
4Maurice evy (28 February 1838 Ribeauvill´e - †30 September 1910 Paris) - engineer, total
strain theory
8.6 Analytical Solutions 303
As considered above the simplest problem of plate bending is the so-called cylin-
drical bending for a plate strip i.e. a very long plate in one direction with such a lat-
eral load and edge support in this direction that the plate problem may be reduced to
a one-dimensional problem and a quasi-beam solution can be used. In the following
we demonstrate analytical solutions for various selected examples.
8.6.1.1 Plate Strip
The model ”plate strip (Fig. 8.4) describes approximately the behavior of a rect-
angular plate with a/b1. The plate dimension ain x1-direction is considered
finite, the other dimension bin x2-direction approximately infinite. The boundary
conditions for the edges x1=0,x1=amay be quite general, but independent of x2
and the lateral load is p3=p3(x1). All derivatives with respect to x2are zero and
the plate equation reduces to a one-dimensional equation. For symmetric laminated
strips Eqs. (8.2.6) and (8.2.9) reduce to
D11w′′′′(x1) = p3(x1),
M1(x1) = D11w′′(x1),
M2(x1) = D12w′′(x1),
M6(x1) = D16w′′(x1)(general case),
M6(x1) = 0 (specially orthotropic case,D16 =0),
Q1(x1) = M
1(x1) = D11w′′′(x1),
Q2(x1) = M
6(x1) = D16w′′′(x1)(general case),
Q2(x1) = 0 (specially orthotropic case)
(8.6.1)
When one compares the differential equation of the strip with the differential equa-
tion bD11w′′′′(x1) = q(x1)of a beam it can be stated that all solutions of the beam
equation can be used for the strip.
For the normal stresses in the layer kthe equations are
Fig. 8.4 Plate strip
x2
x1
a
p3FM
304 8 Modelling and Analysis of Plates
σ
(k)
1(x1,x3) = Q(k)
11 x3
d2w
dx2
1
,
σ
(k)
2(x1,x3) = Q(k)
12 x3
d2w
dx2
1
,
σ
(k)
6(x1,x3) = 0
or
σ
(k)
1(x1,x3) = Q(k)
11
D11
M1(x1)x3,
σ
(k)
2(x1,x3) = Q(k)
12
D12
M2(x1)x3,
σ
(k)
6(x1,x3) = 0
(8.6.2)
and the transverse shear stresses follow from (5.2.19) to
σ
4(x1,x3) = F61 Q1(x1),
σ
5(x1,x3) = F11 Q1(x1)
and
F
F
F(x3) = ˜
B
B
B(x3)D
D
D1
i.e
σ
5
σ
4=F11 F62
F61 F22 Q1
0,(8.6.3)
Considering the solutions of the symmetrical laminated plate strip, we have follow-
ing conclusions:
The solutions for laminate beams and plate strips are very similar, but the calcu-
lation of the strip bending stiffness Di j has to include Poisson’s effects.
Because of including of Poisson’s effect we have the relation
w(x1)strip <w(x1)beam
and M2(x1)6=0.
M1(x1)and Q1(x1)of the strip and the beam are identical. If M6(x1) = 0 then
V1Q1, i.e there is no special effective Kirchhoff transverse force.
The solutions for plate strips with cylindrical bending can be transposed to
lateral loads p3(x1,x2) = x2p(x1). From w(x1,x2) = x2w(x1)it follows that
x2w′′′′(x1) = x2p(x1)/D11 . The displacement w(x1)and the stress resultants
M1(x1),Q1(x)of the plate strip with the lateral load p(x1)have to be multiplied
by the coordinate x2to get the solution for the lateral load x2p(x1). Note that in
contrast to the case above, here M6=2D66w(x1)D16w′′(x1)in the general case
and M6=2D66w(x1)for specially orthotropic strips.
For unsymmetric laminated plate strips the system of three one-dimensional dif-
ferential equations for the displacements u(x1),v(x1)and w(x1)follow with (8.2.6)
as
8.6 Analytical Solutions 305
A11
d2u
dx2
1
+A16
d2v
dx2
1B11
d3w
dx3
1
=0
A16
d2u
dx2
1
+A66
d2v
dx2
1B16
d3w
dx3
1
=0
D11
d4w
dx4
1B11
d3u
dx3
1B16
d3v
dx3
1
=p3
(8.6.4)
These equations can be uncoupled and analytically solved. The first and the second
equation yield
d2u
dx2
1
=B
A
d3w
dx3
1
,d2v
dx2
1
=C
A
d3w
dx3
1
(8.6.5)
with A=A11A66 A2
16,B=A66B11 A16B16,C=A11 B16 A16B11. Differentiating
both equations and substituting the results in the third equation of (8.6.4) we obtain
one differential equation of fourth order in w(x1)
d4w
dx4
1
=A
Dp3,D=D11AB11BB16C(8.6.6)
Equation (8.6.6) can be integrated to obtain w(x1)and than follow with (8.6.5)
d3u
dx3
1
=B
Dp3,d3v
dx3
1
=C
Dp3(8.6.7)
For a transverse load p3=p3(x1)we obtain the analytical solutions for the displace-
ments u(x1),v(x1)and w(x1)as
w(x1) = A
DZZZ Z p3dx1dx1dx1dx1+C1
x3
1
6+C2
x2
1
2+C3x1+C4,
u(x1) = B
DZZZp3dx1dx1dx1+B1
x2
1
2+B2x1+B3,
v(x1) = C
DZZZp3dx1dx1dx1+A1
x2
1
2+A2x1+B3
(8.6.8)
With Eqs. (8.6.5) follows A1=B1=C1and we have eight boundary conditions to
calculate 8 unknown constants, e.g. for clamped supports
u(0) = u(a) = v(0) = v(a) = w(b) = w(a) = 0,w(0) = w(a) = 0,
Eqs. (8.2.12) yield the one-dimensional equations for the forces and moments resul-
tants
N1=A11
du
dx1
+A16
dv
dx1B11
d2w
dx2
1
,
N2=A12
du
dx1
+A26
dv
dx1B12
d2w
dx2
1
,
306 8 Modelling and Analysis of Plates
N6=A16
du
dx1
+A66
dv
dx1B16
d2w
dx2
1
,
M1=B11
du
dx1
+B16
dv
dx1D11
d2w
dx2
1
,
M2=B12
du
dx1
+B26
dv
dx1D12
d2w
dx2
1
,
M6=B16
du
dx1
+B66
dv
dx1D16
d2w
dx2
1
,(8.6.9)
Q1=B11
d2u
dx2
1
+B16
d2v
dx2
1D11
d3w
dx3
1
,
Q2=B16
d2u
dx2
1
+B66
d2v
dx2
1D16
d3w
dx3
1
,
V1=B11
d2u
dx2
1
+B16
d2v
dx2
1D11
d3w
dx3
1
=Q1,
V2=2B16
d2u
dx2
1
+2B66
d2v
dx2
12D16
d3w
dx3
1
=2Q2
The general symmetric case follows with Bi j =0 and for a symmetric cross-ply strip
are Bi j =0 and A16 =0,D16 =0.
Analytical solutions can also be formulated for vibration and buckling of strips
with one-dimensional deformations. The eigen-vibrations of unsymmetrical plate
strips taking account of u(x1,t) = u(x1)ei
ω
t,v(x1,t) = v(x1)ei
ω
t,w(x1,t) = w(x1)ei
ω
t
are mathematically modelled as
A11
d2u
dx2
1
+A16
d2v
dx2
1B11
d3w
dx3
1
ρ
h
ω
2w=0,
A16
d2u
dx2
1
+A66
d2v
dx2
1B16
d3w
dx3
1
ρ
h
ω
2v=0,
D11
d4w
dx4
1B11
d3u
dx3
1B16
d3w
dx3
1
ρ
h
ω
2w=0
(8.6.10)
u,vand ware now functions of x1and t.
If the in-plane inertia effects are neglected the Eqs. (8.6.5) are valid. Differenti-
ating these equations and substituting the result in the third Eq. (8.6.10) lead to the
vibration equation
d4w
dx1A
D
ρ
h
ω
2w=0 (8.6.11)
For a symmetrically laminated cross-ply strip we obtain with A/D=1/D11
d4w
dx1
ρ
h
ω
2
D11
w=0,
ρ
h=
N
k=1
ρ
(k)h(k)(8.6.12)
8.6 Analytical Solutions 307
The analytical solutions correspond to the beam solutions in Sect. 7.6
w(x1) = C1cos
λ
ax1+C2sin
λ
ax1+C3cosh
λ
ax1+C4sinh
λ
ax1,
λ
a4
=
ρ
h
D11
ω
2(symmetric cross-ply strip),
λ
a4
=
ρ
hA
D
ω
2(general unsymmetric strip)
(8.6.13)
For a simply supported strip we have w(0) = w(a) = w′′(0) = w′′(a) = 0 and there-
fore C1=C3=C4=0 and C2sin
λ
aa=0, i.e. with
λ
=n
π
follow
ω
2=n4
π
4
a4
ρ
h
D
A,
ω
=n
π
a2sD
ρ
hA (8.6.14)
Analytical solutions can be calculated for all boundary conditions of the strip.
In analogous manner analytical solutions follow for the buckling behavior of
strips which are subjected to an initial compressive load N1=N0. The third equa-
tion of (8.6.4) is formulated with p3=0 as
D11
d4w
dx4
1B11
d3u
dx3
1B16
d3v
dx3
1N1
d2w
dx2
1
=0 (8.6.15)
and with Eq. (8.6.5) follows
d4w
dx4
1A
DN1
d2w
dx2
1
=0,(general case)
d4w
dx4
11
D11
N1
d2w
dx2
1
=0,(symmetrical cross-ply case)
(8.6.16)
with N1(x1) = N0. The buckling equations correspond again to the beam equation
(7.2.35) and can be solved for all boundary conditions of the strip
w(x1) = C1cos
λ
ax1+C2sin
λ
ax1+C3cosh
λ
ax1+C4sinh
λ
ax1,
λ
a2
=A
DN0
(8.6.17)
For a simply supported strip we have with w(0) = w(a) = w′′(0) = w′′(a) = 0
C2sin
λ
=0,
λ
=n
π
A nonzero solution is obtained if
N0=n2
π
2
a2
D
A(8.6.18)
308 8 Modelling and Analysis of Plates
Thus the critical buckling load follows with to
N0cr =
π
2
a2
D
A(8.6.19)
Summarizing the developments of analytical solutions for unsymmetrical laminated
plate strips we have the following conclusions:
The system of three coupled differential equations for the displacements
u(x1),v(x1)and w(x1)can be uncoupled and reduced to one differential equa-
tion of fourth order for w(x1)and two differential equations of third order for
u(x1)and v(x1), respectively.
Analytical solutions for bending of unsymmetrical laminated plate strips can be
simple derived for all possible boundary conditions. In the general case all stress
resultants (8.6.4) are not equal to zero. The general symmetric case and symmet-
rical cross-ply strips are included as special solutions.
The derivations of bending equations can be expanded to buckling and vibration
problems.
The derivation of analytical solutions for unsymmetrical laminated strips can,
like for the symmetrical case, expanded to lateral loads p3(x1,x2) = x2p(x1).
8.6.1.2 Navier Solution
Figure 8.5 shows a specially orthotropic rectangular plate simply supported at all
edges with arbitrary lateral load p3(x1,x2). In the Navier solution one expands the
deflection w(x1,x2)and the applied lateral load p(x1,x2), respectively, into double
infinite Fourier sine series because that series satisfies all boundary conditions
p3(x1,x2)
x2
x1
a
bBoundary conditions:
w(0,x2) = w(a,x2) = w(x1,0) = w(x1,b) = 0
M1(0,x2) = M1(a,x2) = M2(x1,0) = M2(x1,b) = 0
Fig. 8.5 Rectangular plate, all edges are simply supported, specially orthotropic
8.6 Analytical Solutions 309
p3(x1,x2) =
r=1
s=1
prs sin
α
rx1sin
β
sx2,
prs =4
ab
a
Z
0
b
Z
0
p3(x1,x2)sin
α
rx1sin
β
sx2dx1dx2,
w(x1,x2) =
r=1
s=1
wrs sin
α
rx1sin
β
sx2
(8.6.20)
with
α
r=r
π
/a,
β
s=s
π
/b. The coefficients wrs are to be determined such that the
plate equation (Table 8.1) is satisfied.
Substituting Eqs. (8.6.20) into the plate equation yields
r=1
s=1
wrs D1
α
4
r+2D3
α
2
r
β
2
s+D2
β
4
ssin
α
rx1sin
β
sx2=
r=1
s=1
prs sin
α
rx1sin
β
sx2
(8.6.21)
and we obtain the coefficients wrs
wrs =prs
D1
α
4
r+2D3
α
2
r
β
2
s+D2
β
4
s
=prs
drs
(8.6.22)
The solution becomes
w(x1,x2) =
r=1
s=1
prs
drs
sin
α
rx1sin
β
sx2(8.6.23)
The load coefficients prs one obtains by integrating (8.6.20) for the given lateral
loading p3(x1,x2). For a uniform distributed load p3(x1,x2) = p=const we obtain,
for instance,
prs =16p
π
2rs ,r,s=1,3,5,... (8.6.24)
From Table 8.1, the equations for the moment resultants are:
M1(x1,x2) = D11
2w
x2
1D12
2w
x2
2
=
r=1
s=1
(D11
α
2
r+D12
β
2
s)wrs sin
α
rx1sin
β
sx2,
M2(x1,x2) = D12
2w
x2
1D22
2w
x2
2
=
r=1
s=1
(D12
α
2
r+D22
β
2
s)wrs sin
α
rx1sin
β
sx2,
(8.6.25)
310 8 Modelling and Analysis of Plates
M6(x1,x2) = 2D66
2w
x1
x2
=2D66
r=1
s=1
α
r
β
swrs cos
α
rx1cos
β
sx2,
Q1(x1,x2) =
M1
x1
+
M6
x2
,V1(x1,x2) = Q1+
M6
x2
,
Q2(x1,x2) =
M6
x1
+
M2
x2
,V1(x1,x2) = Q2+
M6
x2
,
and with the stress relation for the klayers
σ
σ
σ
(k)(x1,x2,x3) = Q
Q
Q(k)x3
κ
κ
κ
=x3
Q(k)
11 Q(k)
12 0
Q(k)
12 Q(k)
22 0
0 0 Q(k)
66
2w/
x2
1
2w/
x2
2
2
2w/
x1
x2
(8.6.26)
one obtains the solutions for the in-plane stresses
σ
(k)
1,
σ
(k)
2,
σ
(k)
6
σ
(k)
1
σ
(k)
2
σ
(k)
6
=
r=1
s=1
prs
D1
α
4
r+2D3
α
2
r
β
2
s+D2
β
4
s
(Q(k)
11
α
2
r+Q(k)
12
β
2
s)sin
α
rx1sin
β
sx2
(Q(k)
12
α
2
r+Q(k)
22
β
2
s)sin
α
rx1sin
β
sx2
2Q(k)
66
α
r
β
scos
α
rx1cos
β
sx2
(8.6.27)
With the simplified formula (5.2.19) follows the transverse shear stresses
σ
(k)
4,
σ
(k)
5
"
σ
(k)
5(x1,x3)
σ
(k)
4(x1,x3)#=F11 x3F62 x3
F61 x3F22 x3
r=1
s=1(D11
α
3
r+D12
β
2
s
α
r)wrs cos
α
rx1sin
β
sx2
(D12
α
2
r
β
s+D22
β
3
s)wrs sin
α
rx1cos
β
sx2(8.6.28)
The Navier solution method can be applied to all simply supported specially or-
thotropic laminated rectangular plates in the same way. For a given lateral load
p3(x1,x2)one can obtain the load coefficients prs by integrating (8.6.20), and by
substituting prs in (8.6.3) follows the wrs. Some conclusions can be drawn from the
application of the Navier solution:
The solution convergence is rapid for the lateral deflection w(x1,x2)and uniform
loaded plates. The convergence decreases for the stress resultants and the stresses
and in general with the concentration of lateral loads in partial regions.
The solution convergence is more rapid for the stresses
σ
(k)
1in the fibre direction
but is not as rapid in calculating
σ
(k)
2.
The Navier solutions can be also developed for antisymmetric cross-ply laminate
and for symmetric and antisymmetric angle-ply laminates. For these laminates the
plate equations (8.2.6) are not uncoupled and we have to prescribe in-plane and
out-of-plane boundary conditions. It is easy to review that the Naviers double series
solutions Type 1 and Type 2, i.e.
8.6 Analytical Solutions 311
Type 1:
u(x1,x2) =
r=1
s=1
urs cos
α
rx1sin
β
sx2,
v(x1,x2) =
r=1
s=1
vrs sin
α
rx1cos
β
sx2,
w(x1,x2) =
r=1
s=1
wrs sin
α
rx1sin
β
sx2,
Type 2:
u(x1,x2) =
r=1
s=1
urs sin
α
rx1cos
β
sx2,
v(x1,x2) =
r=1
s=1
vrs cos
α
rx1sin
β
sx2,
w(x1,x2) =
r=1
s=1
wrs sin
α
rx1sin
β
sx2,
α
r=
π
r/a,
β
s=
π
s/asatisfy the following alternative boundary conditions for se-
lected laminated plates:
Simply supported boundary conditions, Type 1
x1=0 and x1=a
w=0,M1=0,v=0,N1=0
x2=0 and x2=b
w=0,M2=0,u=0,N2=0
The Naviers double series Type 1 for u,vand wcan be used only for laminates,
whose stiffness A16,A26 ,B16,B26,D16,D26 are zero, i.e for symmetric or anti-
symmetric cross-ply laminates
Simple supported boundary conditions, Type 2
x1=0 and x1=a
w=0,M1=0,u=0,N6=0
x2=0 and x2=b
w=0,M2=0,v=0,N6=0
The Navier double series solution Type 2 for u,v,wcan be used only for laminate
stacking sequences with A16,A26,B11 ,B12,B22,B66,D16,D26 equal zero, i.e for
symmetric or antisymmetric angle ply laminates.
The Navier solutions can be used for calculating bending, buckling and vibration.
For buckling the edge shear force N6and, respectively, for vibration the in-plane
inertia terms must be necessarily zero.
312 8 Modelling and Analysis of Plates
8.6.1.3 N´
adai-L´
evy Solution
For computing the bending of specially orthotropic rectangular plates with two op-
posite edges simply supported, a single infinite series method can be used. The two
other opposite edges may have arbitrary boundary conditions (Fig. 8.6). N´adai in-
troduced for isotropic plates the solution of the plate equation in the form
w(x1,x2) = wp(x1) + wh(x1,x2),p3=p3(x1),(8.6.29)
where wp(x1)represents the deflection of a plate strip and wh(x1,x2)is the solution
of the homogeneous plate equation ( p3=0). whmust be chosen such that w(x1,x2)
in (8.6.29) satisfy all boundary conditions of the plate. With the solutions for wh,
suggested by L´evy, and wp, suggested by N ´adai,
wh(x1,x2) =
r=1
fr(x2)sin
α
rx1,wp(x1) =
r=1
prsin
α
rx1
D1
α
4
r
(8.6.30)
with
α
r=r
π
/aand
p3(x1) =
r=1
prsin
α
rx1,pr=2
a
a
Z
0
p3(x1)sin
α
rx1dx1
the boundary conditions for x1=0 and x1=aare satisfied.
Substituting (8.6.30) into the plate equation for specially orthotropic plates, Table
8.1, it follow for each term fr(x2)a differential equation of 4th order with constant
coefficients
D2
d4fr(x2)
dx4
22D3
α
2
r
d2fr(x2)
dx2
2
+D1
α
4
rfr(x2) = pr(8.6.31)
or
x2
x1
a
bBoundary conditions:
w(0,x2) = w(a,x2) = 0
M1(0,x2) = M1(a,x2) = 0
For the edges x2=±b/2 may be
arbitrary b.c.
Fig. 8.6 Rectangular specially orthotropic rectangular plate with two opposite edges simply sup-
ported
8.6 Analytical Solutions 313
d4fr(x2)
dx4
22D3
α
2
r
D2
d2fr(x2)
dx2
2
+D1
D2
α
4
rfr(x2) = pr
D2
(8.6.32)
The homogeneous differential equation, i.e. pr=0, can be solved with
frh(x2) = Crexp(
λ
r
α
rx2)(8.6.33)
and yields the characteristic equation for the four roots
λ
4
r2D3
D2
λ
2
r+D1
D2
=0=
λ
2
r=D3
D2±sD3
D22
D1
D2
(8.6.34)
In the case of isotropic plates it follows with D1=D2=D3=Dthere are repeated
roots ±1.
For specially orthotropic laminated plates the form of frh(x2)depends on the
character of the roots of the algebraic equation of 4th order. There are three different
sets of roots:
1. (D3/D2)2>(D1/D2): In this case (8.6.34) leads to four real and different roots
λ
1/2=±
δ
1,
λ
3/4=±
δ
2,
δ
1,
δ
2>0,
frh(x2) = Arcosh
δ
1
α
rx2+Brsinh
δ
1
α
rx2
+Crcosh
δ
2
α
rx2+Drsinh
δ
2
α
rx2
(8.6.35)
2. (D3/D2)2= (D1/D2): In this case (8.6.34) leads to four real and equal roots
λ
1/2= +
δ
,
λ
3/4=
δ
,
δ
>0,
frh(x2) = (Ar+Brx2)cosh
δα
rx2+ (Cr+Drx2)sinh
δα
rx2(8.6.36)
3. (D3/D2)2<(D1/D2): In this case the roots are complex
λ
1/2=
δ
1±i
δ
2,
λ
3/4=
δ
1±i
δ
2,
δ
1,
δ
2>0,
frh(x2) = (Arcos
δ
2
α
rx2+Brsinh
δ
2
α
rx2)cosh
δ
1
α
rx2
+ (Crcos
δ
1
α
rx2+Drsin
δ
1
α
rx2)sinh
δ
1
α
rx2
(8.6.37)
For a given plate for which materials and fibre orientations have been specified only,
one of the three cases exists. However in the design problem, trying to find the best
variant, more than one case may be involved with the consequence of determin-
ing not just four constants Ar,Br,Cr,Dr, but eight or all twelve to calculate which
construction is optimal for the design.
Concerning the particular solution, it is noted that the lateral load may be at most
linear in x2too, i.e p3(x1,x2) = p3(x1)q(x2)with qat most linear in x2. The solution
wpin (8.6.29) is then replaced by
wp(x1,x2) = q(x2)
r=1
prsin
α
rx1
D1
α
4
r
sin
α
rx1(8.6.38)
314 8 Modelling and Analysis of Plates
With the solution w(x1,x2) = wh(x1,x2) + wp(x1,x2)the stress resultants and
stresses can be calculated in the usual way.
The Navier and N´adai-evy solution method can be also applied to eigenvalue
problems. We assume, for instance, that the vibration mode shapes of a laminated
plate with specially orthotropic behavior, which is simply supported at all four
edges, is identical to an isotropic plate. We choose
w(x1,x2,t) =
r=1
s=1
wrs sin
α
rx1sin
α
sx2sin
ω
t(8.6.39)
to represent the expec ted harmonic oscillation and to satisfy all boundary cond itions.
Substituting the expression (8.6.39) into (8.2.16) with p30 yields
[D1
α
4
r+2D3
α
2
r
α
2
s+D2
α
4
s
ρω
2]wrs =0 (8.6.40)
A non-zero value of wrs, i.e. a non-trivial solution, is obtained only if the expression
in the brackets is zero, hence we can find the equation for the natural frequencies
ω
2
rs =
π
4
ρ
hD1r
a4+2D3r
a2s
a2+D2s
a4(8.6.41)
The fundamental frequency corresponds to r=s=1 and is given by
ω
2
11 =
π
4
ρ
ha4D1+2D3a
b2+D2a
b4(8.6.42)
Note that the maximum amplitude wrs cannot be determined, only the vibration
mode shapes are given by (8.6.39). In the case of an isotropic plate the natural
frequencies are with D1=D2=D3=D
ω
2
rs =krs
π
2
a2sD
ρ
h,krs =r2+s2a
b2(8.6.43)
If we consider a buckling problem, e.g. a specially orthotropic laminated plate sim-
ply supported at all edges with a biaxial compression N1and N2, it follows from
(8.2.19) that
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D2
4w
x4
2
=N1
2w
x2
1
+N2
2w
x2
2
(8.6.44)
The Navier solution method yields with (8.6.39)
π
2wrs[D1r4+2D3r2s2
γ
2+D2s4
γ
4] = wrs[N1r2+N2s2
γ
2]a2(8.6.45)
with
γ
=a/b. A non-zero solution of the buckling problem (wrs 6=0) leads to
8.6 Analytical Solutions 315
N1r2+N2s2
γ
2=
π
2
a2[D1r4+2D3r2s2
γ
2+D2s4
γ
4](8.6.46)
We consider the example of uniform compression N1=Nand N2=
κ
N, where
the boundary force Nis positive. Equation (8.6.46) yields
N=
π
2(D1r4+2D3r2s2
γ
2+D2s4
γ
4)
a2(r2+
κ
s2
γ
2)
The critical buckling load Ncr corresponds to the lowest value of N. If
κ
=0 we
have the case of uniaxial compression and the buckling equation simplifies to
N=
π
2
a2r2(D1r4+2D3r2s2
γ
2+D2s4
γ
4)
For a given r, the smallest value of Nis obtained for s=1, because sappears only
in the numerator. To determine which rprovides the smallest value Ncr is not simple
and depends on the stiffness D1,D2,D3, the length-to-width ratio
γ
=a/band r.
However, for a given plate it can be easily determined numerically. Summarizing
the discussion of the classical laminate theory applied to laminate plates we can
formulate the following conclusions:
Specially orthotropic laminate plates can be analyzed with the help of the Navier
solution or the adai-L´evy solution of the theory of isotropic Kirchhoff’s plates,
if all or two opposite plate edges are simply supported. These solution methods
can be applied to plate bending, buckling and vibration.
For more general boundary conditions specially orthotropic plates may be solved
analytically with the help of the variational approximate solutions method of
Rayleigh-Ritz or in a more generalized way based on a variational method of
Kantorovich.
Plates with extensional-bending couplings should be solved numerically, e.g.
with the help of the finite element method, Chap. 11. Note that in special cases
antisymmetric cross-ply respectively symmetric and antisymmetric angle-ply
laminates can be analyzed analytically with Navier‘s solution method.
In this section we illustrated detailed analytical solutions for specially orthotropic
laminates which can predict ”exact” values of deflections, natural frequencies of vi-
bration and critical buckling loads. But even the ”exact” solutions become approxi-
mate because of the truncation of the infinite series solutions or round-off errors in
the solution of nonlinear algebraic equations, etc. However these solutions help one
to understand, at least qualitatively, the mechanical behavior of laminates. Many
laminates with certain fibre orientations have decreasing values of the coefficients
D16,D26 for bending-torsion coupling and they can be analyzed with the help of the
solution methods for specially orthotropic plates.
316 8 Modelling and Analysis of Plates
8.6.2 Shear Deformation Laminate Theory
The analysis of laminated rectangular plates including transverse shear deforma-
tions is much more complicated than in the frame of classical laminate theory. Also
for plate analysis including shear deformations the at least complicated problem is
cylindrical bending, i.e one-dimensional formulations for plate strips.
Unlike to classical plate strips equations only symmetric and unsymmetric cross-
ply laminates can be handled in a unique manner. In the case of two-dimensional
plate equations we restrict the developments of analytical solutions for bending,
buckling and vibrations analogous to Eqs. (8.3.6) - (8.3.8) to midplane symmetric
cross-ply plates with all Bi j =0 and additional A16 =A26 =D16 =D26 =0,A45 =0.
8.6.2.1 Plate Strip
Consider first the cylindrical bending for the plate strip with an infinite length in
the x2-direction and uniformly supported edges x1=0,x1=a, subjected to a load
p3=p(x1). If we restrict the considerations to cross-ply laminated strips the gov-
erning strip equations follow with A16 =A26 =0,B16 =B26 =0, D16 =D26 =0,
A45 =0 and result in a cylindrical deflected middle surface with v=0,
ψ
2=0,
u=u(x1),
ψ
1=
ψ
1(x1),w=w(x1)from (8.3.4) as
A11
d2u
dx2
1
+B11
d2
ψ
1
dx2
1
=0,
B11
d2u
dx2
1
+D11
d2
ψ
1
dx2
1ks
55A55
ψ
1+dw
dx1=0,
ks
55A55 d
ψ
1
dx1
+d2w
dx2
1+p3(x1) = 0
(8.6.47)
The stress resultants Ni(x)1,Mi(x1),i=1,2,6 and Qj,j=1,2 are with (8.3.2) and
(8.3.6)
N1(x1) = A11
du
dx1
+B11
d
ψ
1
dx1
,
N2(x1) = A12
du
dx1
+B12
d
ψ
1
dx1
,
N6(x1) = 0,
M1(x1) = B11
du
dx1
+D11
d
ψ
1
dx1
,
M2(x1) = B12
du
dx1
+D12
d
ψ
1
dx1
,
M6(x1) = 0,
Q1(x1) = ks
55A55
ψ
1+dw
dx1,
Q2(x1) = 0
(8.6.48)
8.6 Analytical Solutions 317
The three coupled differential equations for u,wand
ψ
1can be reduced to one un-
coupled differential equation for
ψ
1. The first equation yields
d2u
dx2
1
=B11
A11
d2
ψ
1
dx2
1
,d3u
dx3
1
=B11
A11
d3
ψ
1
dx3
1
(8.6.49)
Differentiating the second equation and substituting the equation above result in
B2
11
A11
d3
ψ
1
dx3
1
+D11
d3
ψ
1
dx3
1ks
55A55 d
ψ
1
dx1
+d2w
dx2
1=0
or with D11 B2
11
A11 =DR
11
ks
55A55 d
ψ
1
dx1
+d2w
dx2
1=DR
11
d3
ψ
1
dx3
1
(8.6.50)
Substituting Eq. (8.6.50) in the third equation (8.6.47) yield an uncoupled equation
for
ψ
1(x1)
DR
11
d3
ψ
1
dx3
1
=p3(8.6.51)
The uncoupled equations for u(x1)and w(x1)follow then as
d2u
dx2
1
=A11
B11
d2
ψ
1
dx2
1
,dw
dx1
=
ψ
1+DR
11
ks
55A55
d2
ψ
1
dx2
1
(8.6.52)
The three uncoupled equations can be simple integrated
DR
11
ψ
1(x1) = ZZZp3(x1)dx1dx1dx1+C1
x2
1
2+C2x1+C3,
w(x1) = 1
DR
11 "ZZZ Z p3(x1)dx1dx1dx1dx1+C1
x3
1
6+C2
x2
1
2
+C3x1+C4#1
ks
55A55 ZZ p3(x1)dx1dx1+C1x1
=wB(x1) + wS(x1),
u(x1) = A11
B11
1
DR
11 ZZZp3(x1)dx1dx1dx1+C1x1+C5
(8.6.53)
Thus the general analytical solutions for unsymmetric cross-ply laminated strips
are calculated. For symmetrical cross-ply laminated strips the equations yield
DR
11 =D11 and A11u′′(x1) = 0. Restricting to symmetrical cross-ply laminated
strips analytical solutions for buckling or vibrations can be developed analogous to
Timoshenko’s beams or to the classical strip problems.
For a buckling load N1(x1) = N0follow with p3=0
318 8 Modelling and Analysis of Plates
D11
d2
ψ
1
dx2
1ks
55A55
ψ
1+dw
dx1=0,
ks
55A55 d
ψ
1
dx1
+d2w
dx2
1+N0
d2w
dx1
=0
(8.6.54)
The equations can be uncoupled. With
d
ψ
1
dx1
+d2w
dx2
1=D11
ks
55A55
d3
ψ
1
dx3
1
,ks
55A55
d3
ψ
1
dx3
1
=ks
55A55
d4w
dx4
1N0
d2w
dx1
one obtains analogous to Eq. (7.3.23)
D11 1N0
ks
55A55 d4w
dx4
1
+N0
d2w
dx1
=0 (8.6.55)
The general solution for the eigenvalue problem (8.6.55) follows with
w(x1) = Ce
λ
x1(8.6.56)
and the characteristic equation
D11 1N0
ks
55A55
λ
4+N0
λ
2=0 or D11
λ
4+k2
λ
2=0 (8.6.57)
with the solutions
λ
1/2=±ik,
λ
3/4=0
as
w(x1) = C1sin kx1+C2coskx1+C3x1+C4(8.6.58)
If we assume, e.g. simply supported edges x1(0) = 0,x1(a), follow with
w(0) = w(a) = w′′(0) = w′′(a) = 0 the free coefficients C2=C3=C4=0
and C1sinka =0. If C16=0 follow with sin ka =0 the solution k=m
π
/a=
α
m
(m=1,2,...)and k2=
α
2
mand thus
N0
D11 1N0
ks
55A55 =
α
2
m,N0=D11ks
55A55
α
2
m
D11
α
2
m+ks
55A55
The critical buckling load corresponds to the smallest value of N0which is obtained
for m=1
Ncr =D11ks
55A55
π
2
D11
π
2+ks
55A55a2=
π
2D11
a2
1
1+
π
2D11
a2ks
55A55
(8.6.59)
It can be seen that analogous to the Timoshenko’s beam, Sect. 7.3, the including of
shear deformations decreases the buckling loads.
8.6 Analytical Solutions 319
The free vibrations equations of the Timoshenko’s beams were also considered
in Sect. 7.3. For symmetric cross-ply laminated plate strips we obtain comparable
equations
D11
d2
ψ
1
dx2
1ks
55A55
ψ
1+dw
dx1=
ρ
2
2
ψ
t2,
ks
55A55 d
ψ
1
dx1
+d2w
dx2
1=
ρ
0
2w
t2
(8.6.60)
ρ
0and
ρ
2were defined as
ρ
0=
n
k=1
ρ
(k)h(k),
ρ
2=1
3
n
k=1
ρ
(k)x(k)
3
3x(k1)
3
3
and the terms involving
ρ
0and
ρ
2are the translatory and the rotatory inertia terms.
ψ
1and ware functions of x1and tand thus we have partial derivatives. If we assume
again both strip edges simply supported the analytical solution follow with
w(x1,t) = C1mei
ω
mtsin m
π
x1
a,w(0,t) = w(a,t) = 0,
ψ
1(x1,t) = C2mei
ω
mtcos m
π
x1
a,
ψ
1(0,t)
x1
=
ψ
1(a,t)
x1
=0
(8.6.61)
Substituting these solution functions into the vibration equations (8.6.60) follow
D11
α
2
m+ks
55A55
ρ
2
ω
2
mks
55A55
α
m
ks
55A55
α
mks
55A55
α
2
m
ρ
0
ω
2
mC2m
C1m=0
0
The nontrivial solution of the homogeneous algebraic equation yields the eigenfre-
quencies
ω
m
D11
α
2
m+ks
55A55
ρ
2
ω
2
mks
55A55
α
m
ks
55A55
α
mks
55A55
α
2
m
ρ
0
ω
2
m
=0 (8.6.62)
or
ρ
0
ρ
2
ω
4
m(D11
ρ
0
α
m+ks
55A55
ρ
0+ks
55A55
ρ
2
α
2
m)2
ω
2
m+D11ks
55A55
α
4
m=0
A
ω
4
mB
ω
2
m+C=0,
ω
2
m=B
2A±1
2ApB24A2C
The general solution for the vibration equations can be formulated for arbitrary
boundary conditions. For harmonic oscillations we write
w(x1,t) = w(x1)ei
ω
t,
ψ
1(x1,t) =
ψ
1(x1)ei
ω
t(8.6.63)
Substituting w(x1,t)and
ψ
1(x1,t)in the coupled partial differential equations
(8.6.60) yield
320 8 Modelling and Analysis of Plates
D11
d2
ψ
1(x1)
dx2
1ks
55A55
ψ
1+dw(x1)
dx1+
ρ
2
ω
2
ψ
(x1) = 0,
ks
55A55 d
ψ
1(x1)
dx1
+d2w(x1)
dx2
1+
ρ
0
ω
2w(x1) = 0
(8.6.64)
These both equations can be uncoupled. With
ks
55A55
d
ψ
1(x1)
dx1
=
ρ
0
ω
2w(x1)ks
55A55
d2w(x1)
dx1
and
D11
d3
ψ
1(x1)
dx3
1ks
55A55
d2w(x1)
dx2
1ks
55A55
d
ψ
1(x1)
dx1
ρ
2
ω
2d
ψ
1(x1)
dx1
=0
follow
D11
d4w(x1)
dx4
1
+D11
ρ
0
ks
55A55
+
ρ
2
ω
2d2w(x1)
dx2
11
ρ
2
ω
2
ks
55A55
ρ
0
ω
2w(x1) = 0
or
ad4w(x1)
dx4
1
+bd2w(x1)
dx2
1cw(x1) = 0 (8.6.65)
The general solution can be derived as
w(x1) = C1sin
λ
1x1+C2cos
λ
2x1+C3sinh
λ
3x1+C4cosh
λ
4x1(8.6.66)
The
λ
iare the roots of the characteristic algebraic equation of (8.6.65). The deriva-
tions above demonstrated that for any boundary conditions an analytical solution
is possible. Unlike to the classical theory we restricted the considerations in the
frame of the shear deformation theory to cross-ply laminated strips. Summarizing
the derivations we can draw the following conclusions:
Cylindrical bending yields simple analytical solutions for unsymmetrical and
symmetrical cross-ply laminated plate strips.
Restricting to symmetrical laminated cross-ply plate strips we can obtain ana-
lytical solutions for buckling and vibrations problems, but for general boundary
conditions the analytical solution can be with difficulty.
8.6.2.2 Navier Solution
Navier’s double series solution can be used also in the frame of the shear deforma-
tions plate theory. Analogous to Sect. 8.6.1 double series solutions can be obtain for
symmetric and antisymmetric cross-ply and angle-ply laminates with special types
of simply supported boundary conditions. In the interest of brevity the discussion
8.6 Analytical Solutions 321
is limited here to symmetrical laminated cross-ply plates, i.e. specially orthotropic
plates. The in-plane and out-of-plane displacements are then uncoupled.
Rectangular specially orthotropic plates may be simply supported (hard hinged
support) on all four edges.
x1=0,x1=a:w=0,M1=0 respectively
ψ
1
x1
=0,
ψ
2=0,
x2=0,x2=b:w=0,M2=0 respectively
ψ
2
x2
=0,
ψ
1=0
(8.6.67)
The boundary conditions can be satisfied by the following expressions:
w(x1,x2) =
r=1
s=1
wrs sin
α
rx1sin
β
sx2,
ψ
1(x1,x2) =
r=1
s=1
ψ
1rs cos
α
rx1sin
β
sx2,
α
r=r
π
a,
β
s=s
π
b,
ψ
2(x1,x2) =
r=1
s=1
ψ
2rs sin
α
rx1cos
β
sx2
(8.6.68)
The mechanical loading p3(x1,x2)can be also expanded in double Fourier sine se-
ries
p3(x1,x2) =
r=1
s=1
prs sin
α
rx1sin
β
sx2,
prs =4
ab
a
Z
0
b
Z
0
p3(x1,x2)sin
α
rx1sin
β
sx2dx1dx2
(8.6.69)
Now the Navier solution method can be extended to Mindlin’s plates with all edges
simply supported, but the solution is more complex than for Kirchhoff’s plates. Sub-
stituting the expression (8.6.68) and (8.6.69) into the plate differential equations
(8.3.8) gives
L11 L12 L13
L12 L22 L23
L13 L23 L33
ψ
1rs
ψ
2rs
wrs
=
0
0
prs
(8.6.70)
with
L11 =D11
α
2
r+D66
β
2
s+ks
55A55,L12 = (D12 +D66)
α
r
β
s,L13 =ks
55A55
α
r,
L22 =D66
α
2
r+D22
β
2
s+ks
44A44,L33 =ks
55A55
α
2
r+ks
44A44
β
2
s,L33 =ks
44A44
β
s
(8.6.71)
Solving the Eqs. (8.6.65), one obtains
ψ
1rs =L12L23 L22 L13
Det(Li j )prs ,
ψ
2rs =L12L13 L11 L23
Det(Li j)prs,wrs =L11L22 L2
12
Det(Li j)prs
(8.6.72)
Det(Li j)is the determinant of the matrix in (8.6.65).
322 8 Modelling and Analysis of Plates
If the three kinematic values w(x1,x2),
ψ
1(x1,x2),
ψ
2(x1,x2)are calculated the
curvatures
κ
1,
κ
2and
κ
6may be obtained and the stresses in each lamina follow
from (8.3.3) to
σ
1
σ
2
σ
6
(k)
=x3
Q11 Q12 0
Q12 Q22 0
0 0 Q66
(k)
κ
1
κ
2
κ
6
,
σ
5
σ
4(k)
=C55 0
0C44 (k)
ψ
1+
w
x1
ψ
2+
w
x2
(8.6.73)
In a analogous manner natural vibrations and buckling loads can be calculated for
rectangular plates with all edges hard hinged supported.
8.6.2.3 N´
adai-L´
evy Solution
The N´adai-evy solution method can also be used to develop analytical solutions for
rectangular plates with special layer stacking and boundary conditions, respectively,
but the solution procedure is more complicated than in the frame of classical plate
theory. We do without detailed considerations and recommend approximate analyt-
ical solutions or numerical methods to analyze the behavior of general laminated
rectangular plates including shear deformations and supported by any combination
of clamped, hinged or free edges.
Summarizing the discussion of analytical solutions for plates including trans-
verse shear deformations one can formulate following conclusion
Analytical solutions for symmetrical and unsymmetrical laminated plates can be
derived for cylindrical bending, buckling and vibration.
Navier’s double series solutions can be simple derived for specially orthotropic
plates. Navier’s solution method can be also applied to symmetric or antisym-
metric cross-ply and angle-ply laminates, but the solution time needed is rather
high.
Ritz’s, Galerkin’s or Kantorovich’s methods are suited to analyze general lami-
nated rectangular plates with general boundary conditions.
Plates with general geometry or with cut outs etc. should be analyzed by numer-
ical methods
8.7 Problems
Exercise 8.1. A plate strip has the width ain x1-direction and is infinitely long in
the x2-direction. The strip is loaded transversely by a uniformly distributed load
p0and simply supported at x1=0,x1=a. Calculate the deflection w, the resultant
moments M1,M2,M6and the stresses
σ
1,
σ
2,
σ
6
8.7 Problems 323
1. for a symmetrical four layer plate [0/90/90/0],
2. for a unsymmetrical four layer plate [0/0/90/90]
Solution 8.1. The solutions are presented for both stacking sequences separately.
1. The plate strip is a symmetric cross-ply laminate, i.e. Bi j =0, D16 =D26 =0.
The governing differential equations are
d2M1
dx2
1
=p0,dM1
dx1
=Q1,
D11
d2w
dx2
1
=M1,D12
d2w
dx2
1
=M2,M6=0,
D11
d4w
dx4
1
=p0
The vertical deflection w=w(x1)is
w(x1) = 1
D11 p0
x4
1
24 +C1
x3
1
6+C2
x2
1
2+C3x1+C4
Satisfying the boundary conditions
w(0) = 0,w(a) = 0,M1(0) = 0,M1(a) = 0
yield the unknown constants C1-C4
C1=q0a
2,C2=0,C3=q0a3
24 ,C4=0
and as result the complete solution for the deflection w(x1)
w(x1) = p0a4
24D11 x1
a42x1
a3+x1
a
The moment resultants follow as
M1(x1) = p0a2
2x1
a2x1
a,
M2(x1) = D12
D11
M1(x1) = D12
D11
p0a2
2x1
a2x1
a,
M6(x1) = 0
The strains and stresses at any point can be determined as follow
κ
1=d2w
dx1
=p0a2
2x1
ax1
a2,
κ
2=0,
κ
6=0,
ε
1=x3
κ
1=x3
p0a2
2x1
ax1
a2,
ε
2=0,
ε
6=0
324 8 Modelling and Analysis of Plates
The stresses in each layer are
00layers :
σ
1=
σ
1=x3
Q11
D11
p0a2
8x1
ax1
a2,
σ
2=
σ
2=x3
Q12
D11
p0a2
8x1
ax1
a2,
σ
6=0,
900layers :
σ
2=
σ
1=x3
Q11
D11
p0a2
8x1
ax1
a2,
σ
1=
σ
2=x3
Q12
D11
p0a2
8x1
ax1
a2,
σ
6=0
2. The plate strip is an unsymmetric cross-ply laminate, i.e. A16 =A26 =0,B16 =
B26 =0,D16 =D26 =0. The governing equations follow from Eqs. (8.2.6) and
(8.2.12)
A11
d2u
dx2
1B11
d3w
dx3
1
=0,A66
d2v
dx2
1
=0,D11
d4w
dx4
1B11
d3u
dx3
1
=p3,
N1=A11
du
dx1B11
d2w
dx2
1
,N2=A12
du
dx1B12
d2w
dx2
1
,N6=0,
M1=B11
du
dx1D11
d2w
dx2
1
,M2=B12
du
dx1D12
d2w
dx2
1
,M6=0
The equilibrium equations for the stress resultants are
dN1
dx1
=0,d2M1
dx2
1
=p3
The displacement u(x1)and w(x1)are coupled. Substitute
A11
d3u
dx3
1
=B11
d4w
dx4
1
into the second differential equation yield
D11 B2
11
A11 d4w
dx4
=p3
or with
D11 1B2
11
A11D11 =DR
11,p3=p0d4w
dx4
1
=p0
DR
11
For the displacement u(x1)follows
8.7 Problems 325
d3u
dx3
1
=B11
A11
d4w
dx4
1
or with
A11 1B2
11
A11D11 =AR
11,d4w
dx4
1
=p0
DR
11 d3u
dx3
1
=B11
D11
p0
AR
11
These differential equations can be simple integrated
w(x1) = 1
DR
11 q0x4
1
24 +C1
x3
1
6+C2
x2
1
2+C3x1+C4
u(x1) = B11
D11AR
11 q0x3
1
6+C1
x2
1
2+C5x1+C6
Note that with
d2u
dx2
1
=B11
A11
d3w
dx3
1
in both equations there are equal constants C1. The boundary conditions for w
and M1are identically to case 1.
The in-plane boundary conditions are formulated for a fixed-free support, i.e.
u(0) = 0,N1(a) = 0. The boundary conditions lead to the six unknown constants
C1-C6and the solution functions are
u(x1) = B11
D11AR
11
p0a3
12 2x1
a3x1
a2
w(x1) = 1
DR
11
p0a4
24 x1
a42x1
a3+x1
a
The stress and moment resultants follow as
N1(x1) = N6(x1) = 0,
N2(x1) = A12B11
D11AR
11 B12
DR
11 p0a2
2x1
a2x1
a,
M1(x1) = p0a2
2x1
a2x1
a,
M2(x1) = B2
11 D12A11
A11 p0a2
2DR
11 x1
a2x1
a,
M6(x1) = 0
It is interesting to compare the results of case 1. and case 2. The forms of w(x1)
are for the two cases identical except for the magnitude. With
326 8 Modelling and Analysis of Plates
1
DR
11
=1
D11 1B2
11
A11D11 >1
D11
the deflection of the unsymmetric laminate strip will be greater than the deflec-
tion of the symmetric laminate. Note that there is no force resultant N1(x1)in the
unsymmetric case but it is very interesting that there is a force resultant N2as a
function of x1, but N2(0) = N2(a) = 0. With
ε
1=du
dx1
=B11
D11AR
11
p0a2
2x1
a2x1
a,
ε
2=
ε
6=0,
κ
1=d2
dx2
1
=p0a2
2DR
11 x1
a2x1
a,
κ
2=
κ
6=0
follow the strains
ε
1,
ε
2and the stresses
σ
1,
σ
2for the 00and 900-layers in a
similar manner like case 1. With B11 =B12 =0 case 2. yields the symmetrical
case 1.
Exercise 8.2. A plate strip of the width awith a symmetrical cross-ply stacking is
subjected a downward line load q0at x1=a/2. Both edges of the strip are fixed.
Calculate the maximum deflection wmax using the shear deformation theory.
Solution 8.2. With (8.6.51) and (8.6.52) follow
DR
11 D11,D11
d3
ψ
1
dx3
1
=q0
δ
x11
2a,dw
dx1
=
ψ
1+D11
ks
55A55
d2
ψ
1
dx2
1
with
δ
x11
2a=0x16=a/2
1x1=a/2,Z
δ
x11
2adx1=x11
2a0
<x1e>is F¨oppel’s5bracket symbol:
<x1e>n=0x1<e
(x1e)nx1>e,
d
dx1
<x1e>n=n<x1e>n1,
Z<x1e>ndx1=1
1+n<x1e>n+1+C
With (8.6.53) the analytical solutions for
ψ
1and ware given
5August Otto F¨oppl (25 January 1854 Groß-Umstadt †12 August 1924 Ammerlan) - professor
of engineering mechanics and graphical statics
8.7 Problems 327
D11
ψ
1(x1) = 1
2q0<x11
2a>2+C1
x2
1
2+C2x1+C3,
w(x1) = 1
D11 1
6q0<x11
2a>3+C1
x3
1
6+C2
x2
1
2+C3x1+C4
1
ks
55A55
[q0<x11
2a>+C1x1]
ψ
1(0) = 0 : C3=0,
w(0) = 0 : C4=0,
ψ
1(a) = 0 : 1
2q01
2a2
+C1a2
2+C2a=0,
w(a) = 0 : 1
6q01
2a3
+C1a3
6+C2
a2
2D11
ks
55A55 q0
a
2+C1a=0,
C1=q0
2,C2=q0a
8,
ψ
1(x1) = q0a2
8D11 x1
a2x1
a,
w(x1) = q0a3
48D11 3x1
a24x1
a3q0a
2ks
55A55
x1
a=wB+wS,
wmax =q0a3
192D11
+q0a
4ks
55A55
The classical plate theory yields with ks
55A55 the known value
wmax =q0a3
192D11
Exercise 8.3 (Bending of a quadratic sandwich plate). A quadratic sandwich
plate has a symmetric cross-section. The plate properties are a=b=1 m,
hf=0,2875 103m, hc=24,71 103m, Ef=1,42 105MPa,
ν
f=0,3,
Gf=Ef/2(1+
ν
f),Gc=22 MPa. The cover sheet and the core material are
isotropic, hfhc. The transverse uniform distributed load is p=0,05 MPa. The
boundary conditions are hard hinged support for all boundaries. Calculate the max-
imum flexural displacement wmax with the help of a one-term Ritz approximation.
Solution 8.3. The elastic potential
Π
(w,
ψ
1,
ψ
2)of a symmetric and special or-
thotropic Mindlin’s plate is given by (8.3.18). For stiff thin cover sheets and a core
which transmits only transverse shear stresses the bending and shear stiffness for
isotropic face and core materials are (8.4.1)
Di j =hcC(f)
i j =hchQf
i jh(f)¯x(f)
3i=hchf1
2hc+hfQ(f)
i j ,
((i j) = (11),(22),(66),(12)) with
Q11 =Ef
1(
ν
f)2=Q22,Q12 =
ν
fEf
1(
ν
f)2,Q66 =Gf
328 8 Modelling and Analysis of Plates
and
AS
i j =hcCc
i j =hcGc,(i j) = (44),(55)
Π
(w,
ψ
1,
ψ
2) = Z
A(1
2hchf(hc+hf)"Q11
ψ
1
x12
+2Q12
ψ
1
x1
ψ
2
x2
+Q22
ψ
2
x22
+Q66
ψ
1
x2
+
ψ
2
x12#
+kshcGc"
ψ
1+
w
x12
+
ψ
2+
w
x22#)dx1dx2
Zpwdx1dx2
The one-term approximations
w(x1,x2) = a1sin
π
x1
asin
π
x2
a,
ψ
1(x1,x2) = a2cos
π
x1
asin
π
x2
a,
ψ
2(x1,x2) = a3sin
π
x1
acos
π
x2
a
satisfy the boundary conditions. Substituting these approximative functions into
Π
follow ˜
Π
=˜
Π
(a1,a2,a3)and the conditions for a minimum of
Π
, i.e.
˜
Π
/
a1=0,
i=1,2,3 yield the equations for the undetermined coefficients ai
K
K
Ka
a
a=q
q
q
with
a
a
aT= [a1a2a3],q
q
qT= [16p/
π
20 0]
and
K
K
K=
2hcGc
λ
2hcGc
λ
hcGc
λ
hcGc
λ
hchf¯xf
3(Q11 +Q66)
λ
2+hcGchchf¯xf
3(Q12 +Q66)
λ
2
hcGc
λ
hchf¯xf
3(Q12 +Q66)
λ
2hchf¯xf
3(Q22 +Q66)
λ
2+hcGc
with
λ
=
π
/a. The solution of the system of three linear equations leads to
a1=0,0222,a2=a3=0,046 and the maximum displacement follows to
wmax =w(x1=a/2,x2=a/2) = a1=2,22 cm.
Exercise 8.4. A simply supported laminate plate [00/900/00]has the following ma-
terial properties: Em=3.4 GPa, Ef=110 GPa,
ν
m=0.35,
ν
f=0.22, vm=0.4,
vf
φ
=0.6, Gm=Em/2(1+
ν
m) = 1.2593 GPa, Gf=Ef/2(1+
ν
f) = 45.0820
GPa, h(1)=h(2)=h(3)=5 mm, a=b=1 m.
8.7 Problems 329
1. Formulate the equation for the bending surface for a lateral unit load F=1 N
at x1=
ξ
1,x2=
ξ
2using the classical laminate theory.
2. Formulate the equation for the natural frequencies of the laminate plate using the
classical plate theory and neglecting the rotatory inertia.
Solution 8.4. The solutions for both cases can be presented as follows.
1. The stacking sequence of the layers yields a symmetric cross-ply plate which is
specially orthotropic (Table 8.1) Bi j =0,D16 =D26 =0
D11
4w
x4
1
+2(D12 +2D66)
4w
x2
1
x2
2
+D22
4w
x2
2
=p3(x1,x2)
The boundary conditions are (Fig. 8.5)
w(0,x2) = w(a,x2) = w(x1,0) = w(x1,b) = 0,
M1(0,x2) = M1(a,x2) = M2(x1,0) = M2(x1,b) = 0
The Navier’s double infinite series solution (8.6.21) - (8.6.23) leads to
w(x1,x2) =
r=1
s=1
prs
drs
sin
α
rx1sin
β
sx2
with
drs = [D11
α
4
r+2(D12 +2D66)
α
2
r
β
2
s+D22
β
4
s],
α
r=r
π
a,
β
s=s
π
b,
prs =4F
ab sin
α
r
ξ
1sin
β
s
ξ
2
With (section 2.2.1)
E
1=Efvf+Emvm=67,36 GPa,E
2=EfEm
Efvm+Emvf
=8,12 GPa,
G
12 =GfGf
Gfvm+Gmvf
=3,0217 GPa
ν
12 =
ν
fvf+
ν
mvm=0,272,
ν
21 =
ν
12E
2/E
1=0,0328
and (4.1.3)
Q
11 =E
1/(1
ν
12
ν
21) = 67,97 GPa,
Q
22 =E
2/(1
ν
12
ν
21) = 8,194 GPa,
Q
12 =Q
21 =
ν
12Q
22 =2,229 GPa,
Q
66 =G
12 =3,02 GPa
follow (4.2.15) the stiffness
330 8 Modelling and Analysis of Plates
Di j =1
3
3
k=1
Q(k)
i j (x(k)
3)3(x(k1)
3)3,
Q(1)
i j =Q(3)
i j =Q[00]
i j =Q
i j,Q(2)
i j =Q[900]
i j ,Q(2)
11 =Q(1)
22 ,
Q(2)
22 =Q(1)
11 ,Q(2)
66 =Q(1)
66 ,
x(0)
3=7,5 mm,x(1)
3=2,5 mm,
x(2)
3=2,5 mm,x(3)
3=7,5 mm,
D11 =18492 Nm,D22 =2927 Nm,
D12 =D21 =627 Nm,D66 =849 Nm
The equation for the bending surface is
w(x1,x2) = Fa2
π
4
r=1
s=1
sin
α
r
ξ
1sin
β
s
ξ
2
18492r4+4650r2s2+2927s4sin
α
rx1sin
β
sx2
If F=1 N then w(x1,x2)represents the influence surface, i.e. the deflection at
(x1,x2) due to a unit load at (
ξ
1,
ξ
2). This influence function w(x1,x2;
ξ
1,
ξ
2)is
sometimes called Green’s function of the plate with all boundaries simply sup-
ported. In the more general case of a rectangular plate a6=bthe Green’s function
is
w(x1,x2;
ξ
1,
ξ
2) = F
π
4ab
r=1
s=1
sin
α
r
ξ
1sin
β
s
ξ
2
drs
sin
α
rx1sin
β
sx2
The Green’s function can be used to calculate the bending surfaces of sim-
ply supported rectangular plates with any transverse loading. With the solution
w(x1,x2)we can calculate the stress resultants M1,M2,M6,Q1,Q2and the stresses
σ
1,
σ
2,
σ
6,
σ
5and
σ
4using (8.6.27) and (8.6.28).
2. Using (8.6.41) the equation for the natural frequencies of a simply supported
rectangular plate is
ω
2
rs =
π
4
ρ
hD11
α
4
r+2(D12 +2D66)
α
2
r
β
2
s+D22
β
4
s
with
h=
(k)
h(k),
ρ
=1
h
(k)
ρ
(k)x(k)
3x(k1)
3
The fundamental frequency corresponds to r=s=1 and is given by
ω
2
11 =
π
4
ρ
ha4D11 +2(D12 +2D66)a
b2+D22 a
b4
8.7 Problems 331
For a=b=1 m and the given material properties we find the fundamental natural
frequency
ω
11 =1593,5
p
ρ
h
Exercise 8.5. Consider a cylindrically orthotropic circular plate with a midplane
symmetric layer stacking under the conditions of axisymmetric loading and dis-
placements.
1. Develop the differential equations for in-plane loading. Calculate the stress re-
sultants for a solid disk (R,h,Er,E
θ
,
ν
r
θ
) loaded
α
)with a radial boundary force
Nr(R) = Nro and
β
)with a body force h pr=h
ρω
2rcaused by spinning the
disk about the axis with an angular velocity
ω
.
2. Develop the differential equations for transverse loading under the condition of
the first order shear deformation theory. Calculate the stress resultants for a solid
plate (R,h,Er,E
θ
,
ν
r
θ
) loaded by a uniform constant pressure p3(r) p0and
α
)clamped, respectively,
β
)simply supported at the boundary r=R.
Solution 8.5. With Sect. 2.1.6 we obtain x1=xr,x2=
θ
,x3=z,
σ
1=
σ
r,
σ
2=
σθ
,
σ
6=
σ
r
θ
,
ε
1=
ε
r,
ε
2=
εθ
,
ε
6=
ε
r
θ
. For axisymmetric deformations of circular disks
and plates all stresses, strains and displacements are independent of
θ
, i.e. they are
functions of ralone and
σ
6=0,
ε
6=0.
1. For an in-plane loaded cylindrical orthotropic circular disk under the condition
of axisymmetric deformations the equilibrium, constitutive and geometric
equations are:
Equilibrium Equations (Fig. 8.7)
With cos(
π
/2d
θ
/2) = sin(d
θ
/2)d
θ
/2 follow
Fig. 8.7 Disc element
(rdrd
θ
)h
x3
0
d
θ
Nr
pr
Nr+dNr
N
θ
d
θ
2
N
θ
d
θ
2
r+dr
h
dr
332 8 Modelling and Analysis of Plates
d(rNr)
dr N
θ
+prr=0
Constitutive Equations
Nr=A11
ε
r+A12
εθ
,N
θ
=A12
ε
r+A22
εθ
,Nr
θ
=0
Geometric Equations
ε
r=du
dr,
εθ
=u
r,
γ
r
θ
=0
These equations forming the following system of three ordinary differential equa-
tions d(rNr)
dr N
θ
=p0r,
Nr=A11
du
dr+A12
u
r,N
θ
=A12
du
dr+A22
u
r
involving three unknown quantities Nr,N
θ
and u. Substituting the stress resul-
tants in the equilibrium equations yield one uncoupled differential equation for
u(r)
rd2u
dr2+du
dr1
r
δ
2u=rpr
A11
with
δ
2=A22/A11 or
d2u
dr2+1
r
du
dr
δ
2
r2u=pr
A11
α
)Radial boundary force
pr
A11
=0,Nr(R) = Nr0,R1=0,R2=R
The general solution of the differential equations follow with
u(r) = Cr
λ
,
λ
=±
δ
as
u(r) = C1r+
δ
+C2r
δ
With R1=0,R2=Rwe obtain C2=0,C1=Nr0/[(A11
δ
+A12)R
δ
1]and such
Nr(r) = Nr0r
R
δ
1,N
θ
(r) = Nr0
δ
r
R
δ
1
Conclusion 8.1. For
δ
=1 we have an isotropic disk with the well-known solu-
tion Nr=N
θ
=Nr0. For
δ
>0, i.e. the circumferential stiffness exceeds the
radial stiffness, at r=0 we have Nr=N
θ
=0, otherwise for
δ
<0, i.e. the ra-
8.7 Problems 333
dial stiffness exceeds the circumferential, at r=0 we have infinitely high stress
resultants or stresses, respectively.
β
)Body force caused by rotation With pr=
ρω
2rwe obtain the solution of the
inhomogeneous differential equations as
u(r) = C1r
δ
1
A11
ρω
2
9
δ
2r3=C1r
δ
+1
A11
ρω
2
δ
29r3
For
δ
=1 follow the well-known solution
u(r) = Cr 1
ν
2
E
ρω
2
8r3
2. With Fig. 8.8 we obtain:
Equilibrium Equations
d(rMr)
drM
θ
rQr=0,d(rQr)
dr+rp3=0
Constitutive Equations
Mr=D11
κ
r+D12
κθ
,M
θ
=D12
κ
r+D22
κθ
,
Qr=ks
55A55
ψ
r+dw
dr
Geometric Equations
κ
r=d
ψ
r
dr,
κθ
=
ψ
r
r,
γ
rz =
ψ
r+dw
dr
Integrating the second equilibrium equation
Fig. 8.8 Plate element
(rdrd
θ
)h
x3
Qr
Qr+dQr
Mr
M
θ
pz
Mr+dMr
M
θ
334 8 Modelling and Analysis of Plates
Qr(r) = 1
rC1Zp3(r)rdr
and substituting Mr,M
θ
and Qrin the first equilibrium equation yield
rd2
ψ
r
dr2+d
ψ
r
dr1
r
δ
2
p
ψ
r=1
D11 C1Zp3rdr,
δ
2
p=D22
D11
The general solution has again the form
ψ
r(r) = C2r
δ
p+C3r
δ
p+
ψ
p(r)
ψ
p(r)is the particular solution of the inhomogeneous differential equation de-
pending on the form of the loading functions p3(r). The differential equation for
the plate deflection w(r)follows with
ksA55
dw
dr=QrksA55
ψ
r,
dw
dr=1
ksA55 C1
1
r1
rZp3(r)rdrC2r
δ
pC3r
δ
p+
ψ
0,
w(r) = 1
ksA55 C1ln rZ1
rZp3(r)rdrdr
C2
r
δ
p+1
δ
p+1C3
r
δ
p+1
1
δ
p
+C4Z
ψ
pdr
For a constant pressure p3(r) = p0we obtain
w(r) = 1
ksA55 C1ln r+p0r2
4C2
1+
δ
p
r1+
δ
pC3
1
δ
p
r1
δ
p
+C4+C1r2
2D11(
δ
2
p1)+p0r4
8D11(
δ
2
p9),
ψ
r(r) = C2r
δ
p+C3r
δ
pC1r
D11(
δ
2
p1)p0r3
2D11(
δ
2
p9)
This general solution is not valid for
δ
p=1 and
δ
p=3 because the particular
solutions
ψ
pfor theses
δ
p-values include terms coinciding with the fundamental
solutions rand r3. Therefore, the particular solutions must be determined in an-
other form. For
δ
p=1, i.e. for the isotropic case, one can use
ψ
p=Ar ln r+Br3
and for
δ
p=3
ψ
p=Ar +Br3and one obtains the general solutions
δ
p=1
w(r) = 1
ksA55 C1ln r+p0r2
41
2C2r2C3lnr
+C4C1r4
4D11 lnr1
2+p0r4
64D11
,
ψ
r(r) = C2r+C3
1
r+C1
rln r
2D11
+p0r3
16D11
8.7 Problems 335
δ
p=3
w(r) = 1
ksA55 C1ln r+p0r2
41
4C2r4+C3
1
2r2
+C4C1
r2
16D11
+p0r4
48D11 ln r1
4,
ψ
r(r) = C2r3+C3
1
r3C1
rlnr
8D11
+p0
12D11
r3lnr
The constants C1,C2,C3and C4are determined from the boundary conditions at
the inner and outer plate edge. For solid plates with R1=0,R2=Rthe constants
C1and C3must be zero, otherwise
ψ
rand wtend to infinity at the plate center.
For
δ
6=3 the general solution for solid plates is
w(r) = p0r2
4"1
ks
55A55
+r2
2D11(
δ
2
p9)#C2
1+
δ
p
r1+
δ
p+C4,
ψ
r(r) = C2r
δ
pp0r3
2D11(
δ
2
p9)
α
)Clamped solid circular plate (
δ
p6=3)
The boundary conditions are
ψ
1(R) = 0,w(R) = 0 yield the constants C2and C4
and the solution as
w(r) = p0
4ksA55
(R2r2) + p0
2D11(
δ
2
p9)"R3
δ
pr1+
δ
p
1+
δ
pr4
4+R4(
δ
p3)
4(1+
δ
p)#
β
)Simply supported solid circular plate (
δ
p6=3)
We take now the boundary conditions w(R) = 0 and Mr(R) = 0 and have the
solution
w(r) = p0
4ksA55
(R2r2)2
+p0
2D11(
δ
2
p9)"(3D11 +D12)R3
δ
p
(
δ
pD11 +D12)(1+
δ
p)(r1+
δ
pR1+
δ
p) + 1
4(R4r4)#
Note that if the transverse shear deformations are neglected we must put
ks
55A55 . In the particular case ks
55A55 and
δ
p=1 follow the well-known
solutions for the classical theory of isotropic plates, i.e.
α
)w(r) = p0
64D(R2r2)2=p0R4
64Dh1r
Ri2,
β
)w(r) = p0
64D(R2r2)5+
ν
1+
ν
R2r2
=p0R4
64Dh1r
Ri25+
ν
1+
ν
r
R2
336 8 Modelling and Analysis of Plates
Exercise 8.6. A rectangular uniformly loaded symmetric cross-ply plate, Fig. 8.9, is
clamped at the edges x2=±band can be arbitrary supported at the edges x1=±a.
The deflection w(x1,x2)may be represented in separated-variables form w(x1,x2) =
wi j(x1,x2) = fi(x1)gj(x2).
1. Formulate one-term approximate solutions using the Vlasov-Kantorovich
method, (2.2.45) - (2.2.47), based on the variation of the potential energy
Π
(w).
2. Demonstrate for the special case of a plate clamped at all edges the extended
Kantorovich method using the Galerkin’s equations.
Solution 8.6. The differential equation and the elastic potential energy can be for-
mulated, Table 8.1 and Eq. (8.2.24),
D1
4w
x4
1
+2D3
4w
x2
1
x2
2
+D2
4w
x4
2
=p0,
with D1=D11,D2=D22,D3=D12 +2D66,pz=p0
Π
(w) = 1
2
a
Z
a
b
Z
b"D11
2w
x2
12
+D22
2w
x2
22
+2D12
2w
x2
1
2w
x2
2
+4D66
2w
x1
x22
2p0#dx1dx2
The one-term approximate solution ˜w(x1,x2) = wij (x1,x2) = fi(x1)gj(x2)has an
unknown function fi(x1)and a priori chosen trial function gj(x2), which satisfy at
least the geometric boundary conditions at x2=±b.
1. The variation
δΠ
of the elastic potential energy
Π
(w)yields
δΠ
(w) = 1
2
a
Z
a
b
Z
bD11
2w
x2
1
+D12
2w
x2
2
δ
2w
x2
1
+D12
2w
x2
1
+D22
2w
x2
2
δ
2w
x2
2
Fig. 8.9 Rectangular uni-
formly loaded plate, cross-ply
symmetrically laminated,
clamped at the longitudinal
edges x2=±band arbitrary
boundary conditions at the
edges x1=±a
x1
x2
b
b
aa
8.7 Problems 337
+4D66
2w
x1
x2
δ
2w
x1
x2p0
δ
wdx1dx2
Substituting ˜w(x1,x2) = fi(x1)gj(x2)one obtains
δΠ
(w) =
a
Z
aD11A f ′′
i
δ
f′′
i+D12B(f′′
i
δ
fi+fi
δ
f′′
i) + D22C fi
δ
fi
+4D66D f
i
δ
f
i¯p0
δ
fidx1
where
A=
b
Z
b
g2
jdx2,B=
b
Z
b
g′′
jgjdx2,C=
b
Z
b
(g′′
j)2dx2,D=
b
Z
b
(gj)2dx2,¯p0=
b
Z
b
p0gjdx2
Integrating Bby parts yield
B=gjgj
b
b
b
Z
b
(gj)2dx2=D
because gj(±b) = 0 for plates with clamped or simply supported edges x2=±b.
Now we integrated by parts the term
a
Z
a
f′′
i
δ
f′′
idx1
of
δΠ
a
Z
a
f′′
i
δ
f′′
idx1=
a
Z
a
f′′
i(
δ
fi)′′dx1=f′′
i
δ
f
i
a
a
a
Z
a
f′′′
i(
δ
fi)dx1
=f′′
i
δ
f
i
a
af′′′
i
δ
fi
a
a+
a
Z
a
f′′′′
i
δ
fidx1
and the condition
δΠ
=0 yields the ordinary differential equations and the nat-
ural boundary conditions for fi(x1)
D1A f ′′′′
i(x1)2D3D f ′′
i(x1) + D2C fi(x1) = ¯p0,
at x1=±a:[D11A f ′′
i(x1)D12D fi(x1)]
δ
f
i(x1) = 0,
[D11A f ′′′
i(x1)D12D f
i(x1) + 4D66D f
i(x1)]
δ
fi(x1) = 0
If a plate edge is clamped, we have f=0,f=0, if it is simply supported, we
have f=0,f′′ =0 and if it is free, we have
338 8 Modelling and Analysis of Plates
(D11A f ′′
i+D12B fi) = 0,(D11A f ′′′
i+D12B f
i4D66B f
i) = 0
The differential equation for f(x1)can be written in the form
f′′′′(x1)2k2
1f′′(x1) + k4
2f(x1) = kp
with
k2
1=DD3
AD1
,k4
2=CD2
AD1
,kp=¯p0
AD1
The solutions of the differential equation are given in App. E in dependence on
k2
2<k2
1,k2
2=k2
1or k2
2>k2
1in the form
f(x1) =
4
l=1
Cl
Φ
l(x1) + fp
with fp=¯p0/D2C. The solutions can be simplified if the problem is symmetric or
antisymmetric. The constants Clcan be calculated with the boundary conditions
at x1=±a.
2. In the special case of all plate edges are clamped the corresponding boundary
conditions are
x1=±a:w=0,
w
x1
=0,x2=±b:w=0,
w
x2
=0
The one-term deflection approximation is assumed again in the form ˜w(x1,x2) =
wi j(x1,x2) = fi(x1)gj(x2). The Galerkin’s procedure yields
a
Z
a
b
Z
bD1
4wi j
x4
1
+2D3
4wi j
x2
1
x2
2
+D2
4wi j
x4
2p0gjdx2=0
and we obtain
D1
b
Z
b
g2
jdx2
d4fi
dx4
1
+2D3
b
Z
b
d2gj
dx2
2
gjdx2
d2fi
dx2
1
+
D2
b
Z
b
d4gj
dx4
2
gjdx2
fi=
b
Z
b
p0gjdx2
Two of the integral coefficients must be integrated by parts
8.7 Problems 339
b
Z
b
d2gj
dx2
2
gjdx2=dgj
dx2
gj
b
b
b
Z
bdgj
dx2dx2,
b
Z
b
d4gj
dx4
2
gjdx2=d3gj
dx3
2
gj
b
bd2gj
dx2
2
dgj
dx2
b
b+
b
Z
bd2gj
dx2
22
dx2
The results can be simplified because for the clamped edges follow
gj(±b) = 0,dgj
dx2±b=0
and we obtain the same differential equation as in 1.
D1
b
Z
b
g2
jdx2
d4fi
dx4
12D3
b
Z
bdgj
dx22
dx2
d2fi
dx2
1
+
D2
b
Z
b
d2gj
dx2
2
dx2
fi=
b
Z
b
p0gjdx2
To improve the one-term approximative plate solution we present in a second
step now fi(x1)a priori and obtain in a similar manner a differential equation for
an unknown function gj(x2)
D2
a
Z
a
f2
idx1
d4gj
dx4
22D3
a
Z
adfi
dx12
dx2
d2gj
dx2
2
+
D1
a
Z
ad2fi
dx2
22
dx1
gj=
a
Z
a
p0fidx1
In this way we have two ordinary differential equations of the iterative solution
procedure which can be written
D1Ag
d4fi
dx4
12D3Dg
d2fi
dx2
1
+D2Cgfi=¯p0g,
D2Af
d4gj
dx4
22D3Df
d2gj
dx2
2
+D1Cfgj=¯p0f
Both equations can be rearranged in the standard form, App. E
d4fi
dx4
12k2
1g
d2fi
dx2
1
+k4
2gfi=kpg,d4gj
dx4
22k2
1f
d2gj
dx2
2
+k4
2fgj=kp f
with
340 8 Modelling and Analysis of Plates
k2
1g=DgD3
AgD1
,k4
2g=CgD2
AgD1
,kpg =¯p0g
AgD1
,k2
1f=DfD3
AfD1
,k4
2f=CfD2
AfD1
,kp f =¯p0f
AfD1
The solutions of both equations are summarized in App. E and depend on the
relation between k2
2gand k2
1gor k2
2fand k2
1f, respectively.
The iterations start by choosing the first approximation as
w[1]
10 =f1(x1)g1(x2),w[2]
21 =f2(x1)g1(x2),w[3]
22 =f2(x1)g2(x2),...
In the special case under consideration the first approximation is
w10(x1,x2) = f1(x1)(x2
2b2)2
and satisfy the boundary conditions w=0,
w/
x2=0,x2=±b. For a number
of widely used composite material we have k2
2>k2
1. Because the problem is
symmetric we have then the simplified solution
f1(x1) = C1cosh ax1cos bx1+C2sinhax1sinbx1+kpg
k4
2g
The constants C1,C2can be calculated with
f1(±a) = 0,df1
dx1±a=0
and w[0]
10 (x1,x2)is determined. Now one can start the next step
w[1]
11 (x1,x2) = f1(x1)g1(x2)
with the function f1(x1)as the a priori trial function. The iteration steps can be
repeated until the convergence is satisfying. In the most engineering applications
w[0]
11 (x1,x2) = f1(x1)g1(x2)
can be used as satisfying closed analytical solution, i.e. w[1]
11 (x1,x2)is suitable for
engineering analysis of deflection and stresses in a clamped rectangular special
orthotropic plate with uniform lateral load and different aspect ratios.
References
adai A (1925) Die elastischen Platten: die Grundlagen und Verfahren zur Berech-
nung ihrer Form ¨ander ungen und Spannungen, sowie die Anwendu ngen der Theo-
rie der ebenen zweidimensionalen elastischen Systeme auf praktische Aufgaben.
Springer, Berlin
Chapter 9
Modelling and Analysis of Circular Cylindrical
Shells
In the previous Chaps. 7 and 8 we have considered beams and plates, i.e. one- and
two-dimensional structural elements with straight axes and plane reference surfaces.
Thin-walled laminated or sandwich shells can be also modelled as two-dimensional
structural elements but with single or double curved reference surfaces. To cover
shells of general shape a special book is necessary, because a general treatment
of shells of any geometry demands a detailed application of differential geometry
relations.
To give a brief insight into the modelling of shells only the simplest shell ge-
ometry will be selected and the following considerations are restricted to circular
cylindrical shells. The modelling and analysis of circular cylindrical shells fabri-
cated from fibre composite material, i.e. its structural theory, depends on the ra-
dius/thickness ratio R/h. For thin-walled shells, i.e. for R/h1 (R/h>10), either
the classical or the first order shear deformation shell theory is capable of accu-
rately predicting the shell behavior. For thick-walled shells, say R/h<10, a three-
dimensional modelling must be used.
Each single lamina of a filamentary composite material behaves again macro-
scopically as if it were a homogeneous orthotropic material. If the material axes of
all laminae are lined up with the shell-surface principal coordinates, i.e., the axial
and circumferential directions, the shell is said to be special orthotropic or circum-
ferential cross-ply circular cylindrical shell. Since the often used cylindrical shells
with closely spaced ring and/or stringer stiffeners also can be approximated by con-
sidering them to be specially orthotropic, a greater number of analysis have been
carried out for such shell type. If the material-symmetry axes are not lined up with
the shell principal axes, the shell is said to be anisotropic, but since there is no struc-
tural advantage for shells constructed in this way it has been not often subjected to
analysis.
In Chap. 9 there are only a short summarizing section on sandwich shells and
no special section considering hygrothermo-elastic effects. Both problems can be
simple retransmitted from the corresponding sections in Chaps. 7 and 8. Also a spe-
cial discussion of analytical solution methods will be neglected, because no general
shell problems are considered.
341
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_9
342 9 Modelling and Analysis of Circular Cylindrical Shells
9.1 Introduction
Chapter 9 gives a short introduction to the theory of circular cylindrical shells in the
frame of the classical shell theory and the shell theory including transverse shear
deformations. Figure 9.1 shows a laminated circular cylindrical shell with general
layer stacking, the global coordinates x1=x,x2=s=R
ϕ
,x3=z, and the principal
material coordinates 1 =x
1,2=x
2. In the theory of circular cylindrical shells the
most complex problem is the modelling and analysis of laminated shells with an
arbitrary stacking of the layers and arbitrary loading. The at least complex problem
is a mid-plane symmetric cross-ply laminated shell with axially symmetric loads
using the classical shell theory. The mathematically modelling leads in this case to
a
bc
x3=z
x2=R
ϕ
x1=x
u2
u3
u1
h
R
2=x
2
1=x
1
x1
x2
θ
z
h
2
h
2z(2)
z(n1)
z(1)
z(0)=h
2
z(n)=h
2
Fig. 9.1 Circular cylindrical shell. aGeometry, global coordinates x1=x,x2=s=R
ϕ
,bshell
middle surface, principal material coordinates x
1=1,x
2=2, fibre angle
θ
,claminate structure, n
layers, layer coordinates z(k), layer thickness h(k)=z(k)z(k1)
9.2 Classical Shell Theory 343
an ordinary differential equation. This type of stacking and loading will be primary
considered in Chap. 9, because analytical solutions can be derived. Generally as-
sumed is that each layer having a constant angle of wrap, constant volume ratio of
fibre to resin, and the fibre and resin are both isotropic and homogeneous within
themselves. The ply material axes, Fig. 9.1 b, will be rotated away from the global
axes by an angle
θ
, positive in the counterclockwise direction.
9.2 Classical Shell Theory
The following hypotheses are the basis to derivative the equations of the classical
shell theory:
Displacements are small compared to the shell thickness, all strain-displacement
relations may be assumed to be linear.
The Kirchhoff hypothesis is applicable, i.e. line elements normal to the middle
surface before deformation remain straight, normal to the deformed middle sur-
face, and unchanged in length after deformation.
All components of translational inertia are included in modelling vibration prob-
lems, but all components of rotatory inertia are neglected.
The ratio of the shell thickness h to the radius R of the middle surface is small as
compared with unity and Love’s first-approximation shell theory is used which
define a thin or classical shell theory: h/R1 and all terms 1 +(z/R)1. It can
be shown that this relationship is consistent with the neglect of transverse shear
deformation and transverse normal stress.
In addition we assume that each individual layer is considered to behave macro-
scopically as a homogeneous, anisotropic, linear-elastic material, that all layers are
assumed to be bonded together with a perfect bond and that each layer may be of
arbitrary thickness and may be arranged either symmetrically or unsymmetrically
with respect to the middle surface.
9.2.1 General Case
The governing differential equations are formulated in terms of the three middle-
surface displacement components (u1ux,u2us,u3uz)
u1(x1,x2,0) = u(x,s),u2(x1,x2,0) = v(x,s),u3(x1,x2,0) = w(x,s)(9.2.1)
The strain displacement relations for a circular, cylindrical shell of any material,
neglecting the effects of transverse shear deformation and using Love’s first approx-
imation are given by
344 9 Modelling and Analysis of Circular Cylindrical Shells
ε
x=
u
x,
ε
s=
v
s+w
R,
ε
xs =
u
s+
v
x,
κ
x=
2w
x2,
κ
s=
2w
s2+1
R
v
s,
κ
xs =2
2w
x
s+1
R
v
x
(9.2.2)
The total strains at a arbitrary distance zof the middle surface are
ε
x=
ε
x+
κ
xz,
ε
s=
ε
s+
κ
sz,
ε
xs =
ε
xs +
κ
xsz
or
ε
j=
ε
j+
κ
jz,j= (1,2,6)(x,s,xs)(9.2.3)
Each individual layer is assumed to be in a state of generalized plane stress, the
Hooke’s law yields
σ
(k)
i=Q(k)
i j
ε
j,i,j= (1,2,6)(9.2.4)
and in the general anisotropic case the Qi j matrix is full populated (Table 4.2).
Using again the Love’s first approximation 1 + (z/R)1), i.e. neglecting the
difference in the areas above and below the middle surface z=0, the force and
moment resultants, Fig. 9.2, are defined analogous to plates
Ni=
h/2
Z
h/2
σ
idz,Mi=
h/2
Z
h/2
σ
izdz,i= (1,2,6)(x,s,xs)(9.2.5)
Putting Eq. (9.2.4) into (9.2.5) yields the constitutive equations in the known form
N
N
N
M
M
M=A
A
A B
B
B
B
B
B D
D
D
ε
ε
ε
κ
κ
κ
(9.2.6)
with
Ns+
Ns
sds
Qs+
Qs
sds
Nsx +
Nsx
xdx
Nsx +
Nsx
xdx
Qx+
Qx
xdx
Nx+
Nx
xdx
ps
pz
px
Nx
Qx
Nxs
QsNs
Nsx
ds=Rd
ϕ
dx
Msx +
Msx
sds
Ms+
Ms
xdx
Mx+
Mx
xdx
Mxs +
Mxs
sds
Mxs
Mx
Ms
Msx
dx
ds=Rd
ϕ
Fig. 9.2 Positive directions for stress resultants
9.2 Classical Shell Theory 345
(Ai j,Bi j ,Di j ) =
h/2
Z
h/2
(1,z,z2)Qi jdz
i.e. for nlaminate layers
Ai j =
n
k=1
Qi j z(k)z(k1),
Bi j =1
2
n
k=1
Qi j z(k)2z(k1)2,
Di j =1
3
n
k=1
Qi j z(k)3z(k1)3
N
N
NT= [NxNsNxs ],M
M
MT= [MxMsMxs],
ε
ε
ε
T= [
ε
x
ε
s
ε
xs],
κ
κ
κ
T= [
κ
x
κ
s
κ
xs](9.2.7)
The equilibrium equations follow with Fig. 9.2 as
Nx
x+
Nxs
s+px=0,
Mx
x+
Mxs
sQx=0,
Nxs
x+
Ns
s+Qs
R+ps=0,
Mxs
x+
Ms
sQs=0,
Qx
x+
Qs
sNs
R+pz=0
(9.2.8)
The moment equations (9.2.8) can be used to eliminate the transverse shear resul-
tants and one obtains
Nx
x+
Nxs
s+px=0,
Nxs
x+
Ns
s+1
R
Ms
s+
Mxs
x+ps=0,
2Mx
x2+2
2Mxs
x
s+
2Ms
s2Ns
R+pz=0
(9.2.9)
Substituting Eqs. (9.2.6) into (9.2.9) yields a set of three coupled partial differential
equations for the three displacements u,v,w, which can be written in matrix form
L11 L12 L13
L21 L22 L23
L31 L32 L33
u
v
w
=
px
ps
pz
(9.2.10)
The linear differential operators Li j are defined in App. D. For symmetrically ar-
ranged layers the differential operators can be simplified, but the matrix (9.2.10)
stay full populated (App. D).
If we consider natural vibrations of laminated circular cylindrical shells in
Eqs. (9.2.9) and (9.2.10) the distributed loads px,ps,pzare taken zero, i.e.
346 9 Modelling and Analysis of Circular Cylindrical Shells
px=ps=pz=0, but all components of translatory inertia must be included.
Without detailed derivation on obtains
Nx
x+
Nxs
s=
ρ
0
2u
t2,
Nxs
x+
Ns
s+1
R
Ms
s+
Mxs
x=
ρ
0
2v
t2,
2Mx
x2+2
2Mxs
x
s+
2Ms
s2Ns
R=
ρ
0
2w
t2
(9.2.11)
and Eq. (9.2.10) changes to
L11 L12 L13
L21 L22 L23
L31 L32 L33
u
v
w
=
ρ
0
2w
t2
u
v
w
(9.2.12)
with
ρ
0=
n
k=1
z(k)
Z
z(k1)
ρ
(k)
0dz=
n
k=1
ρ
(k)
0h(k)
The stress resultants and the displacement are now functions of x,sand t.
ρ
(k)
0is the
mass density of the kth layer,
ρ
0the mass inertia with respect to the middle surface.
9.2.2 Specially Orthotropic Circular Cylindrical Shells Subjected
by Axial Symmetric Loads
Now we consider cross-ply laminated circular cylindrical shells. The laminate stack-
ing may be not middle-surface symmetric, but the fiber angles are
θ
=00or
θ
=900
and the principal material axes 123 coincide with the structural axes x,s,z, i.e.
the stiffness A16 =A26 =0,D16 =D26 =0. In the case of axial symmetry loading
and deformations there are both, all derivations
/
sand v,Nxs ,Mxs zero. For the
loads per unit of the surface area are the following conditions valid
px=0,ps=0,pz=pz(x)
The equilibrium equations (9.2.8) reduce to
dNx
dx=0,dQx
dxNs
R+pz=0,dMx
dxQx=0
or eliminating Qx
dNx
dx=0,d2Mx
dx2Ns
R=pz(9.2.13)
9.2 Classical Shell Theory 347
The strain-displacement relations follow from (9.2.2)
ε
x=du
dx,
ε
s=w
R,
κ
x=d2w
dx2,
ε
xs =0,
κ
s=
κ
xs =0 (9.2.14)
and the stresses from (9.2.4) and (9.2.14) with Q16 =Q26 =Q66 =0
σ
(k)
x=Q(k)
11 (
ε
x+z
κ
x) + Q(k)
12
ε
s=Q(k)
11 du
dxzd2w
dx2+Q(k)
12
w
R,
σ
(k)
s=Q(k)
12 (
ε
x+z
κ
x) + Q(k)
22
ε
s=Q(k)
12 du
dxzd2w
dx2+Q(k)
22
w
R,
Q(k)
11 =E(k)
x
1
ν
(k)
xs
ν
(k)
sx
,Q(k)
22 =E(k)
s
1
ν
(k)
xs
ν
(k)
sx
,Q(k)
12 =
ν
(k)
sx E(k)
x
1
ν
(k)
xs
ν
(k)
sx
,
ν
(k)
xs
E(k)
x
=
ν
(k)
sx
E(k)
s
(9.2.15)
The constitutive equations (9.2.6) can be written as follow
Nx=A11
ε
x+A12
ε
s+B11
κ
x,
Ns=A12
ε
x+A22
ε
s+B12
κ
x,
Mx=B11
ε
x+B12
ε
s+D11
κ
x,
Ms=B12
ε
x+B22
ε
s+D12
κ
x,
(9.2.16)
with
ε
x,
ε
sand
κ
xfrom Eq. (9.2.14).
Putting (9.2.12) and (9.2.14) in the equilibrium equations (9.2.13) one obtains
after a rearrangement
A11
d2u
dx2+A12
R
dw
dxB11
d3w
dx3=0,
A11D11 B2
11
A11 d4w
dx4+2
RA12B11
A11 B12d2w
dx2
+1
R2A11A22 A2
11
A11 w=pzA12
A11
Nx
R,
and with
DR=A11D11 B2
11
A11
,4
λ
4=1
DRR2
A11A22 A2
11
A11
(9.2.17)
can finally be written
d4w
dx4+2
RDRA12B11
A11 B12d2w
dx2+4
λ
4w=1
DRpzA12
A11
Nx
R(9.2.18)
This is a ordinary differential equation of fourth order with constant coefficients and
can be solved by standard methods .
348 9 Modelling and Analysis of Circular Cylindrical Shells
For the most important case of a symmetrical layer stacking Eq. (9.2.18) can be
reduced with B11 =B12 =0,DR=D11 as
d4w
dx4+4
λ
4w=1
D11 pzA12
A11
Nx
R,4
λ
4=1
D11R2
A11A22 A2
11
A11
(9.2.19)
The inhomogeneous linear differential equation of fourth order has constant coeffi-
cient and can be analytically solved (App. E)
w(x) = wh(x) + wp(x)
The homogeneous solution wh(x) = Ce
α
xyields the characteristic equation
α
4+4
λ
4=0
with the conjugate complex roots
α
14=±
λ
(1±i),i=1 (9.2.20)
and with e±
λ
x=cosh
λ
x±sinh
λ
x, e±i
λ
x=cos
λ
x±isin
λ
xone obtains the solution
of the homogeneous differential equation as
wh(x) = C1cosh
λ
xcos
λ
x+C2cosh
λ
xsin
λ
x
+C3sinh
λ
xcos
λ
x+C4sinh
λ
xsin
λ
x(9.2.21)
or
wh(x) = e
λ
x(C1cos
λ
x+C2sin
λ
x) + e
λ
x(C3cos
λ
x+C4sin
λ
x)(9.2.22)
The particular solution wp(x)of the inhomogeneous equation depends on the load-
ing term.
In solving (9.2.19), another solution form may be utilized, the so-called bending-
layer solution. Note the Eqs. (9.2.16) for the symmetrical case, i.e.
Mx=D11
d2w
dx2,Qx=dMx
dx=D11
d3w
dx3
the solution can be written as:
w(x) = M0
2
λ
2D11
e
λ
x(sin
λ
xcos
λ
x)Q0
2
λ
3D11
e
λ
xcos
λ
x
+ML
2
λ
2D11
e
λ
(Lx)(sin
λ
(Lx)cos
λ
(Lx))
+QL
2
λ
3D11
e
λ
(Lx)cos
λ
(Lx) + wp(x)
(9.2.23)
Instead of the general constants Ci,i=1,2,3,4 the resultant stress moments M0,ML
and resultant stress forces Q0,QLat x=0 respectively x=Lare used as integration
constants. To determine the wp(x)solution one have to consider Nxin Eq. (9.2.19)
9.2 Classical Shell Theory 349
as a constant value following by boundary condition and p(x)have to be restricted
to cases where d4p(x)/dx4=0, what is almost true from view point of practical
applications. It can be easy seen that
wp(x) = 1
4
λ
2D11 p(x)A12
A11
Nx
R(9.2.24)
is a solution of the inhomogeneous differential equation (9.2.19).
The advantage of the solution from (9.2.23) is easily seen. The trigonomet-
ric terms oscillate between ±1 and are multiplied by exponential terms with a
negative exponent which yields to an exponential decay. If we set
λ
x=1.5
π
or
λ
(Lx) = 1.5
π
then is e1.5
π
0.009, i.e. the influence of the boundary
values M0,Q0or ML,QLis strong damped to <1%. With 0 x1.5
π
/
λ
or
0Lx1.5
π
/
λ
bending boundary layers are defined which depend on the shell
stiffness.
The important point is that at each end of the shell a characteristic length LB
can be calculated and the M0- and Q0-terms approach zero at the distance x>LB
from x=0 while the ML- and QL-terms approach zero at the same distance LB
from x=L. In the boundary layer region bending stresses induced from M0,Q0a
ML,QLare superimposed to membrane stresses induced from pz. Looking at a long
shell, Fig. 9.3, with L>LBin the region A-B only M0,Q0and wpare non-zero,
in the region C-D only ML,QLand wpand in the region B-C only the particular
solution wpis nonzero, i.e. in this region only a membrane solution exists. With the
calculated w(x)the first differential equation (9.2.17) with B11 =0 can be solved
and yields the displacement function u(x). It should be noted that only some terms
of u(x)decay away from the boundary edges.
In the case of axially symmetric loading and deformation the bending stresses in
each lamina are given by
σ
x
σ
s(k)
=Q11 Q12
Q12 Q22 (k)
ε
x
ε
s+zQ11 Q12
Q12 Q22 (k)
κ
x
0(9.2.25)
ABCD
LBLB
L
Fig. 9.3 Long circular cylindrical shell: Bending boundary regions (A-B) and (C-D), membrane
region (B-C)
350 9 Modelling and Analysis of Circular Cylindrical Shells
The transverse shear stress
σ
xz follows analogous to the classical beam equations
with (7.2.31).
Summarizing the results of the classical shell equations one can draw the follow-
ing conclusions:
The most general case of laminated circular cylindrical shells is that of arbitrarily
laminated anisotropic layers, i.e. angle-ply layers arbitrarily arranged. The anal-
ysis of these shells is based on approximately analytical methods using Ritz-,
Galerkin- or Kantorovich method and numerical methods , e.g. FEM.
Cross-ply laminated shells, i.e shells with orthotropic layers aligned either ax-
ially or circumferentially and arranged symmetrically with respect to the shell
middle surface have governing shell equations which are the same as those for
a single-layer specially orthotropic shell. For axis symmetrical loading the shell
equations reduce in the static case to ordinary differential equations of the x-
coordinate and can be solved analytically. If the orthotropic layers are arranged
to an unsymmetric laminated cross-ply shell then bending-stretching, coupling is
induced and the governing equations are more complex.
When circular cylindrical shells are laminated of more than one isotropic layer
with each layer having different elastic properties and thickness and the lay-
ers are arranged symmetrically with respect to the middle surface, the govern-
ing equations are identical to those of single layer isotropic shells. However, if
the isotropic layers are arranged unsymmetrically to the middle surface, there is
a coupling between in-surface, i.e stretching and shear, and out-of-surface, i.e
bending and twisting, effects.
Additional to the Kirchhoffs hypotheses all equations of the classical shell the-
ory assumed Love’s first approximation, i.e. the ratio h/Ris so small compared
to 1 that the difference in the areas of shell wall element above and below the
middle surface can be neglected.
9.2.3 Membrane and Semi-Membrane Theories
Thin-walled singe layer shells of revolution can be analyzed in the frame of the so-
called membrane theory. One neglects all moments and transverse stress resultants,
all stresses are considered approximatively constant through the shell thickness i.e.
there are no bending stresses and the coupling and bending stiffness are taken to be
zero in the constitutive equations. In some cases it is possible to use the membrane
theory for structural analysis of laminated shells. The efficient structural behavior of
shells based on the shell curvature that yields in wide regions of shells of revolution
approximately a membrane response upon loading as the basic state of stresses and
strains. The membrane theory is not capable to predict sufficient accurate results in
regions with concentrated loads, boundary constraints or curvature changes, i.e. in
regions located adjacent to each structural, material or load discontinuity. Restrict-
ing the consideration again to circular cylindrical shells with unsymmetric cross-ply
stacking we arrive the following equations
9.2 Classical Shell Theory 351
Mx=Ms=Mxs =0,Qx=Qs=0,
Nx
s+
Nxs
s=px,
Nxs
s+
Ns
s=ps,Ns=Rpz,
ε
x=
u
x,
ε
s=
v
s+w
R,
ε
xs =
u
s+
v
x=
γ
xs,
Nx=A11
ε
x+A12
ε
s,Ns=A12
ε
x+A22
ε
s,Nxs =A66
ε
xs
(9.2.26)
The membrane theory yield three equilibrium conditions to calculate three unknown
stress resultants, i.e the membrane theory is statically determined. The membrane
theory is the simplest approach in shell analysis and admit an approximative an-
alytical solution that is very convenient for a first analysis and design of circular
cylindrical shells.
But the problems which can be solved by the membrane theory are unfortunately
limited. To avoid generally to use the more complex bending theory we can con-
sider a so-called semi-membrane theory of circular cylindrical shells. The semi-
membrane theory is slightly more complicated than the membrane theory but more
simpler than the bending theory. The semi-membrane theory was first developed by
Vlasov on the basis of statically and kinematically hypotheses.
If one intends to construct a semi-membrane theory of composite circular cylin-
drical shells bearing in mind the hypotheses underlying the classical single layer
shell theory and the characteristics of the composite structure. The semi-membrane
theory for composite circular cylindrical shells introduces the following assump-
tions:
The shell wall has no stiffness when bended but in axial direction and when
twisted, i.e. D11 =D66 =0,B11 =B66 =0.
The Poisson’s effect is neglected, i.e. A12 =0,B12 =0,D12 =0.
With the assumptions above follow Mx=Mxs =0,Qx=0.
The cross-section contour is inextensible, i.e.
ε
s=
v
s+w
R=0
The shear stiffness of composite shells can be small. Therefore, the assumption
of the classical single layer semi-membrane theory that the shear stiffness is in-
finitely large, is not used.
Taking into account the assumptions above, one obtains the following set of eleven
equations for eleven unknown functions.
Nx
x+
Nxs
s=0,
Nxs
x+
Ns
s+Qs
R+ps=0,
Ms
sQs=0,
Qs
sNs
Rpz=0,
ε
x=
u
x,
ε
xs =
v
x+
u
s,
ε
s=
v
s+w
R,(9.2.27)
352 9 Modelling and Analysis of Circular Cylindrical Shells
Nx=A11
ε
x,Nxs =A66
ε
xs,Ms=D22
κ
s,Qs=ks
44A44
ψ
s+
w
s
The system (9.2.27) can be reduced. For the circular cylindrical shell the unknown
functions and loads can be represented with trigonometric series and after some
manipulations we obtained one uncoupled ordinary differential equation of fourth
order for wn(x),n=0,1,2,.... The detailed derivation of the governing solutions
shall not be considered.
9.3 Shear Deformation Theory
Analogous to plates, considered in Chap. 8, the classical shell theory is only suffi-
ciently accurate for thin shells. For moderately thick shells we have to take, at least
approximately, the transverse shear deformation effects into account. The Kirch-
hoff’s hypotheses are again relaxed in one point: the transverse normals do not
remain perpendicular to the middle-surface after deformation, but a line element
through the shell thickness perpendicular to the middle-surface prior loading, un-
dergoes at most a translation and rotation upon the load applications, no stretching
or curvature.
The following considerations are restricted to axial symmetrical problems of
symmetrical laminated cross-ply circular cylindrical shells including transverse
shear deformation. We start with a variational formulation including the trapeze
effect, i.e. Love’s first approximation is not valid. For axial symmetrical problems
we have the following simplifications of the shell equations:
All derivatives
/
s(...)are zero and for the strains, stress resultants and loads we
assume
ε
xs =0,
κ
s=0,
κ
xs =0,
Nxs =0,Mxs =0,
ps=0,px=0,pz=pz(x)
(9.3.1)
The kinematical assumptions yield with (5.1.2) the shell displacements
ux(x,z) = u(x) + z
ψ
x(x),
us(x,z) = 0,
uz(x,z) = w(x)
(9.3.2)
The strain-displacement relations are
ε
x=du
dx+zd
ψ
x
dx,
ε
s=w
R+z,
ε
xz =
ψ
x+dw
dx
(9.3.3)
9.3 Shear Deformation Theory 353
and the stresses in the kth layer of the shell are
σ
x
σ
s
σ
xz
(k)
=
Q11 Q12 0
Q12 Q22 0
0 0 Q55
(k)
du
dx+zd
ψ
x
dx
w
r+z
ψ
x+dw
dx
(9.3.4)
For a special orthotropic shell the Qi j are
Q11 =Ex
1
ν
xs
ν
sx
,Q22 =Es
1
ν
xs
ν
sx
,
Q12 =
ν
xsEx
1
ν
xs
ν
sx
,Q55 =Gxz
(9.3.5)
The stress resultant forces and couples are defined as
Nx=
n
k=1Z
h(k)
σ
(k)
x1+z
Rdz,Ns=
n
k=1Z
h(k)
σ
(k)
sdz,
Mx=
n
k=1Z
h(k)
σ
(k)
xz1+z
Rdz,Ms=
n
k=1Z
h(k)
σ
(k)
szdz,
Qx=
n
k=1Z
h(k)
σ
(k)
xz 1+z
Rdz,
(9.3.6)
and one obtains with (9.3.4) and (9.3.6)
Nx=A11
du
dx+A12
w
R+1
RD11
d
ψ
x
dx,
Ns=A12
du
dx+T22w,
Mx=D11 1
R
du
dx+d
ψ
x
dx,
Ms=D12
d
ψ
x
dx+˜
T22w,
Qx=A55
ψ
x+dw
dxor Qx=ks
55A55
ψ
x+dw
dx
(9.3.7)
The stiffness coefficients are
354 9 Modelling and Analysis of Circular Cylindrical Shells
(Ai j,Di j ) =
n
k=1Z
h(k)
Q(k)
i j (1,z2)dz,(i j) = (11),(12),
T22 =
n
k=1Z
h(k)
Q(k)
22
dz
R+z=
n
k=1
Q(k)
22 hln1+z
Rih/2
h/2
1
RA22 +1
R3D22,
˜
T22 =
n
k=1Z
h(k)
Q(k)
22
zdz
R+z=
n
k=1
Q(k)
22 Rhz
Rln1+z
Rih/2
h/2
1
R2D22
(9.3.8)
and ks
55 is the shear correction factor.
The variational formulation for the axial symmetrically circular cylindrical shell
with symmetrically laminated
θ
=00and
θ
=900laminae and coincided principal
material and structural axes is given as
Π
(u,w,
ψ
x) =
Π
i
Π
a,
Π
i(u,w,
ψ
x) = 1
2
h/2
Z
h/2
π
Z
0
L
Z
0
(
σ
x
ε
x+
σ
s
ε
s+
σ
xz
ε
xz)dx(R+z)d
ϕ
dz
=1
2
n
k=1Z
h(k)
π
Z
0
L
Z
0Q(k)
11
ε
2
x
+2Q(k)
12
ε
x
ε
s+Q(k)
22
ε
2
s+Q(k)
55
ε
2
xz(R+z)dxd
ϕ
dz
(9.3.9)
Using Eq. (9.3.3) one obtains
Π
i=1
2
2
π
Z
0
L
Z
0(R"A11 du
dx2
+D11 d
ψ
x
dx2
+2D11
du
dx
d
ψ
x
dx
+2A12wdu
dx+T22w2+Rks
55A55"
ψ
2
x+2
ψ
x
dw
dx+dw
dx2#)dxd
ϕ
,
Π
a=
2
π
Z
0
L
Z
0
pz(x)wRdxd
ϕ
(9.3.10)
Equation (9.3.10) can be used for solving shell problems by the variational meth-
ods of Ritz, Galerkin or Kantorovich. It can be also used to derive the differential
equations and boundary conditions but this will be done later on the direct way.
Hamilton’s principle is formulated to solve vibration problems. The potential
energy function
Π
is given with (9.3.10) but all displacements are now functions of
xand the time t. If we analyze natural vibrations the transverse load pzis taken zero.
The kinetic energy follows as
9.3 Shear Deformation Theory 355
T=
n
k=1
T(k)=1
2
n
k=1Z
h(k)
2
π
Z
0
L
Z
0
ρ
(k)"
u
t2
+2z
u
t
ψ
x
t
+z2
ψ
x
t2
+
w
t2#Rd
ϕ
dxdz
=1
2
2
π
Z
0
L
Z
0(R
ρ
0"
u
t2
+
w
t2#
+2R
ρ
1
u
t
ψ
x
t+R
ρ
2
ψ
x
t2)d
ϕ
dx
(9.3.11)
In Eq. (9.3.11)
ρ
(k)is the mass density of the kth layer,
ρ
0and
ρ
2are the mass
and the moment of inertia with respect to the middle surface per unit area and
ρ
1
represents the coupling between extensional and rotational motions.
ρ
1does not
appear in equations for homogeneous shells.
Now with the Lagrange function L(u,w,
ψ
x) = T(u,w,
ψ
x)
Π
(u,w,
ψ
x)the
Hamilton’s principle is obtained as
δ
t2
Z
t1
L(u,w,
ψ
x)dt=0 (9.3.12)
The direct derivation of the differential equations for symmetrical cross-ply circular
cylindrical shells follow using the constitutive, kinematics and equilibrium equa-
tions. The stiffness matrix is defined as
N
N
N
M
M
M=A
A
A0
0
0
0
0
0D
D
D
ε
ε
ε
κ
κ
κ
,Q
Q
Qs=A
A
As
ε
ε
ε
s,
N
N
NT= [NxNsNxs ],M
M
MT= [MxMsMxs],Q
Q
QsT= [QxQs],
ε
ε
ε
T=
u
x
v
s+w
R
u
s+
v
x,
κ
κ
κ
T=
ψ
x
x
ψ
s
s
ψ
x
s+
ψ
s
x,
ε
ε
ε
sT=
ψ
x+
w
x
ψ
s+
w
sv
R
(9.3.13)
For symmetric cross-ply shells all Bi j are zero and also the Ai j ,Di j with
(i j) = (16),(26)and (45). Such we have
A
A
A=
A11 A12 0
A12 A22 0
0 0 A66
,D
D
D=
D11 D12 0
D12 D22 0
0 0 A66
,A
A
As=ks
55A55 0
0ks
44A44
356 9 Modelling and Analysis of Circular Cylindrical Shells
The static equilibrium equations are identical with (9.2.8). For vibration analysis
inertia terms have to be added and one can formulate
Nx
x+
Nxs
s=px+
ρ
0
2u
t2+
ρ
1
2
ψ
x
t2,
Nxs
x+
Ns
s+Qs
R=ps+
ρ
0
2v
t2+
ρ
1
2
ψ
s
t2,
Qx
x+
Qs
sNs
R=pz+
ρ
0
2w
t2,
Mx
x+
Mxs
sQx=
ρ
2
2
ψ
x
t2+
ρ
1
2u
t2,
Mxs
x+
Ms
sQs=
ρ
2
2
ψ
s
t2+
ρ
1
2v
t2
(9.3.14)
ρ
0,
ρ
1,
ρ
2are like in (9.3.11) generalized mass density and are defined in (8.3.9).
Putting the constitutive equations (9.3.12) in the equilibrium equations (9.3.14) the
equations can be manipulated in similar manner to those of the classical theory and
one obtains the simultaneous system of differential equations
˜
L11 ˜
L12 ˜
L13 ˜
L14 ˜
L15
˜
L21 ˜
L22 ˜
L23 ˜
L24 ˜
L25
˜
L31 ˜
L32 ˜
L33 ˜
L34 ˜
L35
˜
L41 ˜
L42 ˜
L43 ˜
L44 ˜
L45
˜
L51 ˜
L52 ˜
L53 ˜
L54 ˜
L55
u
v
ψ
x
ψ
s
w
=
px
ps
0
0
pz
+
ρ
00
ρ
10 0
0
ρ
00
ρ
10
ρ
10
ρ
20 0
0
ρ
10
ρ
20
0 0 0 0
ρ
0
2
t2
u
v
ψ
x
ψ
s
w
(9.3.15)
The linear differential operators are defined in App. D.2.
For free vibrations the loads px,ps,pzare zero and the shell will perform sim-
ple harmonic oscillations with the circular frequency
ω
. Corresponding to simple
supported conditions on both ends of the cylinder, i.e
Nx=0,v=0,w=0,Mx=0,
ψ
s=0,
the spatial dependence can be written as products of two trigonometric functions
and the complete form of vibrations can be taken as
u(x,
ϕ
,t) =
r=1
s=1
Ursei
ω
rstcos
α
mxcosn
ϕ
,
v(x,
ϕ
,t) =
r=1
s=1
Vrsei
ω
rstsin
α
mxsinn
ϕ
,
w(x,
ϕ
,t) =
r=1
s=1
Wrsei
ω
rstsin
α
mxcos n
ϕ
,(9.3.16)
ψ
x(x,
ϕ
,t) =
r=1
s=1
Ψ
rsei
ω
rstcos
α
mxcosn
ϕ
,
9.3 Shear Deformation Theory 357
ψ
s(x,
ϕ
,t) =
r=1
s=1
˜
Ψ
rsei
ω
rstsin
α
mxsinn
ϕ
,
where Urs,Vrs,Wrs,
Ψ
rs,˜
Ψ
rs denote amplitudes,
α
m=m
π
/l,m,nare the longitudinal
and the circumferential wave numbers. Substituting Eqs. (9.3.16) into (9.3.15) re-
sults in a homogeneous algebraic system and its solutions for a particular pair (m,n)
gives the frequency and amplitude ratio corresponding to these wave numbers. For
arbitrary boundary conditions the Ritz’ or Galerkin’s method can be recommended
to obtain the characteristic equations for solving the eigenvalue problem. Then the
natural frequencies and the mode shapes can be calculated. The solution process is
manageable, but involved.
If one restricts the problem to statics and to axially symmetrical loading
pz=pz(x)the Eqs. (9.3.13) - (9.3.15) can be simplified:
Equilibrium equations
dNx
dx=0,dMx
dxQx=0,dQx
dxNs
R+pz=0 (9.3.17)
Strain-displacement equations
ε
x=du
dx,
ε
s=w
R,
κ
x=d
ψ
x
dx,
ε
xz =
ψ
x+dw
dx(9.3.18)
Constitutive equations
Nx=A11
ε
x+A12
ε
s,
Ns=A12
ε
x+A22
ε
s,
Mx=D11
d
ψ
x
dx,
Qx=ks
55A55
ε
xz
(9.3.19)
All derivatives
/
s(...)and v,
ε
xs,
κ
s,
ε
sz,Nxs ,Mxs,Qsare zero. The stress resultant-
displacement relations follow as
Nx=A11
du
dx+A12
w
R,
Ns=A12
du
dx+A22
w
R,
Mx=D11
d
ψ
x
dx,
Qx=ks
55A55
ψ
x+dw
dx
(9.3.20)
With dNx/dx=0 we have Nx=const =N0and one obtains
du
dx=1
A11 N0A12
w
R(9.3.21)
The equilibrium equation (9.3.17) for Qxyields
358 9 Modelling and Analysis of Circular Cylindrical Shells
d
dx
ψ
x+dw
dx=1
ks
55A55 Ns
Rpz
=1
ks
55A55 1
RA12
du
dx+A22
w
Rpz(9.3.22)
and the equilibrium equation dMx/dxQx=0
D11
d2
ψ
x
dx2ks
55A55
ψ
x+dw
dx=0 (9.3.23)
After some manipulations follow with (9.3.22), (9.3.23) two differential equations
for
ψ
xand was
d
ψ
x
dx=d2w
dx2+1
ks
55A55 1
RA12
A11
N0A2
12 A11A22
A11
w
Rpz,
dw
dx=D11
ks
55A55
d2
ψ
x
dx2
ψ
x
(9.3.24)
Differentiating the second equation (9.3.24) and eliminating
ψ
xthe first equation
leads to one uncoupled differential equation of fourth order for w(x)
d4w
dx41
ks
55A55
1
R2
A11A22 A2
12
A11
d2w
dx2+1
R2
A11A22 A2
12
D11A11
w
=1
D11 A12
A11
N0
R+pz1
ks
55A55
d2pz
dx2,
d
ψ
x
dx=d2w
dx2+1
ks
55A55
A11A22 A2
12
R2A11
w+1
ks
55A55 A12
A11
N0
Rpz
(9.3.25)
With ks
55A55 Eqs. (9.3.25) simplify to the corresponding equations of the clas-
sical shell theory (9.2.19).
The governing equation of the axisymmetric problem for or circular cylindrical
shell in the frame of the shear deformation theory can be written as
d4w
dx42k1
d2w
dx2+k4
2w=kp(9.3.26)
with
k2
1=1
ks
55A55
1
R
A11A22 A2
12
A11
,k4
2=1
ks
55A55
1
R2
A11A22 A2
12
D11A11
,
kp=1
D11 A12
A11
N0
R+pz1
D11ks
55A55
d2pz
dx2
The differential equation can be analytical solved
w(x) = wh(x) + wp(x)
9.3 Shear Deformation Theory 359
The particular solution wp(x)has e.g. with d2pz/dx2=0, the form
wp(x) = kp
k4
2
=R2pzRA12N0
A11A22 A2
12
(9.3.27)
The homogeneous solution wh(x) = Ce
α
xyield the characteristic equation
α
42k2
1
α
2+k4
2=0 (9.3.28)
with the roots
α
14=±rk2
1±qk4
1k4
1
which can be conjugate complex, real or two double roots depending on the relations
of the constants k1and k2.
The general solution can be written as (App. E)
w(x) =
4
i=1
Ci
Φ
i(x) + wp(x)
The functions
Φ
i(x)are given in different forms depending on the roots of the char-
acteristic equation (9.3.28). The roots and the functions
Φ
i(x)are summarized in
App. E. The most often used solution form in engineering application is given for
k2
2>k2
1.
For short shells with edges affecting one another, the
Φ
i(x)involving the hy-
perbolic functions are convenient. If there are symmetry conditions to the middle
cross-section x=L/2 the solution can be simplified, for we have
Φ
3=
Φ
4=0.
For long shells with ends not affecting one another applying the
Φ
i(x)that involve
exponential functions.
Analogous to the classical shell solution for long shells a bending-layer solution
can be applied. Only inside the bending-layer region with the characteristic length
LBthe homogen eous part whand the particular part wpof the general solution whave
to superimposed. Outside the bending-layer region, i.e for x>LBor (Lx)>LB
only wpcharacterizes the shell behavior.
Summarizing the results of the shear deformation shell theory one can say
If one restricts the consideration to symmetrical cross-ply circular cylindrical
shells subjected to axially symmetric loadings the modelling and analysis is most
simplified and correspond to the classical shell theory.
In more general cases including static loading and vibration and not neglecting
the trapeze effect the variational formulation is recommended and approximative
analytical or numerical solutions should be applied.
Circular cylindrical shells are one of the most used thin-walled structures of con-
ventional or composite material. Such shells are used as reservoirs, pressure ves-
sels, chemical containers, pipes, aircraft and ship elements. This is the reason for
a long and intensive study to model and analyze circular cylindrical shells and as
result efficient theories and solutions methods are given in literature.
360 9 Modelling and Analysis of Circular Cylindrical Shells
9.4 Sandwich Shells
Sandwich shells are widely used in many industrial branches because sandwich con-
structions often results in designs with lower structural weight then constructions
with other materials. But there is not only weight saving interesting, but in several
engineering applications the core material of a sandwich construction can be also
used as thermal insulator or sound absorber. Therefore one can find numerous lit-
erature on modelling and analysis for sandwich shells subjected static, dynamic or
environmental loads .
But as written in Sects. 7.4 and 8.4 sandwich constructions are, simply consid-
ered, laminated constructions involving three laminae: the lower face, the core, and
the upper face. And by doing so, one can employ all methods of modelling and
analysis of laminated structural elements.
It was discussed in detail in Sect. 8.4 that, considering sandwich structural ele-
ments, we have to keep in mind the assumptions on the elastic behavior of sand-
wiches. Such there are differences in the expressions for the flexural bending and
transverse shear stiffness in comparison with laminated circular cylindrical shells
and essential differences in the stress distribution over the thickness of the shell
wall. The stiffness parameter for sandwich shells depend on the modelling of sand-
wiches having thin or thicker faces, in the same manner as for plates, Eqs. (8.4.1)
and (8.4.2).
For sandwich constructions generally the ratio of the in-plane moduli of elasticity
to the transverse shear moduli is high and transverse shear deformations are mostly
included in its structural modelling. For this reason, the first order shear deformation
theory of laminated shells is used in priority for sandwich shells. But for thin-walled
sandwich shells with a higher shear stiffness approximately the classical sandwich
theory can be used.
The correspondence between laminated and sandwich shells is for vibration or
buckling problems limited and only using for overall buckling and vibration. There
are some special local problems like face wrinkling and core shear instability in
buckling or the face must be additional considered as a shell of elastic foundation
on the core and also shear mode vibration can occur where each face is vibrating
out of phase with the other face. These problems are detailed discussed in a number
of special papers and can not considered in this book.
9.5 Problems
Exercise 9.1. A circular cylindrical sandwich shell has two unequal faces with the
reduced stiffness Qf1
11,Qf3
11 and the thicknesses hf1,hf3. The shell has an orthotropic
material behavior and the material principal axes shall coincide with the structural
axes x,s. The core with the thickness hcdoes not contribute significantly to the
extensional and the flexural shell stiffness. The lateral distributed load is pz=pz(x).
9.5 Problems 361
Formulate the differential equation using a perturbation constant to characterize the
asymmetry of the sandwich and find the perturbation solution way.
Solution 9.1. The shell problem is axially symmetric. In the frame of the classical
shell theory one can use the differential equation (9.2.18)
d4w
dx4+2
RDRA12B11
A11 B12d2w
dx2+4
λ
4w=1
DRpzA12
A11
Nx
R
The stiffness parameter are calculated for sandwiches with thin faces, Sect. 4.3.2,
hhc
A11 =Qf1
11hf1+Qf3
11hf3=Qf1
11hf1 1+Qf3
11hf3
Qf1
11hf1!,
D11 =h
22
Qf1
11hf1+h
22
Qf3
11hf3=h
22
Qf1
11hf1 1+Qf3
11hf3
Qf1
11hf1!,
B11 =h
2Qf1
11hf1+h
2Qf3
11hf3=h
2Qf1
11hf1 1+Qf3
11hf3
Qf1
11hf1!
B12 and D12 can be calculated analogous. A asymmetry constant can be defined as
η
=B11
D11A11
=1+Qf3
11hf3
Qf1
11hf1
1+Qf3
11hf3
Qf1
11hf1
For a symmetric sandwich wall is B11 =0 and so
η
=0. For an infinite stiffness of
face 1 follows
η
1 and of face 3
η
+1, i.e. the constant
η
is for any sandwich
construction given as
1<
η
<+1
The differential equation (9.2.18) can be written with the constant
η
as
d4w
dx4+2
RDRA12D11
A11 A11B12 D11
B11
η
d2w
dx2+4
λ
4w=1
DRpzA12
A11
Nx
R
Since |
η
|<1 one can find w(x)in the form of a perturbation solution
w(x) =
n=0
wn(x)
η
n
For n=0 follow
d4w0
dx44
λ
4w0=1
DRpzA12
A11
Nx
R
and for n1
362 9 Modelling and Analysis of Circular Cylindrical Shells
d4wn
dx4+4
λ
4wn=2
RDRA12D11
A11 A11B12 D11
B11
η
d2wn1
dx2
The left hand side corresponds to the middle-surface symmetric shell with axially
symmetric loading. The right hand side corresponds to the second derivation of the
previously obtained w-solution.
Conclusion 9.1. The perturbation solution yield the solution of the differential equa-
tion as a successive set of solutions of axially symmetric problems of which many
solutions are available. The perturbation solution converges to the exact solution. In
many engineering applications w(x) = wo(w) +
η
w1(x)will be sufficient accurate.
Exercise 9.2. A symmetrical cross-ply circular cylindrical shell is loaded at the
boundary x=0 by an axially symmetric line pressure Q0and line moment M0.
Calculate the ratio M0/Q0that the boundary shell radius does not change if the shell
is very long.
Solution 9.2. We use the solution (9.2.23) with wp=0 and neglect for the long shell
the influence of MLand QL
w(x) = M0
2
λ
2D11
e
λ
x(sin
λ
xcos
λ
x)Q0
2
λ
3D11
cos
λ
x
The condition of no radius changing yields
w(x=0) = 0 M0
2
λ
2D11 Q0
2
λ
3D11
=0M0
Q0
=1
λ
Exercise 9.3. For a long fluid container, Fig. 9.4 determinate the displacement w(x)
and the stress resultants Ns(x)and Mx(x). The container has a symmetrical cross-ply
layer stacking and can be analyzed in the frame of the classical laminate theory.
Solution 9.3. For a long circular cylindrical shell the solution for wh(x), Eq.
(9.2.22), can be reduced to the first term with the negative exponent
Fig. 9.4 Long fluid container,
L>LB
L
p(x)
R R
x
9.5 Problems 363
wh(x) = e
λ
x(C1sin
λ
x+C2cos
λ
x)
The particular solution wp(x)follow with
p(x) = p01x
L
and Eq. (9.2.24) as
wp(x) = p0
4
λ
4D11 1x
L
The boundary constraints are
w(0) = 0C2=p0
4
λ
4D11
,dw(0)
dx=0C1=p0
4
λ
4D11 1x
L
and we obtain the solutions
w(x) = p0
4
λ
4D11 n1x
Lhcos
λ
x+1x
Lsin
λ
xie
λ
xo
In addition,
Ns=A11
ε
x+A22
ε
s=A11
du
dx+A22
w
R
i.e. with
ε
x=0
Ns=A22
Rw(x),Mx=D11
κ
x=D11
d2w
dx2,Qx=dMx
dx=D11
d3w
dx3
Exercise 9.4. Consider a cantilever circular cylindrical shell, Fig. 9.5. The normal
and shear forces Nxand Nxs as are distributed along the contour of the cross-section
x=Lthat they can reduced to the axial force FH, the transverse force FV, the bending
moment MBand the torsion moment MT. Calculate the resultant membrane stress
forces with the membrane theory.
Solution 9.4. With (9.2.26) we have the following equations
x
L
FV
FH
MB
MT
ϕ
s=R
ϕ
RNxs
Fig. 9.5 Tension, bending and torsion of a cantilever circular cylindrical shell
364 9 Modelling and Analysis of Circular Cylindrical Shells
Ns=Rpz,
Nxs
x=ps,
Nx
x=px
Nxs
s
and pz=ps=px=0 yield
Ns=0,Nxs =const,Nx=const
The distributions of FH,FV,MBand MTover the cross-section contour x=Lcan be
represented as
Nx(x=L) = 1
2
π
RFH+2MB
Rcos
ϕ
,
Nrs(x=L) = 1
2
π
RMT
R+2FVsin
ϕ
and yield the reduced forces FV,FMand moments MB,MT
2
π
Z
0
Nx(x=L)Rd
ϕ
=FH,
2
π
Z
0
Nxs(x=L)sin
ϕ
Rd
ϕ
=1
π
R
π
Z
0
2F
Vsin2
ϕ
Rd
ϕ
=F
V,
4
π
/2
Z
0
Nx(x=L)Rcos
ϕ
d
ϕ
=4
π
MB
π
/2
Z
0
cos2
ϕ
d
ϕ
=MB,
2
π
/2
Z
0
RNxs(x=L)Rd
ϕ
=1
R
π
π
/2
Z
0
MTRd
ϕ
=MT
The equilibrium equations yield
Nx(x) = FH
2
π
R[MB+FV(Lx)] cos
ϕ
π
R,
Ns(x) = 0,
Nsx =MT
2
π
R12 +FV
π
Rsin
ϕ
Part IV
Modelling and Analysis of Thin-Walled
Folded Plate Structures
The fourth part (Chap. 10) includes the modelling and analysis of thin-walled folded
plate structures or generalized beams. This topic is not normally considered in stan-
dard textbooks on structural analysis of laminates and sandwiches, but it is included
here because it demonstrates the possible application of Vlasov’s theory of thin-
walled beams and semi-membrane shells on laminated structural elements.
Chapter 10
Modelling and Analysis of Thin-walled Folded
Structures
The analysis of real structures always is based on a structural and mathematical
modelling. It is indispensable for obtaining realistic results that the structural model
represents sufficiently accurate the characteristic structure behavior.
Generally the structural modelling can be divided into three structure levels
Three-dimensional modelling. It means structural elements, their dimensions in
all three directions are of the same order, we have no preferable direction.
Two-dimensional modelling. One dimension of a structural element is significant
smaller in comparison with the other both, so that we can regard it as a quasi
two-dimensional element. We have to distinguish plane and curved elements e.g.
discs, plates and shells.
One-dimensional modelling. Here we have two dimensions (the cross-section)
in the same order and the third one (the length) is significant larger in com-
parison with them, so that we can regard such a structural element as quasi one-
dimensional. We call it ro d, column,bar, beam or arch and can d istinguish straight
and curved forms also.
The attachment of structural elements to one of these classes is not well defined
rather it must be seen in correlation with the given problem.
Many practical problems, e.g., in mechanical or civil engineering lead to the
modelling and analysis of complex structures containing so-called thin-walled ele-
ments. As a result of the consideration of such structures a fourth modelling class
was developed, the modelling class of thin-walled beams and so-called beam shaped
shells including also folded plate structures. In this fourth modelling class it is typ-
ical that we have structures with a significant larger dimension in one direction
(the length) in comparison with the dimensions in transverse directions (the cross-
section) and moreover a significant smaller thickness of the walls in comparison
with the transverse dimensions.
In Chap. 7 the modelling of laminate beams is given in the frame of the
Bernoulli’s and Timoshenko’s beam theory which cannot applied generally to thin-
walled beams. The modelling of two-dimensional laminate structures as plates and
shells was the subject of the Chaps. 8 and 9. In the present Chap. 10 the investi-
367
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_10
368 10 Modelling and Analysis of Thin-walled Folded Structures
gation of beams with thin-walled cross-sections and beam shaped shells especially
folded structures is carried out. Chapter 10 starting in Sect. 10.1 from a short recall
of the classical beam models. In Sect. 10.2 a generalized beam model for prismatic
thin-walled folded plate structures is introduced, including all known beam models.
Section 10.3 discusses some solution procedures and in Sect. 10.4 selected problems
are demonstrated.
10.1 Introduction
Analyzing thin-walled structures it can be useful to distinguish their global and local
structural behavior. Global bending, vibration or buckling is the response of the
whole structure to external loading and is formulated in a global coordinate system.
A typical example for global structure behavior is the deflection of a ship hull on
the waves. But the deflections and stresses in a special domain of the ship e.g. in the
region of structure loading or deck openings or the vibrations or buckling of single
deck plates represent typical local effects.
A necessary condition for a global analysis is that the geometry of the structure
allows its description in a global co-ordinate system, i.e. the thin-walled structure is
sufficient long how it is given in case of a quasi one-dimensional structure.
Of course there are interactions between global and local effects, and in the most
cases these interactions are nonlinear. Usually the global analysis is taken as the
basic analysis and its results are the boundary conditions for local considerations
by using special local co-ordinate systems. The reactions of local to global effects
whereas are neglected.
From this point of view the global analysis of thin-walled beams and beam
shaped shells can be done approximately by describing them as one-dimensional
structures with one-dimensional model equations. For such problems the classical
beam model of J. Bernoulli was used. This model is based on three fundamental
hypotheses:
There are no deformations of the cross-sectional contour.
The cross-section is plane also in case of deformed structures.
The cross-section remain orthogonally to the deformed system axis.
As a result of bending without torsion we have normal stresses
σ
and strains
ε
only
in longitudinal direction. Shear deformations are neglected. The shear stresses
τ
caused by the transverse stress resultants are calculated with the help of the equilib-
rium equations, but they are kinematically incompatible.
The Bernoulli’s beam model can be used for beams with compact and sufficient
stiff thin-walled cross-sections. In case of thin-walled cross-section it is supposed
that the bending stresses
σ
band the shear stresses
τ
qare distributed constantly over
the thickness t. If we have closed thin-walled cross-sections a statically indetermi-
nate shear flow must be considered.
10.1 Introduction 369
A very important supplement to Bernoulli’s beam model was given by Saint-
Venant1for considering the torsional stress. Under torsional stresses the cross-
sections have out-of-plane warping, but assuming that these are the same in all cross-
sections and they are not constrained we have no resulting longitudinal strains and
normal stresses. In this way we have also no additional shear stresses. The distribu-
tion of the so-called Saint-Venant torsional shear stresses is based on a closed shear
flow in the cross-section. For closed thin-walled cross-sections the well-known ele-
mentary formulae of Bredt2can be used.
The Timoshenko’s beam model is an extension of the Bernoulli’s beam model.
It enables to consider the shear deformations approximately. The first two basic hy-
potheses of the Bernoulli’s model are remained. The plane cross-section stays plane
in this case but is not orthogonally to the system axes in the deformed structure. For
the torsional stress also the relationships of Saint-Venant are used.
Rather soon the disadvantages of this both classical beam models were evident
for modelling and analysis of general thin-walled beam shaped structures. Espe-
cially structures with open cross-section have the endeavor for warping, and be-
cause the warping generally is not the same in all cross-sections, there are additional
normal stresses, so-called warping normal stresses and they lead to warping shear
stresses too. Therefore the torsional moment must be divided into two parts, the
Saint-Venant part and the second part caused by the warping shear stresses.
Very fundamental and general works on this problem were done by Vlasov. Be-
cause his publications are given in Russian language they stayed unknown in west-
ern countries for a long time. In 1958 a translation of Vlasov’s book ”General Shell
Theory and its Application in Technical Sciences” into German language was pub-
lished (Wlassow, 1958) and some years later his book on thin-walled elastic beams
was published in English (Vlasov, 1961). By Vlasov a general and systematic ter-
minology was founded, which is used now in the most present papers.
The Vlasov’s beam model for thin-walled beams with open cross-sections is
based on the assumption of a rigid cross-section contour too, but the warping ef-
fects are considered. Neglecting the shear strains of the mid-planes of the walls the
warping of the beam cross-section are given by the so-called law of sectorial areas.
The application of this Vlasov beam model to thin-walled beams with closed cross-
sections leads to nonsatisfying results, because the influences of the cross-sectional
contour deformations and of the mid-plane shear strains in the walls are significant
in such cases.
Therefore a further special structural model was developed by Vlasov in form
of the so-called semi-moment shell theory, in which the longitudinal bending mo-
ments and the torsional moments in the plates of folded structures with closed cross-
sections are neglected. By this way we have in longitudinal direction only membrane
stresses and in transversal direction a mixture of membrane and bending stresses.
This two-dimensional structural model can be reduced to a one-dimensional one
by taking into account the Kantorovich relationships in form of products of two
1Adh´emar Jean Claude Barr´e de Saint-Venant (23 August 1797 Villiers-en-Bi`ere †6 January
1886 Saint-Ouen) - mechanician and mathematician
2Rudolf Bredt (17 April 1842 Barmen - †18 May 1900 Wetter) - mechanical engineer
370 10 Modelling and Analysis of Thin-walled Folded Structures
functions. One of them describes a given deformation state of the cross-section,
considered as a plane frame structure and the other is an unknown function of the
longitudinal co-ordinate.
In 1994 the authors of this textbook published a monograph on thin-walled folded
plate structures in German (Altenbach et al, 1994). Starting from a general struc-
tural model for isotropic structures also a short outlook to anisotropic structures was
given. The general model equations including the semi-moment shell model and all
classical and generalized linear beam models could be derived by neglecting special
terms in the elastic energy potential function or by assuming special conditions for
the contour deformation states. In Sect. 10.2 the derivation of generalized folded
structural model is given for anisotropic plates, e.g. off-axis loaded laminates. The
derivations are restricted to prismatic systems with straight system axes only.
Summarizing one can conclude from the above discussion there are several rea-
sons why for thin-walled structures must be given special consideration in design
and analysis. In thin-walled beams the shear stresses and strains are relatively much
larger than those in beams with solid, e.g. rectangular, cross-sections. The assump-
tions of Bernoulli’s or Timoshenko’s beam theory can be violated e.g. by so-called
shear lag effects, which result in a non-constant distribution of normal bending
stresses which are different from that predicted by the Bernoulli hypotheses for
beams carrying only bending loads. When twisting also occurs warping effects, e.g.
warping normal and shear stresses, have to add to those arising from bending loads.
The warping of the cross-section is defined as its out-of-plane distorsion in the direc-
tion of the beam axis and violated the Bernoulli’s hypotheses and the Timoshenko’s
hypotheses too.
Because of their obvious advantages fibre reinforced laminated composite beam
structures are likely to play an increasing role in design of the present and, espe-
cially, of future constructions in the aeronautical and aerospace, naval or automotive
industry. In addition to the known advantages of high strength or high stiffness to
weight ratio, the various elastic and structural couplings, which are the result of the
directional nature of composite materials and of laminae-stacking sequence, can be
successfully exploited to enhance the response characteristics of aerospace or naval
vehicles.
In order to be able to determine the behavior of these composite beam structures,
consistent mechanical theories and analytical tools are required. So a Vlasov type
theory for fiber-reinforced beams with thin-walled open cross-sections made from
mid-plane symmetric fiber reinforced laminates was developed but in the last 30
years many improved or simplified theories were published.
Because primary or secondary structural configurations such as aircraft wings,
helicopter rotor blades, robot arms, bridges and other structural elements in civil en-
gineering can be idealized as thin- or thick-walled beams, especially as box beams,
beam models appropriate for both thin- and thick-walled geometries which include
the coupled stiffness effects of general angle-ply laminates, transverse shear defor-
mation of the cross-section and the beam walls, primary and secondary warping,
etc. were developed. But nearly all governing equations of thin- and thick-walled
composite beams adopt the basic Vlasov assumption:
10.2 Generalized Beam Models 371
Fig. 10.1 Thin-walled prismatic folded plate structures with open or closed cross-sections
The contours of the original beam cross-section do not deform in their own
planes.
This assumption implies that the normal strain
ε
sin the contour direction is small
compared to the normal strain
ε
zparallel to the beam axis. This is particular valid
for thin-walled open cross-sections, for thin-walled closed cross-section with stiff-
eners (transverse sheets) and as the wall thickness of closed cross-sections increase.
Chapter 10 focuses the considerations to a more general model of composite thin-
walled beams which may be include the classical Vlasov assumptions or may be
relax these assumptions, e.g. by including the possibility of a deformation of the
cross-section in its own plane, etc.
In the following a special generalized class of thin-walled structures is consid-
ered, so-called folded plate structures. A folded plate structure shall be defined as
a prismatic thin-walled structure which can be formed by folding a flat rectangu-
lar plate or joining thin plate strips along lines parallel to their length. Figure 10.1
demonstrates thin-walled structures of the type defined above. The plate strips can
be laminates.
10.2 Generalized Beam Models
Section 10.2 defines the outline of modelling beam shaped, thin-walled prismatic
folded plate structures with open, one or multi-cell closed or mixed open-closed
cross-sections. The considerations are limited to global structural response. Assum-
ing the classical laminate theory for all laminated plate strips of the beam shaped
structure the elastic energy potential function is formulated. The energy potential
is a two-dimensional functional of the coordinate xof the structure axis and the
cross-section contour coordinate s.
Following the way of Vlasov-Kantorovich the two-dimensional functional is re-
duced to an approximate one-dimensional one. A priori fixed generalized coordi-
372 10 Modelling and Analysis of Thin-walled Folded Structures
nate functions describing the cross-section kinematics are introduced. Generalized
displacement functions which depend on the system coordinate xonly are the inde-
pendent functions of the reduced variational statement which leads to a system of
matrix differential equations, the Euler equations of the variational statement, and
to the possible boundary equations.
The general structural model can be simplified by neglecting selected terms in the
energy formulation or by restricting the number of the generalized coordinate func-
tions, i.e. the cross-section kinematics. All results are discussed under the viewpoint
of a sufficient general structural model for engineering applications. A general struc-
tural model is recommended which includes all above noted forms of cross-sections
and enables to formulate efficient numerical solution procedures.
10.2.1 Basic Assumptions
A prismatic system is considered, its dimensions are significant larger in one di-
rection (the length) in comparison with these in transverse directions. The system
consists of nplane thin-walled strip elements; it means their thickness is significant
smaller than the strip width, i.e. tidi. Rigid connections of the plate strips along
their length lines are supposed. Closed cross-sections as well as open cross-sections
and combined forms are possible. In Fig. 10.2 a general thin-walled folded structure
is shown. There is a global co-ordinate system x,y,zwith any position. In each strip
we have a local co-ordinate system x,si,ni, the displacements are ui,vi,wi. We re-
strict our considerations to prismatic structures only and neglect the transverse shear
strains in the strips normal to their mid-planes, it means the validity of the Kirchhoff
hypotheses is supposed or we use the classical laminate theory only. All constants
of each strip are constant in x-direction. For the displacements we can write
n1,w1
s1,v1
x,u1
n2,w2
s2,v2
x,u2
ni,wisi,vi
x,ui
x
y
z
ti
di
l
Fig. 10.2 Thin-walled folded structure geometry and co-ordinate systems
10.2 Generalized Beam Models 373
ui=ui(x,si),vi=vi(x,si),wi=wi(x,si)(10.2.1)
uiand viare the displacements in the mid-plane and wiis the deflection normal to
the mid-plane of the ith strip. As loads are considered:
surface forces, distributed on the unit of the mid-plane
pxi=pxi(x,si),psi=psi(x,si),pni=pni(x,si)(10.2.2)
line forces, distributed on the length unit of the boundaries of the structure
qxi|x=0=qxi(0,si),qsi|x=0=qsi(0,si),qni|x=0=qni(0,si)
qxi|x=l=qxi(l,si),qsi|x=l=qsi(l,si),qni|x=l=qni(l,si)(10.2.3)
If a linear anisotropic material behavior is supposed, for each strip we can use the
constitutive relationship given as
N
N
N
M
M
Mi
=A
A
A B
B
B
B
B
B D
D
Di
ε
ε
ε
κ
κ
κ
i
(10.2.4)
or
Nxi
Nsi
Nxsi
Mxi
Msi
Mxsi
=
A11iA12iA16iB11iB12iB16i
A12iA22iA26iB12iB22iB26i
A16iA26iA66iB16iB26iB66i
B11iB12iB16iD11iD12iD16i
B12iB22iB26iD12iD22iD26i
B16iB26iB66iD16iD26iD66i
ε
xi
ε
si
ε
xsi
κ
xi
κ
si
κ
xsi
The following steps are necessary for calculating the elements of the matrices
A
A
A,B
B
B,D
D
Dfor the ith strip:
Calculate the reduced stiffness matrix Q
Q
Qfor each lamina (k)of the strip (i)by
using the four elastic moduli EL,ET,
ν
LT,GLT, Eqs. (4.1.2) and (4.1.3).
Calculate the values of the transformed reduced stiffness matrix Q
Q
Qfor each lam-
ina (k)of the strip (i)(Table 4.2).
Considering the stacking structure, it means, considering the positions of all lam-
inae in the ith strip calculate the matrix elements Akli,Bkli,Dkli, (4.2.15).
It must be noted that the co-ordinates x1,x2,x3used in Sect. 4.1.3 are corresponding
to the coordinates x,si,niin the present chapter and the stresses
σ
1,
σ
2,
σ
6here are
σ
xi,
σ
si,
σ
xsi. For the force and moment resultants also the corresponding notations
Nxi,Nsi,Nxsi,Mxi,Msi,Mxsiare used and we have to take here:
Nxi=
ti/2
Z
ti/2
σ
xidni,Nsi=
ti/2
Z
ti/2
σ
sidni,Nxsi=
ti/2
Z
ti/2
σ
xsidni,
(10.2.5)
374 10 Modelling and Analysis of Thin-walled Folded Structures
Mxi=
ti/2
Z
ti/2
σ
xinidni,Msi=
ti/2
Z
ti/2
σ
sinidni,Mxsi=
ti/2
Z
ti/2
σ
xsinidni
In Fig. 10.3 the orientations of the loads, see Eqs. (10.2.2) and (10.2.3), and the
resultant forces and moments in the ith wall are shown. In the frame of the classical
laminate theory the transverse force resultants Nsniand Nxnifollow with the help
of the equilibrium conditions for a strip element.
In the same way here we have the following definitions for the elements of the
deformation vector [
ε
1
ε
2
ε
6
κ
1
κ
2
κ
6]T[
ε
xi
ε
si
ε
xsi
κ
xi
κ
si
κ
xsi]Twith
ε
xi=
ui
x=u
i,
ε
si=
vi
si
=v
i,
ε
xsi=
ui
si
+
vi
x=u
i+v
i,
κ
xi=
2wi
x2=w′′
i,(10.2.6)
Fig. 10.3 Loads and resultant
forces and moments in the ith
strip
ni
x
si
Nsi
Nxsi
Nxi
Nxsi
Nsi
Nxsi
pxi
psi
qsi|x=0
qxi|x=0
x
si
ni
Mxsi
Nsni
Msi
Nxni
Mxsi
Mxi
Msi
Mxsi
Nsni
qni|x=0
pni
10.2 Generalized Beam Models 375
κ
si=
2wi
s2
i
=w••
i,
κ
xsi=2
2wi
x
si
=2w
i
10.2.2 Potential Energy of the Folded Structure
The potential energy of the whole folded structure can be obtained by summarizing
the energy of all the nstrips
Π
=1
2
(i)
l
Z
0
di
Z
0N
N
NTM
M
MTi
ε
ε
ε
κ
κ
κ
i
dsidxWa(10.2.7)
With equation (10.2.4) the vectors of the resultant forces and moments can be ex-
pressed and we obtain
Π
=1
2
(i)
l
Z
0
di
Z
0
ε
ε
ε
κ
κ
κ
T
iA
A
A B
B
B
B
B
B D
D
Di
ε
ε
ε
κ
κ
κ
i
dsidxWa(10.2.8)
The external work of the loads is also the sum of all the nstrips
Wa=
i(1
2
l
Z
0
di
Z
0
2(pxiui+psivi+pniwi)dsidx
+
di
Z
0(qxiui+qsivi+qniwi)x=0
+ (qxiui+qsivi+qniwi)x=ldsi)(10.2.9)
After some steps considering the Eqs. (10.2.4) and (10.2.6) Eq. (10.2.9) leads to
Π
=
(i)(1
2
l
Z
0
di
Z
0A11iu2
i+2A12iu
iv
i+2A16iu
i(u
i+v
i)
+A22iv
i
2+2A26iv
i(u
i+v
i) + A66iu
i+v
i2
2B11iu
iw′′
i2B12iu
iw••
i2B12iv
iw′′
i
4B16iu
iw
i2B16i(u
i+v
i)w′′
i2B22iv
iw••
i4B26iv
iw
i
2B26i(u
i+v
i)w••
i4B66i(u
i+v
i)w
i
+D11iw′′2
i+2D12iw′′
iw••
i+4D16iw′′
iw
i(10.2.10)
+D22iw••
i
2+4D26iw••
iw
i+4D66iw
i2
2(pxiui+psivi+pniwi)dsidx
376 10 Modelling and Analysis of Thin-walled Folded Structures
di
Z
0
[(qxiui+qsivi+qniwi)|x=0+ (qxiui+qsivi+qniwi)|x=l]dsi)
10.2.3 Reduction of the Two-dimensional Problem
Equation (10.2.10) represents the complete folded structure model, because it con-
tains all the energy terms of the membrane stress state and of the bending/torsional
stress state under the validity of the Kirchhoff hypotheses. An analytical solution of
this model equations is really impossible with the exception of some very simple
cases. Therefore here we will take another way. As the main object of this section
we will find approximate solutions by reducing the two-dimensional problem to an
one-dimensional one taking into account the so-called Kantorovich separation rela-
tionships (Sect. 2.2).
For the displacements ui,vi,wiin the ith strip we write the approximative series
solutions
ui(x,si) =
(j)
Uj(x)
ϕ
i j (si) = U
U
UT
ϕ
ϕ
ϕ
=
ϕ
ϕ
ϕ
TU
U
U,
vi(x,si) =
(k)
Vk(x)
ψ
ik(si) = V
V
VT
ψ
ψ
ψ
=
ψ
ψ
ψ
TV
V
V,
wi(x,si) =
(k)
Vk(x)
ξ
ik(si) = V
V
VT
ξ
ξ
ξ
=
ξ
ξ
ξ
TV
V
V
(10.2.11)
Here the
ϕ
i j(si),
ψ
ik(si),
ξ
ik(si)are a priori given trial functions of the co-ordinates
siand Uj(x),Vk(x)unknown coefficient functions of the longitudinal co-ordinate x.
Vlasov defined the
ϕ
i j(si),
ψ
ik(si),
ξ
ik(si)as the generalized co-ordinate functions
and the Uj(x),Vk(x)as the generalized displacement functions. Of course it is very
important for the quality of the approximate solution, what kind and which number
of generalized co-ordinates
ϕ
i j (si),
ψ
ik(si),
ξ
ik(si)are used.
Now we consider a closed thin-walled cross-section, e.g. the cross-section of a
box-girder, and follow the Vlasov’s hypotheses:
The out-of-plane displacements ui(si)are approximately linear functions of si. In
this case there are nlinear independent trial functions
ϕ
i j.nis the number of
parallel strip edge lines of cross-section.
The strains
ε
si(si)can be neglected, i.e.
ε
si0. The trial functions
ψ
ik(si)are
then constant functions in all strips and we have n∗∗ linear independent
ψ
ik(si)
and
ξ
ik(si)with n∗∗ =2nm.mis the number of strips of the thin-walled
structure and nis defined above.
The generalized co-ordinate functions can be obtained as unit displacement states
in longitudinal direction (
ϕ
)and in transversal directions (
ψ
,
ξ
). Usually however
generalized co-ordinate functions are used, which allow mechanical interpretations.
In Fig. 10.4, e.g., the generalized co-ordinate functions for a one-cellular rectangu-
lar cross-section are shown.
ϕ
1characterizes the longitudinal displacement of the
10.2 Generalized Beam Models 377
y
z
x
w1u
v1
u
w2
v2u
v3
w3
u
w4
v4
d1
d4
d3
d2
t1
t2
t3
t4
1
a
a
a
a
a
b
b
b
b
bc
ϕ
1
ϕ
2
ϕ
3
ϕ
4
ψ
1
ψ
2
ψ
3
ψ
4
1
1
1
1
1
1
11
ξ
1
ξ
2
ξ
3
ξ
4
d1=d3=dS
d2=d4=dG
a=dS/2,b=dG/2
c=dSdG/4
Fig. 10.4 Generalized coordinate functions of an one-cellular rectangular cross-section
378 10 Modelling and Analysis of Thin-walled Folded Structures
whole cross-section,
ϕ
2and
ϕ
3its rotations about the global y- and z-axes.
ϕ
1,
ϕ
2,
ϕ
3represent the plane cross-section displacements, while
ϕ
4shows its warping.
ψ
2and
ψ
3characterize the plan cross-section displacements in z- and y-direction
and
ψ
1the rotation of the rigid cross-section about the system axis x.
ψ
4defines a
cross-sectional contour deformation, e.g. a distorsion. The generalized co-ordinate
functions
ξ
1,
ξ
2,
ξ
3,
ξ
4represent displacements of the strips corresponding to
ψ
1,
ψ
2,
ψ
3,
ψ
4. For the example of a box-girder cross-section there is n=4, m=4
and n∗∗ =84=4.
In the following more general derivations the strains
ε
siwill be included, we will
take into account more complicated forms of warping functions and therefore there
are no restrictions for the number of generalized co-ordinate functions. After the
input of Eq. (10.2.11) into (10.2.10) and with the definition of the 28 matrices
ˆ
A
A
A1=
(i)
di
Z
0
A11i
ϕ
ϕ
ϕϕ
ϕ
ϕ
Tdsi,ˆ
A
A
A2=
(i)
di
Z
0
A16i
ϕ
ϕ
ϕ
ϕ
ϕ
ϕ
Tdsi,
ˆ
A
A
A3=
(i)
di
Z
0
A66i
ϕ
ϕ
ϕ
ϕ
ϕ
ϕ
Tdsi,ˆ
A
A
A4=
(i)
di
Z
0
A66i
ψ
ψ
ψψ
ψ
ψ
Tdsi,
ˆ
A
A
A5=
(i)
di
Z
0
A26i
ψ
ψ
ψ
ψ
ψ
ψ
Tdsi,ˆ
A
A
A6=
(i)
di
Z
0
A22i
ψ
ψ
ψ
ψ
ψ
ψ
Tdsi,
ˆ
A
A
A7=
(i)
di
Z
0
D11i
ξ
ξ
ξξ
ξ
ξ
Tdsi,ˆ
A
A
A8=
(i)
di
Z
0
D16i
ξ
ξ
ξ
ξ
ξ
ξ
Tdsi,
ˆ
A
A
A9=
(i)
di
Z
0
D66i
ξ
ξ
ξ
ξ
ξ
ξ
Tdsi,ˆ
A
A
A10 =
(i)
di
Z
0
D12i
ξ
ξ
ξ
••
ξ
ξ
ξ
Tdsi,
ˆ
A
A
A11 =
(i)
di
Z
0
D26i
ξ
ξ
ξ
••
ξ
ξ
ξ
Tdsi,ˆ
A
A
A12 =
(i)
di
Z
0
D22i
ξ
ξ
ξ
••
ξ
ξ
ξ
••Tdsi,
ˆ
A
A
A13 =
(i)
di
Z
0
A16i
ϕ
ϕ
ϕψ
ψ
ψ
Tdsi,ˆ
A
A
A14 =
(i)
di
Z
0
A12i
ϕ
ϕ
ϕψ
ψ
ψ
Tdsi,(10.2.12)
ˆ
A
A
A15 =
(i)
di
Z
0
A66i
ϕ
ϕ
ϕ
ψ
ψ
ψ
Tdsiˆ
A
A
A16 =
(i)
di
Z
0
A26i
ϕ
ϕ
ϕ
ψ
ψ
ψ
Tdsi,
ˆ
A
A
A17 =
(i)
di
Z
0
B11i
ϕ
ϕ
ϕξ
ξ
ξ
Tdsi,ˆ
A
A
A18 =
(i)
di
Z
0
B16i
ϕ
ϕ
ϕξ
ξ
ξ
Tdsi,
10.2 Generalized Beam Models 379
ˆ
A
A
A19 =
(i)
di
Z
0
B12i
ϕ
ϕ
ϕξ
ξ
ξ
••Tdsi,ˆ
A
A
A20 =
(i)
di
Z
0
B16i
ϕ
ϕ
ϕ
ξ
ξ
ξ
Tdsi,
ˆ
A
A
A21 =
(i)
di
Z
0
B66i
ϕ
ϕ
ϕ
ξ
ξ
ξ
Tdsi,ˆ
A
A
A22 =
(i)
di
Z
0
B26i
ϕ
ϕ
ϕ
ξ
ξ
ξ
••Tdsi,
ˆ
A
A
A23 =
(i)
di
Z
0
B16i
ψ
ψ
ψξ
ξ
ξ
Tdsi,ˆ
A
A
A24 =
(i)
di
Z
0
B66i
ψ
ψ
ψξ
ξ
ξ
Tdsi,
ˆ
A
A
A25 =
(i)
di
Z
0
B26i
ψ
ψ
ψξ
ξ
ξ
••Tdsi,ˆ
A
A
A26 =
(i)
di
Z
0
B12i
ψ
ψ
ψ
ξ
ξ
ξ
Tdsi,
ˆ
A
A
A27 =
(i)
di
Z
0
B26i
ψ
ψ
ψ
ξ
ξ
ξ
Tdsi,ˆ
A
A
A28 =
(i)
di
Z
0
B22i
ψ
ψ
ψ
ξ
ξ
ξ
••Tdsi
and the load vectors
f
f
fx=
(i)
di
Z
0
pxi
ϕ
ϕ
ϕ
dsi,r
r
rx=
(i)
di
Z
0
qxi
ϕ
ϕ
ϕ
dsi,
f
f
fs=
(i)
di
Z
0
psi
ψ
ψ
ψ
dsi,r
r
rs=
(i)
di
Z
0
qsi
ψ
ψ
ψ
dsi,
f
f
fn=
(i)
di
Z
0
pni
ξ
ξ
ξ
dsi,r
r
rn=
(i)
di
Z
0
qni
ξ
ξ
ξ
dsi
(10.2.13)
the potential energy in matrix form is obtained as follows
Π
=1
2
l
Z
0U
U
UTˆ
A
A
A1U
U
U+V
V
VTˆ
A
A
A6V
V
V+U
U
UTˆ
A
A
A3U
U
U+2U
U
UTˆ
A
A
A15V
V
V
+V
V
VTˆ
A
A
A4V
V
V+V
V
V′′Tˆ
A
A
A7V
V
V′′ +V
V
VTˆ
A
A
A12V
V
V+4V
V
VTˆ
A
A
A9V
V
V
+2U
U
UTˆ
A
A
A14V
V
V+2U
U
UTˆ
A
A
A2U
U
U+2U
U
UTˆ
A
A
A13V
V
V2U
U
UTˆ
A
A
A17V
V
V′′
2U
U
UTˆ
A
A
A19V
V
V4U
U
UTˆ
A
A
A18V
V
V+2U
U
UTˆ
A
A
A16V
V
V+2V
V
VTˆ
A
A
A5V
V
V
2V
V
VTˆ
A
A
A26V
V
V′′ 2V
V
VTˆ
A
A
A28V
V
V4V
V
VTˆ
A
A
A27V
V
V2U
U
UTˆ
A
A
A20V
V
V′′
2V
V
VTˆ
A
A
A23V
V
V′′ 2U
U
UTˆ
A
A
A22V
V
V2V
V
VTˆ
A
A
A25V
V
V4U
U
UTˆ
A
A
A21V
V
V
4V
V
VTˆ
A
A
A24V
V
V+2V
V
VTˆ
A
A
A10V
V
V′′ +4V
V
VTˆ
A
A
A8V
V
V′′ +4V
V
VTˆ
A
A
A11V
V
V
2(U
U
UTf
f
fx+V
V
VTf
f
fs+V
V
VTf
f
fn)dx
(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=0(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=l
(10.2.14)
380 10 Modelling and Analysis of Thin-walled Folded Structures
The variation of the potential energy function (10.2.14) and using
Π
U
U
Ud
dx
Π
U
U
U=0
0
0,
Π
V
V
Vd
dx
Π
V
V
V+d2
dx2
Π
V
V
V′′ =0
0
0,
δ
U
U
UT
Π
U
U
Ux=0,l
=0,
δ
V
V
VT
Π
V
V
Vd
dx
Π
V
V
V′′ x=0,l
=0,
δ
V
V
VT
Π
V
V
V′′ x=0,l
=0
(10.2.15)
leads to a system of matrix differential equations and matrix boundary conditions of
the complete thin-walled folded plate structure
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
U
+ˆ
A
A
A17V
V
V′′′ (ˆ
A
A
A13 2ˆ
A
A
A18 +ˆ
A
A
A20)V
V
V′′
+(ˆ
A
A
A14 +ˆ
A
A
A15 +ˆ
A
A
A19 2ˆ
A
A
A21)V
V
V+ ( ˆ
A
A
A16 ˆ
A
A
A22)V
V
V=f
f
fx
ˆ
A
A
AT
17U
U
U′′′ (ˆ
A
A
AT
13 2ˆ
A
A
AT
18 +ˆ
A
A
AT
20)U
U
U′′
+( ˆ
A
A
AT
14 ˆ
A
A
AT
15 ˆ
A
A
AT
19 +2ˆ
A
A
AT
21)U
U
U+ ( ˆ
A
A
AT
16 ˆ
A
A
AT
22)U
U
U
+ˆ
A
A
A7V
V
V′′′′ + (2ˆ
A
A
AT
82ˆ
A
A
A8+ˆ
A
A
A23 A
A
AT
23)V
V
V′′′
(ˆ
A
A
A4+4ˆ
A
A
A9ˆ
A
A
A10 ˆ
A
A
AT
10 4ˆ
A
A
A24 +ˆ
A
A
A26 +ˆ
A
A
AT
26)V
V
V′′
+( ˆ
A
A
A5ˆ
A
A
AT
5+2ˆ
A
A
A11 2ˆ
A
A
AT
11 +ˆ
A
A
A25 ˆ
A
A
AT
25 2ˆ
A
A
A27 +2ˆ
A
A
AT
27)V
V
V
+( ˆ
A
A
A6+ˆ
A
A
A12 2ˆ
A
A
A28)V
V
V=f
f
fs+f
f
fn
(10.2.16)
δ
U
U
UT[ˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
Uˆ
A
A
A17V
V
V′′ + ( ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V
+( ˆ
A
A
A14 ˆ
A
A
A19)V
V
V±r
r
rx]x=0,l=0
δ
V
V
VT[ˆ
A
A
AT
17U
U
U′′ + ( ˆ
A
A
AT
13 2ˆ
A
A
AT
18 +ˆ
A
A
AT
20)U
U
U+ ( ˆ
A
A
AT
15 2ˆ
A
A
AT
21)U
U
U
ˆ
A
A
A7V
V
V′′′ + (2ˆ
A
A
A82ˆ
A
A
AT
8ˆ
A
A
A23 +ˆ
A
A
AT
23)V
V
V′′
+( ˆ
A
A
A4+4ˆ
A
A
A9ˆ
A
A
AT
10 4ˆ
A
A
A24 +ˆ
A
A
AT
26)V
V
V
+( ˆ
A
A
AT
5+2ˆ
A
A
AT
11 ˆ
A
A
A25 2ˆ
A
A
AT
27)V
V
V±r
r
rs±r
r
rn]x=0,l=0
δ
V
V
VT[ˆ
A
A
AT
17U
U
Uˆ
A
A
AT
20U
U
U
+ˆ
A
A
A7V
V
V′′ + (2ˆ
A
A
AT
8ˆ
A
A
AT
23)V
V
V+ ( ˆ
A
A
AT
10 ˆ
A
A
AT
26)V
V
V]x=0,l=0
(10.2.17)
In Eqs. (10.2.17) the upper sign (+) is valid for the boundary x=0 of the structure
and the lower one () for the boundary x=l. This convention is also valid for all
following simplified models.
10.2 Generalized Beam Models 381
10.2.4 Simplified Structural Models
Starting from the complete folded structure model two ways of derivation simplified
models are usual:
Neglecting of special terms in the potential energy function of the complete
folded plate structure.
Restrictions of the cross-section kinematics by selection of special generalized
co-ordinate functions.
For the first way we will consider the energy terms caused by
the longitudinal curvatures
κ
xi,
the transversal strains
ε
si,
the shear deformations of the mid-planes
ε
xsiand
the torsional curvatures
κ
xsi
in the strips. But not all possibilities for simplified models shall be taken into ac-
count. We will be restricted the considerations to:
A a structure model with neglected longitudinal curvatures
κ
xionly,
B a structure model with neglected longitudinal curvatures
κ
xiand neglected tor-
sional curvatures
κ
xsi,
C a structure model with neglected longitudinal curvatures
κ
xiand neglected
transversal strains
ε
si,
D a structure model with neglected longitudinal curvatures
κ
xi, neglected transver-
sal strains
ε
siand torsional curvatures
κ
xsi, and
E a structure model with neglected longitudinal curvatures
κ
xi, neglected transver-
sal strains
ε
siand neglected shear strain
ε
xsiof the mid-planes of the strips.
In Fig. 10.5 is g iven an overview on the development of structural simplified models.
10.2.4.1 Structural Model A
The starting point is the potential energy equation (10.2.10), in which all terms
containing w′′
ihave to vanish. Together with (10.2.11) and (10.2.12) we find that in
this case
ˆ
A
A
A7=0
0
0,ˆ
A
A
A8=0
0
0,ˆ
A
A
A10 =0
0
0,ˆ
A
A
A17 =0
0
0,ˆ
A
A
A20 =0
0
0,ˆ
A
A
A23 =0
0
0,ˆ
A
A
A26 =0
0
0
The matrix differential equations (10.2.16) and the boundary conditions (10.2.17)
change then into
382 10 Modelling and Analysis of Thin-walled Folded Structures
κ
si6=0,
ε
si6=0,
ε
xsi6=0,
κ
xsi6=0
κ
xi=0
ε
si6=0
ε
si=0
ε
xsi6=0
ε
xsi=0
ε
xsi6=0
ε
xsi=0
κ
xsi6=0
κ
xsi=0
κ
xsi6=0
κ
xsi=0
κ
xsi6=0
A B C D E
Fig. 10.5 Overview to the derivation of usual simplified models for thin-walled folded plate struc-
tures
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
U(ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V′′
+(ˆ
A
A
A14 +ˆ
A
A
A15 +ˆ
A
A
A19 2ˆ
A
A
A21)V
V
V+ ( ˆ
A
A
A16 ˆ
A
A
A22)V
V
V=f
f
fx,
(ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U′′ + ( ˆ
A
A
AT
14 ˆ
A
A
AT
15 ˆ
A
A
AT
19 +2ˆ
A
A
AT
21)U
U
U
+( ˆ
A
A
AT
16 ˆ
A
A
AT
22)U
U
U(ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V′′
+( ˆ
A
A
A5ˆ
A
A
AT
5+2ˆ
A
A
A11 2ˆ
A
A
AT
11 +ˆ
A
A
A25 ˆ
A
A
AT
25 2ˆ
A
A
A27 +2ˆ
A
A
AT
27)V
V
V
+( ˆ
A
A
A6+ˆ
A
A
A12 2ˆ
A
A
A28)V
V
V=f
f
fs+f
f
fn,
(10.2.18)
δ
U
U
UThˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ ( ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V
+( ˆ
A
A
A14 ˆ
A
A
A19)V
V
V±r
r
rxix=0,l=0,
δ
V
V
VTh(ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U+ ( ˆ
A
A
AT
15 2ˆ
A
A
AT
21)U
U
U
+( ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V
+( ˆ
A
A
AT
5+2ˆ
A
A
AT
11 ˆ
A
A
A25 2ˆ
A
A
AT
27)V
V
V±r
r
rs±r
r
rnix=0,l=0
(10.2.19)
10.2 Generalized Beam Models 383
10.2.4.2 Structural Model B
Here the longitudinal curvatures
κ
xiand the torsional curvatures
κ
xsiare neglected
and therefore in the potential energy additionally to w′′
i0 in model A all terms
containing w
ihave to vanish. Additionally to the case of model A now also the ma-
trices ˆ
A
A
A9,ˆ
A
A
A11,ˆ
A
A
A18,ˆ
A
A
A21,ˆ
A
A
A24,ˆ
A
A
A27 are null-matrices. This leads to the following matrix
differential equations and the corresponding boundary conditions:
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
Uˆ
A
A
A13V
V
V′′
+(ˆ
A
A
A14 +ˆ
A
A
A15 +ˆ
A
A
A19)V
V
V+ ( ˆ
A
A
A16 ˆ
A
A
A22)V
V
V=f
f
fx,
ˆ
A
A
AT
13U
U
U′′ + ( ˆ
A
A
AT
14 ˆ
A
A
AT
15 ˆ
A
A
AT
19)U
U
U+ ( ˆ
A
A
AT
16 ˆ
A
A
AT
22)U
U
Uˆ
A
A
A4V
V
V′′
+( ˆ
A
A
A5ˆ
A
A
AT
5+ˆ
A
A
A25 ˆ
A
A
AT
25)V
V
V+ ( ˆ
A
A
A6+ˆ
A
A
A12 ˆ
A
A
A28)V
V
V=f
f
fs+f
f
fn,
(10.2.20)
δ
U
U
UThˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ˆ
A
A
A13V
V
V+ ( ˆ
A
A
A14 ˆ
A
A
A19)V
V
V±r
r
rxix=0,l=0,
δ
V
V
VThˆ
A
A
AT
13U
U
U+ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V+ ( ˆ
A
A
AT
5ˆ
A
A
A25)V
V
V±r
r
rs±r
r
rnix=0,l=0
(10.2.21)
10.2.4.3 Structural Model C
In this structure model the longitudinal curvatures
κ
xiand the strains
ε
siare ne-
glected. Therefore in this case in the potential energy function (10.2.10) all terms
containing w′′
iand v
ihave to vanish. Considering the equations (10.2.11) and
(10.2.12) we find, that additionally to the case of the structure model A here the
matrices ˆ
A
A
A5,ˆ
A
A
A6,ˆ
A
A
A14,ˆ
A
A
A16,ˆ
A
A
A27,ˆ
A
A
A28 are null-matrices and in this way we obtain the
following matrix differential equations and the corresponding boundary conditions:
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
U(ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V′′
+( ˆ
A
A
A15 +ˆ
A
A
A19 2ˆ
A
A
A21)V
V
Vˆ
A
A
A22V
V
V=f
f
fx,
(ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U′′ + (ˆ
A
A
AT
15 ˆ
A
A
AT
19 +2ˆ
A
A
AT
21)U
U
Uˆ
A
A
AT
22U
U
U
(ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V′′
+(2ˆ
A
A
A11 2ˆ
A
A
AT
11 +ˆ
A
A
A25 ˆ
A
A
AT
25)V
V
V+ˆ
A
A
A12V
V
V=f
f
fs+f
f
fn,
(10.2.22)
δ
U
U
UThˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ ( ˆ
A
A
A13 2ˆ
A
A
A18)V
V
Vˆ
A
A
A19V
V
V±r
r
rxix=0,l=0,
δ
V
V
VTh(ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U+ ( ˆ
A
A
AT
15 2ˆ
A
A
AT
21)U
U
U
+( ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V+ (2ˆ
A
A
AT
11 ˆ
A
A
A25)V
V
V±r
r
rs±r
r
rnix=0,l=0
(10.2.23)
384 10 Modelling and Analysis of Thin-walled Folded Structures
10.2.4.4 Structural Model D
This structure model neglects the longitudinal curvatures
κ
xi, the strains
ε
siand
the torsional curvatures
κ
xsi. In the potential energy function all terms containing
w′′
i,v
i,w
ihave to vanish and we find together with (10.2.11) and (10.2.12) that
additionally to the structure model C the matrices ˆ
A
A
A9,ˆ
A
A
A11,ˆ
A
A
A18,ˆ
A
A
A21,ˆ
A
A
A24 are null-
matrices. We obtain the matrix differential equations and the boundary conditions
in the following form:
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
U
ˆ
A
A
A13V
V
V′′ + ( ˆ
A
A
A15 +ˆ
A
A
A19)V
V
Vˆ
A
A
A22V
V
V=f
f
fx,
ˆ
A
A
AT
13U
U
U′′ + (ˆ
A
A
AT
15 ˆ
A
A
AT
19)U
U
Uˆ
A
A
AT
22U
U
U
ˆ
A
A
A4V
V
V′′ + ( ˆ
A
A
A25 ˆ
A
A
AT
25)V
V
V+ˆ
A
A
A12V
V
V=f
f
fs+f
f
fn,
(10.2.24)
δ
U
U
UThˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ˆ
A
A
A13V
V
Vˆ
A
A
A19V
V
V±r
r
rxix=0,l=0,
δ
V
V
VThˆ
A
A
AT
13U
U
U+ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
Vˆ
A
A
A25V
V
V±r
r
rs±r
r
rnix=0,l=0(10.2.25)
10.2.4.5 Structural Model E
Now the longitudinal curvatures
κ
xi, the transversal strains
ε
siand the shear strains
ε
xsiof the mid-planes shall be neglected. Therefore in the potential energy function
all the terms containing w′′
iand v
ivanish again. The neglecting of the shear strains
of the mid-planes leads with Eq. (10.2.6) to
ε
xsi=
ui
si
+
vi
x=u
i+v
i=0
and we can see that the generalized displacement functionsU
U
U(x)and V
V
V(x)and also
the generalized co-ordinate functions
ϕ
ϕ
ϕ
(si)and
ψ
ψ
ψ
(si)are no more independent from
each other
u
i=v
i,U
U
UT
ϕ
ϕ
ϕ
=V
V
VT
ψ
ψ
ψ
,
ϕ
ϕ
ϕ
=
ψ
ψ
ψ
,U
U
U=V
V
V,U
U
U=V
V
V′′,U
U
U′′ =V
V
V′′′ (10.2.26)
Therefore the potential energy function must be reformulated before the variation
of U
U
Uand V
V
V. Considering the vanishing terms w′′
i,v
iand (10.2.26) we obtain the
potential energy function in the following form:
10.2 Generalized Beam Models 385
Π
=1
2
l
Z
0V
V
V′′Tˆ
A
A
A1V
V
V′′ +V
V
VTˆ
A
A
A12V
V
V+2V
V
V′′Tˆ
A
A
A19V
V
V+4V
V
V′′Tˆ
A
A
A18V
V
V
+4V
V
VTˆ
A
A
A9V
V
V+4V
V
VTˆ
A
A
A11V
V
V2(U
U
UTf
f
fx+V
V
VTf
f
fs+V
V
VTf
f
fn)dx
(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=0(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=l
(10.2.27)
The variation of the potential energy, see also (10.2.15), leads to the matrix differ-
ential equations and the boundary conditions for the structural model E
ˆ
A
A
A1V
V
V′′′′ +2(ˆ
A
A
A18 ˆ
A
A
AT
18)V
V
V′′′ + (4ˆ
A
A
A9+ˆ
A
A
A19 +ˆ
A
A
AT
19)V
V
V′′
+2(ˆ
A
A
A11 ˆ
A
A
AT
11)V
V
V+ˆ
A
A
A12V
V
V=f
f
f
x+f
f
fs+f
f
fn,
(10.2.28)
δ
V
V
VThˆ
A
A
A1V
V
V′′′ +2(ˆ
A
A
A18 +ˆ
A
A
AT
18)V
V
V′′ + (4ˆ
A
A
A9ˆ
A
A
A19)V
V
V
+2ˆ
A
A
AT
11V
V
V+f
f
fx±r
r
rx±r
r
rnx=0,l=0,
δ
V
V
VThˆ
A
A
A1V
V
V′′ +2ˆ
A
A
A18V
V
V+ˆ
A
A
A19V
V
Vr
r
rxix=0,l=0
(10.2.29)
10.2.4.6 Further Special Models by Restrictions of the Cross-Section
Kinematics
All the five given simplified structure models include the neglecting of the longitu-
dinal curvatures
κ
xiin the strips. Because in the case of a beam shaped thin-walled
structure the influence
κ
xion the deformation state and the stresses of the whole
structure can be seen as very small, its neglecting is vindicated here. The main ad-
vantage of the given five simplified structure models however is that by neglecting
the longitudinal curvatures in the strips we have a decreasing of the order of deriva-
tions of the generalized displacement functions U
U
Uand V
V
Vin the potential energy.
This is an important effect for practical solution strategies of the model equations.
The structure models A and C can be used for the analysis of thin-walled beam
shaped structures with open or closed cross-sections. The difference exists only in
the including or neglecting of the strains
ε
siin the strips. Usually they can be ne-
glected, if we have not temperature loading or concentrated transversal stiffeners in
the analyzed structure. The structure models B and D are valid only for structures
with closed cross-sections, because there the torsional curvatures and the torsional
moments Mxsiare very small. The use of the structure model E is vindicated only
for beam shaped structures with open cross-sections. There the shear strains of the
mid-planes of the plate strips have only small influence on the displacements and
the stress state of the structure how in opposite to the case of a closed cross-section.
Further for each of the five considered models we can develop model variants
restricting the cross-section kinematics by selection of special sets of generalized
co-ordinate functions
ϕ
ϕ
ϕ
,
ψ
ψ
ψ
,
ξ
ξ
ξ
. For example, in the structure model B the number of
generalized co-ordinate functions is unlimited. In model D in contrast the number
386 10 Modelling and Analysis of Thin-walled Folded Structures
of
ψ
ψ
ψ
,
ξ
ξ
ξ
-co-ordinates is limited to n∗∗, see Sect. 10.2.3. Restricting in this model ad-
ditionally the
ϕ
ϕ
ϕ
-co-ordinates to n, the semi-moment shell theory for an anisotropic
behavior of the strips is obtained. The Eqs. (10.2.24) and (10.2.25) stay unchanged.
A symmetric stacking sequence in all the strips leads for this model to a further
simplification, because then we have no coupling between stretching and bending
in the strips, all the elements of the coupling matrix B
B
Bvanish and therefore all the
matrices ˆ
A
A
A17 ˆ
A
A
A28 are null-matrices. In this special case the following matrix dif-
ferential equations and boundary conditions are valid
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ˆ
A
A
A3U
U
Uˆ
A
A
A13V
V
V′′ +ˆ
A
A
A15V
V
V=f
f
fx,
ˆ
A
A
AT
13U
U
U′′ ˆ
A
A
AT
15U
U
Uˆ
A
A
A4V
V
V′′ +ˆ
A
A
A12V
V
V=f
f
fs+f
f
fn,
(10.2.30)
δ
U
U
UThˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ˆ
A
A
A13V
V
V±r
r
rxx=0,l=0,
δ
V
V
VTˆ
A
A
AT
13U
U
U+ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V±r
r
rs±r
r
rnix=0,l=0(10.2.31)
If we have symmetric cross-ply laminates in all the plate strips and one of the main
axes of them is identical with the global x-axis, there is no stretching/shearing or
bending/twisting coupling and therefore additionally to the B
B
B-matrix the elements
A16 =A26 =0,D16 =D26 =0. With these the matrices ˆ
A
A
A2,ˆ
A
A
A13, are null-matrices,
Eq. (10.2.12). In this case the Eqs. (10.2.30) and (10.2.31) lead to
ˆ
A
A
A1U
U
U′′ +ˆ
A
A
A3U
U
U+ˆ
A
A
A15V
V
V=f
f
fx,
ˆ
A
A
A15U
U
Uˆ
A
A
A4V
V
V′′ +ˆ
A
A
A12V
V
V=f
f
fs+f
f
fn,(10.2.32)
δ
U
U
UThˆ
A
A
A1U
U
U±r
r
rxix=0,l=0,
δ
V
V
VThˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V±r
r
rs±r
r
rnix=0,l=0(10.2.33)
and we find that we have equations of the same type how in case of the classical
semi-moment shell theory of Vlasov. Here however the matrices ˆ
A
A
Aconsider the
anisotropic behavior of the strips.
Otherwise if starting from the structure model E then the generalized co-ordinate
functions
ψ
ψ
ψ
,
ξ
ξ
ξ
are restricted to three co-ordinates, representing the rigid cross-
section, i.e. (
ψ
1,
ξ
1), the rotation of the cross-section about the x-axis, (
ψ
2,
ξ
2) and
(
ψ
3,
ξ
3) the displacements in the global y- and z-direction. If we additionally assume
only four
ϕ
-functions, then the first three representing the plane cross-section, i.e.
ϕ
1the displacement in x-direction,
ϕ
2,
ϕ
3the rotations about the y- and z-axes and
ϕ
4is a linear warping function, the so-called unit warping function according to
the sectorial areas law. Therefore we have a structural model similar the classical
Vlasov beam model. The difference is only that in the classical Vlasov beam model
isotropic material behavior is assumed and here the anisotropic behavior of the plate
strips is considered. Note that for comparison of these both models usually the fol-
lowing correlations should be taken into account:
10.2 Generalized Beam Models 387
U1(x) = u(x)displacements of the plane cross-section in x-direction,
U2(x) =
ϕ
y=w(x)rotation of the plane cross-section about the y-axis,
U3(x) =
ϕ
z=v(x)rotation of the plane cross-section about the z-axis,
U4(x) =
ωθ
(x)the warping with
ω
as the unit warping function,
V1(x) =
θ
(x)rotation of the rigid cross-section about the x-axis,
V2(x) = v(x)displacement of the rigid cross-section in y-direction,
V3(x) = w(x)displacement of the rigid cross-section in z-direction
More details about the equations shall not be given here.
If we in this anisotropic Vlasov beam model suppress the warping of the
cross-section and use only the functions
ϕ
1,
ϕ
2,
ϕ
3representing the plane cross-
section kinematics and do not take into account the torsion, we obtain a special-
ized Bernoulli beam model for laminated beams with thin-walled cross-sections
and anisotropic material behavior. In a similar way a specialized Timoshenko beam
model can be obtained, if we restrict the generalized co-ordinate functions
ϕ
ito the
three functions for the plane cross-section kinematics and take into account only
the two
ψ
-co-ordinates for the displacements in y- and z-direction, here however
starting from the structural model D.
The both above discussed quasi beam models are specialized for structures with
cross-sections consisting of single thin plate strips without any rule of their arrange-
ment in the cross-section. The curvatures
κ
xiare generally neglected. The special-
ized beam equations described above cannot be compared directly with the beam
equations in Chap. 7 because the derivation there is not restricted to thin-walled
folded plate cross-sections.
10.2.5 An Efficient Structure Model for the Analysis of General
Prismatic Beam Shaped Thin-walled Plate Structures
Because in the following only beam shaped thin-walled structures are analyzed the
neglecting of the influence of the longitudinal curvatures
κ
xiin the single plate
strips is vindicated. How it was mentioned above, we have in this case a decreasing
of the order of derivations of the generalized displacement functions in the potential
energy, and this is very important for the solution procedures.
The selected structure model shall enable the analysis of thin-walled structures
with open, closed and mixed open/closed cross-sections. Because the influence of
κ
xsiis small only for closed cross-sections but the influence of
ε
xsican be neglected
for open cross-sections only, the selected structure model for general cross-sections
have to include the torsional curvatures and the shear strains of the mid-planes.
The strains
ε
sihave in the most cases only a small influence and could be ne-
glected generally. But we shall see in Chap. 11 that including
ε
siin the model equa-
tions leads an effective way to define the shape functions for special finite elements
and therefore also the
ε
siare included in the selected structure model.
388 10 Modelling and Analysis of Thin-walled Folded Structures
Summarizing the above discussion the structure model A is selected as an uni-
versal model for the modelling and analysis of beam shaped thin-walled plate struc-
tures. An extension of the equations to eigen-vibration problems is given in Sect.
10.2.6.
10.2.6 Free Eigen-Vibration Analysis, Structural Model A
Analogous to the static analysis, the eigen-vibration analysis also shall be restricted
to global vibration response. Local vibrations, e.g. vibration of single plates, are
excluded. A structure model neglecting the longitudinal curvatures cannot describe
local plate strip vibrations. Further only free undamped vibrations are considered.
The starting point is the potential energy function, Eq. (10.2.10), but all terms
including
κ
xiare neglected. With the potential energy
Π
(ui,vi,wi)
Π
=
(i)
1
2
l
Z
0
di
Z
0A11iu2
i+2A12iu
iv
i+2A16iu
i(u
i+v
i)
+A22iv
i
2+2A26iv
i(u
i+v
i) + A66i(u
i+v
i)2
2B12iu
iw••
i4B16iu
iw
i2B22iv
iw••
i4B26iv
iw
i
2B26i(u
i+v
i)w••
i4B66i(u
i+v
i)w
i
+D22iw••
i
2+4D26iw••
iw
i+4D66iw
i2dsidx
(10.2.34)
and the kinetic energy T(ui,vi,wi)
T(u
u
u) = 1
2
(i)
l
Z
0
di
Z
0
ρ
iti"
ui
t2
+
vi
t2
+
wi
t2#dsidx,(10.2.35)
where
ρ
iis the average density of the ith plate strip
ρ
i=1
ti
n
k=1
ρ
(k)
it(k)
i(10.2.36)
Because we have thin plate strips only, rotational terms of the kinetic energy can be
neglected.
The reduction of the two-dimensional problem is carried out again with the gen-
eralized co-ordinate functions
ϕ
ϕ
ϕ
(si),
ψ
ψ
ψ
(si),
ξ
ξ
ξ
(si), but we must remark that the gener-
alized displacement functions U
U
U,V
V
Vare time-dependent and therefore they are writ-
ten in the following with a tilde. The reduction relationships are
10.2 Generalized Beam Models 389
ui(x,si,t) =
(j)
˜
Uj(x,t)
ϕ
i j (si) = ˜
U
U
UT
ϕ
ϕ
ϕ
=
ϕ
ϕ
ϕ
T˜
U
U
U,
vi(x,si,t) =
(k)
˜
Vk(x,t)
ψ
ik(si) = ˜
V
V
VT
ψ
ψ
ψ
=
ψ
ψ
ψ
T˜
V
V
V,
wi(x,si,t) =
(k)
˜
Vk(x,t)
ξ
ik(si) = ˜
V
V
VT
ξ
ξ
ξ
=
ξ
ξ
ξ
T˜
V
V
V
(10.2.37)
Additionally to the ˆ
A
A
Amatrices equation (10.2.12) the following matrices are defined:
ˆ
B
B
B1=
(i)
di
Z
0
ρ
iti
ϕ
ϕ
ϕϕ
ϕ
ϕ
Tdsi,ˆ
B
B
B2=
(i)
di
Z
0
ρ
iti
ψ
ψ
ψψ
ψ
ψ
Tdsi,ˆ
B
B
B3=
(i)
di
Z
0
ρ
iti
ξ
ξ
ξξ
ξ
ξ
Tdsi(10.2.38)
and we obtain the so-called Lagrange function L=T
Π
L=
(i)
1
2
l
Z
0˙
˜
U
U
UTˆ
B
B
B1˙
˜
U
U
U+˙
˜
V
V
VTˆ
B
B
B2˙
˜
V
V
V+˙
˜
V
V
VTˆ
B
B
B3˙
˜
V
V
V
(˜
U
U
UTˆ
A
A
A1˜
U
U
U+2˜
U
U
UTˆ
A
A
A14 ˜
V
V
V+2˜
U
U
UTˆ
A
A
A2˜
U
U
U+2˜
U
U
UTˆ
A
A
A13 ˜
V
V
V
+˜
V
V
VTˆ
A
A
A6˜
V
V
V+2˜
U
U
UTˆ
A
A
A16 ˜
V
V
V+2˜
V
V
VTˆ
A
A
A5˜
V
V
V+˜
U
U
UTˆ
A
A
A3˜
U
U
U
+2˜
U
U
UTˆ
A
A
A15 ˜
V
V
V+˜
V
V
VTˆ
A
A
A4˜
V
V
V2˜
U
U
UTˆ
A
A
A19 ˜
V
V
V4˜
U
U
UTˆ
A
A
A18 ˜
V
V
V
2˜
V
V
VTˆ
A
A
A28 ˜
V
V
V4˜
V
V
VTˆ
A
A
A27 ˜
V
V
V2˜
U
U
UTˆ
A
A
A22 ˜
V
V
V2˜
V
V
VTˆ
A
A
A25 ˜
V
V
V
4˜
U
U
UTˆ
A
A
A21 ˜
V
V
V4˜
V
V
VTˆ
A
A
A24 ˜
V
V
V
+˜
V
V
VTˆ
A
A
A12 ˜
V
V
V+4˜
V
V
VTˆ
A
A
A11 ˜
V
V
V+4˜
V
V
VTˆ
A
A
A9˜
V
V
V)dx
(10.2.39)
The time derivations of the generalized displacement functions are written with the
point symbol
˙
˜
U
U
U=
˜
U
U
U
t,˙
˜
V
V
V=
˜
V
V
V
t(10.2.40)
The Hamilton principle yields the variational statement
δ
t2
Z
t1
Ldt=
t2
Z
t1
δ
Ldt=0,L=L(x,t,˜
U
U
U,˜
V
V
V,˜
U
U
U,˜
V
V
V,˙
˜
U
U
U,˙
˜
V
V
V)(10.2.41)
and we obtain two differential equations
L
˜
U
U
Ud
dx
L
˜
U
U
Ud
dt
L
˙
˜
U
U
U=0
0
0,
L
˜
V
V
Vd
dx
L
˜
V
V
Vd
dt
L
˙
˜
V
V
V=0
0
0
(10.2.42)
390 10 Modelling and Analysis of Thin-walled Folded Structures
If further harmonic relationships for the generalized displacement functions are as-
sumed
˜
U
U
U(x,t) = U
U
U(x)sin
ω
0t,˜
V
V
V(x,t) = V
V
V(x)sin
ω
0t(10.2.43)
and we obtain after some steps the following matrix differential equations:
ˆ
A
A
A1U
U
U′′ + ( ˆ
A
A
A2ˆ
A
A
AT
2)U
U
U+ ( ˆ
A
A
A3
ω
2
0ˆ
B
B
B1)U
U
U(ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V′′
+(ˆ
A
A
A14 +ˆ
A
A
A15 +ˆ
A
A
A19 2ˆ
A
A
A21)V
V
V+ ( ˆ
A
A
A16 ˆ
A
A
A22)V
V
V=0
0
0,
(ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U′′ + ( ˆ
A
A
AT
14 ˆ
A
A
AT
15 ˆ
A
A
AT
19 +2ˆ
A
A
AT
21)U
U
U
+( ˆ
A
A
AT
16 ˆ
A
A
AT
22)U
U
U(ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V′′
+( ˆ
A
A
A5ˆ
A
A
AT
5+2ˆ
A
A
A11 2ˆ
A
A
AT
11 +ˆ
A
A
A25 ˆ
A
A
AT
25 2ˆ
A
A
A27 +2ˆ
A
A
AT
27)V
V
V
+( ˆ
A
A
A6+ˆ
A
A
A12 2ˆ
A
A
A28
ω
2
0ˆ
B
B
B2
ω
2
0ˆ
B
B
B3)V
V
V=0
0
0,
(10.2.44)
δ
U
U
UT[ˆ
A
A
A1U
U
U+ˆ
A
A
AT
2U
U
U+ ( ˆ
A
A
A13 2ˆ
A
A
A18)V
V
V+ ( ˆ
A
A
A14 ˆ
A
A
A19)V
V
V]x=0,l=0
δ
V
V
VT[( ˆ
A
A
AT
13 2ˆ
A
A
AT
18)U
U
U+ ( ˆ
A
A
AT
15 2ˆ
A
A
AT
21)U
U
U+ ( ˆ
A
A
A4+4ˆ
A
A
A94ˆ
A
A
A24)V
V
V
+( ˆ
A
A
AT
5+2ˆ
A
A
AT
11 ˆ
A
A
A25 2ˆ
A
A
AT
27)V
V
V]x=0,l=0
(10.2.45)
With these equations given above the global free vibration analysis of prismatic
beam shaped thin-walled plate structures can be done sufficient exactly.
10.3 Solution Procedures
Two general kinds of solution procedures may be taken into account
analytic solutions and
numerical solutions
The consideration below distinguish exact and approximate analytical solution pro-
cedures. In the first case an exact solution of the differential equations is carried
out. In the other case, the variational statement of the problem is, e.g., solved by
the Ritz or Galerkin method, and in general, the procedures yield in an approximate
analytical series solution.
Numerical solution procedures essentially consist of methods outgoing from the
differential equation or from the corresponding variational problem. The numeri-
cal solutions of differential equations may include such methods as finite difference
methods, Runge3-Kutta4methods and transfer matrix methods. The main represen-
tative for the second way is the finite element method (FEM). After a few remarks
3Carl David Tolm´e Runge (30 August 1856 Bremen †3 January 1927 ottingen) - mathemati-
cian and physicist
4Martin Wilhelm Kutta (3 November 1867 Pitschen †25 December 1944 F¨urstenfeldbruck) -
mathematician
10.3 Solution Procedures 391
in Sect. 10.3.1 about analytic solution possibilities for the here considered prob-
lems, the numerical solution procedure using the transfer matrix method is consid-
ered in detail in Sect. 10.3.2. The application of the FEM and the development of
special one-dimensional finite elements for beam shaped thin-walled structures are
discussed in Chap. 11.
10.3.1 Analytical Solutions
For the generalized beam models given in Sect. 10.2 only for simplified special
cases analytical solutions are possible. If we use, for example, the structure model
D in connection with a symmetric cross-ply stacking in all plates, what means
that the differential equations are from the same type as in case of the isotropic
semi-moment shell theory of Vlasov, analytical solutions can be developed for spe-
cial cross-sections geometry. It is very useful to choose orthogonal generalized co-
ordinate functions
ϕ
ϕ
ϕ
,
ψ
ψ
ψ
,
ξ
ξ
ξ
, because it yields the possibility of decomposition of the
system of differential equations into some uncoupled partial systems. For example,
the generalized co-ordinate functions
ϕ
ϕ
ϕ
in Fig. 10.4 are completely orthogonal and
in this way the matrix ˆ
A
A
A1is a diagonal matrix. Therefore some couplings between
the single differential equations vanish.
A suitable method for construction an exact solution is the Krylow5method or
the so-called method of initial parameters. The first step for the application of this
method is to convert the system of differential equations into an equivalent differen-
tial equation of n-th order.
L[y(x)] =
n
ν
=0
a
ν
y(
ν
)(x) = r(x)(10.3.1)
Its homogeneous solution shall be written as
yh(x) = y(0)K1(x) + y(0)K2(x) + ...+y(n1)(0)Kn(x)(10.3.2)
The free constants of the solution are expressed by the initial constants, i.e. the
function y(x)and its derivatives till the (n1)th order at x=0. A particular solution
can be obtained with
yp(x) =
x
Z
0
Kn(xt)r(t)dt(10.3.3)
or in case that r0(x)is not defined for x0 but for xx0
5Aleksei Nikolajewitsch Krylow (3 August 1863jul./15 August 1863greg.Wisjaga - †26 October
1945 Leningrad) - naval engineer, applied mathematician
392 10 Modelling and Analysis of Thin-walled Folded Structures
yp(x) =
x>x0
x
Z
x0
Kn(xt)r0(t)dt(10.3.4)
Equation (10.3.4) is a quasi closed analytical solution for the differential equation
of the structure model D and different functions ri(x)for respectively x>xi,i=
0,1,...,n
y(x) = y(0)K1(x) + y(0)K2(x) + ...+y(n1)(0)Kn(x)
+
x>x0
x
Z
x0
Kn(xt)r0(t)dt+
x>x1
x
Z
x1
Kn(xt)r1(t)dt(10.3.5)
+
x>x2
x
Z
x2
Kn(xt)r2(t)dt+...
Complete closed analytical solutions for isotropic double symmetric thin-walled
box-girders and general loads one can find in (Altenbach et al, 1994). Also analytical
solution for a two-cellular box-girder including shear lag effects is given there. But
in the majority of engineering applications refer to general laminated thin-walled
structures, an analytical solution has to be ruled out.
10.3.2 Transfer Matrix Method
The differential equations and their boundary conditions are the starting point of a
numerical solution by transfer matrix method. At first the system of higher order dif-
ferential equations has to transfer into a system of differential equations of first order
using the natural boundary conditions as definitions of generalized cross-sectional
forces.
For sake of simplicity this solution method shall be demonstrated for the structure
model D and for a symmetric cross-ply stacking in all plates of the structure. Then
the following system of differential equations and boundary conditions are valid,
see also Sect. 10.2.4.6 and Eqs. (10.2.32) and (10.2.33)
ˆ
A
A
A1U
U
U′′ ˆ
A
A
A3U
U
Uˆ
A
A
A15V
V
V+f
f
fx=0
0
0,
ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V′′ ˆ
A
A
A12V
V
V+f
f
fs+f
f
fn=0
0
0,(10.3.6)
δ
U
U
UThˆ
A
A
A1U
U
U±r
r
rxix=0,l=0,
δ
V
V
VThˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V±r
r
rs±r
r
rnix=0,l=0(10.3.7)
Equations (10.3.7) leads to the definitions of the generalized cross-sectional forces,
i.e. generalized longitudinal forces and transverse forces.
10.3 Solution Procedures 393
P
P
P=ˆ
A
A
A1U
U
U,(10.3.8)
Q
Q
Q=ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V,(10.3.9)
It can be shown that we have with Eqs. (10.3.8), (10.3.9) really the definitions of
generalized forces
P
P
P=
(i)
di
Z
0
ϕ
ϕ
ϕσ
xitidsi=
(i)
di
Z
0
ϕ
ϕ
ϕ
Nxidsi,
Q
Q
Q=
(i)
di
Z
0
ψ
ψ
ψτ
xsitidsi=
(i)
di
Z
0
ψ
ψ
ψ
Nxsidsi
In the here considered structure model we have only membrane stresses
σ
xiand
τ
xsi
because the longitudinal curvatures and longitudinal bending moments are neglected
in all plates. Additional with cross-ply stacking are A16 =A26 =0. Therefore and
with Eqs. (10.2.4), (10.2.6), (10.2.11) we can write
Nxi=A11i
ε
xi=A11iu
i=A11i
ϕ
ϕ
ϕ
TU
U
U,
Nxsi=A66i
ε
xsi=A66i(u
i+v
i) = A66i(
ϕ
ϕ
ϕ
TU
U
U+
ψ
ψ
ψ
TV
V
V)
Considering (10.2.12) we obtain again the definitions of the generalized forces given
in (10.3.8) and (10.3.9)
P
P
P=
(i)
di
Z
0
A11i
ϕ
ϕ
ϕϕ
ϕ
ϕ
TdsiU
U
U=ˆ
A
A
A1U
U
U,
Q
Q
Q=
(i)
di
Z
0
A66i(
ψ
ψ
ψϕ
ϕ
ϕ
TU
U
U+
ψ
ψ
ψψ
ψ
ψ
TV
V
V)dsi=ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V
The inversion of the Eqs. (10.3.8) and (10.3.9) leads to
U
U
U=ˆ
A
A
A1
1P
P
P,(10.3.10)
V
V
V=ˆ
A
A
A1
4ˆ
A
A
AT
15U
U
U+ˆ
A
A
A1
4Q
Q
Q(10.3.11)
With the first derivativesof Eqs. (10.3.8) and (10.3.9) and after the input of (10.3.10)
and (10.3.11) into (10.3.6) we obtain the following system of differential equations
of first order
U
U
U=ˆ
A
A
A1
1P
P
P,
V
V
V=ˆ
A
A
A1
4ˆ
A
A
AT
15U
U
U+ˆ
A
A
A1
4Q
Q
Q,
P
P
P= ( ˆ
A
A
A3ˆ
A
A
A15 ˆ
A
A
A1
4ˆ
A
A
AT
15)U
U
U+ˆ
A
A
A15 ˆ
A
A
A1
4Q
Q
Qf
f
fx,
Q
Q
Q=ˆ
A
A
A12V
V
Vf
f
fsf
f
fn
(10.3.12)
respectively written in matrix notation
394 10 Modelling and Analysis of Thin-walled Folded Structures
U
U
U
V
V
V
P
P
P
Q
Q
Q
1
=
0
0
0 0
0
0ˆ
A
A
A1
10
0
0o
o
o
ˆ
A
A
A1
4ˆ
A
A
AT
15 0
0
0 0
0
0ˆ
A
A
A1
4o
o
o
(ˆ
A
A
A3ˆ
A
A
A15 ˆ
A
A
A1
4ˆ
A
A
AT
15)0
0
0 0
0
0ˆ
A
A
A15 ˆ
A
A
A1
4f
f
fx
0
0
0ˆ
A
A
A12 0
0
0 0
0
0f
f
fsf
f
fn
o
o
oTo
o
oTo
o
oTo
o
oT0
U
U
U
V
V
V
P
P
P
Q
Q
Q
1
y
y
y=B
B
By
y
y(10.3.13)
B
B
Bis the system matrix and y
y
ythe so-called state vector containing all the generalized
displacement functions U
U
Uand V
V
Vand all the generalized forces P
P
Pand Q
Q
Q. 0
0
0 and o
o
oin
the B
B
B-matrix are null matrices and vectors.
The next step is a discretization of the one-dimensional problem, see Fig. 10.6.
Between the state vectors at the point j+1 and the point jwe have generally the
relationship
y
y
yj+1=W
W
Wjy
y
yj(10.3.14)
where W
W
Wjis the transfer matrix for the structure section j(j+1).
A first order differential equation
y(x) = b y(x),b=const
has the solution
y(x) = Cebx
and with
y(x0) = Cebx0C=y(x0)ebx0,
we obtain
y(x) = y(x0)eb(xx0)
In the same way the solution of the matrix differential equations is
y
y
y(x) = B
B
By
y
y(x),
y
y
y(x) = y
y
y(x0)eB(xx0)
j
x=x0x=xN
0N
j1jj+1
Fig. 10.6 Discretization of the one-dimensional structure
10.3 Solution Procedures 395
and we find that eB
B
B(xx0)can be defined as the transfer matrix from the point x0to x.
Therefore the transfer matrix between two points xjand xj+1generally is obtained
as
W
W
Wj=eB
B
B(xj+1xj)(10.3.15)
The numerical calculation of transfer matrices can be carried out by series develop-
ment of the exponential function
W
W
Wj=I
I
I+jB
B
B+2
j
2! B
B
B2+3
j
3! B
B
B3+... (10.3.16)
and also by using a Runge-Kutta method
W
W
Wj=I
I
I+j
6(M
M
M1j+2M
M
M2j+2M
M
M3j+M
M
M4j),(10.3.17)
M
M
M1j=B
B
B(xj),
M
M
M2j=B
B
B(xj+1
2j)(I
I
I+1
2jM
M
M1j),
M
M
M3j=B
B
B(xj+1
2j)(I
I
I+1
2jM
M
M2j),
M
M
M4j=B
B
B(xj+j)(I
I
I+jM
M
M3j)
In both equations I
I
Iare unit matrices of the same rank as the system matrix. The
boundary conditions of the problem can be expressed by a matrix equation
y
y
y0=A
A
Ax
x
x(10.3.18)
Here A
A
Ais the so-called start matrix containing the boundary conditions at x=x0
and x
x
xis the vector of the unknown boundary values there. In the last column of the
start matrix the known boundary values are included. For the unknown boundary
values the last column elements are zero, and by a unit in the corresponding row
the unknown value is associated with an element of the unknown vector x
x
x. For
example, in Eq. (10.3.19) a start matrix is shown in case of a free structure end, it
means all the displacements are unknown and all forces are given.
396 10 Modelling and Analysis of Thin-walled Folded Structures
U1
...
Um
V1
...
Vn
P
1
...
P
m
Q1
...
Qn
1
0
=
1 0
... 0
1 0
1 0
... 0
1 0
P
10
...
P
m0
Q10
...
Qn0
1
x
1
...
...
...
...
x
m+n
1
,
y
y
y0=A
A
A0x
x
x
(10.3.19)
Now the multiplications with the transfer matrices can be carried out over all sec-
tions (xj,xj+1)until x=xN. With the equation
S
S
Sy
y
yN=0
0
0 (10.3.20)
the boundary conditions are formulated at x=xN.S
S
Sis the so-called end matrix con-
taining in its last column the negative values of the given displacements or forces. A
unit in an other column of each row yields the association to an element of the state
vector y
y
yN. Equation (10.3.21) shows the end matrix for a clamped end, where all
displacements are given. This matrix equation leads to a system of linear equations
for the unknowns in the vector x
x
x
1... U1N
... ... ...
1... UmN
1... V1N
... ... ...
1... VnN
U1
...
Um
V1
...
Vn
P
1
...
P
m
Q1
...
Qn
1
N
=
0
...
0
0
...
0
(10.3.21)
The real parts of the eigenvalues of the system matrix B
B
Blead to numerical instable
solutions especially for long beam structures. From the mechanical point of view it
means that the influence of the boundary conditions of both structure ends to each
other are very low and with this we have a nearly singular system of linear equa-
10.3 Solution Procedures 397
tions. For the consolidation of this problem intermediate changes of the unknowns
are carried out, by formulation of a new start matrix A
A
Aat such an intermediate point.
Usually the generalized displacements are chosen as the new unknowns. The fol-
lowing equations show the general procedure schedule
y
y
y0=A
A
A0x
x
x
0
y
y
y1=W
W
W0y
y
y0=W
W
W0A
A
A0x
x
x
0
...
y
y
yi=W
W
Wi1W
W
Wi2...W
W
W0A
A
A0x
x
x
0=F
F
Fix
x
x
0
y
y
yi=A
A
A1x
x
x
1first change of unknowns
y
y
yi+1=W
W
Wiy
y
yi=W
W
WiA
A
A1x
x
x
1
...
y
y
yj=F
F
Fjx
x
x
l1
y
y
yj=A
A
Alx
x
x
llth change of unknowns
y
y
yj+1=W
W
WjA
A
Alx
x
x
l
...
y
y
yk=F
F
Fkx
x
x
n1
y
y
yk=A
A
Anx
x
x
nnth change of unknowns
y
y
yk+1=W
W
WkA
A
Anx
x
x
n
... (10.3.22)
y
y
yN=F
F
FNx
x
x
n
S
S
Sy
y
yN=0
0
0 system of linear equations for the solution of the unknowns x
n
The multiplications of the state vector y
y
y0with transfer matrices are carried out until
the first intermediate change of unknowns. The product of the transfer matrices and
the start matrix makes the matrix F
F
Fi. A new unknown vector x
x
x
1is defined by the new
start matrix A
A
A1and this procedure is repeated at the following intermediate points.
General for the l-th intermediate change of unknowns the Eq. (10.3.23) is current
F
F
Fjx
x
x
l1=A
A
Alx
x
x
l(10.3.23)
With a segmentation of the state vector y
y
yjinto the sub-vectors y
y
yvfor the displace-
ments and y
y
ykfor the forces
398 10 Modelling and Analysis of Thin-walled Folded Structures
y
y
yj=
U
U
U
V
V
V
P
P
P
Q
Q
Q
1
j
=
y
y
yv
y
y
yk
1
j
,(10.3.24)
we obtain a separated form of Eq. (10.3.23)
F
F
F1jf
f
f1j
F
F
F2jf
f
f2j
o
o
oT1
˜
x
x
x
l1
1=
A
A
A1la
a
a1l
A
A
A2la
a
a2l
o
o
oT1
˜
x
x
x
l
1(10.3.25)
With the assumption that the displacements are the new unknowns we find that the
sub-matrix A
A
A1lis a unit matrix and the sub-vector a
a
a1lis a null vector
A
A
A1l=I
I
I,a
a
a1l=o
o
o(10.3.26)
This leads to
F
F
F1j˜
x
x
x
l1+f
f
f1j=˜
x
x
x
l,(10.3.27)
˜
x
x
x
l1=F
F
F1
1j(˜
x
x
x
lf
f
f1j),x
x
x
l1=F
F
F1
1j(˜
x
x
x
lf
f
f1j)
1(10.3.28)
and than the second equation of (10.3.25) yields the structure of the new start matrix
F
F
F2jF
F
F1
1j(˜
x
x
x
lf
f
f1j) + f
f
f2j=A
A
A2l˜
x
x
x
l+a
a
a2l,
A
A
A2l=F
F
F2jF
F
F1
1j,a
a
a2l=f
f
f2jF
F
F2jF
F
F1
1jf
f
f1j,(10.3.29)
A
A
Al=
I
I
I o
o
o
F
F
F2jF
F
F1
1jf
f
f2jF
F
F2jF
F
F1
1jf
f
f1j
o
o
oT1
(10.3.30)
At such an intermediate change point it is also possible to consider the introduction
of concentrated generalized forces or the disposition of supports with given general-
ized displacements. Than the new start matrix must be modified additionally, in the
first case by a modification of the sub-vector a
a
a2land in the second case by consid-
eration of the jump behavior of the forces at this point. But more details about this
shall not be given here.
With the end matrix S
S
Sand the end state vector y
y
yNthe relationship S
S
Sy
y
yNyields
a system of linear equations for the last unknown vector x
x
x
nand after this all the
unknown vectors can be calculated by repeatedly using Eq. (10.3.28).
The transfer matrix method with intermediate changes of the unknown state vec-
tors yields in contrast to the classical transfer method a numerical stable procedure
also for long beam structures. From the mechanical point of view correspond each
intermediate change x=x
la substitution of the structure section 0 xx
lby
generalized elastic springs.
10.4 Problems 399
The transfer matrix procedure is also applicable to the analysis of eigen-
vibrations. There we have a modified system matrix B
B
Bcontaining frequency de-
pendent terms
y
y
y=B
B
B(
ω
0)y
y
y(10.3.31)
Therefore, the transfer matrices can be calculated only with assumed values for the
frequencies. The end matrix leads to a homogenous system of linear equations. Its
coefficient determinant must be zero. The assumed frequencies are to vary until this
condition is fulfilled sufficiently.
The transfer matrix method with numerical stabilization was applied successfully
to several isotropic thin-walled box-beam structures. The structure model D con-
sidered above has for a symmetrical cross-ply stacking of all plates an analogous
mathematical model structure as isotropic semi-moment shell structures. Therefore,
the procedure can be simply transferred to such laminated thin-walled beam struc-
tures. An application to other structure models, Sect. 10.2, is in principle possible
but rather expansive and not efficient.
The development and application of special finite elements and their implemen-
tation in a FEM-program system is more generally and more efficiently. FEM will
be discussed in detail in Chap. 11.
10.4 Problems
Exercise 10.1. Establish the system of differential equations for the box-girder with
a rectangular cross-section, which is shown in Fig. 10.4. It shall be supposed that its
dimensions are symmetric to both axes and therefore we have here
t1=t3=tS,
t2=t4=tG,
d1=d3=dS,
d2=d4=dG,
Further we have a cross-ply stacking in all the plate strips. The stiffness of both
horizontally arranged strips (index G) are the same, but they are different from the
stiffness of the vertically arranged strips (index S), what means that
A11 1 =A11 3 =A11 S ,
A11 2 =A11 4 =A11 G ,
A66 1 =A66 3 =A66 S ,
A66 2 =A66 4 =A66 G ,
D22 1 =D22 3 =D22 S,
D22 2 =D22 4 =D22 G
For the calculation of this box girder the simplified structure model D shall be used
and because we have cross-ply stacking, the Eqs. (10.2.32) are valid
400 10 Modelling and Analysis of Thin-walled Folded Structures
ˆ
A
A
A1U
U
U′′ ˆ
A
A
A3U
U
Uˆ
A
A
A15V
V
V+f
f
fx=0
0
0,
ˆ
A
A
AT
15U
U
U+ˆ
A
A
A4V
V
V′′ ˆ
A
A
A12V
V
V+f
f
fs+f
f
fn=0
0
0
Solution 10.1. At first we have to calculate the matrices ˆ
A
A
A1,ˆ
A
A
A3,ˆ
A
A
A4,ˆ
A
A
A12 and ˆ
A
A
A15,
their definitions are given in Eq. (10.2.12)
ˆ
A
A
A1=
(i)
di
Z
0
A11 i
ϕ
ϕ
ϕϕ
ϕ
ϕ
Tdsi,
ˆ
A
A
A3=
(i)
di
Z
0
A66 i
ϕ
ϕ
ϕ
ϕ
ϕ
ϕ
Tdsi,
ˆ
A
A
A4=
(i)
di
Z
0
A66 i
ψ
ψ
ψψ
ψ
ψ
Tdsi,
ˆ
A
A
A12 =
(i)
di
Z
0
D22 i
ξ
ξ
ξ
••
ξ
ξ
ξ
••Tdsi,
ˆ
A
A
A15 =
(i)
di
Z
0
A66 i
ϕ
ϕ
ϕ
ψ
ψ
ψ
Tdsi,
The co-ordinate functions
ϕ
ϕ
ϕ
,
ψ
ψ
ψ
,
ξ
ξ
ξ
are also shown in Fig. 10.4. For solving the inte-
grals to obtain the ˆ
A
A
A-matrices, the functions
ϕ
ϕ
ϕ
,
ψ
ψ
ψ
,
ξ
ξ
ξ
must be written as functions of
the co-ordinates siof each strip. In accordance with Fig. 10.4 we find
ϕ
1(si) = +1,i=1,2,3,4;
ϕ
1(si) = 0,i=1,2,3,4;
ϕ
2(s1) = dS
212s1
dS;
ϕ
2(si) =
ψ
2(si),i=1,2,3,4;
ϕ
2(s2) = dS
2;
ϕ
2(s3) = + dS
212s3
dS;
ϕ
2(s4) = dS
2;
ϕ
3(s1) = dG
2;
ϕ
3(si) =
ψ
3(si),i=1,2,3,4;
ϕ
3(s2) = + dG
212s2
dG;
ϕ
3(s3) = dG
2;
ϕ
3(s4) = dG
212s4
dG;
10.4 Problems 401
ϕ
4(s1) = + dSdG
412s1
dS;
ϕ
4(si) =
ψ
4(si),i=1,2,3,4;
ϕ
4(s2) = dSdG
412s2
dG;
ϕ
4(s3) = + dSdG
412s3
dS;
ϕ
4(s4) = dSdG
412s4
dG;
ψ
1(s1) = dG
2;
ψ
2(s1) = +1;
ψ
1(s2) = dS
2;
ψ
2(s2) = 0;
ψ
1(s3) = dG
2;
ψ
2(s3) = 1;
ψ
1(s4) = dS
2;
ψ
2(s4) = 0;
ψ
3(s1) = 0;
ψ
4(s1) = dG
2;
ψ
3(s2) = 1;
ψ
4(s2) = + dS
2;
ψ
3(s3) = 0;
ψ
4(s3) = dG
2;
ψ
3(s4) = +1;
ψ
4(s4) = + dS
2;
ξ
1(s1) = dS
212s1
dS;
ξ
••
1(si) = 0,i=1,2,3,4;
ξ
1(s2) = dG
212s2
dG;
ξ
1(s3) = dS
212s3
dS;
402 10 Modelling and Analysis of Thin-walled Folded Structures
ξ
1(s4) = dG
212s4
dG;
ξ
2(s1) = 0;
ξ
••
2(si) = 0,i=1,2,3,4;
ξ
2(s2) = +1;
ξ
2(s3) = 0;
ξ
2(s4) = 1;
ξ
3(s1) = +1;
ξ
••
3(si) = 0,i=1,2,3,4;
ξ
3(s2) = 0;
ξ
3(s3) = 1;
ξ
3(s4) = 0;
Some additional considerations are necessary to determine the functions
ξ
4(si). The
generalized co-ordinate function
ξ
4is corresponding to
ψ
4and represents therefore
a double antisymmetric deflection state of the cross-section. The cross-section is
double symmetric in its geometry and in the elastic behavior. Therefore we must
have an antisymmetric function
ξ
4(si)in each strip. It means that the following
conditions are valid
ξ
4(si=0) =
ξ
i0,
ξ
4(si=di) =
ξ
i0,
ξ
4(si=0) =
α
0,
ξ
4(si=di) =
α
0,
ξ
••
4(si=0) =
κ
si0,
ξ
••
4(si=di) =
κ
si0
Supposing a polynomial function of the third order, we can write
ξ
••
4(si) =
κ
si012si
di,
ξ
4(si) =
κ
si0di"si
disi
di2#+
α
0,
ξ
4(si) =
κ
si0d2
i
6"3si
di2
2si
di3#+
α
0di
si
di
+
ξ
i0
The condition
ξ
4(si=di) =
ξ
i0leads to
α
0di=
κ
si0d2
i
6+2
ξ
i0
and than we obtain
ξ
4(si) =
κ
si0d2
i
6"3si
di2
2si
di3
si
di#+
ξ
i012si
di
With the antisymmetric properties mentioned above we find
10.4 Problems 403
κ
s10=
κ
s30=
κ
S0,
κ
s20=
κ
s40=
κ
G0
The continuity of the rotation angles at the corners and the equilibrium equation
ξ
4(s1=dS) =
ξ
4(s2=0),
ξ
4(s2=dG) =
ξ
4(s3=0),
ξ
4(s3=dS) =
ξ
4(s4=0),
ξ
4(s4=dG) =
ξ
4(s1=0),
D22S
κ
S0 =D22G
κ
G0
lead to the unknown curvatures
κ
S0 =12D22G
dGD22S +dSD22G
,
κ
G0 = + 12D22S
dGD22S +dSD22G
and we obtain
ξ
4(s1) = 2D22Gd2
S
dGD22S +dSD22G "3s1
dS2
2s1
dS3
s1
dS#
+dS
212s1
dS,
ξ
4(s2) = 2D22Sd2
G
dGD22S +dSD22G "3s2
dG2
2s2
dG3
s2
dG#
dG
212s2
dG,
ξ
4(s3) = 2D22Gd2
S
dGD22S +dSD22G "3s3
dS2
2s3
dS3
s3
dS#
+dS
212s3
dS,
ξ
4(s4) = 2D22Sd2
G
dGD22S +dSD22G "3s4
dG2
2s4
dG3
s4
dG#
dG
212s4
dG,
ξ
••
4(s1) = 12D22G
dGD22S +dSD22G 12s1
dS,
ξ
••
4(s2) = + 12D22S
dGD22S +dSD22G 12s2
dG,
ξ
••
4(s3) = 12D22G
dGD22S +dSD22G 12s3
dS,
ξ
••
4(s4) = + 12D22S
dGD22S +dSD22G 12s4
dG
404 10 Modelling and Analysis of Thin-walled Folded Structures
Now all elements of the matrices can be calculated. Here only the calculation of the
element ˆ
A122 of the matrix ˆ
A
A
A1shall be derived in a detailed manner
ˆ
A111 =
4
i=1
di
Z
0
A11i
ϕ
1(si)
ϕ
1(si)dsi=2(A11SdS+A11GdG),
ˆ
A112 =
4
i=1
di
Z
0
A11i
ϕ
1(si)
ϕ
2(si)dsi=0,
ˆ
A113 =
4
i=1
di
Z
0
A11i
ϕ
1(si)
ϕ
3(si)dsi=0,
ˆ
A114 =
4
i=1
di
Z
0
A11i
ϕ
1(si)
ϕ
4(si)dsi=0,
ˆ
A122 =
4
i=1
di
Z
0
A11i
ϕ
2(si)
ϕ
2(si)dsi
=A11S
d2
S
4
dS
Z
012s1
dS2
ds1+A11G
d2
S
4
dG
Z
0
ds2
+A11S
d2
S
4
dS
Z
012s3
dS2
ds3+A11G
d2
S
4
dG
Z
0
ds4
=A11S
d3
S
12 +A11G
d2
S
4dG+A11S
d3
S
12 +A11G
d2
S
4dG
=d2
S
6(A11SdS+3A11GdG),
ˆ
A123 =
4
i=1
di
Z
0
A11i
ϕ
2(si)
ϕ
3(si)dsi=0,
ˆ
A124 =
4
i=1
di
Z
0
A11i
ϕ
2(si)
ϕ
4(si)dsi=0,
ˆ
A133 =
4
i=1
di
Z
0
A11i
ϕ
3(si)
ϕ
3(si)dsi=d2
G
6(3A11SdS+A11GdG),
ˆ
A134 =
4
i=1
di
Z
0
A11i
ϕ
3(si)
ϕ
4(si)dsi=0,
ˆ
A144 =
4
i=1
di
Z
0
A11i
ϕ
4(si)
ϕ
4(si)dsi=d2
Sd2
G
24 (A11SdS+A11GdG)
10.4 Problems 405
With all elements ˆ
A1i j we obtain the matrix ˆ
A
A
A1to
ˆ
A
A
A1=
A111 0 0 0
0A122 0 0
0 0 A133 0
0 0 0 A14 4
with
A111 =2(A11SdS+A11GdG),
A122 =d2
S
6(A11SdS+3A11GdG),
A133 =d2
G
6(3A11SdS+3A11GdG),
A144 =d2
Sd2
G
24 (A11SdS+3A11GdG)
One can see that the generalized co-ordinate functions
ϕ
iare orthogonal to each
other and therefore the matrix ˆ
A
A
A1is a diagonal matrix
In the same way the other ˆ
A
A
Aimatrices are obtained
ˆ
A
A
A3=
0 0 0 0
0 2A66SdS0 0
002A66GdG0
0 0 0 dSdG
2(A66SdG+3A66GdS)
,
ˆ
A
A
A4=
A0 0 0
0 2A66SdS0 0
0 0 2A66GdG0
B0 0 A
,
ˆ
A
A
A12 =
0 0 0 0
0 0 0 0
0 0 0 0
dSdG
2(A66SdGA66GdS)0 0 96D22GD22S
dGD22S+dSD22G
,
ˆ
A
A
A15 =
0 0 0 0
0 2A66SdS0 0
0 0 2A66GdG0
B0 0 A
with
A=dSdG
2(A66SdG+A66GdS),B=dSdG
2(A66SdGA66GdS)
Now the system of differential equations can be developed with the help of Eq.
(10.2.32).
406 10 Modelling and Analysis of Thin-walled Folded Structures
2(A11SdS+A11GdG)U′′
1=fx1,
d2
S
6(A11SdS+3A11GdG)U′′
22A66SdS(U2+V
2) = fx2,
2A66SdS(U
2+V′′
2) = (fs2+fn2),
d2
G
6(3A11SdS+A11GdG)U′′
32A66GdG(U3+V
3) = fx3,
2A66GdG(U
3+V′′
3) = (fs3+fn3),
d2
Sd2
G
24 (A11SdS+A11GdG)U′′
4
dSdG
2(A66SdG+A66GdS)(U4+V
4)
dSdG
2(A66SdGA66GdS)V
1=fx4,
dSdG
2(A66SdGA66GdS)(U
4+V′′
4)
+dSdG
2(A66SdG+A66GdS)V′′
1=(fs1+fn1),
dSdG
2(A66SdG+A66GdS)(U
4+V′′
4)
+dSdG
2(A66SdGA66GdS)V′′
1
96D22GD22S
dGD22S+dSD22G
V4=(fs4+fn4)
We can see, that the system of differential equations is divided into four decou-
pled partial systems. The first equation describes the longitudinal displacement, the
second and the third partial systems represent the bending about the global y- and
z-axes and the fourth - the torsion, the warping and the contour deformation of the
cross-section. An analytic solution of the fourth partial system is more difficult like
the solutions of the first three partial systems but it is possible too. The analytical
solution of an analogous system for an isotropic box girder is given in detail by
Vlasov and by the authors of this book, see also the remarks in 10.3.1.
References
Altenbach J, Kissing W, Altenbach H (1994) unnwandige Stab- und Stabschalen-
tragwerke. Vieweg-Verlag, Brauschweig/Wiesbaden
Vlasov VZ (1961) Thin-walled elastic beams. National Science Foundation and De-
partment of Commerce, Arlington, VI
Wlassow WS (1958) Allgemeine Schalentheorie und ihre Anwendung in der Tech-
nik. Akademie-Verlag, Berlin
Part V
Finite Classical and Generalized Beam
Elements, Finite Plate Elements
The fifth part (Chap. 11) presents a short introduction into the finite element proce-
dures and developed finite classical and generalized beam elements and finite plate
elements in the frame of classical and first order shear deformation theory. Selected
examples demonstrate the possibilities of finite element analysis.
Chapter 11
Finite Element Analysis
The Finite Element Method (FEM) is one of the most effective methods for the nu-
merical solution of field problems formulated in partial differential equations. The
basic idea of the FEM is a discretization of the continuous structure into substruc-
tures. This is equivalent to replacing a domain having an infinite number of degrees
of freedom by a system having a finite number of degrees of freedom. The actual
continuum or structure is represented as an assembly of subdivisions called finite
elements. These elements are considered to be interconnected at specified joints
which are called nodes. The discretization is defined by the so-called finite element
mesh made up of elements and nodes.
We assume one-dimensional elements, when one dimension is very large in com-
parison with the others, e.g. truss or beam elements, two-dimensional elements,
when one dimension is very small in comparison with the others, e.g. plate or shell
elements, and volume elements. From the mechanical point of view the nodes are
coupling points of the elements, where the displacements of the coupled elements
are compatible. On the other hand from the mathematical point of view the nodes are
the basic points for the approximate functions of the displacements inside a finite el-
ement and so at these nodes the displacements are compatible. It must be noted here
that all considerations are restricted to the displacement method. The force method
or hybrid methods are not considered in this book.
An important characteristic of the discretization of a structure is the number of
degrees of freedom. To every node, a number of degrees of freedom will be assigned.
These are nodal constants which usually (but not necessarily) have a mechanical or
more general physical meaning. The number of degrees of freedom per element is
defined by the product of the number of nodes per element and degrees of freedom
per node. The number of degrees of freedom in the structure is the product of the
number of nodes and the number of degrees of freedom per node.
Chapter 11 contains an introduction to the general procedure of finite element
analysis in a condensed form (Sect. 11.1). For more detailed information see the
vast amount of literature. In Sects. 11.2 and 11.3 the development of finite beam
elements and finite plate elements for the analysis of laminate structures is given.
Section 11.4 contains the development of generalized finite beam elements based on
409
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0_11
410 11 Finite Element Analysis
a generalized structure model for beam shaped thin-walled folded structures given
in Sect. 10.2. In Sect. 11.5 the results of some numerical applications show the
influences of chosen parameters on the behavior of laminate structures.
11.1 Introduction
The principle of the total minimum potential energy and the Hamilton‘s principle
are given in Sect. 2.2.2 in connection with analytical variational approaches, they are
also the theoretical basis of the FEM solutions of elastostatic and of dynamic prob-
lems. In this way we have variational problems. For such problems the Ritz method
may be used as a so-called direct solution method (see Sect. 2.2.3). In the classical
Ritz method the approximation functions are defined for the whole structure, and
so for complex geometries it is difficult to realize the requirements of satisfying
the boundary conditions and of the linear independence and completeness of these
functions.
One way to overcome these difficulties is by the discretization of the structure
into a number of substructures, if possible of the same kind (finite elements). Then
the approximation functions can be defined for the elements only and they must sat-
isfy the conditions of geometrical compatibility at the element boundaries. Because
it is usual to define different types of finite elements, we have special types of ap-
proximation functions for each element type. Here the approximation functions are
denoted Ni, the so-called shape functions. They are arranged in a matrix N
N
N, the ma-
trix of the shape functions of the particular element type. The following introduction
to the FEM procedure is given in a general but condensed form and illustrates that
the step-by-step finite element procedure can be stated as follows:
Discretization of the structure,
Selection of a suitable element displacement model,
Derivation of element stiffness matrices and load vectors,
Assembly of element equations to obtain the system equations,
Calculation of the system equations for the unknown nodal displacements,
Computation of element strains and stresses
11.1.1 FEM Procedure
The starting point for elastostatic problems is the total potential energy given in Eq.
(2.2.28). In accordance with the Ritz method the approximation
˜
u
u
u(x
x
x) = N
N
N(x
x
x)v
v
v(11.1.1)
is used for the displacement field vector u
u
u. Here N
N
Nis the matrix of the shape func-
tions, they are functions of the position vector x
x
x, and v
v
vis the element displacement
11.1 Introduction 411
vector. The matrix N
N
Nhas the same number of rows as the displacement vector u
u
u
has components and the same number of columns as the element displacement vec-
tor v
v
vhas components. If the element has nKE nodes and the degree of freedom
for each node is nF, the element displacement vector v
v
vcontains nKE subvectors
v
v
vi,i=1,...,nKE with nFcomponents in each, and so v
v
vhas nKE nFcomponents. The
number of components of the displacement field vector u
u
uis nu. Then the structure
of the matrix N
N
Nis generally
N
N
N= [N1I
I
InuN2I
I
Inu...N˜nI
I
Inu],˜n=nKE nF
nu
(11.1.2)
with I
I
Inuas unit matrices of the size (nu,nu). Therefore the size of N
N
Nis generally
(nu,nKE nF). In dependence on the kind of continuity at the element boundaries, the
so-called C(0)- or C(1)-continuity, see below, two cases can be distinguished. In the
case of C(0)-continuity nFequals nuand therefore ˜nis equal nKE, we have only nKE
shape functions Ni, whereas we can have up to nKE nFshape functions in the case of
C(1)-continuity.
For the stresses and the strains we obtain from (11.1.1)
σ
σ
σ
(x
x
x) = C
C
C
ε
ε
ε
(x
x
x) = C
C
CD
D
DN
N
N(x
x
x)v
v
v,
ε
ε
ε
(x
x
x) = D
D
Du
u
u(x
x
x) = D
D
DN
N
N(x
x
x)v
v
v=B
B
B(x
x
x)v
v
v(11.1.3)
With the approximation (11.1.1) the total potential energy is a function of all the
nodal displacement components arranged in the element displacement vector v
v
v, e.g.
Π
=
Π
(v
v
v). The variation of the total potential energy
δΠ
=
δ
v
v
vT
Z
V
B
B
BTC
C
CB
B
Bv
v
vdVZ
V
N
N
NTp
p
pdVZ
Aq
N
N
NTq
q
qdA
(11.1.4)
leads with
δΠ
=0 to
δ
v
v
vT(K
K
Kv
v
vf
f
fpf
f
fq) = 0 (11.1.5)
K
K
Kis the symmetric stiffness matrix with the size (nK E nF,nKE nF)
K
K
K=Z
V
B
B
BTC
C
CB
B
BdV(11.1.6)
and f
f
fpand f
f
fqare the vectors of the volume forces and the surface forces
f
f
fp=Z
V
N
N
NTp
p
pdV,f
f
fq=Z
Aq
N
N
NTq
q
qdA(11.1.7)
If the components of
δ
v
v
vare independent of each other, we obtain from (11.1.5) a
system of linear equations
K
K
Kv
v
v=f
f
f,f
f
f=f
f
fp+f
f
fq(11.1.8)
412 11 Finite Element Analysis
For elastodynamic problems, we have to consider that forces and displacements are
also dependent on time and the Hamilton‘s principle is the starting point for the
FEM procedure. Assuming again the independence of the components of
δ
v
v
vthe
matrix equation is
M
M
M¨
v
v
v(t) + K
K
Kv
v
v(t) = f
f
f(t)(11.1.9)
for elastic systems without damping effects. M
M
Mis symmetric mass matrix
M
M
M=Z
V
ρ
N
N
NTN
N
NdV(11.1.10)
and f
f
f(t)the vector of the time dependent nodal forces. Assuming the damping pro-
portional to the relative velocities, an additional term C
C
CD˙
v
v
v(t)can be supplemented
formally in Eq. (11.1.9)
M
M
M¨
v
v
v(t) +C
C
CD˙
v
v
v(t) + K
K
Kv
v
v(t) = f
f
f(t),(11.1.11)
where C
C
CDis the damping matrix. C
C
CDhas the same size as the matrices K
K
Kand M
M
M
and usually it is formulated approximately as a linear combination of K
K
Kand M
M
M. The
factors
α
and
β
can be chosen to give the correct damping at two frequencies
C
C
CD
α
M
M
M+
β
K
K
K(11.1.12)
In selecting the shape functions N
N
Ni(x
x
x)it must be remembered that these functions
must be continuous up to the (n1)th derivative, if we have derivatives of the nth
order in the variational problem, i.e. in the total potential energy or in the Hamil-
ton’s function. In this case only the results of FEM approximations converge to the
real solutions by increasing the number of elements. For more-dimensional finite
elements in this way it is to realize that the displacements are compatible up to the
(n1)th derivative at the boundaries of adjacent elements, if they are compatible at
the nodes.
In plane stress or plane strain problems and in general three-dimensional prob-
lems the vector u
u
ucontains displacements only (no rotations) and the differential
operator D
D
Dis of the 1st order. In this way we must only satisfy the displacements
compatibility at the element boundaries that means the so-called C(0)-continuity.
By using beam or plate models especially of the classical Bernoulli beam model
or the classical Kirchhoff plate model, the rotation angles are expressed by deriva-
tives of the displacements of the midline or the midplane and the differential op-
erator D
D
Dis of the second order. Then we have to satisfy the compatibility of dis-
placements and rotations at the element boundaries. In such cases we speak about a
C(1)-continuity and finding the shape functions Niis more difficult.
Because we have no differential operator in connection with the mass matrix M
M
M,
it would be possible to use other, more simple functions N
ifor it. In such a case
the mass matrix would have another population, e.g. a diagonal matrix structure is
possible. Then we speak about a so-called condensed mass matrix, otherwise we
have a consistent mass matrix. By using the condensed mass matrix we have less
11.1 Introduction 413
computational expense than by using the consistent mass matrix, but a decreasing
convergence to the real results is possible.
All equations considered above are only valid for a single element and strictly
they should have an additional index E. For example, we have the inner element
energy
UE=1
2v
v
vT
EZ
VE
B
B
BTC
C
CB
B
BdVv
v
vE=1
2v
v
vT
EK
K
KEv
v
vE(11.1.13)
with the element stiffness matrix
K
K
KE=Z
VE
B
B
BTC
C
CB
B
BdV(11.1.14)
Since the energy is a scalar quantity, the potential energy of the whole structure can
be obtained by summing up the energies of the single elements. Previously a system
displacement vector containing the displacements of all nodes of the whole system
must be defined. By a so-called coincidence matrix L
L
LEthe correct position of each
single element is determ ined. L
L
LEis a Boolean matrix of the size (nKE nF,nKnF) with
nKas the number of nodes of the whole structure.
The element displacement vector v
v
vEis positioned into the system displacement
vector v
v
vby the equation
v
v
vE=L
L
LEv
v
v(11.1.15)
and we obtain the system equation by summing up over all elements
i
L
L
LT
iEK
K
KiEL
L
LiE !v
v
v="
i
L
L
LiE (f
f
fiE p +f
f
fiEq )#
K
K
Kv
v
v=f
f
f
(11.1.16)
The system stiffness matrix is also symmetric, but it is a singular matrix, if the
system is not fixed kinematically, i.e., we have no boundary conditions constraining
the rigid body motion. After consideration of the boundary conditions of the whole
system, K
K
Kbecomes a positive definite matrix and the system equation can be solved.
Then with the known displacements v
v
vthe stresses and deformations are calculated
using the element equations (11.1.1) and (11.1.3).
For elastodynamic problems, the system stiffness matrix and the system mass
matrix are obtained in the same manner and we have the system equation
M
M
M¨
v
v
v(t) +C
C
CD˙
v
v
v(t) + K
K
Kv
v
v(t) = f
f
f(t)(11.1.17)
For investigation of the eigen-frequencies of a system without damping harmonic
vibrations are assumed and with
v
v
v(t) = ˆ
v
v
vcos(
ω
t+
ϕ
)(11.1.18)
and C
C
CD=0
0
0,f
f
f(t) = o
o
othe matrix eigen-value problem follows
414 11 Finite Element Analysis
(K
K
K
ω
2M
M
M)v
v
v=o
o
o(11.1.19)
and the eigen-frequencies and the eigen-vectors characterizing the mode shapes can
be calculated.
11.1.2 Problems
Exercise 11.1. A plane beam problem is given. The beam is divided into three plane
two-node beam elements. The number of nodal degrees of freedom is three (u,w,
ϕ
):
1. What size are the element stiffness matrix and the system stiffness matrix before
the consideration of the boundary conditions?
2. Show the coincidence matrix L
L
L2of the second element lying between the nodes
2 and 3!
3. Show the population of the system stiffness matrix and the boundary conditions,
if the beam is fixed at node 1 (cantilever beam)! Do the same as in the previous
case but consider that the beam is simply supported (node 1 is constrained for the
deflections uand wand node 4 only for the deflection w)!
Solution 11.1. For the plane beam problem one gets
1. With nKE =2 and nF=3 the element stiffness matrix has the size (6,6). Be-
cause we have 4 nodes (nK=4) the size of the system stiffness matrix before the
consideration of the boundary conditions is (12,12).
2. The coincidence matrices in this case have the size (6,12). Because it must be
v
v
v2=L
L
L2v
v
v
we obtain the coincidence matrix for the element Nr. 2
L
L
L2=
0 0 0 1 0 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0 0 0 0 0
0 0 0 0 0 1 0 0 0 0 0 0
0 0 0 0 0 0 1 0 0 0 0 0
0 0 0 0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 0 1 0 0 0
3. The system stiffness matrix without consideration of the boundary conditions is
defined by
K
K
K=
4
i=1
L
L
LT
iK
K
KiL
L
Li
In this case we obtain the following population of the matrix K
K
K
11.2 Finite Beam Elements 415
u1w1
ϕ
1u2w2
ϕ
2u3w3
ϕ
3u4w4
ϕ
4
[(v)] [(v)] [(v)] [(v)] [(v)] [(v)] [(0)] [(0)] [(0)] [(0)] [(0)] [(0)]
[(v)] [(v)] [(v)] [(v)] [(v)] [(v)] [(0)] [(0)] [(0)] [(0)] [(0)] [(0)]
[(v)] [(v)] (v) (v) (v) (v) (0) (0) (0) (0) (0) (0)
[(v)] [(v)] (v)v+x v +x v +x x x x 0[(0)] 0
[(v)] [(v)] (v)v+x v +x v +x x x x 0[(0)] 0
[(v)] [(v)] (v)v+x v +x v +x x x x 0[(0)] 0
[(0)] [(0)] (0)x x x x +z x +z x +z z [z]z
[(0)] [(0)] (0)x x x x +z x +z x +z z [z]z
[(0)] [(0)] (0)x x x x +z x +z x +z z [z]z
[(0)] [(0)] (0)0 0 0 z z z z [z]z
[(0)] [(0)] [(0)] [(0)] [(0)] [(0)] [z] [z] [z] [z] [z] [z]
[(0)] [(0)] [(0)] 0 0 0 z z z z [z]z
u1
w1
ϕ
1
u2
w2
ϕ
2
u3
w3
ϕ
3
u4
w4
ϕ
4
v- components of the stiffness matrix of element No. 1, x- components of the
stiffness matrix of element No. 2, z- components of the stiffness matrix of ele-
ment No. 3.
Considering the boundary conditions for a cantilever beam clamped at node 1
(u1=0,w1=0,
ϕ
1=0) we have to cancel the first three rows and the first three
columns in the obtained matrix - characterized by brackets (...). If we have a
simply supported beam with u1=0,w1=0 and w4=0, the first two rows and
columns and the row and the column No. 11 must be deleted - characterized by
square brackets [...].
11.2 Finite Beam Elements
A beam is a quasi one-dimensional structure, the dimensions of the cross-section
of it are very small in comparison to its length. The connection of the centers of
the cross-sectional areas is called the midline of the beam. We distinguish between
straight beams and beams with an in-plane or spatial curved midline, respectively.
Here we consider beams with a straight midline only.
Generally such a beam can be loaded by tension/compression, one- or two-axial
bending and torsion. Especially with respect to the use of laminate beams the fol-
lowing investigations are restricted to tension/compression and one-axial bending.
For two-axial bending and torsion, laminate beams are not so predestined.
Laminate beams consist of UD-laminae mostly have a rectangular cross-section
of the dimension b(width) and h(hight) and very often the laminae are arranged
symmetrically to the midline. We will assume this special case for the following
development of finite laminate beam elements. In this way we have no coupling of
tension and bending and we can divide our considerations into the development of
laminate elements for tension/compression, so-called laminate truss elements, and
laminate beam elements for bending only.
416 11 Finite Element Analysis
11.2.1 Laminate Truss Elements
The laminate truss element is a very simple element. It is assumed to be a straight
structure of the length lwith a constant cross-sectional area A. The nodal degree of
freedom is one - the displacement uin axial direction (Fig. 11.1). In the potential
energy we have only the first derivative and so we can use a two-node truss element
with linear shape functions Ni(x1)and Nj(x1), which satisfies C(0)-continuity
u(x1) = N
N
Nv
v
vE,v
v
vT
E= [uiuj],N
N
N= [Ni(x1)Nj(x1)] (11.2.1)
The two shape functions (see also Fig. 11.2) are
Ni(x1) = 1x1
l,Nj(x1) = x1
l(11.2.2)
With the stress resultant
N(x1) = Z
A
σ
dA=A11
ε
1(x1) = A11
du
dx1
,A11 =b
n
k=1
C(k)
11 hk(11.2.3)
and the longitudinal load per length n(x1)the total potential energy can be written
as
Π
(u) = 1
2
l
Z
0
A11u2dx1
l
Z
0
n(x1)udx1(11.2.4)
and for the element stiffness matrix we obtain
Fig. 11.1 Laminate truss
element l
x1
n(x1)
uiuj
x1
l
x1
l
0
0
0
0
1.0
1.0
1.0
1.0
0.5
0.5
NiNj
Fig. 11.2 Shape functions of the two-node truss element
11.2 Finite Beam Elements 417
K
K
KE=A11
l
Z
0
N
N
NTN
N
Ndx1=A11
l11
1 1 (11.2.5)
The element force vector is defined as
f
f
fnE =
l
Z
0
N
N
NTn(x1)dx1
If we assume that n(x1)is a linear function with niand njas the intensities at the
nodes
n(x1) = N
N
Nni
nj
then
f
f
fnE =
l
Z
0
N
N
NTN
N
Ndx1ni
nj=l
62 1
1 2 ni
nj(11.2.6)
In case of nodal forces f
f
fPE = [FiFj]T, the vector f
f
fPE must be added to the vector
f
f
fnE
f
f
fE=f
f
fnE +f
f
fPE (11.2.7)
The system equation can be obtained in dependence on the structure of the whole
system, defined by a coincidence matrix together with the transformation of all el-
ement equations into a global coordinate system. Considering the boundary con-
ditions, the system equation can be solved and with the known displacements the
stresses can be calculated for each element.
For vibration analysis, the element mass matrix (11.1.10) has to be used
M
M
ME=Z
V
ρ
N
N
NTN
N
NdV=
l
Z
0
ρ
N
N
NTN
N
Ndx1,
ρ
=1
h
n
k=1
ρ
(k)h(k)
All parts of the cross-section have the same translation uand the corresponding
acceleration ¨umultiplied by the distributed mass produces a distributed axial inertia
force. Instead of handling the distributed mass directly, we generate fictitious nodal
masses contained in the consistent mass matrix
M
M
ME=
ρ
Al
62 1
1 2 (11.2.8)
With the system equation, obtained in the same manner as for elastostatic problems,
the eigen-frequencies and mode shapes can be calculated.
418 11 Finite Element Analysis
11.2.2 Laminate Beam Elements
For the analysis of laminate beams in this book two theories are considered, the
classical laminate theory and the shear deformation theory. The classical laminate
theory is based on the Bernoulli beam model and the shear deformation theory on
the Timoshenko beam model. The Bernoulli beam model neglects the shear strains
in the bending plane and so it seems to be less realistic for the calculation of laminate
beams. Therefore it is better to use the Timoshenko beam model, which includes the
shear strains in a simple form (Chap. 7).
In the following discussion, only the shear deformation theory is used and we
assume a simple rectangular cross-section with a symmetric arrangement of the UD-
laminae. This means that we have no coupling of tension and bending. The main
advantage of the shear deformation theory in comparison with the Bernoulli theory
is that the cross-sectional rotation angle
ψ
is independent of the displacement wand
therefore the differential operator D
D
Din the strain energy is of the 1st order. In this
way we can use elements with C(0)-continuity, and a two-node element with linear
shape functions is possible. The nodal degrees of freedom are 2 (w,
ψ
). In Fig. 11.3
such a two-node beam element is shown. The element displacement vector is
v
v
vT
E= [wi
ψ
iwj
ψ
j](11.2.9)
For the displacement vector u
u
uthe approximation (11.2.11) is used
u
u
u(x1) = w(x1)
ψ
(x1)=N
N
Nv
v
vE,(11.2.10)
where the matrix of the shape functions is
N
N
N=Ni(x1)1 0
0 1 Nj(x1)1 0
0 1  (11.2.11)
with the shape functions (11.2.2), see also Fig. 11.2.
A better element accuracy can be expected, if we consider a three-node element,
as shown in Fig. 11.4. Then the element displacement vector is
v
v
vT
E= [wi
ψ
iwj
ψ
jwk
ψ
k](11.2.12)
and for the matrix N
N
Nwe obtain
Fig. 11.3 Two-node beam
element l
x1
x3
wiwj
ψ
i
ψ
j
11.2 Finite Beam Elements 419
Fig. 11.4 Three-node beam
element
l/2
l/2
x1
x3
wiwjwk
ψ
k
ψ
i
ψ
j
N
N
N=Ni(x1)1 0
0 1 Nj(x1)1 0
0 1 Nk(x1)1 0
0 1  (11.2.13)
with the shape functions
Ni(x1) = 13x1
l+2x2
1
l2,Nj(x1) = 4x1
l4x2
1
l2,Nk(x1) = x1
l+2x2
1
l2,(11.2.14)
which are shown in Fig. 11.5. A further increase in the element accuracy can be
achieved with a four-node element, see Fig. 11.6. Here the element displacement
vector and the matrix of the shape functions are
v
v
vT
E= [wi
ψ
iwj
ψ
jwk
ψ
kwl
ψ
l](11.2.15)
x1
l
x1
l
x1
l
0
0
0
0
0
0
1.0
1.0
1.0
1.0
1.0
1.0
0.5
0.5
0.5
-0.125
-0.125
Ni
Nk
Nj
Fig. 11.5 Shape functions of the three-node element
420 11 Finite Element Analysis
Fig. 11.6 Four-node beam
element
x1
x3,w
l
wiwjwkwl
l/3l/3l/3
ψ
i
ψ
j
ψ
k
ψ
l
N
N
N=Ni(x1)1 0
0 1Nj(x1)1 0
0 1Nk(x1)1 0
0 1Nl(x1)1 0
0 1 (11.2.16)
with the shape functions
Ni=111
2
x1
l+9x1
l29
2x1
l3,Nj=9x1
l45
2x1
l2+27
2x1
l3,
Nk=9
2
x1
l+18 x1
l227
2x1
l3,Nl=x1
l9
2x1
l2+9
2x1
l3
(11.2.17)
which are shown in Fig. 11.7. The three types of beam elements given above show
the possibility of using elements of different accuracy. Of course, using the element
with higher number of nodes means that less elements and a more coarse mesh
can be used, but the calculations of the element stiffness matrices will be more
computationally expensive.
The further relationships are developed formally independent of the chosen num-
ber of element nodes. The element stiffness matrix is obtained with (11.1.6)
K
K
KE=
l
Z
0
B
B
BTC
C
CB
B
Bdx1=
l
Z
0
N
N
NTD
D
DTC
C
CD
D
DN
N
Ndx1(11.2.18)
Here
D
D
D=
0d
dx1
d
dx1
0
,C
C
C=D11 0
0ksA55 (11.2.19)
with the stiffness
D11 =b
3
n
k=1
C(k)
11 x(k)3
3x(k1)3
3,A55 =b
n
k=1
C(k)
55 h(k)(11.2.20)
and the shear correction factor ksgiven in (7.3.20).
For calculation of the element force vector, we assume that the element is loaded
by a distributed transverse load per length q(x1)and we can write the external work
as
WE=
l
Z
0
q(x1)w(x1)dx1=v
v
vT
Ef
f
fE
11.2 Finite Beam Elements 421
and with
w(x1) = [w
ψ
]1
0=u
u
uTR
R
R=v
v
vT
EN
N
NTR
R
R(11.2.21)
the element force vector f
f
fEis obtained
f
f
fE=
l
Z
0
N
N
NTR
R
Rq(x1)dx1(11.2.22)
If single nodal forces or moments are acting, they must be added.
The system equation can be obtained in dependence on the structure of the whole
system defined by a coincidence matrix together with the transformation of all el-
ement equations into a global coordinate system. After considering the boundary
conditions, the system equation can be solved. After this the stress resultants are
obtained for all elements
0.00.0
0.00.0
0.20.2
0.20.2
0.40.4
0.40.4
0.60.6
0.60.6
0.80.8
0.80.8
1.01.0
1.01.0
x1
l
x1
l
x1
l
x1
l
0.2
0.2
0.2
0.2
0.0
0.0
0.0
0.0
0.2
0.2
0.2
0.2
0.4
0.4
0.4
0.4
0.6
0.6
0.6
0.6
0.8
0.8
0.8
0.8
1.0
1.0
1.0
1.0
0.4
0.4
1.2
1.2
NiNj
NkNl
i
i
i
i
j
j
j
j
k
k
k
k
l
l
l
l
Fig. 11.7 Shape functions of the four-node element
422 11 Finite Element Analysis
σ
σ
σ
E=M
QE
=C
C
CD
D
DN
N
Nv
v
vE(11.2.23)
For elastodynamic problems the mass matrix must be calculated. The Timoshenko
beam model includes in the general case axial, transversal and rotational inertia
forces and moments. So it must be noted that the laminae of the beam have different
velocities in x1-direction
TE=1
2
l
Z
0
b/2
Z
b/2
x(n)
3
Z
x(0)
3
ρ
(˙u2+˙w2)dx1dx2dx3,˙u=˙u0+x3˙
ψ
(11.2.24)
After integration with respect to dx2and dx3follow
TE=1
2
l
Z
0
ρ
0(˙u2
0+˙w2) + 2
ρ
1˙u0˙
ψ
+
ρ
2˙
ψ
2dx1(11.2.25)
with the so-called generalized densities
ρ
0=b
n
k=1
ρ
(k)x(k)
3x(k1)
3,
ρ
1=b1
2
n
k=1
ρ
(k)x(k)
3
2x(k1)
3
2,
ρ
2=b1
3
n
k=1
ρ
(k)x(k)
3
3x(k1)
3
3
(11.2.26)
ρ
(k)is the density of the kth lamina.
Because we assumed a symmetric arrangement of the laminae in the cross-
section it follows that
ρ
1=0,˙u0=0
and therefore
TE=1
2
l
Z
0
(
ρ
0˙w2+
ρ
2˙
ψ
2)dx1=1
2
l
Z
0
˙
u
u
uTR
R
R0˙
u
u
udx1(11.2.27)
with the matrix R
R
R0
R
R
R0=
ρ
00
0
ρ
2(11.2.28)
Using (11.1.1)
TE=1
2
l
Z
0
˙
v
v
vT
EN
N
NTR
R
R0N
N
N˙
v
v
vEdx1
the element mass matrix M
M
MEis obtained
11.2 Finite Beam Elements 423
M
M
ME=
l
Z
0
N
N
NTR
R
R0N
N
Ndx1(11.2.29)
The system equation is established in the same manner as for elastostatic problems,
and with the assumption of harmonic vibrations, the eigen-frequencies and the mode
shapes can be calculated.
11.2.3 Problems
Exercise 11.2. Let us assume a two-node beam element.
1. Calculate the element stiffness matrix for a two-node beam element by analytical
integration!
2. Calculate the element force vector for a two-node beam element, loaded by a
linear distributed transverse load per length q(x1). The intensities at the nodes
are qiand qj!
3. Calculate the element mass matrix for a two-node beam element!
Solution 11.2. The three solutions are;
1. In the case of a two-node beam element the matrix of the shape functions is
N
N
N=Ni(x1)1 0
0 1 Nj(x1)1 0
0 1 
with Ni(x1) = 1(x1/l),Nj(x1) = x1/l. The element stiffness matrix is defined
by (11.2.18)
K
K
K=
l
Z
0
N
N
NTD
D
DTCDN
CDN
CDNdx1,
where in (11.2.19) are given
D
D
D=
0d
dx1
d
dx1
0
,C
C
C=¯
D11 0
0ks¯
A55
with ¯
D11 and ks¯
A55 in according to (11.2.20). After execution the matrix opera-
tions we obtain for the stiffness matrix
424 11 Finite Element Analysis
K
K
KE=
l
Z
0
ks¯
A55 dNi
dx12
0ks¯
A55
dNi
dx1
dNj
dx1
0
0¯
D11 dNi
dx12
0¯
D11
dNi
dx1
dNj
dx1
ks¯
A55
dNi
dx1
dNj
dx1
0ks¯
A55 dNj
dx12
0
0¯
D11
dNi
dx1
dNj
dx1
0¯
D11 dNj
dx12
dx1
and finally
K
K
KE=1
l
ks¯
A55 0ks¯
A55 0
0¯
D11 0¯
D11
ks¯
A55 0ks¯
A55 0
0¯
D11 0¯
D11
2. The element force vector to calculate with respect to (11.2.22)
f
f
fE=
l
Z
0
N
N
NTR
R
Rq(x1)dx1with R
R
R=1
0
For the loading function q(x1)we can write
q(x1) = [Ni(x1)Nj(x1)]qi
qj
and then we find
f
f
fE=
l
Z
0
Ni(x1)2Ni(x1)Nj(x1)
0 0
Ni(x1)Nj(x1)Nj(x1)2
0 0
qi
qjdx1=l
6
2qi+qj
0
qi+2qj
0
3. The element mass matrix for such a two-node beam element we find in according
to (11.2.29)
M
M
ME=
l
Z
0
N
N
NTR
R
R0N
N
Ndx1,R
R
R0=
ρ
00
0
ρ
2
with the generalized densities
ρ
0and
ρ
2(11.2.26).
Inserting the matrix of the shape functions given above and executing the matrix
operations we obtain
l
Z
0
Ni(x1)2
ρ
00Ni(x1)Nj(x1)
ρ
00
0Ni(x1)2
ρ
20Ni(x1)Nj(x1)
ρ
2
Ni(x1)Nj(x1)
ρ
00Nj(x1)2
ρ
00
0Ni(x1)Nj(x1)
ρ
20Nj(x1)2
ρ
2
dx1
11.3 Finite Plate Elements 425
and after integration the element mass matrix is in this case
M
M
ME=l
6
2
ρ
00
ρ
00
0 2
ρ
20
ρ
2
ρ
00 2
ρ
00
0
ρ
20 2
ρ
2
11.3 Finite Plate Elements
Plates are two-dimensional structures that means that one dimension, the thickness,
is very small in comparison to the others and in the unloaded state they are plane.
Usually, the midplane between the top and the bottom plate surfaces is defined as
the reference plane and is taken as the plane of xy. The z-direction corresponds
to the thickness direction. To avoid double indexes in the following relationships in
this section we will use the coordinates x,y,zinstead of x1,x2,x3. Laminate plates
consist of a number of bonded single layers. We assume that the single layer as
quasi-homogeneous and orthotropic. In each layer we can have different materials,
different thicknesses and especially different angle orientations of the fibres. The
whole plate is assumed to be a continuous structure. The stacking sequence of the
single layers has a great influence on the deformation behavior of the plate. Plates
can be loaded by distributed and concentrated loads in all directions, so called in-
plane and out of plane loading. In a special case of laminate plates, if we have an
arrangement of the single layers symmetric to the midplane, the in-plane and out of
plane states are decoupled.
In Chap. 8 the modelling of laminate plates is given and it distinguishes between
the classical laminate theory and the shear deformation theory like the modelling
of beams. The plate model based on the classical laminate theory usually is called
Kirchhoff plate with its main assumption that points lying on a line orthogonal to
the midplane before deformation are lying on such a normal line after deformation.
This assumption is an extended Bernoulli hypothesis of the beam model to two-
dimensional structures.
The application of the classical laminate theory should be restricted to the anal-
ysis of very thin plates only. For moderate thick plates it is better to use the shear
deformation theory. The plate model based on this theory is called the Mindlin plate
model. The following development of finite laminate plate elements will be carried
out for both models. Here we will be restricted to symmetric laminate plates in both
cases, it means that we have no coupling of membrane and bending/twisting states
and we will consider bending only.
In both cases we consider a triangular finite plate element. The approximation
of complicated geometric forms, especially of curved boundaries, can be done
easily with triangular elements. Usually special coordinates are used for triangu-
lar elements. The triangle is defined by the coordinates of the three corner points
P
i(xi,yi),i=1,2,3. A point P(x,y)within the triangle is also defined by the natural
426 11 Finite Element Analysis
triangle coordinates L1,L2,L3,P(L1,L2,L3), see Fig. 11.8. There
L1=A1
A
,L2=A2
A
,L3=A3
A
(11.3.1)
with the triangle area Aand the partial areas A1,A2,A3,A=A1+A2+A3. There-
fore
L1+L2+L3=1 (11.3.2)
The areas A1,A2,A3,Acan be expressed by determinants
A=
1x1y1
1x2y2
1x3y3
,A1=
1x y
1x2y2
1x3y3
,
A2=
1x1y1
1x y
1x3y3
,A3=
1x1y1
1x2y2
1x y
(11.3.3)
and for the coordinates L1,L2,L3of the point P(x,y)
L1=1
2A
[(x2y3x3y2) + (y2y3)x+ (x3x2)y],
L2=1
2A
[(x3y1x1y3) + (y3y1)x+ (x1x3)y],
L3=1
2A
[(x1y2x2y1) + (y1y2)x+ (x2x1)y]
(11.3.4)
and the coordinates x,ycan be expressed by
x=x1L1+x2L2+x3L3,y=y1L1+y2L2+y3L3(11.3.5)
Considering Eq. (11.3.2) we obtain for cartesian coordinates
Fig. 11.8 Natural triangle
coordinates x
y
A1
A2
A3
1(i)
2(j)
3(k)
P
11.3 Finite Plate Elements 427
x=L1(x1x3) + L2(x2x3) + x3,y=L1(y1y3) + L2(y2y3) + y3(11.3.6)
In Fig. 11.9 the natural triangle coordinates L1,L2,L3are illustrated for some special
points: the corner points and the points in the middle of the sides.
Because the shape functions Niused for the approximation of the deformation
field in a triangular plate element are usually written as functions of the natural
element coordinates, it is necessary to find relationships for the derivatives of the
shape functions with respect to the global cartesian coordinates. At first the deriva-
tives of the shape functions Niare given by the natural triangle coordinates L1and
L2.L3depends from L1and L2, see (11.3.2). So we consider only two independent
coordinates. Here we have
Ni
L1
Ni
L2
=
x
L1
y
L1
x
L2
y
L2
Ni
x
Ni
y
=J
J
J
Ni
x
Ni
y
(11.3.7)
J
J
Jis the Jacobi matrix of the coordinate transformation
J
J
J=
x
L1
y
L1
x
L2
y
L2
=cjbj
cibi(11.3.8)
and the expressions bi,bj,ci,cjare
bi=y2y3,bj=y3y1,ci=x3x2,cj=x1x3(11.3.9)
With
DetJ
J
J=cjbibjci=
(11.3.10)
Fig. 11.9 Natural triangle
coordinates of special points x
y
1(i)
(1; 0; 0)2(j)
(0; 1; 0)
3(k)(0; 0; 1)
6(n)
(0.5; 0; 0.5)5(m)
(0; 0.5; 0.5)
4(l)
(0.5; 0.5; 0)
428 11 Finite Element Analysis
it follows that
J
J
J1=1
bibj
cicj(11.3.11)
and then we obtain for the derivatives of the shape functions Niwith respect to the
cartesian coordinates xand y
Ni
x
Ni
y
=J
J
J1
Ni
L1
Ni
L2
(11.3.12)
In case of the classical laminate theory, the second partial derivatives
2Ni
L2
1
,
2Ni
L1L2
,
2Ni
L2
2
are also required. For this we must put the result for
Ni/
L1instead of Niinto the
first row of Eq. (11.3.7), and we obtain
2Ni
L2
1
=
x
L12
2Ni
x2+2
x
L1
y
L1
2Ni
x
y+
y
L12
2Ni
y2(11.3.13)
In the same manner we can do so with
Ni/
L2and the second row of Eq. (11.3.7)
and for the mixed second partially derivative with
Ni/
L2and the first row or vice
versa. The three relationships obtained can be written in matrix form
J
J
J
2Ni
x2
2Ni
x
y
2Ni
y2
=
2Ni
L2
1
2Ni
L1
L2
2Ni
L2
2
(11.3.14)
where J
J
Jis a modified or extended Jacobi matrix
J
J
J=
x
L12
2
x
L1
y
L1
y
L12
x
L1
x
L2
x
L2
y
L1
+
x
L1
y
L2
y
L1
y
L2
x
L22
2
x
L2
y
L2
y
L22
(11.3.15)
11.3 Finite Plate Elements 429
Now the second partial derivatives of the shape functions by the cartesian coordi-
nates can be calculated
2Ni
x2
2Ni
x
y
2Ni
y2
=J
J
J1
2Ni
L2
1
2Ni
L1
L2
2Ni
L2
2
(11.3.16)
Of course, by consequently using the natural triangle coordinates it follows that the
integrands in the energy terms are functions of these coordinates. Therefore we have
to consider for the variables of integration the relationship
dA=dxdy=DetJ
J
JdL1dL2=
dL1dL2(11.3.17)
In Sects. 11.3.1 and 11.3.2 the development of triangular finite plate elements will
be shown in a condensed way for the classical laminate theory and for the shear
deformation theory, respectively.
11.3.1 Classical Laminate Theory
The starting point is the total potential energy of an symmetric laminate plate, see
also (8.2.24)
Π
(w) = 1
2Z
A"D11
2w
x22
+D22
2w
y22
+2D12
2w
x2
2w
y2+4D66
2w
x
y2
+4D16
2w
x2+D26
2w
y2
2w
x
y#dAZ
A
pzwdA
(11.3.18)
with the stiffness Di j,i,j=1,2,6, see Table 8.3. The strain energy simplifies the
couplings, if we assume special orthotropic laminates (e.g. cross-ply-laminates).
We have no bending-twisting coupling, i.e. D16 =D26 =0. Supposing in other cases
these coupling terms as very small, especially if we have a great number of very thin
layers, we use the following simplified strain energy approximately
430 11 Finite Element Analysis
Π
(w) = 1
2Z
A"D11
2w
x22
+D22
2w
y22
+2D12
2w
x2
2w
y2+4D66
2w
x
y2
2pzw#dA
(11.3.19)
The total potential energy of the classical plate model contains second derivatives
and so we have to realize C(1)-continuity at the element boundaries. This means,
continuity of the deflections and the derivatives in normal direction to the bound-
aries. It must be noted that we do not have C(1)-continuity, if the first derivatives
at the corner points of adjacent elements are equal because we have to guarantee
the continuity of the derivatives in the normal direction at all boundary points of
adjacent elements.
It can be shown that we have to use a polynomial with minimum of 18 coeffi-
cients, and because we want to have a complete polynomial, we choose a polynomial
of fifth order with 21 coefficients. Therefore we define a triangular finite plate ele-
ment with 6 nodes as shown in Fig. 11.10. At the corner nodes 1,2,3(i,j,k)we have
6 degrees of freedom, the deflection, the first derivatives in both directions and the
three curvatures, but at the mid-side nodes the first derivatives in normal direction
only.
It is a disadvantage when using this element in a general program system that we
have a different number of degrees of freedom at the nodes. Therefore an elimina-
tion, a so-called static condensation of the nodal constants of the mid-side nodes,
can be done and then we have only 18 degrees of freedom for the element. The el-
ement is converted into a three-node element, the nodes 4,5,6(l,m,n)vanish. The
polynomial approximation of the displacement field in the finite element is given by
a special 5th order polynomial, it contains however a complete polynomial of 4th
order. In this way we obtain 18 shape functions Ni(L1,L2,L3),i=1,2,...,18 which
are not illustrated here. Because the coordinates L1,L2,L3are not independent, see
(11.3.2), L3usually is eliminated by
L3=1L1L2(11.3.20)
Fig. 11.10 Six-node plate
element x
y
1(i)
2(j)
3(k)
6(n)5(m)
4(l)
11.3 Finite Plate Elements 431
According to (11.1.1) we have the approximation
w(x,y) = N
N
N(L1,L2)v
v
v(11.3.21)
with N
N
Nas the matrix of the 18 shape functions (here it has only one row) and the el-
ement displacement vector v
v
vincluding 18 components. For the differential operator
D
D
DOP must be written
D
D
DOP =
2
x2
2
y22
2
x
yT
(11.3.22)
and after this, see also Eq. (11.1.3), the matrix B
B
Bleads to
B
B
B=D
D
DOPN
N
N(11.3.23)
Since the shape functions are functions of the natural triangle coordinates L1and L2,
for the derivatives by the cartesian coordinates we have to take into consideration
(11.3.16). The element stiffness matrix follows according to (11.1.6)
K
K
KE=Z
AE
B
B
BTDB
DB
DBdA
and with the substitution of the integration variable Eq. (11.3.17)
K
K
KE=
1
Z
0
1L1
Z
0
B
B
BTDB
DB
DBdL2dL1(11.3.24)
Here D
D
Dis the matrix of the plate stiffness, the coupling of bending and twisting is
neglected (D16 =D26 =0)
D
D
D=
D11 D12 0
D12 D22 0
0 0 D66
According to (11.1.7) we obtain the element force vector
f
f
fE=Z
AE
N
N
NTpdA=
1
Z
0
1L1
Z
0
N
N
NTpdL2dL1(11.3.25)
where p(x,y) = p(L1,L2)is the element surface load.
For the flexural vibration analysis of plates the element mass matrix must be
calculated. According to (11.1.10), the element mass matrix reduces to
M
M
ME=Z
VE
ρ
N
N
NTN
N
NdV=
1
Z
0
1L1
Z
0
ρ
N
N
NTN
N
NhdL2dL1(11.3.26)
432 11 Finite Element Analysis
with
ρ
as an average density
ρ
=1
h
n
k=1
ρ
(k)h(k)(11.3.27)
Note that the classical laminate theory does not consider the rotary kinetic energy.
The integrations in the (11.3.24) for the element stiffness matrix K
K
KE, (11.3.25)
for the element force vector f
f
fEand (11.3.26) for the element mass matrix must
be carried out numerically. Only the force vector f
f
fEcan be calculated analytically,
if we have a constant surface loading p(x,y) = const. For the numerical solutions
it is recommended that integration formulae of the same order are used like the
polynomials for the shape functions, in this case of the fifth order.
11.3.2 Shear Deformation Theory
The Mindlin plate model, which is based on the first order shear deformation theory,
considers the shear deformation in a simplified form. In the Mindlin plate model
the Kirchhoff’s hypotheses are relaxed. Transverse normals to the midplane do not
remain perpendicular to the middle surface after deformation. In Sect. 8.3 the basic
equations are given for this plate model.
Here the starting point is the total potential energy, and if we restrict ourselves to
symmetric and special orthotropic laminates, we have
Π
(w,
ψ
1,
ψ
2) = 1
2Z
A"D11
ψ
1
x2
+2D12
ψ
1
x
ψ
2
y+D22
ψ
2
y2
+D66
ψ
1
y+
ψ
2
x2
+ks
55A55
ψ
1+
w
x2
+ks
44A44
ψ
2+
w
y2#dxdyZ
A
pzwdxdy
δΠ
(w,
ψ
1,
ψ
2) = 0
(11.3.28)
or written in matrix form
Π
(w,
ψ
1,
ψ
2) = 1
2Z
A
(
κ
κ
κ
TD
D
D
κ
κ
κ
+
ε
ε
ε
sTA
A
As
ε
ε
ε
s)dxdyZ
A
p3wdxdy(11.3.29)
The matrices of the plate stiffness for this case (D16 =D26 =0) and the shear stiff-
ness with A45 =0 are, see also (8.3.7),
11.3 Finite Plate Elements 433
D
D
D=
D11 D12 0
D12 D22 0
0 0 D66
,A
A
As=ks
55A55 0
0ks
44A44 (11.3.30)
The stiffness are given in detail in (4.2.15) and for the shear correction factor see
Sect. 8.3. Note that we have, in the elastic potential three independent deformation
components (the deflection wand the rotations
ψ
1and
ψ
2), so the displacement field
vector u
u
uhas three components (nu=3), see also (11.1.2).
For the curvatures and the shear strains we have
κ
κ
κ
=D
D
Dbu
u
u,
ε
ε
ε
s=D
D
Dsu
u
u(11.3.31)
where D
D
Dband D
D
Dsare the matrices of the differential operators
D
D
Db=
0
x0
0 0
y
0
y
x
,D
D
Ds=
x1 0
y0 1
(11.3.32)
The most important property of the elastic potential however is that it contains first
derivatives only. Therefore, we have to guarantee onlyC(0)-continuity at the element
boundaries and it will be possible to take a three-node finite element with linear
shape functions, but it shall be not done here.
Due to the better approximation properties we will choose a six-node element
with polynomials of the second order as shape functions. The six-node element with
its nodal degrees of freedom is shown in Fig. 11.11. Then we have the nodal and the
element displacement vectors
v
v
vT
i= [wi
ψ
xi
ψ
yi],v
v
vT
E= [v
v
vT
iv
v
vT
jv
v
vT
kv
v
vT
lv
v
vT
mv
v
vT
n](11.3.33)
and according to (11.1.2) with nu=nF,˜n=nKEthe matrix of the shape functions is
given by
N
N
N= [NiI
I
I3NjI
I
I3NkI
I
I3NlI
I
I3NmI
I
I3NnI
I
I3],(11.3.34)
Fig. 11.11 Six-node finite
plate element with nodal
degrees of freedom x
y
z
ψ
1i
ψ
2i
wi
ilj
m
k
n
434 11 Finite Element Analysis
where I
I
I3are unit matrices of the size (3,3). The shape functions are
Ni= (2L11)L1,Nj= (2L21)L2,Nk= (2L31)L3,
Nl=4L1L2,Nm=4L2L3,Nn=4L1L3
(11.3.35)
They are functions of the natural triangle co-ordinates L1,L2,L3, see (11.3.1) -
(11.3.4).
The curvatures and the shear strains in (11.3.29) can be expressed by
κ
κ
κ
=D
D
Dbu
u
u=D
D
DbN
N
Nv
v
vE=B
B
Bbv
v
vE,B
B
Bb=D
D
DbN
N
N
ε
ε
ε
s=D
D
Dsu
u
u=D
D
DsN
N
Nv
v
vE=B
B
Bsv
v
vE,B
B
Bs=D
D
DsN
N
N(11.3.36)
and consideration of (11.3.12) leads to the element stiffness matrix, see also (11.1.6)
consisting of two parts
K
K
KE=K
K
Kb
E+K
K
Ks
E,K
K
Kb
E=Z
AE
B
B
BbTD
D
DB
B
Bbdxdy,K
K
Ks
E=Z
AE
B
B
BsTA
A
AsB
B
Bsdxdy(11.3.37)
Because the shape functions in N
N
Nare functions of the natural triangle co-ordinates,
the integration variables must be substituted by (11.3.17), and then we find
K
K
Kb
E=
1
Z
0
1L1
Z
0
B
B
BbTD
D
DB
B
Bb
dL2dL1,K
K
Ks
E=
1
Z
0
1L1
Z
0
B
B
BsTA
A
AsB
B
Bs
dL2dL1(11.3.38)
To obtain the element force vector f
f
fEa load vector q
q
qmust be defined with the same
number of components as the displacement field vector u
u
u. Because only surface
loading p(x,y)is considered here, it leads to
q
q
qT= [p0 0]
and then the element force vector is
f
f
fE=Z
AE
N
N
NTq
q
qdxdy,f
f
fE=
1
Z
0
1L1
Z
0
N
N
NTq
q
q
dL2dL1(11.3.39)
with the substitution of integration variables.
The integrations in (11.3.39) can be done analytically only in the case of constant
surface loading p=const. In the other cases it must be calculated numerically. For
the numerical integration it is recommended to apply integration formulae of the
same order as used for shape polynomials, here of the second order. It must be done
in this manner for the first part K
K
Kb
Eof the stiffness matrix, for the second part of K
K
Ks
E
a lower order can be used. Such a different kind of integration for the two parts of
the stiffness matrix is called selective integration.
For dynamic analysis the element mass matrix M
M
MEmust also be calculated. For
the shear deformation theory the rotatory kinetic energy is usually taken into con-
11.3 Finite Plate Elements 435
sideration. The kinetic energy of an element is then
TE=1
2Z
VE
ρ
˙
u
u
uT˙
u
u
udV=1
2Z
AE
h
2
Z
h
2
ρ
(˙w2+˙
ψ
2
1+˙
ψ
2
2)dzdA(11.3.40)
If the so called generalized densities are used
ρ
0=
n
k=1
ρ
(k)[z(k)z(k1)] =
n
k=1
ρ
(k)h(k),
ρ
1=1
2
n
k=1
ρ
(k)[z(k)2z(k1)2],
ρ
2=1
3
n
k=1
ρ
(k)[z(k)3z(k1)3]
(11.3.41)
and it is noted here that
ρ
1=0, because we have assumed symmetric laminates only,
then for the kinetic energy we obtain
TE=1
2Z
AE
˙
v
v
vTR
R
R0˙
v
v
vdA(11.3.42)
R
R
R0is a matrix of the generalized densities
R
R
R0=
ρ
00 0
0
ρ
20
0 0
ρ
2
(11.3.43)
Using the approximation for the displacement field vector according to Eq. (11.1.1)
we obtain
TE=1
2v
v
vT
EZ
AE
N
N
NTR
R
R0N
N
NdAv
v
vE(11.3.44)
and the element mass matrix is
M
M
ME=Z
AE
N
N
NTR
R
R0N
N
NdA,M
M
ME=
1
Z
0
1L1
Z
0
N
N
NTR
R
R0N
N
N
dL1dL2(11.3.45)
with substitution of the integration variables.
The finite laminate plate element developed above is called PL18, where the
number 18 gives the degrees of freedom of all element nodes. This element can be
used only for laminate plates with laminae arranged symmetrically to the midplane,
where we have no coupling of membrane and bending/twisting states and we have
no in-plane loading.
436 11 Finite Element Analysis
In many cases we have nonsymmetric laminates and we have a coupling of mem-
brane and bending/twisting states or there are in-plane and out-of-plane loadings.
Then an element is necessary where the nodal degrees of freedom also include the
deflections in x- and y-direction u,v. For such an element, assuming six nodes again,
the nodal and the element displacement vectors are
v
v
vT
i= [uiviwi
ψ
xi
ψ
yi],v
v
vT
E= [v
v
vT
iv
v
vT
jv
v
vT
kv
v
vT
lv
v
vT
mv
v
vT
n](11.3.46)
The structure of the matrix of the shape functions N
N
Nis in this more general case
N
N
N= [NiI
I
I5NjI
I
I5NkI
I
I5NlI
I
I5NmI
I
I5NnI
I
I5](11.3.47)
with I
I
I5as unit matrices of the size (5,5), the shape functions remain unchanged.
The total potential energy for this case is, see also (8.3.15),
Π
(u,v,w,
ψ
1,
ψ
2) = 1
2Z
A
(
ε
ε
ε
TA
A
A
ε
ε
ε
+
κ
κ
κ
TB
B
B
ε
ε
ε
+
ε
ε
ε
TB
B
B
κ
κ
κ
+
κ
κ
κ
TD
D
D
κ
κ
κ
+
ε
ε
ε
sTA
A
As
ε
ε
ε
s)dx1dx2Z
A
pzwdxdy
(11.3.48)
and we have to take into consideration the membrane stiffness matrix A
A
Aand the
coupling matrix B
B
Badditionally, the element stiffness matrix consists of four parts
K
K
KE=K
K
Km
E+K
K
Kmb
E+K
K
Kb
E+K
K
Ks
E(11.3.49)
representing the membrane state (K
K
Km
E), the coupling of membrane and bending states
(Kmb
E), the bending state (K
K
Kb
E) and the transverse shear state (K
K
Ks
E).
The general form for the element force vector (11.3.39) is unchanged, it must be
noted that the loading vector q
q
qhere has another structure containing loads in three
directions
q
q
qT= [pxpypz0 0](11.3.50)
the general form for the element mass matrix is the same as in (11.3.45), but here
the matrix of the generalized densities R
R
R0is
R
R
R0=
ρ
00 0
ρ
10
0
ρ
00 0
ρ
1
0 0
ρ
00 0
ρ
10 0
ρ
20
0
ρ
10 0
ρ
2
(11.3.51)
The remarks about the realization of the integrations remains unchanged here. Of
course they are all more complicated for this element. Such an extended element
would be called PL30, because the degree of freedom of all nodal displacements is
30. Further details about this extended element are not given here.
11.4 Generalized Finite Beam Elements 437
11.4 Generalized Finite Beam Elements
In civil engineering and also in mechanical engineering a special kind of struc-
tures are used very often structures consisting of thin-walled elements with sig-
nificant larger dimensions in one direction (length) in comparison with the dimen-
sions in the transverse direction. They are called beam shaped shell structures. Beam
shaped shell structures include folded plate structures as the most important class
of such structures. In Chap. 10 the modelling of folded plate structures was con-
sidered and there a generalized beam model was developed by the reduction of the
two-dimensional problem to an one-dimensional one following the way of Vlasov-
Kantorovich. This folded structure model contains all the energy terms of the mem-
brane stress state and of the bending/twisting stress state under the validity of the
Kirchhoff hypotheses. Outgoing from this complete folded structure model some
simplified structure models were developed (see Sect. 10.2.4) by neglecting of se-
lected energy terms in the potential function e.g. the terms caused by the longitudinal
curvatures
κ
xi, the shear strains
ε
xsi, the torsional curvatures
κ
xsior the transversal
strains
ε
siof the strips. Because the influence of the longitudinal curvatures
κ
xiof
the single strips to the deformation state and the stress state of the whole structure
is very small for beam shaped structures, they are neglected generally. The shear
strains
ε
xsiof the strips can be neglected for structures with open cross-sections,
but not in the case of closed cross-sections. In opposite to this the torsional curva-
tures
κ
xsican be neglected for closed cross-sections, but not for open cross-sections.
Therefore, because we had in mind to find a generalized beam model as well the
shear strains as the torsional curvatures are considered. Although the influence of
the transversal strains in most cases is very small, they are considered too, because
with this we have a possibility to define the generalized co-ordinate functions for a
general cross-section systematically. Therefore as a generalized structure model for
beam shaped thin-walled folded plate structures the structure model A (see Sects.
10.2.4 and 10.2.5) is chosen, in which only the longitudinal curvatures
κ
xiof the
strips are neglected.
11.4.1 Foundations
The starting point for the development of generalized finite beam elements is the
potential energy, see equation (10.2.10). Because in all strips the longitudinal curva-
tures
κ
xiare neglected all terms containing w′′
ihave to vanish. It leads together with
equations (10.2.11), (10.2.12) and with ˆ
A
A
A7=0
0
0, ˆ
A
A
A8=0
0
0, ˆ
A
A
A10 =0
0
0, ˆ
A
A
A17 =0
0
0, ˆ
A
A
A20 =0
0
0,
ˆ
A
A
A23 =0
0
0, ˆ
A
A
A26 =0
0
0 to a simplification of the potential energy equation (10.2.10)
438 11 Finite Element Analysis
Π
=1
2
l
Z
0U
U
UTˆ
A
A
A1U
U
U+V
V
VTˆ
A
A
A6V
V
V+U
U
UTˆ
A
A
A3U
U
U+2U
U
UTˆ
A
A
A25V
V
V
+V
V
VTˆ
A
A
A4V
V
V+V
V
VTˆ
A
A
A12V
V
V+4V
V
VTˆ
A
A
A9V
V
V
+2U
U
UTˆ
A
A
A14V
V
V+2U
U
UTˆ
A
A
A2U
U
U+2U
U
UTˆ
A
A
A13V
V
V
2U
U
UTˆ
A
A
A19V
V
V4U
U
UTˆ
A
A
A18V
V
V+2U
U
UTˆ
A
A
A16V
V
V
+2V
V
VTˆ
A
A
A5V
V
V2V
V
VTˆ
A
A
A28V
V
V4V
V
VTˆ
A
A
A27V
V
V
2U
U
UTˆ
A
A
A22V
V
V2V
V
VTˆ
A
A
A25V
V
V4U
U
UTˆ
A
A
A21V
V
V
4V
V
VTˆ
A
A
A24V
V
V+4V
V
VTˆ
A
A
A11V
V
V2(U
U
UTf
f
fx+V
V
VTf
f
fs+V
V
VTf
f
fn)dx
(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=0(U
U
UTr
r
rx+V
V
VTr
r
rs+V
V
VTr
r
rn)|x=l
(11.4.1)
We can see that the one-dimensional energy function contains only derivatives of
the first order.
11.4.2 Element Definitions
Outgoing from Eq. (11.4.1) a one-dimensional finite element can be defined. Be-
cause we have no higher derivatives than of the first order in the potential energy
only a C(0)continuity is to satisfy at the element boundaries and therefore it would
be possible to use a two-node element with linear shape functions. To have a better
accuracy here we will take a three-node element using second order polynomials as
shape functions, Fig. 11.12. The shape functions are again like (11.2.15)
N1(x) = 13x
l+2x2
l2,N2(x) = 4x
l4x2
l2,N3(x) = x
l+2x2
l2(11.4.2)
They are shown in Fig. 11.5.
Because a generalized finite beam element with a general cross-section shall be
developed at first we must find a rule to define the cross-section topology. We will
use for it the profile node concept. For this we will see the midlines of all strips as the
cross-sections profile line. The start- and the endpoints of each strip on this profile
line are defined as the so-called main profile nodes. In the middle of each strip there
Fig. 11.12 Three-node gener-
alized beam element
123
x
l
11.4 Generalized Finite Beam Elements 439
are additional profile nodes, they are called secondary profile nodes. Figure 11.13
shows an example for it. The topology of the thin-walled cross-section is described
sufficiently by the co-ordinates of the main profile nodes. Additionally the stiffness
parameters of each strip must be given. The connections of the strips in the main
profile nodes are supposed as rigid.
For the generation of the generalized deflection co-ordinate functions
ϕ
,
ψ
,
ξ
is
assumed that a main profile node has four degrees of freedom, the displacements
in the directions of the global co-ordinate axes x,y,zand the rotation about the
global xaxis, see Fig. 11.14. The displacements of the main profile nodes lead linear
generalized co-ordinate functions
ϕ
,
ψ
and cubic functions
ξ
between the adjacent
nodes. For an increasing the accuracy the activation of the secondary profile node
degrees of freedom is optional, they are shown in Fig. 11.15. In this case
ϕ
and
ψ
are
quadratic and
ξ
polynomials of 4th and 5th order between the adjacent main nodes.
Therefore a more complex deformation kinematics of the cross-section is consid-
erable. The generalized co-ordinate functions for any thin-walled cross-section are
here defined as follows:
1. Main node displacements or rotations result in non-zero co-ordinate functions
only in the adjacent intervals of the profile line
Fig. 11.13 Description of a
general cross-section
11
2
2
3
344
5
56
6
7 7
8
z
y
y1s2s4s5s6
s3s7
s8
- main profile nodes (MPN)
- secondary profile nodes SPN)
deflection in x-direction deflection in y-direction
deflection in z-direction rotation about x-axis
Fig. 11.14 Main profile node degrees of freedom
440 11 Finite Element Analysis
second order v-deflection second order u-deflection
fourth order w-deflection fifth order w-deflection
Fig. 11.15 Secondary profile node degrees of freedom
2. Secondary node displacements or rotations result in non-zero co-ordinate func-
tions only in the interval between the adjacent main nodes.
In Fig. 11.16 the generalized coordinate functions for axial parallel arranged strips
are shown. Figure 11.17 gives the supplements for slanting arranged strips.
11.4.3 Element Equations
In the case of non-activated degrees of freedom of the secondary profile nodes we
have a degree of freedom of an element node of four times the number of main
profile nodes (4 nMP N ) and the element displacement vector consists of 12 nMPN
components
v
v
vT= [v
v
v1v
v
v2v
v
v3],v
v
vj=¯
u
u
uj
¯
v
v
vj(11.4.3)
¯
u
u
uj,¯
v
v
vjcontain the values of the generalized displacement functions at the node j.
The displacement vector u
u
u(x)contains here the generalized displacement functions
U
U
U(x)and V
V
V(x)and in accordance with Eq. (11.1.1) we obtain for the interpolation
u
u
u(x) = U
U
U(x)
V
V
V(x)=Nv
Nv
Nv (11.4.4)
The matrix of the shape functions for the chosen three-node element is
N
N
N= [N1(x)I
I
I N2(x)I
I
I N3(x)I
I
I],(11.4.5)
where I
I
Iare the unit matrices of the size 4 nMPN and with this the matrix N
N
Nhas
the format (4 nMPN , 12 nM PN ). Following the equation (11.4.4) for the generalized
11.4 Generalized Finite Beam Elements 441
1 3 2
x,u
n,ws,v
d
main profile node functions secondary profile node functions
u-deflections
v-deflections
w-deflections
1
1
11
1
1
11 0.25
0.25
1/16
1/8d
ϕ
1=1s
d
ϕ
2=s
d
ϕ
3=s
ds
d2
ψ
1=1s
d
ξ
1=0
ψ
2=s
d
ξ
2=0
ψ
3=s
ds
d2
ξ
3=0
ξ
4=13s
d2+2s
d3
ψ
4=0
ξ
5=3s
d22s
d3
ψ
5=0
ξ
8=s
d22s
d3+s
d4
ψ
8=0
ξ
6=s12s
d+s
d2
ψ
6=0
ξ
7=ss
d+s
d2
ψ
7=0
ξ
9=s
d24s
d3
+5s
d42s
d5
ψ
4=0
Fig. 11.16 Generalized co-ordinate functions for axial parallel arranged strips
442 11 Finite Element Analysis
1
2
3
x
z
y
n,w
x,u
s,v
α
ψ
1=1s
dcos
α
ξ
1=13s
d2+2s
d3sin
α
ψ
2=s
dcos
α
ξ
2=3s
d22s
d3sin
α
ψ
4=1s
dsin
α
ξ
4=13s
d2+2s
d3cos
α
ψ
5=s
dsin
α
ξ
5=3s
d22s
d3cos
α
Fig. 11.17 Supplements for slanting arranged strips
11.4 Generalized Finite Beam Elements 443
displacement functions we have to write
U
U
U(x) = L
L
LT
10u
u
u(x) = L
L
LT
10Nv
Nv
Nv,V
V
V(x) = L
L
LT
01u
u
u(x) = L
L
LT
01Nv
Nv
Nv (11.4.6)
with the matrices
L
L
LT
10 = [I
I
I0
0
0],L
L
LT
01 = [0
0
0I
I
I](11.4.7)
In the first case (L
L
L10)I
I
Iis a unit matrix of the size nMNP and the null matrix has the
format (nMPN , 3 nM PN ), in the second case (L
L
L01)I
I
Iis a unit matrix of the size 3 nMPN
and the null matrix has the format (3 nM PN ,nMPN ).
Of course in the case of activated degrees of freedom of the secondary pro-
file nodes all the dimensions given above are increased correspondingly. Inserting
the generalized displacement functions (11.4.6) into the potential energy equation
(11.4.1) we obtain
Π
=1
2v
v
vTKv
Kv
Kv f
f
fTv
v
v(11.4.8)
The condition
Π
v
v
v=0
0
0 (11.4.9)
leads to the element equation
Kv
Kv
Kv =f
f
f(11.4.10)
with the symmetric element stiffness matrix
K
K
K=
K
K
K11 K
K
K12 K
K
K13
K
K
KT
12 K
K
K22 K
K
K23
K
K
KT
13 K
K
KT
23 K
K
K33
(11.4.11)
For the sub-matrices Kmn we find the general equation
Kmn =
3
h=1
ˆ
A
A
A1hImn1
+ˆ
A
A
A2hImn2
+ˆ
A
A
AT
2hImn3
+ˆ
A
A
A3hImn4
(ˆ
A
A
A13h2ˆ
A
A
A18h)Imn1
+( ˆ
A
A
A15h2ˆ
A
A
A21hImn2
+( ˆ
A
A
A14hˆ
A
A
A19h)Imn3
+( ˆ
A
A
A16hˆ
A
A
A22h)Imn4
(ˆ
A
A
AT
13h2ˆ
A
A
AT
18h)Imn1
+( ˆ
A
A
AT
15h2ˆ
A
A
AT
21h)Imn2
+( ˆ
A
A
AT
14hˆ
A
A
AT
19h)Imn3
+( ˆ
A
A
AT
16hˆ
A
A
AT
22h)Imn4
(ˆ
A
A
A4h+ˆ
A
A
A9h2ˆ
A
A
A24h2ˆ
A
A
AT
24h)Imn1
+( ˆ
A
A
A5h+2ˆ
A
A
A11hˆ
A
A
AT
25h2ˆ
A
A
A27h)Imn2
+( ˆ
A
A
AT
5h+2ˆ
A
A
A11hˆ
A
A
A25h2ˆ
A
A
AT
27h)Imn3
+( ˆ
A
A
A6h+ˆ
A
A
A12hˆ
A
A
A28hˆ
A
A
AT
28h)Imn4
(11.4.12)
with
444 11 Finite Element Analysis
Imn1=
l
Z
0
NhN
mN
ndx,Imn2=
l
Z
0
NhNmN
ndx,
Imn3=
l
Z
0
NhN
mNndx,Imn4=
l
Z
0
NhNmNndx
To include approximately slight non-prismatic structures the matrices of the stiff-
ness parameters ˆ
A
A
Ai, see Eq. (10.2.12), are interpolated in the element in the same
manner as the displacements
ˆ
A
A
Ai=
3
h=1
ˆ
A
A
AihNh(11.4.13)
ˆ
A
A
Aih are the matrices at the nodes h=1,2,3.
The element force vector is obtained as
f
f
f=
f
f
f1
f
f
f2
f
f
f3
(11.4.14)
with the sub-vectors
f
f
fm=
3
h=1
f
f
fxh
l
Z
0
NhNmdx
(
(
(fsh +fnh)
l
Z
0
NhNmdx
(11.4.15)
Here f
f
fxh,f
f
fsh,f
f
fnh are the generalized load vectors, see Eq. (10.2.13), at the nodes
h=1,2,3.
11.4.4 System Equations and Solution
The system equations can be obtained by using the Eqs. (11.1.15) and (11.1.16) with
the coincidence matrices, determining the position of each element in the whole
structure. In the so founded system stiffness matrix the boundary conditions of the
whole structure are to consider, otherwise this matrix is singular, if the structure
is not fixed kinematically. The solution of the system equations lead to the nodal
displacements and with them the strains and curvatures in the single strips of each
element can be calculated, see Eqs. (10.2.6) and (10.2.11),
11.4 Generalized Finite Beam Elements 445
ε
x(x,si) =
3
h=1
N
h¯
u
u
uT
h
ϕ
ϕ
ϕ
,
ε
s(x,si) =
3
h=1
Nh¯
v
v
vT
h
ψ
ψ
ψ
,
ε
xs(x,si) =
3
h=1
(Nh¯
u
u
uT
h
ϕ
ϕ
ϕ
+N
hv
v
vT
h
ψ
ψ
ψ
),
κ
s(x,si) =
3
h=1
Nh¯
v
v
vT
h
ξ
ξ
ξ
••,
κ
xs(x,si) = 2
3
h=1
N
h¯
v
v
vT
h
ξ
ξ
ξ
(11.4.16)
Now we can obtain the stress resultants in the kth lamina, which has the distance nk
from the mid plane of the strip
Nxk
Nsk
Nxsk
=
A11kA12kA16k
A12kA22kA26k
A16kA26kA66k
ε
x
ε
s+nk
κ
s
ε
xs +nk
κ
xs
(11.4.17)
These stress resultants are related on the strip co-ordinate axes xand si. Therefore, it
is necessary to transform them into the material co-ordinate system of the kth lamina
(for the transformation relationship see Table 4.1)
NLk
NTk
NLTk
=
cos2
α
ksin2
α
k2sin
α
kcos
α
k
sin2
α
kcos2
α
k2sin
α
kcos
α
k
sin
α
kcos
α
ksin
α
kcos
α
kcos2
α
ksin2
α
k
Nxk
Nsk
Nxsk
(11.4.18)
Than the stresses of the kth lamina are obtained
σ
Lk=NLk
tk
,
σ
Tk=NTk
tk
,
τ
LTk=NLTk
tk
(11.4.19)
In some cases the strains in the kth lamina related to the material co-ordinate system
are important for the failure assessment of the lamina. Then they can be calculated
with help of the following matrix equation
ε
Lk
ε
Tk
ε
LTk
=Q
Q
Q
k1
NLk
NTk
NLTk
(11.4.20)
There Q
Q
Qis the reduced stiffness matrix of the kth lamina.
11.4.5 Equations for the Free Vibration Analysis
The variation statement given by the Hamilton’s principle, see Eq. (10.2.41) leads
with the Lagrange function (10.2.39) and the assumption of harmonic vibrations for
the considered generalized beam element to the element equation
(K
K
K
ω
2M
M
M)v
v
v=0
0
0 (11.4.21)
446 11 Finite Element Analysis
K
K
Kis the element stiffness matrix, see Eqs. (11.4.11) and (11.4.12), and M
M
Mis the ele-
ment mass matrix. The element mass matrix is obtained with the matrices ˆ
B
B
B1,ˆ
B
B
B2,ˆ
B
B
B3,
see Eqs. (10.2.38)
M
M
M=
M
M
M11 M
M
M12 M
M
M13
M
M
MT
12 M
M
M22 M
M
M23
M
M
MT
13 M
M
MT
23 M
M
M33
(11.4.22)
with
M
M
Mmn =
3
h=1ˆ
B
B
B1hImn40
0
0
0
0
0(ˆ
B
B
B2h+ˆ
B
B
B3h)Imn4(11.4.23)
There the ˆ
B
B
Bmatrices are also interpolated in the element by using the shape func-
tions. In this way slight non-prismatic structures are considerable too. The system
equations can be developed in a similar way as it was done for a static analysis. Here
we have to find a system stiffness matrix and a system mass matrix. After consider-
ation the boundary conditions the eigen-value problem can be solved and the mode
shapes can be estimated.
11.5 Numerical Results
Additional to a great number of special FEM programs general purpose FEM pro-
gram systems are available. The significance of universal FEM program packages
is increasing. In universal FEM program systems we have generally the possibility
to consider anisotropic material properties, e.g. in the program system COSMOS/M
we can use volume elements with general anisotropic material behavior and plane
stress elements can have orthotropic properties.
Laminate shell elements are available e.g. in the universal FEM program systems
ANSYS, NASTRAN or COSMOS/M. In many program systems we have no spe-
cial laminate plate elements, the laminate shell elements are used also for the anal-
ysis of laminate and sandwich plates. Perhaps, because of the higher significance
of two-dimensional laminate structures in comparison with beam shaped structures
laminate beam elements are missing in nearly all universal FEM program systems.
The generalized beam elements, Sect. 11.4, are e.g. implemented only in the FEM
program system COSAR.
For the following numerical examples the program system COSMOS/M is used.
In COSMOS/M a three node and a four node thin laminate shell element are
available (SHELL3L; SHELL4L). Each node has 6 degrees of freedom. The element
can consist of up to 50 layers. Each layer can have different material parameters,
different thicknesses and especially different angles of fibre directions. We have no
restrictions in the stacking structure, symmetric, antisymmetric and nonsymmetric
structures are possible. The four nodes of the SHELL4L element must not arranged
in-plane. By the program in such case a separation is done into two or four triangular
partial elements. Further there is a SHELL9L element available. It has additional
11.5 Numerical Results 447
nodes at the middles of the four boundaries and in the middle of the element. For
the following examples only the element SHELL4L is used.
11.5.1 Examples for the Use of Laminated Shell Elements
By the following four examples the application of the laminate shell element
SHELL4L shall be demonstrated. At first a thin-walled beam shaped laminate struc-
ture with L-cross-section under a concentrated force loading is considered, and the
second example is a thin-walled laminate pipe under torsional loading. In both cases
the influence of the fibre angles in the layers is tested. The use of the laminate shell
element for the static and dynamic analysis of a sandwich plate is shown in the
third example. A buckling analysis of a laminate plate is demonstrated by the fourth
example. In all 4 cases a selection of results is given.
11.5.1.1 Cantilever Beam
A cantilever beam with L-cross-section consists of 3 layers with the given material
parameters Ex,Ey,
ν
xy,
ν
yx,Gxy . It is loaded by a concentrated force F, see Fig. 11.18.
The material parameters are
L
F
+
α
+
α
α
3
3
4
300
400
10
Fig. 11.18 Cantilever beam: cross-section and stacking structure (F= 4.5 kN, L= 4 m, all other
geometrical values in mm)
448 11 Finite Element Analysis
Ex=1.53 ·104kN/cm2,Ey=1.09 ·103kN/cm2,Gxy =560kN/cm2,
ν
xy =0.30,
ν
yx =0.021
The fibre angle
α
shall be varied:
α
=0,10,20,30,40.
The FEM model after the input of all properties into COSMOS/M is illustrated
in Fig. 11.19. The computing yields a lot of results. In Fig. 11.20, e.g., is shown
the deformed shape for a fibre angle of
α
=30. Here should be selected only
the displacements of the corner node at the free edge (node No. 306 in our FE
model) in y- and z-direction and the maximal stresses in fibre direction (
σ
x) and
perpendicular to it (
σ
y) for the left side of the vertical part of the cross-section (layer
No. 1, bottom):
v306,y= -2,204 cm, v306,z= -1,805 cm,
σ
lay1,max,x= 7,487 kN/cm2,
σ
lay1,max,y= 0,824 kN/cm2
Similar the displacements and stresses for the fibre angles
α
=0,10,20,40are
calculated, and the results are shown in Figs. 11.21 and 11.22. The results show
that for such a beam shaped structure the main stresses are lying in the longitudinal
direction and therefore the fibre angle 0leads to the most effective solution.
11.5.1.2 Laminate Pipe
A laminate pipe consisting of 2 layers with the given material parameters Ex,Ey,
ν
xy,
ν
yx,Gxy is fixed at left end and loaded by a torsional moment, see Fig. 11.23.
The material parameters are the same as in the previous example:
Ex=1,53 104kN/cm2,Ey=1,09 103kN/cm2,Gxy =560 kN/cm2,
ν
xy =0,30,
ν
yx =0,021
The fibre angle
α
shall be varied:
α
=0,15,30,45. After the input of all parame-
Fig. 11.19 FE-model of can-
tilever beam in COSMOS/M
(650 elements, 714 nodes)
x
y
z
Fig. 11.20 Cantilever beam
deformed shape
x
y
z
11.5 Numerical Results 449
vyvz
Fibre Angle
Displacements/cm
010203040
0.5
1
1.5
2
2.5
3
3.5
4
4.5
Fig. 11.21 Displacements of the corner point at the free edge
σ
x
σ
y
Fibre Angle
Stresses kN/cm2
010203040
0
2
4
6
8
10
Fig. 11.22 Maximal stresses at the bottom of layer No. 1
450 11 Finite Element Analysis
L
Mt
+
α
α
3
3
D
6
Fig. 11.23 Laminate pipe: geometry, cross-section and stacking sequence (Mt=1200 kNcm,
L=2 m, D=200 mm)
ters and properties into COSMOS/M the FEM model can be illustrated (Fig. 11.24).
From the results of the analysis only the twisting angle of the free edge shall
be considered here. For this we have to list the results for the displacements in y-
direction of two nodes at the free edge, lying in opposite to each other, e.g. the nodes
255 and 663 in our FE model. The twisting angle is calculated by
ϕ
= (vy,255 vy,663 )/D
Carrying out the analysis for all fibre angles we obtain the results, given in Fig.
11.25. The diagram demonstrates the well known fact that in case of pure shear
loading the main normal stresses are lying in a direction with an angle of 45to the
shear stresses. Therefore here the fibre angels of +45/45to the longitudinal
axis are the most effective arrangements, because these fibre angels yield the greatest
shear rigidity.
Fig. 11.24 FE-model of Lam-
inate Pipe in COSMOS/M
(800 elements, 816 nodes)
x
y
z
11.5 Numerical Results 451
Fibre Angle
Twisting Angle
010203040
1
2
4
6
8
Fig. 11.25 Twisting angle of the free edge
11.5.1.3 Sandwich Plate
The sandwich plate (Fig. 11.26) is clamped at both short boundaries and simply
supported at one of the long boundaries. The cover sheets consist of an aluminium
alloy and the core of foam of polyurethan. The material parameters are:
AlZnMgCu0.5F450:
ρ
=2.7·103kg/m3,E=7.0·1010 N/m2,
ν
=0.34
polyurethan foam:
ρ
= 150 kg/m3,E=4.2·107N/m2,
ν
=0.30
Additional to a stress analysis of the plate under constant pressure loading p,
a vibration analysis will be performed is asked. We have to calculate the 4 lowest
eigenfrequencies and the mode shapes, respectively. Note that we use in this exam-
ple only the basic units of the SI-system, so we avoid the calculation of correction
factors for the obtained eigenfrequencies.
The FE-model is given in Fig. 11.27. The static analysis leads the displacements
and stresses. We consider only the stresses of the bottom of the lower cover sheet
(layer 3, top). The Figs. 11.28 and 11.29 show the plots of stress distributions for the
flexural stresses
σ
xand
σ
z, Fig. 11.30 the distribution of the von Mises equivalent
stress. The lowest 4 eigenfrequencies and their 4 mode shapes are shown in the
following Fig. 11.31. The static and frequency computations confirm the successful
application of the SHELL4L element for a sandwich plate.
452 11 Finite Element Analysis
6.0 m
4.0 m
p=1200 N/m2
aluminium alloy
polyurethan foam
24 30 mm
Fig. 11.26 Sandwich plate
Fig. 11.27 FE-model of Sandwich Plate in COSMOS/M (600 elements, 651 nodes)
11.5.1.4 Buckling Analysis of a Laminate Plate
For a rectangular laminate plate consisting of 4 layers with the given material pa-
rameters a buckling analysis shall be carried out. The plate is simply supported at
11.5 Numerical Results 453
x
y
z
2.3717E+007
1.4703E+007
5.6887E+006
-3.3255E+006
-1.2340E+007
-2.1354E+007
-3.0368E+007
-3.9382E+007
-4.8396E+007
σ
x
Fig. 11.28 Stresses in x-direction for the bottom of the lower cover sheet
x
y
z
1.1711E+007
8.1902E+006
4.6695E+006
1.1488E+006
-2.3719E+006
-5.8926E+006
-9.4133E+006
-1.2934E+007
-1.6455E+007
σ
z
Fig. 11.29 Stresses in z-direction for the bottom of the lower cover sheet
454 11 Finite Element Analysis
x
y
z
4.4196E+007
3.9218E+007
3.4240E+007
2.9262E+007
2.4284E+007
1.9306E+007
1.4328E+007
9.3496E+006
4.3715E+006
von Mises
Fig. 11.30 Von Mises stress for the bottom of the lower cover sheet
Fig. 11.31 Mode shapes for the lowest four eigenfrequencies: f1=5,926 Hz (top-left),
f2=12,438 Hz (top-right), f3=13,561 Hz (bottom-left), f4=19,397 Hz (bottom-right)
11.5 Numerical Results 455
all boundaries and loaded by a uniaxial uniform load, see Fig. 11.32. Material pa-
rameters are again the same as in the previous examples
Ex= 1.53 104kN/cm2,Ey= 1.09 103kN/cm2,Gxy = 560 kN/cm2,
ν
xy =0.30,
ν
yx =0.021
For the stacking structure two cases shall be considered, a symmetric (case I) and
a antisymmetric (case II) laminate structure (Fig. 11.32). The fibre angle is to vary:
α
=0,15,30,45,60,75,90. For the buckling analysis in COSMOS/M a unit
pressure loading must be created, and the program calculates a buckling factor
ν
B
to multiply the unit loading for obtaining the buckling load.
The FE-model created in COSMOS/M by the input of all properties and param-
eters is shown in Fig. 11.33. The calculation for
α
=30leads to a buckling factor
ν
B=1,647 and to the buckling mode shown in Fig. 11.34. In the same manner the
calculations for the other fibre angels and for the antisymmetric laminate were per-
formed. The results for the buckling factors are shown in a diagram in Fig. 11.35.
The buckling modes are symmetric to the symmetric axis in loading direction. For
the symmetric laminates the buckling modes for
α
=0,15,30are nearly the
same, see Fig. 11.34. For fibre angles 45,60,75,90the buckling modes have
different shapes, they are shown in the following figures. The buckling modes for
the antisymmetric laminate are very similar but not identical to the buckling modes
1.5 m
1.0 m
+
α
+
α
+
α
+
α
α
α
α
α
4×25
case I case II
Fig. 11.32 Rectangular laminate plate
Fig. 11.33 FE-model of the
laminate plate in COSMOS/M
(600 elements, 651 nodes)
456 11 Finite Element Analysis
Fig. 11.34 Buckling modes for symmetric laminates
α
=30(top-left),
α
=45(top-right),
α
=
60(middle-left),
α
=75(middle-right),
α
=90(bottom)
of the symmetric laminates. They are not given here. A fibre angle near 45leads to
the highest buckling load for a quadratic plate. It shall be noted that the antimetric
stacking sequence of the laminate improved the buckling stability.
11.5.2 Examples of the Use of Generalized Beam Elements
Generalized finite elements for the analysis of thin-walled beam shaped plate struc-
tures, Sect. 11.4, were implemented and tested in the frame of the general purpose
FEM-program system COSAR. The real handling of the FEM-procedures are not
given here, but two simple examples shall demonstrate the possibilities of these el-
ements for global static or dynamic structure analysis.
Figure 11.36 shows thin-walled cantilever beams with open or closed cross-
sections and different loadings. All these beam structure models have equal length,
hight and width and also the total thicknesses of all laminate strips are equal, inde-
pendent of the number of the layers.
11.5 Numerical Results 457
n, symm. n, antim.
Fibre Angle
Buckling Value
010 20 30 40 50 60 70 80 90
1
1.2
1.4
1.6
1.8
2
2.2
Fig. 11.35 Results of the Buckling Analysis
The stacking structure may be symmetric or antisymmetric. Figure 11.37 shows
the two considered variants: case A with three laminae and symmetric stacking and
case B with two laminae and antisymmetric stacking. The fibre reinforced material
is characterized again by the following effective moduli
EL=153000 N/mm2,
ν
LT =0.30,
ET=10900 N/mm2,
ν
TL =0.021,
GLT =5600 N/mm2,
ρ
=2 g/cm3
The fibre angles shall be varied.
Figure 11.38 shows the profile nodes. There are four main profile nodes for both
cross-sections but three secondary profile nodes for the open and four for the closed
cross-section. The numerical analysis shall demonstrate the influence of the stacking
structure. Figure 11.39 illustrates the relative changes of the cantilever beam in the
loaded point, if the symmetric stacking structure (wA) is change to the antisymmetric
one (wB). The antisymmetric layer stacking leads to higher values of the vertical
deflections wBin comparison to the wAvalues in the case of symmetric stacking.
Generally, only two degrees of freedom of secondary profile nodes were activated.
In a separate analysis the influence of a higher degrees of freedom in the sec-
ondary profile nodes was considered. As a result it can be recommended that for
antisymmetric layer structures and for open profiles more than two degrees of free-
458 11 Finite Element Analysis
F=260 kN
F
F
x
x
x
x
x
y
y
y
y
y
z
z
z
z
z
400 400
800
800
5000
Fig. 11.36 Cantilever beams, geometry and loading
11.5 Numerical Results 459
ab
ss
nn
α
α
+
α
+
α
+
α
Fig. 11.37 Stacking structure of the laminates. aSymmetric sandwich, btwo-layer
44
4
3
333
2
2
22
111
1
Fig. 11.38 Cross-sections with main profile nodes () and secondary profile nodes (×)
dom should be activated. Ignoring the activation of secondary profile node degrees
of freedom leads to nonrealistic structure stiffness. The structure model is to stiff
and therefore the deflections are to small.
The second example concern ed the eigen-vibration analysis. For the closed cross-
section with symmetric layer stacking the influences of the degree of freedom of
secondary profile nodes and of the variation of the fibre angles were considered. As
a result it can be stated that the influence of higher degrees of freedom of the sec-
ondary profile nodes is negligible but the influence of the fibre angles is significant.
Figure 11.40 illustrates the influence of the fibre angle variations on the eigenfre-
quencies of the beam, which can be used for structure optimization.
Summarizing Sect. 11.5 one have to say that only a small selection of one- and
two-dimensional finite elements was considered. Many finite plate and shell ele-
ments were developed using equivalent single layer theories for laminated struc-
tures but also multi-layered theories are used. Recent review articles give a detailed
overview on the development, implementation and testing of different finite lami-
nate and sandwich elements.
460 11 Finite Element Analysis
0
2
4
6
8
1 0
1 2
1 4
0 ° 1 0 ° 2 0 ° 3 0 ° 4 0 ° 5 0 ° 6 0 ° 7 0 ° 8 0 ° 9 0 °
w
%
Fig. 11.39 Relative changes (wBwA)/wA=
w(
α
)of the vertical deflections wAand wB
0
2 0
4 0
6 0
8 0
1 0 0
1 2 0
1 4 0
0 ° 1 0 ° 2 0 ° 3 0 ° 4 0 ° 5 0 ° 6 0 ° 7 0 ° 8 0 ° 9 0 °
Hz
1. EF
2. EF
3. EF
4. EF
Fig. 11.40 Influence of the fibre angle on the first four eigen-frequencies of the cantilever box-
beam
Part VI
Appendices
This part is focussed on some basics of mathematics and mechanics like matrix
operations (App. A), stress and strain transformations (App. B), differential opera-
tors for rectangular plates (App. C) and differential operators for circular cylindrical
shells (App. D). In addition, the Krylow functions as solution forms of a special
fourth order ordinary differential equation are discussed (App. E) and some mate-
rial’s properties are given in App. F. In the last one section like always in this book
material or constitutive parameters are used as usual. Note that any material param-
eter is a parameter since there are dependencies on temperature, time, etc. Last but
not least there are given some references for further reading (App. G).
Appendix A
Matrix Operations
The following short review of the basic matrix definitions and operations will pro-
vide a quick reference and ensure that the particular use of vector-matrix notations
in this textbook is correct understood.
A.1 Definitions
1. Rectangular matrix
A
A
A=
a11 a12 ······ a1n
a21 a22 ······ a2n
.
.
..
.
..
.
..
.
..
.
.
am1am2······ amn
= [ai j]
Rectangular matrix with i=1,2,...,mrows and j=1,2,...,ncolumns, is a
rectangular-ordered array of quantities with mrows and ncolumns. m×nor
often (m,n)is the order of the matrix, ai j is called the (i,j)-element of A
A
A. There
are two important special cases
a
a
a=
a1
.
.
.
am
= [ai]
is a m×1 matrix or column vector, while
a
a
aT=a1···an= [ai]T
is a 1 ×nmatrix or row vector.
With a
a
arespectively a
a
aTa matrix A
A
Acan be written
463
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
464 A Matrix Operations
A
A
A=a
a
a1···a
a
an,a
a
aj=
a1j
.
.
.
am j
,j=1,...,n
or
A
A
A=
a
a
aT
1
···
a
a
aT
m
,a
a
aT
j=aj1···ajn ,j=1,...,m
If n=mthe matrix is square of the order n×n. For a square matrix the elements
ai j with i=jdefine the principal matrix diagonal and are located on it.
2. Determinant of a square matrix A
A
A
|A
A
A|=
a11 a12 ······a1n
a21 a22 ······a2n
.
.
..
.
..
.
..
.
..
.
.
an1an2······ann
=|ai j|=detA
A
A
The determinant of a matrix A
A
Awith elements ai j is given by
|A
A
A|=a11M11 a12M12 +a13M13 ...(1)1+na1nM1n,
where the minor Mi j is the determinant of the matrix |A
A
A|with missing row iand
column j. Note the following properties of determinants:
Interchanging two rows or two columns changing the sign of |A
A
A|.
If all elements in a row or a column of A
A
Aare zero then |A
A
A|=0.
Multiplication by a constant factor cof all elements in a row or column of A
A
A
multiplies |A
A
A|by c.
Adding a constant multiple of row or column kto row or column ldoes not
change the determinant.
If one row kis a linear combination of the rows land mthen the determinant
must be zero.
3. Regular matrix
A square matrix A
A
Ais regular if |A
A
A| 6=0.
4. Singular matrix
A square matrix A
A
Ais singular if |A
A
A|=0.
5. Trace of a matrix
The trace of a square matrix A
A
Ais the sum of all elements of the principal diagonal,
i.e.
trA
A
A=
m
k=1
akk
6. Rank of a matrix
The rank rk(A
A
A)of a m×nmatrix A
A
Ais the largest value of rfor which there exist a
A.2 Special Matrices 465
r×rsubmatrix of A
A
Athat is non-singular. Submatrices are smaller arrays of k×k
elements ai j of the matrix A
A
A, i.e. kmif mnor knif nm.
A.2 Special Matrices
In the following the
δ
i j denotes the Kronecker symbol
δ
i j =0i6=j
1i=j
1. Null matrix 0
0
0
All elements ai j of a m×nmatrix are identically equal zero
ai j 0,i=1,...,m,j=1,...,n
2. Diagonal matrix D
D
D=diag[aii] = diag[ai j
δ
i j]
A diagonal matrix is a square matrix in which all elements are zero except those
on the principal diagonal
ai j =0,i6=j,ai j 6=0,i=j
3. Unit matrix I
I
I= [
δ
i j]
A unit or identity matrix is a special case of the diagonal matrix for which aij =1
when i=jand ai j =0 when i6=j.
4. Transpose A
A
ATof a matrix A
A
A
The transpose of a matrix A
A
Ais found by interchanging rows and columns. If
A
A
A= [ai j]follow A
A
AT= [aT
i j]with aT
i j =aji . A transposed matrix is denoted by a
superscript T. Note (A
A
AT)T=A
A
A
5. Symmetric matrix A
A
AS
A square matrix A
A
Ais said to be symmetric if for all i6=j ai j =aji , i.e. A
A
A=A
A
AT. A
symmetric matrix is denoted by a superscript S.
6. Skew-symmetric matrix A
A
AA
A square matrix A
A
Ais said to be skew-symmetric if all principal diagonal elements
are equal zero and for all i6=j aij =aji , i.e. A
A
A=A
A
AT. A skew-symmetric
matrix is denoted by a superscript A.
7. Any matrix can be decomposed in a symmetric and a skew-symmetric part in a
unique manner
A
A
A=A
A
AS+A
A
AA
Proof. Since
A
A
AS=1
2(A
A
A+A
A
AT)
and
A
A
AA=1
2(A
A
AA
A
AT)
466 A Matrix Operations
the sum A
A
AS+A
A
AAis equal to A
A
A.
A.3 Matrix Algebra and Analysis
1. Addition and subtraction
Am×nmatrix A
A
Acan be added or subtracted to a m×nmatrix B
B
Bto form a m×n
matrix C
C
C
A
A
A±B
B
B=C
C
C,ai j ±bi j =ci j,i=1,...,m,j=1,...,n
Note A
A
A+B
B
B=B
B
B+A
A
A,A
A
AB
B
B=(B
B
BA
A
A) = B
B
B+A
A
A,(A
A
A±B
B
B)T=A
A
AT±B
B
BT.
2. Multiplication
Multiplication the matrix A
A
Aby a scalar
α
involves the multiplication of all
elements of the matrix by the scalar
α
A
A
A=A
A
A
α
= [
α
ai j],
(
α
±
β
)A
A
A=
α
A
A
A±
β
A
A
A,
α
(A
A
A±B
B
B) =
α
A
A
A±
α
B
B
B
The product of a (1×n)matrix (row vector a
a
aT) and a (n×1)matrix (column
vector b
b
b) forms a (1×1)matrix, i.e. a scalar
α
a
a
aTb
b
b=b
b
bTa
a
a=
α
,
α
=
n
k=1
akbk
The product of a (m×n)matrix A
A
Aand a (n×1)column vector b
b
bforms a
(m×1)column vector c
c
c
Ab
Ab
Ab =c
c
c,ci=
n
j=1
ai jbj=ai1b1+ai2b2+...+ainbn,i=1,2,...,m
The forgoing product is only possible if the number of columns of A
A
Ais equal
the number of rows of b
b
b.
Note A.1. b
b
bTA
A
AT=c
c
cT
If A
A
Ais a (m×n)matrix and B
B
Ba(p×q)matrix the product AB
AB
AB =C
C
Cexists if
n=p, in which case C
C
Cis a (m×q)matrix. For n=pthe matrix A
A
Aand B
B
Bare
said to be confor mable for multiplication. The elements of the matrix C
C
Care
ci j =
n=p
k=1
aikbk j ,i=1,2,...,m,j=1,2,...,q
Note A.2. AB
AB
AB 6=BA
BA
BA,A
A
A(B
B
B±C
C
C) = AB
AB
AB ±AC
AC
AC,(AB
AB
AB)T=B
B
BTA
A
AT
A.3 Matrix Algebra and Analysis 467
(l×m)
A
A
A
(m×n)
B
B
B
(n×p)
C
C
C=
(l×p)
D
D
D
3. Inversion and division
AI
AI
AI =IA
IA
IA =A
A
A,A
A
A1A
A
A=A
A
AA
A
A1=I
I
I,A
A
A11=A
A
A
The matrix inversion is based on the existence of a n×nunit matrix I
I
Iand a square
n×nmatrix A
A
A.A
A
A1is the inverse of A
A
Awith respect to the matrix multiplication
A
A
AA
A
A1=A
A
A1A
A
A=I
I
I. If A
A
A1exist, the matrix A
A
Ais invertible or regular, otherwise
non-invertible or singular. Matrix division is not defined.
Note A.3. (AB
AB
AB)1=B
B
B1A
A
A1,(ABC
ABC
ABC)1=C
C
C1B
B
B1A
A
A1,...
Cofactor matrix
With the minors Mi j introduced above to define the determinant |A
A
A|of a matrix
A
A
Aa so-called cofactor matrix A
A
Ac= [Ai j]can be defined, where
Ai j = (1)i+jMi j
The cofactor matrix is denoted by the superscript c.
Adjoint or adjugate matrix
The adjoint matrix of the square matrix A
A
Ais the transpose of the cofactor
matrix
adjA
A
A= (A
A
Ac)T
Note A.4. Because symmetric matrices possess symmetric cofactor matrices
the adjoint of a symmetric matrix is the cofactor matrix itself
adjA
A
AS=A
A
ASc
It can be shown that
A
A
A(adjA
A
A) = |A
A
A|I
I
I
i.e. A
A
A(adjA
A
A)
|A
A
A|=I
I
I=A
A
AA
A
A1A
A
A1=adjA
A
A
|A
A
A|
Inverse matrices have some important properties
A
A
AT1=A
A
A1T
and if A
A
A=A
A
AT
A
A
A1=A
A
A1T
i.e. the inverse matrix of a symmetric matrix A
A
Ais also symmetric.
Note A.5. Symmetric matrices posses symmetric transposes, symmetric cofac-
tors, symmetric adjoints and symmetric inverses.
468 A Matrix Operations
4. Powers and roots of square matrices
If n×nmatrix A
A
Ais conformable with itself for multiplication, one may define its
powers
A
A
An=A
A
AA
A
A...A
A
A,
and for symmetric positive semidefinite matrices
A
A
A1
n=n
A
A
A,A
A
An=A
A
A1n
and if A
A
Ais regular
(A
A
Am)n=A
A
Amn,A
A
AmA
A
An=A
A
Am+n,
5. Matrix eigenvalue problems
The standard eigenvalue problem of a quadratic n×nmatrix A
A
Ais of the form:
find (
λ
,x
x
x)with 6=
6=
6=0
0
0 such that
A
A
Ax
x
x=
λ
x
x
xor (A
A
A
λ
I
I
I)x
x
x=0
0
0
K
K
K= [A
A
A
λ
I
I
I]is called the characteristic matrix of A
A
A, detK
K
K=0 is called the char-
acteristic determinant or equation of A
A
A. The characteristic determinant produces
a characteristic polynomial with powers of
λ
up to
λ
nand therefore when it set
equal zero having nroots which are called the eigenvalues. If the characteristic
equation has ndistinct roots, the polynomial can be factorized in the form
(
λ
λ
1)(
λ
λ
2)...(
λ
λ
n) = 0
If we put
λ
=0 in the characteristic equation we get
detA
A
A=
λ
1
λ
2...
λ
n
Inserting any root
λ
iinto the standard eigenvalue equation leads to
[A
A
A
λ
iI
I
I]x
x
xi=0
0
0,i=1,2,...,n
x
x
xiare the eigendirections (eigenvectors) which can be computed from the last
equation considering the orthogonality condition. A nontrivial solution exists if
and only if
det[A
A
A
λ
iI
I
I] = 0
Note A.6. If we have the 3x3 symmetric matrix the eigendirection x
x
xican be com-
puted for each
λ
ifrom the polynomial of third order. Three different solutions
are possible:
all solutions
λ
iare distinct - three orthogonal eigendirections can be computed
(but their magnitudes are arbitrary),
one double solution and one distinct solutions - only one eigendirection can
be computed (its magnitudes is arbitrary), and
all solutions are identically - no eigendirections can be computed.
A.3 Matrix Algebra and Analysis 469
Anyway, the orthogonality condition x
x
xT·x
x
x=1 should be taken into account.
The general eigenvalue problem is given in the form
A
A
Ax
x
x=
λ
B
B
Bx
x
x
which can be premultiplied by B
B
B1to produce the standard form
B
B
B1A
A
Ax
x
x=B
B
B1
λ
B
B
Bx
x
x= (B
B
B1A
A
A)x
x
x=
λ
I
I
Ix
x
x=
λ
x
x
x
resulting in
(B
B
B1A
A
A
λ
I
I
I)x
x
x=0
0
0
Note A.7. In the case of non-symmetric matrix A
A
Athe eigenvalue can be complex.
6. Differentiating and integrating
To differentiate a matrix one differentiates each matrix element ai j in the con-
ventual manner.
To integrate a matrix one integrates each matrix element ai j in the conventual
manner. For definite integrals, each term is evaluated for the limits of integra-
tion.
7. Partitioning of matrices
A useful operation with matrices is partitioning into submatrices. These subma-
trices may be treated as elements of the parent matrix and manipulated by the
standard matrix rules reviewed above. The partitioning is usually indicated by
dashed partitioning lines entirely through the matrix
M
M
M= [mi j] =
A
A
A.
.
.B
B
B
·········
C
C
C.
.
.D
D
D
For a m×nmatrix M
M
Mwe may have submatrices A
A
A(i×j),B
B
B(i×p),C
C
C(k×j),
D
D
D(k×p)with i+k=m,j+p=n, i.e.
M
M
Mm×n=
A
A
Ai×j
.
.
.B
B
Bi×(nj)
··· ··· ···
C
C
C(mi)×j
.
.
.D
D
D(mi)×(nj)
,
A
A
A.
.
.B
B
B
·········
C
C
C.
.
.D
D
D
±
E
E
E.
.
.F
F
F
·········
G
G
G.
.
.H
H
H
=
A
A
A±E
E
E.
.
.B
B
B±F
F
F
··· ··· ···
C
C
C±G
G
G.
.
.D
D
D±H
H
H
,
470 A Matrix Operations
A
A
A.
.
.B
B
B
·········
C
C
C.
.
.D
D
D
E
E
E.
.
.F
F
F
·········
G
G
G.
.
.H
H
H
=
AE
AE
AE +BG
BG
BG .
.
.AF
AF
AF +BH
BH
BH
··· ··· ···
CE
CE
CE +DG
DG
DG .
.
.CF
CF
CF +DH
DH
DH
The multiplications are only defined if the correspondent matrices are con-
formable for multiplication
A
A
A.
.
.B
B
B
·········
C
C
C.
.
.D
D
D
T
=
A
A
AT.
.
.C
C
CT
··· ··· ···
B
B
BT.
.
.D
D
DT
If the matrix
M
M
M=
A
A
A.
.
.B
B
B
·········
C
C
C.
.
.D
D
D
is symmetric (M
M
M=M
M
MT), it follows A
A
A=A
A
AT,D
D
D=D
D
DT,B
B
B=C
C
CT,C
C
C=B
B
BT
Appendix B
Stress and Strain Transformations
Stress and strain transformations under general orthogonal coordinate transforma-
tion e
e
e=R
R
Re
e
eor e
i=Ri jej:
1.
σ
p=T
σ
pq
σ
q. The matrix [T
σ
pq]is defined by
R2
11 R2
12 R2
13 2R12R13 2R11 R13 2R11R12
R2
21 R2
22 R2
23 2R22R23 2R21 R23 2R21R22
R2
31 R2
32 R2
33 2R32R33 2R31 R33 2R31R32
R21R31 R22 R32 R23R33 R22R33 +R23 R32 R21R33 +R23R31 R21R32 +R22 R31
R11R31 R12 R32 R13R33 R12R33 +R13 R32 R11R33 +R13R31 R11R32 +R12 R31
R11R21 R12 R22 R13R23 R12R23 +R13 R22 R11R23 +R13R21 R11R22 +R12 R21
2.
ε
p=T
ε
pq
ε
q. The matrix [T
ε
pq]is defined by
R2
11 R2
12 R2
13 R12R13 R11 R13 R11R12
R2
21 R2
22 R2
23 R22R23 R21 R23 R21R22
R2
31 R2
32 R2
33 R32R33 R31 R33 R31R32
2R21R31 2R22 R32 2R23R33 R22 R33+R23 R32 R21R33 +R23R31 R21R32 +R22 R31
2R11R31 2R12 R32 2R13R33 R12 R33+R13 R32 R11R33 +R13R31 R11R32 +R12 R31
2R11R21 2R12 R22 2R13R23 R12 R23+R13 R22 R11R23 +R13R21 R11R22 +R12 R21
3. Rotation about the e
e
e1-direction, Fig. B.1:
[Ri j] =
1 0 0
0c s
0s c
,e
e
e=1
R
R
R e
e
e
471
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
472 B Stress and Strain Transformations
x1,x
1
x2
x3
x
2
x
3
e
e
e1,e
e
e
1
e
e
e2
e
e
e3
e
e
e
2
e
e
e
3
Φ
Φ
Fig. B.1 Rotation about the e
e
e1-direction
[
1
T
σ
pq] =
1 0 0 0 0 0
0c2s22cs 0 0
0s2c22cs 0 0
0cs cs c2s20 0
0 0 0 0 cs
0 0 0 0 s c
,
σ
σ
σ
=
1
T
T
T
σσ
σ
σ
[
1
T
ε
pq] =
1 0 0 0 0 0
0c2s2cs 0 0
0s2c2cs 0 0
02cs 2cs c2s20 0
0 0 0 0 cs
0 0 0 0 s c
,
ε
ε
ε
=
1
T
T
T
εε
ε
ε
Appendix C
Differential Operators for Rectangular Plates
Below two cases will be discussed
the classical plate theory and
the shear deformation theory.
C.1 Classical Plate Theory
1. General unsymmetric laminates
L11 L12 L13
L22 L23
sym L33
u
v
w
=
0
0
p
,
L11 =A11
2
x2
1
+2A16
2
x1
x2
+A66
2
x2
2
,
L22 =A22
2
x2
2
+2A26
2
x1
x2
+A66
2
x2
1
,
L33 =D11
4
x4
1
+4D16
4
x3
1
x2
+2(D16 +2D66)
4
x2
1
x2
2
+4D26
4
x1
x3
2
+D22
4
x4
2
,
L12 =A16
2
x2
1
+ (A12 +A66)
2
x1
x2
+A26
2
x2
2
,
L13 =B11
3
x3
1
+3B16
3
x2
1
x2
+ (B12 +2B66)
3
x1
x2
2
+B26
3
x3
2,
473
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
474 C Differential Operators for Rectangular Plates
L23 =B22
3
x3
2
+3B26
3
x1
x2
2
+ (B12 +2B66)
3
x2
1
x2
+B16
3
x3
1
2. General symmetric laminates
L11 L12 0
L22 0
sym L33
u
v
w
=
0
0
p
Bi j =0, i.e. L13 =L31 =0,L23 =L32 =0. L11,L22 ,L33,L12 as above in 1.
3. Balanced symmetric laminates
L11 L12 0
L22 0
sym L33
u
v
w
=
0
0
p
In addition to 2. both A16 and A26 are zero, i.e. L13 =L31 =0, L23 =L32 =0 and
L11,L22 and L33 simplify with A16 =A26 =0.
4. Cross-ply symmetric laminates
L11 L12 0
L22 0
sym L33
u
v
w
=
0
0
p
In addition to 3. both D16 and D26 are zero, i.e. L33 simplifies.
5. Balanced unsymmetric laminates
L11 L12 L13
L22 L23
sym L33
u
v
w
=
0
0
p
With A16 =A26 =0 only the operators L11 ,L22 and L12 of 1. can be simplified.
6. Cross-ply unsymmetric laminates
L11 L12 L13
L22 L23
sym L33
u
v
w
=
0
0
p
In addition to 5., D16,D26 ,B16 and B26 are zero, i.e. all operators of 1. can be
simplified.
C.2 Shear Deformation Theory 475
C.2 Shear Deformation Theory
1. General unsymmetrical laminates
˜
L11 ˜
L12 ˜
L13 ˜
L14 0
˜
L22 ˜
L23 ˜
L24 0
˜
L33 ˜
L34 ˜
L35
˜
L44 ˜
L45
S Y M ˜
L55
u
v
ψ
1
ψ
2
w
=
0
0
0
0
p
with ˜
L11 =L11,˜
L22 =L22,˜
L12 =L12 (the Li j can be taken from Appendix C.1)
and
˜
L33 =D11
2
x2
1
+2D16
2
x1
x2
+D66
2
x2
2ks
55A55,
˜
L44 =D66
2
x2
1
+2D26
2
x1
x2
+D22
2
x2
2ks
44A44,
˜
L55 =ks
55A55
2
x2
1
+ks
45A45
2
x1
x2
+ks
44A44
2
x2
2,
˜
L13 =˜
L31 =B11
2
x2
1
+2B26
2
x1
x2
+B66
2
x2
2
,
˜
L14 =˜
L41 =˜
L23 =˜
L32 =B16
2
x2
1
+ (B12 +B66)
2
x1
x2
+B26
2
x2
2
,
˜
L24 =˜
L42 =B66
2
x2
1
+2B26
2
x1
x2
+B22
2
x2
2
,
˜
L34 =˜
L43 =D16
2
x2
1
+ (D12 +D66)
2
x1
x2
+D26
2
x2
2
,
˜
L35 =˜
L53 =ks
55A55
x1
+ks
45A45
x2,
˜
L45 =˜
L54 =ks
45A45
x1
+ks
44A44
x2
with ks
45 =pks
44ks
55.
2. General symmetric laminates
Bi j =0, i.e. ˜
L13 =˜
L31,˜
L14 =˜
L41,˜
L23 =˜
L32 and ˜
L24 =˜
L42 are zero
˜
L11 ˜
L12
˜
L12 ˜
L22 u
v=0
˜
L33 ˜
L34 ˜
L35
˜
L34 ˜
L44 ˜
L45
˜
L53 ˜
L54 ˜
L55
ψ
1
ψ
1
w
=
0
0
p
476 C Differential Operators for Rectangular Plates
3. Cross-ply symmetric laminates
In addition to 2. both D16 ,D26 and A16,A26 ,A45 are zero.
Appendix D
Differential Operators for Circular Cylindrical
Shells
Below two cases will be considered
the classical case and
the first order shear deformation theory.
D.1 Classical Shell Theory
1. General unsymmetrical laminates
L11 L12 L13
L22 L23
SYM L33
u
v
w
=
px
ps
pz
L11 =A11
2
x2+2A16
2
x
s+A66
2
s2,
L12 = (A16 +R1B16)
2
x2
1
+ (A12 +R1B12 +A66 +R1B66)
2
x
s
+ (A26 +R1B26)
2
s2,
L13 =R1A16
x+R1A26
sB11
3
x33B16
3
x2
s
(B12 +B66)
3
x
s2B26
3
s3,
L22 = (A66 +2R1B66 +R2D66)
2
x2
477
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
478 D Differential Operators for Circular Cylindrical Shells
+2(A26 +2R1B26 +2R2D26)
2
x
s
+ (A22 +2R1B22 +R2D22)
2
s2,
L23 =R1(A26 +R1B26)
x+R1(A22 +R1B22)
s
(B16 +R1D16)
3
x3[B12 +2B66 +R1(D12 +2D66)]
3
x2
s
3(B26 +R1D26)
3
x
s2(B22 +R1D22)
3
s3,
L33 =R2(A22 +R1B22) + 2R1B12
2
x2+4R1B26
2
x
s+2R1B22
2
s2
D11
4
x44D16
4
x3
s2(D12 +2D66)
4
x2
s2
4D26
4
x
s3D22
4
s4
2. General symmetrical laminates
All Bi j =0, but the matrix [Li j ]is full populated, i.e. all [Li j ]are nonequal zero.
Note that for general symmetrically laminated circular cylindrical shells there is
a coupling of the in-plane and out-of-plane displacements and stress resultants.
3. Cross-ply symmetrical laminates
Bi j =0,A16 =A26 =0,D16 =D26 =0
4. Cross-ply antisymmetrical laminates
B22 =B11,all other Bi j =0,A16 =A26 =0,D16 =D26 =0
5. Axisymmetric deformations of symmetrical cross-ply laminates
Additional to 3. all derivative
/
sand the displacement vare taken zero and
yield
L12 =L13 =L23 0
D.2 Shear Deformation Theory 479
D.2 Shear Deformation Theory
1. General unsymmetrical laminates
˜
L11 ˜
L12 ˜
L13 ˜
L14 ˜
L15
˜
L21 ˜
L22 ˜
L23 ˜
L24 ˜
L25
˜
L31 ˜
L32 ˜
L33 ˜
L34 ˜
L35
˜
L41 ˜
L42 ˜
L43 ˜
L44 ˜
L45
˜
L51 ˜
L52 ˜
L53 ˜
L54 ˜
L55
u
v
ψ
1
ψ
2
w
=
px
ps
0
0
pz
,
˜
L11 =A11
2
x2+2A16
2
x
s+A66
2
s2,
˜
L12 =A16
2
x2+ (A12 +A66)
2
x
s+A26
2
s2,
˜
L13 =B11
2
x2+2B16
2
x
s+B66
2
s2,
˜
L14 =B16
2
x2+ (B12 +B66)
2
x
s+B26
2
s2,
˜
L15 =1
RA12
x+1
RA26
s,
˜
L22 =A66
2
x2+2A26
2
x
s+A22
2
s21
R2ks
44A44 ,
˜
L23 =B16
2
x2+ (B16 +B66)
2
x
s+B26
2
s2+1
Rks
45A45 ,
˜
L24 =B66
2
x2+2B26
2
x
s+B22
2
s2+1
Rks
44A44,
˜
L25 = (A12 +ks
55A55)1
R
x+ (A26 +ks
45A45)1
R
s,
˜
L33 =D11
2
x2+2D16
2
x
s+D66
2
s2ks
55A55
˜
L34 =D16
2
x2+ (D12 +D66)
2
x
s+D26
2
s2ks
45A45,
˜
L35 =B12
1
RAs
55
x+B26
1
Rks
45A45
s,
˜
L44 =D66
2
x2+2D26
2
x
s+D22
2
s2ks
44A44,
˜
L45 =B26
1
Rks
45A45
x+B22
1
Rks
44A44
s,
˜
L55 =A55
2
x2+2ks
45A45
2
x
sA22)1
R2
480 D Differential Operators for Circular Cylindrical Shells
and
˜
L51 =˜
L15,˜
L52 =˜
L25,˜
L53 =˜
L35,˜
L54 =˜
L45,
ks
45 =qks
44ks
55
2. Cross-ply symmetrical laminates
Bi j =0,i,j=1,2,6,
A16 =A26 =A45 =0,D16 =D26 =0
3. Cross-ply antisymmetrical laminates
B22 =B11,all other Bi j =0,i,j=1,2,6,
A16 =A26 =A45 =0,D16 =D26 =0
Appendix E
Krylow-Functions as Solution Forms of a Fourth
Order Ordinary Differential Equation
The solutions of the following fourth order ordinary differential equation
w′′′′ k2
1w′′ +k4
2w=0
can be presented in the form of so-called Krylow functions (Filonenko-Boroditsch,
1952):
1. k2
2>k2
1
α
1=
α
2=a+ib,
α
3=
α
4=aib,
a=r1
2(k2
2+k2
1),b=r1
2(k2
2k2
1)
and the solutions are
Φ
1=cosh ax cosbx,
Φ
2=sinhaxsinbx,
Φ
3=cosh ax sinbx,
Φ
4=sinhaxcosbx
or
Φ
1=eax cosbx,
Φ
2=eax sinbx,
Φ
3=eax sinbx,
Φ
4=eax cosbx
2. k2
2<k2
1
α
1=
α
2=a,
α
3=
α
4=b,
a=rk2
1qk4
1k4
2,b=rk2
1+qk4
1k4
2
and the solutions are
Φ
1=cosh ax,
Φ
2=coshbx,
Φ
3=sinhax,
Φ
4=sinhbx
or
Φ
1=eax,
Φ
2=ebx,
Φ
3=eax,
Φ
4=ebx
481
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
482 E Krylow-Functions as Solution Forms of a Fourth Order Ordinary Differential Equation
3. k2
2=k2
1
α
1=
α
2=a,
α
3=
α
4=a
and the solutions are
Φ
1=coshax,
Φ
2=xsinhax,
Φ
3=sinhax,
Φ
4=xcoshax
or
Φ
1=eax,
Φ
2=xeax,
Φ
3=eax,
Φ
4=xeax
References
Filonenko-Boroditsch MM (1952) Festigkeitslehre, vol II. Verlag Technik, Berlin
Appendix F
Material’s Properties
Below material properties for classical materials, for the constituents of various
composites and for unidirectional layers are presented. The information about the
properties were taken from different sources (see the Handbooks, Textbooks and
Monographs at the end of this appendix).
Note that the presentation of material data in a unique way is not so easy due to
the incompleteness of material data in the original sources. This means that there are
some empty places in the above following tables. The authors of this textbook were
unable to fill out these places. Another problem is connected with the different unit
systems in the original sources. For recalculation approximate relations are used
(e.g. 1 kgf 10 N).
With respect to the quick changes in application composite materials all material
data are only examples showing the main tendencies. Every year new materials are
developed and for the material data one have to contact directly the companies.
References
Czichos H, Hennecke M (eds) (2012) H ¨utte - das Ingenieurwissen, 34th edn.
Springer, Berlin, Heidelberg
Grote KH, Feldhusen J (eds) (2014) Dubbel - Taschenbuch f¨ur den Maschinenbau,
24th edn. Springer Vieweg, Berlin, Heidelberg
Hyer M (1998) Stress Analysis of Fiber-Reinforced Composite Materials. McGraw-
Hill, Boston et al.
Vasiliev V, Morozov E (2001) Mechanics and Analysis of Composite Materials.
Elsevier, Amsterdam
483
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
484 F Material’s Properties
Table F.1 Material properties of conventional materials at room temperature (bulk form), after
Grote and Feldhusen (2014)] and Czichos and Hennecke (2012)
Density Young’s Maximum Ultimate Maximum Coefficient
modulus specific strength specific of thermal
modulus strength expansion
ρ
E E/
ρ σ
u
σ
u/
ρ α
(g/cm3) (GPa) (MNm/kg) (MPa) (kNm/kg) (106/K)
Steel 7.8-7.85 180-210 27 340-2100 270 13
Gray cast iron 7.1-7.4 64-181 25 140-490 69 9-12
Aluminium 2.7-2.85 69-72 27 140-620 230 23
Titanium 4.4-4.5 110 25 1000-1200 273 11
Magnesium 1.8 40 22 260 144 26
Beryllium 1.8-1.85 300-320 173 620-700 389
Nickel 8.9 200 22 400-500 56 13
Zirconium 6.5 100 15 390 60 5.9
Tantalum 16.6 180 11 275 17 6.5
Tungsten 19.3 350 18 1100-4100 212 6.5
Glass 2.5 70 28 700-2100 840 3.5-5.5
References 485
Table F.2 Material properties of fibre materials, after Hyer (1998)
Density Young’s Maximum Ultimate Maximum Coefficient Diameter
modulus specific strength specific of thermal
modulus strength expansion
- fibre
direction
ρ
E E/
ρ σ
u
σ
u/
ρ α
d
(g/cm3) (GPa) (MNm/kg) (MPa) (kNm/kg) (106/K (
µ
m)
E-Glass 2.54 72.4 29 3450 1358 5 8-14
C-Glass 2.49 68.9 28 3160 1269 7.2
S-Glass 2.49 85.5 34 4600 1847 5.6 10
Carbon
Intermediate 1.78-1.82 228-276 155 2410-2930 1646 -0.1- -0.5 8-9
modulus
High modulus 1.67-1.9 331-400 240 2070-2900 1736 -1- -4 5-7
High strength 1.85 240 130 3500 1892 -1- -4 5-7
Polymeric fibres
Kevlar-29 1.44 62 43 2760 1917 -2 12
Aramid (Kevlar-49) 1.48 131 89 2800-3792 2562 -2 12
Spectra 900 0.97 117 121 2580 2660 38
Boron 2.63 385 146 2800 1065 4 100-140
Boron Carbide 2.5 480 192 2100-2500 1000 50
Boron Nitride 1.9 90 47 1400 737 7
Titanium Carbide 4.9 450 92 1500 306 280
486 F Material’s Properties
Table F.3 Material properties of matrix and core materials, after Hyer (1998)
Density Young’s Shear Young’s Ultimate Coefficient
modulus modulus modulus strength of thermal
(tension) (compression) expansion
ρ
EtG Ec
σ
u
α
(g/cm3) (GPa) (GPa) (GPa) (MPa) (106/K)
Thermosetting polymers
Polyester 1.2-1.3 3-4.2 0.7-2 90-250 40-90 80-150
Vinyl ester 1.15 3-4 127 65-90 80-150
Bismaleimide 1.32 3.6 1.8 200 48-78 49
Polyimide 1.43-1.89 3.1-4.9 70-120 90
Epoxy 1.1-1.6 3-6 1.1-1.2 100-200 30-100 45-80
Thermoplastic polymers
PEEK 1.32 3.6 1.38 140 92-100 47
PPS 1.34 2.5 70-75 54-100
Polysulfone 1.24 2.5 70-75 56-100
Polypropylene 0.9 1-1.4 0.38-0.54 25-38 110
Nylon 1.14 1.4-2.8 0.54-1.08 34 60-75 90
Polcarbonate 1.06-1.2 2.2-2.4 86 45-70 70
Ceramics
Borosilicate glass 2.3 64 26.4 100 3
Balsa wood 0.1-0.19 2-6 8-18
Polystyrene 0.03-0.07 0.02-0.03 0.25-1.25
References 487
Table F.4 Material properties of selected unidirectional composites
E-Glass/ S-Glass/ Kevlar/ Boron/ Carbon
epoxy epoxy epoxy epoxy epoxy
Fibre volume fraction vf0.55 0.50 0.6 0.5 0.63
Density (g/cm3)2.1 2.0 1.38 2.03 1.58
Longitudinal modulus EL(GPa) 39 43 87 201 142
Transverse modulus ET(GPa) 8.6 8.9 5.5 21.7 10.3
In-plane shear modulus GLT (GPa) 3.8 4.5 2.2 5.4 7.2
Major Poisson’s ratio
ν
LT 0.28 0.27 0.34 0.17 0.27
Minor Poisson’s ratio
ν
TL 0.06 0.06 0.02 0.02 0.02
Longitudinal ultimate stress
σ
Lu (MPa) 1080 1280 1280 1380 2280
Transverse ultimate stress
σ
Tu (MPa) 39 49 30 56 57
In-plane ultimate shear stress
σ
LTu (MPa) 89 69 49 62 71
Longitudinal thermal
expansion coefficients
α
L(106/K) 7 5 -2 6.1 -0.9
Transverse thermal
expansion coefficient
α
T(106/K) 21 26 60 30 27
488 F Material’s Properties
Table F.5 Typical properties of unidirectional composites as functions of the fibre volume fraction,
after Vasiliev and Morozov (2001)
Glass/ Carbon/ Carbon/ Aramid/ Boron/ Boron/ Carbon/ Al2O3/
epoxy epoxy PEEK epoxy epoxy aluminium carbon aluminium
Fibre volume
fraction vf0.65 0.62 0.61 0.6 0.5 0.5 0.6 0.6
Density
ρ
(g/cm3) 2.1 1.55 1.6 1.32 2.1 2.65 1.75 3.45
Longitudinal
modulus EL(GPa) 60 140 140 95 210 260 170 260
Transverse
modulus ET(GPa) 13 11 10 5.1 19 140 19 150
In-plane shear
modulus GLT (GPa) 3.4 5.5 5.1 1.8 4.8 60 9 60
Major Poisson’s
ratio
ν
LT 0.3 0.27 0.3 0.34 0.21 0.3 0.3 0.24
Longitudinal ultimate
tensile stress
σ
t
Lu (GPa) 1.8 2 2.1 2.5 1.3 1.3 0.34 0.7
Longitudinal ultimate
compressive stress
σ
c
Lu (GPa) 0.65 1.2 1.2 0.3 2 2 0.18 3.4
Transverse ultimate
tensile stress
σ
t
Tu (GPa) 0.04 0.05 0.075 0.03 0.07 0.14 0.007 0.19
Transverse ultimate
compressive stress
σ
c
Tu (GPa) 0.09 0.17 0.25 0.13 0.3 0.3 0.05 0.4
In-plane ultimate
shear stress
σ
LTu (GPa) 0.05 0.07 0.16 0.03 0.08 0.09 0.03 0.12
Appendix G
References
G.1 Comprehensive Composite Materiala
1. Editors-in-chief Kelly, A. and Zweben, C.: Comprehensive Composite Materials.
Pergamon, Oxford, 2000.
Vol. 1: Fiber Reinforcements and General Theory of Composites (ed. by T.-W.
Chou)
Vol. 2: Polymer Matrix Composites (ed. by R. Talreja & J.-A. E. M˚anson)
Vol. 3: Metal Matrix Composites (ed. by T. W. Clyne)
Vol. 4: Carbon/Carbon, Cement, and Ceramic Matrix Composites (ed. R. War-
ren)
Vol. 5: Test Methods, Nondestructive Evaluation, and Smart Materials (ed. by
L. Carlsson, R.L. Crane & K. Uchino)
Vol. 6: Design and Applications (ed. by M.G. Bader, K.T. Kedward & Y.
Sawada)
2. Editors-in-chief Zweben, C. and Beaumont, P.: Comprehensive Composite Ma-
terials II. Elsevier, 2017.
Vol. 1: Reinforcements and General Theories of Composites (ed. by E.E. Gd-
outos)
Vol. 2: Polymer Matrix Composites: Fundamentals (ed. by R. Talreja)
Vol. 3: Polymer Matrix Composites: Manufacture and Applications (ed. by A.
Poursartip)
Vol. 4: Metal Matrix Composites (ed. by T.W. Clyne)
Vol. 5: Ceramic and Carbon Matrix Composites (ed. by M.B. Ruggles-Wrenn)
Vol. 6: Nanocomposites and Multifunctional Materials (ed. by T. Peijs and
E.T. Thostenson)
Vol. 7: Testing, Nondestructive Evaluation and Structural Health Monitoring
(ed by R. Crane)
Vol. 8: Design and Analysis of Composite Structures (ed. by A. Johnson and
C. Soutis)
489
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
490 G References
G.2 Selected Textbooks and Monographs on Composite
Mechanics
1. Agarwal, B.D., L.J. Broutman, K. Chandrashekhara: Analysis and Performance
of Fiber Composites. John Wiley & Sons, Hoboken (NJ), 3. ed., 2006.
2. Altenbach, H., J. Altenbach, R. Rikards: Einf ¨
uhrung in die Mechanik der
Laminat- und Sandwichtragwerke - Modellierung und Berechnung von Balken
und Platten aus Verbundwerkstoffen. Dt. Verl. f¨ur Grundstoffindustrie, Stuttgart,
1996.
3. Altenbach, H., J. Altenbach, W. Kissing: Structural Analysis of Laminate and
Sandwich Beams and Plates. Lubelskie Towarzystwo Naukove, Lublin, 2001.
4. Altenbach, H., W. Becker (Eds): Modern Trends in Composite Laminates Me-
chanics. CISM Courses and Lectures. Springer, Wien, New York, 2003.
5. Ashbee K.H.G.: Fundamental Principles of Reinforced Composites. Technomic,
Lancaster et al., 2. ed., 1993.
6. Becker, W.: Beitr¨
age zur analytischen Behandlung ebener Laminate. Habil.-
Schrift, TH Darmstadt 1993.
7. Bergmann, H.W.: Konstruktionsgrundlagen f¨
ur Faserverbundbauteile. Springer,
Berlin u.a., 1992.
8. Berthelot, J.-M.: Composite Materials. Mechanical Behaviour and Structure
Analysis. Springer, New York et al., 1999.
9. Bogdanovich, A.E., C.M. Pastore: Mechanics of Textile and Laminated Compos-
ite. With Applications to Structural Analysis. Chapman & Hall, London, 1996.
10. Buhl, H. (Ed.): Advanced Aerospace Materials. Materials Research and Engi-
neering. Springer, Berlin, Heidelberg, 1992.
11. Carlsson, L.A., D.F. Adams, D.F., R.B. Pipes: Experimental Characterization of
Advanced Composite Materials. CRC Press, Boca Raton, 4rd edition, 2014.
12. Chawla, K.K.: Composite Materials. Science and Engineering. Springer, New
York, 2012.
13. Chung, D.D.L.: Composite Materials: Functional Materials for Modern Tech-
nologies. Springer, London, 2003.
14. Chung, D.D.L.: Composite Materials: Science and Applications. Springer, Lon-
don, 2010.
15. Daniel, I.M., O. Ishai: Engineering Mechanics of Composite Materials. Oxford
University Press, New York, Oxford, 2nd ed., 2006.
16. Davies, J.M. (Ed.): Lightweight Sandwich Construction. Blackwell Science,
Oxford et al., 2001.
17. Decolon, C.: Analysis of Composite Structures. HPS, London, 2002.
18. Delhaes, P. (Ed.): Fibres and Composites. Taylor & Francis, London, 2003.
19. Dimitrienko, Yu.I.: Thermomechanics of Composites under High Temperature.
Springer Netherlands, Dordrecht, 2016.
20. Ehrenstein, G.W.: Faserverbund-Kunststoffe. Hanser, M¨unchen, Wien, 2nd ed.,
2006.
G.2 Selected Textbooks and Monographs on Composite Mechanics 491
21. Friedrich, K. (Ed.): Application of Fracture Mechanics to Composite Materials.
Bd. 6 Composite Material Series Elsevier, Amsterdam, 1989.
22. Gay, D.: Composite Materials: Design and Applications. CRC Press, Boca
Raton, 3rd. ed., 2014.
23. Geier, M.H.: Quality Handbook for Composite Materials. Chapman & Hall,
London et al., 1994.
24. Gibson, R.F.: Principles of Composite Material Mechanics. CRC Press, Boca
Raton, 4th ed., 2016.
25. Gibson, R.F.: Dynamic Mechanical Behavior of Composite Materials and Struc-
tures. CRC Press, Boca Raton, 2002.
26. G ¨urtal, Z., Haftka, R.T., Hajela, P.: Design and Optimization of Laminated Com-
posite Materials. John Willy & Sons Inc. New-York, 1999.
27. Harper, C.A. (Ed.): Handbook of Plastics, Elastomers, and Composites.
McGraw-Hill, New York et al., 4th edition, 2002.
28. Harris, B.: Engineering Composite Materials. IOM Communications Ltd., Lon-
don, 2nd edition, 1999.
29. Hoa, S.V.: Analysis for Design of Fiber Reinforced Plastic Vessels and Pipes.
Technomic, Lancaster, Basel, 1991.
30. Hoa, S.V., Wei Fang: Hybrid Finite Element Method for Stress Analysis of Lam-
inated Composites. Kluwer Academic Publishers. Dordrecht, 1998.
31. Hollaway, L.: Polymer Composites for Civil and Structural Engineering. Blackie
Academic & Professional, London et al., 1993.
32. Hult, J., F.G. Rammersdorfer (Eds): Engineering Mechanics of Fibre Rein-
forced Polymers and Composite Structures. CISM Courses and Lectures No.
348. Springer, Wien, New York, 1994.
33. Hyer, M.W.: Stress Analysis of Fibre-Reinforced Composite Materials. DEStech
Publ., Lancaster (PA), Uptadet ed., 2009.
34. Hull, D., T.W. Clyne: An Introduction to Composite Materials. Cambridge Uni-
versity Press. 2nd ed., 2003.
35. Jones, R.M.: Mechanics of Composite Materials. Taylor & Francis, London,
1999.
36. Kachanov, L.M.: Delamination Buckling of Composite Materials. Mechanics of
Elastic Stability, Vol. 14. Kluwer, Dordrecht, Boston, London, 1988.
37. Kalamkarov, A.L.: Composite and Reinforced Elements of Construction. Wiley
& Sons, Chichester et al., 1992.
38. Kalamkarov, A.L., A.G. Kolpakov: Analysis, Design and Optimization of Com-
posite Structures. Wiley & Sons, Chichester et al., 1997.
39. Kaw, A.K.: Mechanics of Composite Materials. CRC Press, Boca Rotan, New
York, 2nd ed., 2006.
40. Kim, D.-H.: Composite Structures for Civil and Architectural Engineering.
E&FN SPON, London et al., 1995.
41. Kollar, L.P., G.S. Springer: Mechanics of Composite Structures. Cambridge
University Press, Cambridge, 2003.
42. Matthews, F.L., R.D. Rawlings: Composite Materials: Engineering and Science.
Woodhead Publishing, Cambridge, 1999.
492 G References
43. Matthews, F.L., Davies, G.A.O., Hitching, D., Soutis, C.: Finite Element Mod-
elling of Composite Materials and Structures. CRC Press. Woodhead Publishing
Limited, Cambridge, 2000.
44. McCullough, R.L.: Micromechanical Materials Modelling. Delaware Compos-
ites Design Encyclopedia, Vol. 2. Technomic, Lancaster, Basel, 1991.
45. Michaeli, W., D. Huybrechts, M. Wegener: Dimensionieren mit Faserverbund-
kunststoffen: Einf¨
urung und praktische Hilfen. Hanser, M ¨unchen, Wien, 1995.
46. Milton, G. W.: The Theory of Composites. Cambridge University Press, Cam-
bridge, 2002.
47. Mittelstedt, C., W. Becker: Strukturmechanik ebener Laminate. TU Darmstadt,
Darmstadt, 2016.
48. Moser, K.: Faser-Kunststoff-Verbund: Entwurfs- und Berechnungsgrundlagen.
VDI-Verlag, D ¨usseldorf, 1992.
49. Newaz, G.M. (Ed.): Delamination in Advanced Composites. Technomic, Lan-
caster, 1991.
50. Nethercot, D.A.: Composite Construction. Spon Press, New York, 2003.
51. Ochoa, O.O., J.N. Reddy: Finite Element Analysis of Composite Laminates.
Solid Mechanics and its Applications, Vol. 7. Kluwer, Dordrecht, Boston, Lon-
don, 1992.
52. Plantema, F.J.: Sandwich Constructions. John Wiley & Sons, New York, 1966.
53. Powell, P.C.: Engineering with Fibre-polymer Laminates. Chapman & Hall,
London et al., 1994.
54. Puck, A.: Festigkeitsanalyse an Faser-Matrix-Laminaten: Realistische Bruchkri-
terien und Degradationsmodelle. Hanser, M”unchen, 1996.
55. Reddy, J.N.: Mechanics of Laminated Composite Plates - Theory and Analysis.
CRC Press, Boca Rotan et al., 1997.
56. Reddy, J.N.: Mechanics of Laminated Composite Plates and Shells: Theory and
Analysis. CRC Press, Boca Rotan et al., 2004.
57. Reddy, J.N., A. Miravete: Practical Analysis of Composite Laminates. CRC
Press, Boca Rotan et al., 1995.
58. Rohwer, K.: Modelle und Methoden zur Berechnung von Laminaten aus uni-
direktionalen Faserverbunden. Fortschritt-Berichte VDI: Reihe 1 Konstruktion-
stechnik, Maschinenelemente Nr. 264. VDI-Verlag, D¨usseldorf, 1996.
59. Schulte, K., B. Fiedler:: Structure and Properties of Composite Materials.
TUHH-Technologie GmbH, Hamburg, 2. Aufl., 2005.
60. Sih, G.C., A. Carpinteri, G. Surace (Eds): Advanced Technology for Design
and Fabrication of Composite Materials and Structures. Engng. Appl. of Fract.
Mech., Vol.14. Kluwer Academic Publ., Dordrecht, Boston, London, 1995.
61. Sih, G.C., A.M. Skudra (Eds): Failure Mechanics of Composites. Handbook of
Composites, Bd. 3. North-Holland, Amsterdam, New York, Oxford, 1985.
62. Talreja, R. (Ed.): Damage Mechanics of Composite Materials. Composite Ma-
terials Series, Vol. 9. Elsevier, Amsterdam et al., 1994.
63. Tarnopol’ski, J.M., T. Kincis: Test Methods for Composites. Van Nostrand Rein-
hold, New York, 1985.
64. Tsai, S.W.: Composites Design. Think Composites, Dayton, Paris, Tokyo, 1988.
G.3 Supplementary Literature for Further Reading 493
65. Turvey, G.J., I.H. Marshall (Eds.): Buckling and Postbuckling of Composite
Plates. Chapman & Hall, London, 1995.
66. Vasiliev, V.V., Jones, R.M. (Engl. Ed. Editor): Mechanics of Composite Struc-
tures. Taylor & Francis, Washington, 1993.
67. Vasiliev, V.V., Morozov, E.V.: Mechanics and Analysis of Composite Materials.
Elsevier, London, 2001.
68. Vasiliev, V.V., Morozov, E.V.: Advanced Mechanics of Composite Materials and
Structural Elements. Elsevier, London, 3rd ed., 2013.
69. Vinson, J.R.: The Behavior of Shells Composed of Isotropic and Composite
Materials. Solid Mechanics and its Applications, Vol. 18. Kluwer, Dordrecht,
Boston, London, 1993.
70. Vinson, J.R., R.L. Sierakowski: The Behavior of Structures Composed of Com-
posite Materials. Springer, Dordrecht et al., 3rd ed., 2008.
71. Whitney, J.M.: Structural Analysis of Laminated Anisotropic Plates. Technomic
Publishing Co. Inc., Lancaster, 1987.
72. Ye, J.: Laminated Composite Plates and Shells: 3D Modelling. Springer, London
et al., 2003.
73. Zweben, C., H.T. Hahn, T.-W. Chou: Mechanical Behavior and Properties of
Composite Materials. Delaware Composites Design Encyclopedia, Bd. 1. Tech-
nomic, Lancaster, Basel, 1989.
G.3 Supplementary Literature for Further Reading
1. Altenbach, H., J. Altenbach, K. Naumenko: Ebene Fl¨
achentragwerke - Grund-
lagen der Modellierung und Berechnung von Scheiben und Platten. Springer
Vieweg, Berlin, Heidelberg, New York, 2. Aufl., 2016.
2. Altenbach, H.: Kontinuumsmechanik - Einf¨
uhrung in die materialunabh¨
angigen
und materialabh ¨
angigen Gleichungen. Springer Vieweg, Berlin, Heidelberg, 3.
Aufl., 2015.
3. Altenbach, J., W. Kissing, H. Altenbach: D¨
unnwandige Stab- und Stabschalen-
tragwerke. Vieweg-Verlag, Braunschweig/Wiesbaden, 1994.
4. Altenbach, H., J. Altenbach, A. Zolochevsky: Erweiterte Deformationsmodelle
und Versagenskriterien der Werkstoffmechanik. Deutscher Verlag f ¨ur Grund-
stoffindustrie, Stuttgart, 1995.
5. Ambarcumyan, S.A.: Theory of Anisotropic Plates: Strength, Stability, and Vi-
brations. Hemisphere Publishing, Washington, 1991.
6. Betten, J.: Kontinuumsmechanik. Springer-Verlag. Berlin, Heidelberg, New York,
2nd edition, 2001.
7. Betten, J. Finite Elemente f¨
ur Ingenieure - Grundlagen, Matrixmethoden, Kon-
tinuum. Springer-Verlag, Berlin, Heidelberg, New York, 2nd edition, 2003.
8. Haupt, P.: Continuum Mechanics and Theory of Materials. Springer-Verlag.
Berlin, Heidelberg, New York, 2nd edition, 2002
494 G References
9. Lekhnitskii, S.G.: Anisotropic Plates. Gordon and Breach Science Publishers,
London, 3rd print, 1987.
10. Lekhnitskij, S.G.: Theory of Elasticity of an Anisotropic Body. Mir Publishers,
Moscow, 1981.
11. Lewi ´nski, T., J.J. Telega: Plates, Laminates and Shells Asymptotic Analysis and
Homogenization. World Scientific, Singapore, 2000.
12. Wlassow, V.S.: Allgemeine Schalentheorie und ihre Anwendung in der Technik.
Akademie-Verlag, Berlin, 1958.
13. Zienkiewicz, O.C., R.L. Taylor: The Finite Element Method, Vol. 2: Solid Me-
chanics. McGraw Hill, Oxford, 5th ed., 2000.
G.4 Selected Review Articles
1. Altenbach, H.: Modellierung des Deformationsverhaltens mehrschichtiger Fl ¨
a-
chentragwerke - ein ¨
Uberblick zu Forschungsrichtungen und -tendenzen. Wiss.
Ztschr. TH Magdeburg 32(4): 86 94, 1988.
2. Bert, C.W.A. A critical evaluation of new plate theories applied to laminated
composites. Comp. Struc. 2, 329 347, 1984.
3. Carrera, E., L. Demasi, M. Manganello: Assessment of plate elements on bendi-
jng and vibration of composite structures. Mech. of Adv. Mat. and Struct. 9, 333
357, 2002.
4. Christenson, R.M.: A survey of and evaluation methodology for Fiber Composite
Material Failure Theories. In ”Mechanics for a New Millennium”, Eds H. Aref
and J.W. Philips, 25 40, 1998.
5. Ha, K.H.: Finite element analysis of sandwich construction: a critical review.
Sandwich Construction 1, 69 85, 1989.
6. Failure criteria in fibre-reinforced polymer composites. Special Issue of Com-
posites Science and Technology 58, 1998.
7. Hashin, Z.: Analysis of composite materials - A Survey. Trans. ASME. J. Appl.
Mech. 50: 481 505, 1983.
8. Hohe, J., W. Becker: Effective stress-strain relations for two-dimensional cel-
lular sandwich cores: Homogenization, material models, and properties. Appl.
Mech. Rev. 54: 61 87, 2001.
9. Irschik, H.: On vibration of layered beams and plates. ZAMM 73 (4-5), T34
T45, 1993.
10. Leissa, W.W.: A review of laminated composite plate buckling. Appl. Mech. Rev.
40 (5), 575 590, 1987.
11. Lui, M., P. Habip: A survey of modern developments in the analysis of sandwich
structures. Appl. Mech. Rev. 18(2), 93 98, 1965.
12. Mallikarguwa, T. Kant: A critical review and some results of recently developed
refined theories of fiber-reinforced laminated composites and sandwiches. Comp.
Structures 23, 293 312, 1993.
G.4 Selected Review Articles 495
13. Nahas, M.N.: Survey of failure and post-failure theories of laminated fibre-
reinforced composites. J. Composites Technology & Research 8: 138 153,
1986.
14. Naumenko, K., J. Altenbach, H. Altenbach, V.K. Naumenko: Closed and ap-
proximate analytical solutions for rectangular Mindlin plates. Acta Mechanica
147: 153 172, 2001.
15. Noor, A.K., W.S. Burton: Assessment of shear deformation theories for multilay-
ered composite plates. Appl. Mech. Rev. 41(1): 1 13, 1989.
16. Noor, A.K., W.S. Burton, C.W. Bert: Computational models for sandwich panels
and shells. Appl. Mech. Rev. 49(3), 155 199, 1996.
17. Noor, A.K., W.S. Burton, J.U. Peters: Assessment of computational models for
multilayered composite cylinders. Int. J. Solids Structures, 27 (10), 1269 1286,
1991.
18. Vinson, J.R.: Sandwich structures. Appl. Mech. Rev. 54(3), 201 214, 2001.
19. Reddy, J.N.: A review of refined theories of laminated composite plates. Shock
Vibr. Dig. 22: 3 17, 1990.
20. Reddy, J.N.: An evaluation of equivalent-single-layer and layer theories of com-
posite laminates. Composite Structures 25, 21 35, 1993.
21. Reddy, J.N., Robbins Jr., D.H. Theories and computational models for composite
laminates. Appl. Mech. Rev. 47 (6) 147 169, 1994.
22. Rohwer, K. Computational models for laminated composites. Z. Flugwiss. Wel-
traumforsch. 17, 323 330, 1993.
Index
Airy stress function, 290
analytical solution, 306
anisotropy
curvilinear, 51
cylindrical, 51
rectilinear, 51
approximation function, 414
auxetic material, 22
basic modes of failure of a lamina, 210
beam, 419
beam element, 413
laminate, 450
beam equations, 251
elementary, 233
beam resultants, 236
beam shaped shell, 371
beam shaped thin-walled folded structure, 414
beam theory
Bernoulli-Euler, 232
classical, 232
elementary, 232
generalized
Vlasov , 232
Timoshenko, 232, 246
bending
cylindrical, 294
bending stress, 126
bending-layer solution, 352
Bernoulli
Jakob I., 77
Bernoulli beam, 416
Bernoulli beam model, 422
Saint-Venant supplement, 373
specialized, 391
Boole
George, 190
boundary condition, 296
boundary conditions, 283
Bredt
Rudolf, 373
buckling, 286, 288, 297, 304, 305, 310, 318,
321
buckling equation, 242, 251
bulk modulus, 43
Christensen
Richard M., 223
classical beam model, 372
classical laminate theory, 181, 281, 286, 293,
303, 306, 376, 422, 432
closed thin-walled cross-section, 380
CLT, 181
coincidence matrix, 417
collocation method, 73
column, 232
compatibility conditions, 60
compliance hypermatrix, 143
compliance matrix, 27, 30, 32, 36
bending, 157
coupling, 157
extensional, 157
flexural, 125
monoclinic material, 33
off-axis extensional, 124
off-axis flexural, 125
off-axis in-plane, 123
on-axis extensional, 124
on-axis flexural, 125
orthotropic material, 34
transversely isotropic material, 35
compliance modulus, 22
compliance submatrix, 143
concentrated transversal stiffener, 389
497
© Springer Nature Singapore Pte Ltd. 2018
H. Altenbach et al., Mechanics of Composite Structural
Elements, https://doi.org/10.1007/978-981-10-8935-0
498 Index
constitutive equations, 62
constitutive equations of the lamina resultants,
126
continuity
element boundaries, 415
contraction, 22
coordinate functions, 70
coupling inertia term, 298
curvature, 281, 282
curvatures
longitudinal, 385
torsional, 385
cylindrical bending, 306, 320
d’Alambert principle, 67
damping matrix, 416
delamination, 209
density, 392
differential equation of flexure, 241
differential matrix, 62
discretization of the structure, 414
displacement
virtual, 64
displacement method, 66, 413
distorsion strain energy criteria of Tsai-Hill,
211
distortional energy criterion, 214
effective compliance, 24
effective hygrothermal coefficients, 112
effective moduli, 21
effective stiffness, 24
effective thermal expansion coefficients, 112
eigen-vibration, 403
eigenvalue problem, 318, 322
elastic parameters, 43
elasticity modulus, 22
elasticity tensor, 26
end matrix, 400
energy principles, 63
engineering parameters, 36
equilibrium equations
dynamic, 59, 62
plate, 281
static, 59, 62
surface, 62
equivalent single layer theory
higher order, 182
equivalent single-layer theory, 181, 183
ESLT, 181
Euler
Leonhard, 65
extended Vlasov-Kantorowich method, 75
extensional strains, 21, 24
external virtual work, 64
extremal principles, 63
oppl
August Otto, 330
face wrinkling, 263
finite element, 413
beam, 413
generalized, 413
one-dimensional, 442
plate, 413
laminate, 439
triangular, 429, 434
finite element procedure, 414
first order shear deformation theory, 246, 251,
280, 295, 301, 436
first-order shear deformation theory, 181, 183
first-ply failure, 211
fixed boundary, 296
flexibility matrix, 27
flexibility modulus, 22
flexure equations, 251
folded plate structure, 371, 375
force method, 66, 413
force resultant, 348, 377
forced vibration, 242, 251
forces
in-plane, 122
transverse, 122
four-node element, 423
fracture
interlaminar, 206
intralaminar, 206
translaminar, 206
fracture modes, 207
free boundary, 296
free edge, 297
free vibration, 242, 251
FSDT, 181
Galerkin
Boris Grigorjewitsch, 73
Galerkin’s method, 73
general three-dimensional problem, 416
generalized beam, 441
generalized co-ordinate function, 380, 392
generalized coordinates, 70
generalized densities
matrix, 439, 440
generalized density, 426
generalized displacement function, 392
generalized mass density, 298
Hamilton
Index 499
William Rowan, 68
Hamilton principle, 68, 299, 393
Hamilton’s principle, 449
Hamilton‘s principle, 414, 416
harmonic oscillation, 318
higher order theory, 295
Hill
Rodney, 211
Hoffman criterion, 217
Hooke
Robert, 20
Hooke’s law, 21
generalized, 26
inverse form, 22
hybrid criteria, 210
hybrid method, 413
hygrothermal loading, 263
hypotheses
Bernoulli, 186
Kirchhoff, 186
Love, 186
in-plane strain, 282
in-plane stress resultant force vector, 122
inner energy
element, 417
interactive criteria, 210, 214
interactive tensor polynomial criterion of
Tsai-Wu, 211
internal virtual work, 64
iso-strain condition, 90
iso-stress condition, 91
Jacobi matrix, 431
Kantorovich
Leonid Vitaliyevich, 74
Kantorovich separation relationships, 380
kinematic equations, 60, 62
kinetic energy, 392
rotatory, 438
Kirchhoff
Gustav Robert, 181
Kirchhoff hypotheses, 281, 376
Kirchhoff plate, 416, 429
Kirchhoff shear force resultant, 283
Krylov method, 395
Krylow
Aleksei Nikolajewitsch, 395
Krylow functions, 485
Kutta
Martin Wilhelm, 394
evy
Maurice, 306
Lagrange
Joseph-Louis, 65
Lagrange function, 393
Lagrangian function, 68
Lam´e
Gabriel eon Jean Baptiste, 43
Lam´e coefficients, 43
lamina, 11
laminae
UD-laminae, 109
unidirectional, 109
laminate, 11
angle-ply, 139, 293, 294
antisymmetric, 138, 156, 285
balanced, 153
cross-ply, 152
asymmetric, 138
balanced, 139, 156, 285
beam element, 419
cross-ply, 139, 156, 285, 293, 294
general, 139, 156
isotropic layers, 293
nonsymmetric, 440
quasi-isotropic, 157
special orthotropic, 436
specially orthotropic, 280
symmetric, 138, 156, 285, 436
angle-ply, 138
balanced, 150
cross-ply, 149
general case, 289
regular, 138
special case with isotropic layers, 290
with isotropic layers, 148
truss element, 419
unsymmetric, 138
with isotropic layers, 157
laminate beam
symmetric, 252
laminate code, 138
laminate plate
arbitrary stacking, 280
symmetric, 288
laminate theory
classical, 183
second order, 183
third order, 183
laminates
specially orthotropic, 290
last-ply failure, 211
layer, 11
Layerwise theory, 197
layerwise theory, 182
500 Index
least-squares method, 73
limit criteria, 210
Love
Augustus Edward Hough, 186
Love’s first-approximation shell theory, 347
macro-mechanical level, 11
macro-mechanical modelling, 108
macroscopic approach, 87
macroscopic level, 13
mass matrix
condensed, 416
consistent, 416
element, 439, 450
symmetric, 416
material
anisotropic, 4
arranged in parallel, 23
arranged in series, 23
ceramics, 4
composite
advantages, 15
classification, 5
examples, 5
limitations, 16
matrix, 7
reinforcement, 5
heterogeneous, 4
homogeneous, 4
inhomogeneous, 4
isotropic, 4, 36, 42
metals, 4
monoclinic, 32, 43
monolithic, 5
orthotropic, 34
polymers, 4
transversely isotropic, 35, 40
triclinic, 32
maximum strain theory, 213
maximum stress theory, 211
membrane stress, 126
membrane theory, 354
mesh, 413
method of initial parameters, 395
micro-mechanical level, 11
microscopic approach, 87
microscopic level, 13
mid-plane strain, 123
Mindlin
Raymond David, 295
Mindlin plate, 300, 325, 429, 436
Mindlin plate theory, 295
Mises
Richard Edler von, 214
modelling
one-dimensional, 371
three-dimensional, 371
two-dimensional, 371
moment
resultant, 122, 124
moment resultant, 348, 377
multidirectional laminate, 136
adai
´
Arp´ad, 306
adai-L´evy solution, 306, 316, 326
nanoscale level, 13
Navier
Claude Louis Marie Henri, 306
Navier solution, 306, 312, 324
nodal force
vector
time dependent, 416
node, 413
normal stress, 21
normal stresses, 24
number of degrees of freedom, 413
off-axis case, 110
off-axis loaded UD-lamina, 113
off-axis stretching, 123
on-axis case, 109
one-dimensional element, 413
one-dimensional structure, 419
optimal global laminate behavior, 137
plane beam problem, 418
plane strain problem, 416
plane strain state, 48
plane stress problem, 416
plane stress state, 45
plate, 279, 429
bending, 293
buckling, 293
moderately thick, 295
specially orthotropic, 288, 298
symmetric
general case, 287
vibration, 293
plate element, 413
laminate, 450
plate strip, 232, 307, 320
symmetrical laminated, 308
unsymmetric laminated, 308
plate theory
classical
Kirchhoff, 181
ply, 11
Index 501
Poisson
Sim´eon Denis, 15
Poisson effect, 308
Poisson’s ratio, 22, 43
major, 92
minor, 93
potential energy, 383, 392, 420, 436, 440
folded structure, 379
single element, 417
whole structure, 417
prebuckling, 304
principle of complementary virtual work, 63,
65
principle of minimum potential energy, 299
principle of minimum total complementary
energy, 66
principle of minimum total potential energy, 65
principle of the total minimum potential
energy, 414
principle of virtual displacements, 66
principle of virtual forces, 66
principle of virtual work, 63, 64
prismatic structure, 376
profile node
main, 442
profile node concept, 442
profile nodes
secondary, 443
Puck
Alfred, 223
Rayleigh, 69
Rayleigh-Ritz method, 69
reduced compliances, 48
reduced stiffness, 47
Reissner
Eric (Max Erich), 63
Reissner plate theory, 295
Reissner theory, 195
Reissner’s functional, 67
Reissner’s variational theorem, 63, 66
resultant
in-plane force, 282, 288
moment, 282
transverse shear force, 282
resultant moment vector, 124
Reuss
Andr´as (Endre), 24, 88
Reuss estimate, 92
Reuss model, 24, 91, 93, 94
Ritz
Walter, 69
Ritz approximation, 70, 72
Ritz method, 69, 414
rod, 232
rotation matrix, 28
rotational term, 392
rotatory inertia, 251, 286, 323
rule of mixture, 90
rule of mixtures, 88, 94, 96
inverse, 88, 92
Runge
Carl David Tolm ´e, 394
Sachs
Oscar, 88
Saint-Venant
Adh´emar Jean Claude Barr´e de, 373
sandwich, 11
symmetric
thick cover sheets, 302
thin cover sheets, 301
sandwich beam
dissimilar faces, 258
symmetric, 252
sandwich composites, 172
sandwich plate, 301
Schapery
Richard Allan, 112
selective integration, 438
semi-empirical solution of Halpin and Tsai, 96
semi-membrane theory, 355
shape function, 414, 415, 423, 424, 431, 434,
442
linear, 420
shear correction coefficient, 300
shear correction factor, 142, 195, 246, 251
shear deformation theory, 306
shear deformations of the mid-planes, 385
shear lag effect, 374, 396
shear modulus, 22, 43
shear rigid theory, 187
shear strains, 21
engineering, 25
tensorial, 25
shear stress, 21
shell, 345
circular cylindrical, 345
circumferential cross-ply, 345
special orthotropic, 345
classical theory, 345, 346
hypotheses, 347
first order shear deformation theory, 345
moderately thick, 356
thin-walled, 345
transverse shear deformations, 346
shell element, 413
laminate, 450
502 Index
simply supported boundary, 297
six-node element, 437
stacking
symmetric, 390
stacking codes of laminates, 137
stiffness
plate, 288
reduced, 111
shear, 194
stiffness matrix, 27, 30, 32, 36
bending, 141
coupling, 141
element, 417, 450
extensional, 123, 141
flexural, 125
isotropic material, 36
monoclinic material, 33
off-axis extensional, 124
off-axis flexural, 125
on-axis extensional, 124
on-axis flexural, 125
orthotropic material, 34
reduced, 377
symmetric, 415
transverse shear, 142
transversely isotropic material, 35
stiffness submatrix, 143
strain tensor, 24, 25
strain vector, 25
strain-displacement relations, 60
stress resultant, 283, 299, 449
plate, 281
stress resultants, 233
stress tensor, 24, 25
stress vector, 25
structural behavior
global, 372
local, 372
structural level, 13
Strutt
John William, 69
submatrix
bending, 148
coupling, 147
extensional, 148
surface force
vector, 415
Taylor
Geoffrey Ingram, 88
tensile compliance, 22
tensile flexibility, 22
tensile stiffness, 22
theorem of Castigliano, 66
thin-walled beam, 371
three-node element, 422, 437, 442
Timoshenko
Stepan Prokopovich, 80
Timoshenko beam model, 373, 422
specialized, 391
total virtual work, 64
transfer matrix, 398
transfer matrix method, 396
transformation matrix, 28, 29, 62
translatory inertia, 251, 323
transversal strains, 385
transverse force resultant, 378
transverse shear, 376
transverse shear deformation, 246
transverse shear resultant, 125
transverse shear stress, 295
trial function, 380
triangle co-ordinates
natural, 438
triangle coordinates
natural, 430
truss element, 413
Tsai
Stephen Wei-Lun, 96, 211
Tsai-Hill criterion, 215
Tsai-Wu criterion, 216
two-dimensional element, 413
two-dimensional structural element, 345
two-dimensional structure, 429
two-node element, 442
beam, 422
truss, 420
unknown coefficient function, 380
variational formulation
axial symmetrically circular cylindrical
shell, 358
variational iteration method, 75
variational operations, 64
vector of curvature, 125
vibration, 286, 310, 318, 421
forced, 297
forced transversal, 286
free, 323, 392
Vlasov
Vasily Zakharovich, 74
Vlasov beam model, 373, 391
Vlasov hypotheses, 380
Voigt
Woldemar, 23, 88
Voigt estimate, 90
Voigt model, 23, 90, 92, 94
Index 503
volume element, 413
volume force
vector, 415
warping, 373
weak form of the model equations, 70
weighted residual methods, 73
weighted-residual methods, 69
Wu
Edward Ming-Chi, 211
Young
Thomas, 14
Young’s modulus, 22, 43
... The stress field contains squared root singularity at the interface crack tip. The principle mechanism of fracture failures in composite materials is explained in detail in many monographs and research articles [1,2]. The presence of composite materials in engineering structures and their applications has encouraged many researchers to study the interfacial cracks in dissimilar isotropic materials [3][4][5][6]. ...
... Owning to the geometry of the problem and the loading condition, the boundary conditions along the interface for the considered problem are given by (1) ...
Article
The present article considers an anti‐plane stress problem of three cracks at different orthotropic materials' interfaces. According to the geometry of the problem, the governing equations and mixed boundary conditions have been formulated. Fourier integral transformation is used to convert the mixed boundary value problem into dual integral equations, which gives two equations containing infinite series. The investigation of the problem concerning anti‐plane cracks subjected to static loadings is done with the help of the Schmidt method to satisfy the given boundary conditions. The difference in displacements is expanded to proceed further in the problem, which becomes zero outside the cracks. Numerical computations are carried out for the graphical representation of stress intensity factors (SIFs) at all cracks' tips. The Interaction among the cracks as those are in close proximity to each other or move away are represented pictorially. Detailed numerical results and discussion are done for the considered materials, which include aluminium, epoxy and graphite epoxy. The novelty of the present article is the numerical analysis and pictorial presentation of SIFs at the tips of interfacial offset parallel cracks for various crack lengths and normalised heights for different combinations of materials. The authors have obtained variations in SIFs for the cracks at the interfaces of dissimilar composite materials.
... [9][10][11] The fibers can be natural or synthetic, short or long, and have varied orientations. 12,13 In the preparation of the composites, several manufacturing techniques are used with specific contour conditions, in addition to that superficial treatments can be performed, seeking to improve the properties of the resulting material. 14,15 The attempt to use natural fibers to replace synthetic fibers can be explained by the growing worldwide concern for the preservation of the environment. ...
... Arpitha et al. 60 realized that the flexural strength and modulus increase with the increase of the glass portion of the composite. 57 Sisal/Glass with red mud 2 wt.% and particle size 4 mm 142.69 -Sisal/Glass with red mud 4 wt.% and particle size 4 mm 140.78 -Sisal/Glass with red mud 6 wt.% and particle size 4 mm 153.39 -Sisal/Glass with red mud 8 wt.% and particle size 4 mm 162.51 -Sisal/Glass with red mud 2 wt.% and particle size 13 mm 126.43 -Sisal/Glass with red mud 4 wt.% and particle size 13 mm 131.66 -Sisal/Glass with red mud 6 wt.% and particle size 13 mm 138.42 -Sisal/Glass with red mud 8 wt.% and particle size 13 57 Sisal/Glass with red mud 2 wt.% and particle size 4 mm 334.53 -Sisal/Glass with red mud 4 wt.% and particle size 4 mm 358.50 -Sisal/Glass with red mud 6 wt.% and particle size 4 mm 360.67 -Sisal/Glass with red mud 8 wt.% and particle size 4 mm 417.17 -Sisal/Glass with red mud 2 wt.% and particle size 13 mm 360.67 -Sisal/Glass with red mud 4 wt.% and particle size 13 mm 468.59 -Sisal/Glass with red mud 6 wt.% and particle size 13 mm 487.04 -Sisal/Glass with red mud 8 wt.% and particle size 13 mm 501.77 -Sisal/Glass (S: 2 wt.% and G: 2 fix layer) 87.34 -Rana et al. 59 In addition, the sisal/glass hybrid composites have values in this property greater than those of pure sisal and smaller than that of pure glass. Aslan et al. 61 noted that the flexural strength and modulus perform best in the sisal/glass configuration with 10 vol.% sisal and 17 vol.% ...
Preprint
Full-text available
Currently, there is a tendency to develop materials that have an increasingly specific purpose for a particular application and the composites can be part of these types of materials. In addition, there is a search for replacement of synthetic fibers with natural ones, because of concerns about the environment. Hybridization of glass fiber composites with sisal fibers can result in improved performance by combining the benefits of each fiber with various applications. Many studies are developed on these types of composites, sisal/glass hybrids, varying certain parameters, such as matrix, orientation of fibers, treatment of fibers, stacking of layers, composition of elements, method of manufacture, among others, characterizing the various types of properties. The present work organizes the researches on the hybrid composites of sisal/glass, in chronological order, highlighting the specifications and parameters used and the main results obtained.
... The overall mean for the S/N ratio of the erosion rate is found to be - 16 ...
... Forms of Reinforcement[16] ...
Research
Erosive wear of mechanical parts caused by abrasive particles is a remarkable issue. Polymer composites are regularly utilized as designing/auxiliary parts where erosive wear happens. Because of the operational necessities in dusty situations, the investigation of solid particle erosion characteristics of the polymeric composites turns out to be very significant. Yet, even today, the impact on erosion characteristics of the fiber reinforced and particulate filled polyester composites has remained a significantly less contemplated zone. Research in this regard is required especially with the incorporation of fillers both in perspective of the logical furthermore, business significance. Also, a full comprehension of the impacts of all framework factors on the wear rate is important keeping in mind the end goal to attempt suitable ventures in the outline of machine or auxiliary part and in the selection of materials to decrease/control wear. The air jet erosion wear behavior of fabricated composites has been conducted under various operating conditions such as filler loading, impact velocity, impingement angle, erodent discharge rate. Air jet erosion wear characteristics of these fabricated composites are successfully analyzed using Taguchi’s experimental design method. The analytical and experimental investigation into the erosion behavior of glass fiber reinforced epoxy composites shows the positive response. Experiments can be successfully reduce using Taguchi method rather than using traditional method. Taguchi method, in spite of being a simple, efficient and systematic approach to optimize designs for performance, quality and cost, is used only in a limited number of applications worldwide. Its implementation in parametric appraisal of wear processes has hardly been reported. To this purpose, a program of experiments is constructed according to Taguchi L9 experimental design orthogonal matrices, to reduce the number of experiments without reducing the information is produced. Analysis of Variance (ANOVA) is also done to find out the most important factors in the experiment. The tensile strength and flexural strength of marble dust filled glass fiber reinforced epoxy composite by Universal Testing Machine (UTM). The analytical and experimental investigation indicate that impact velocity, Filler loading and impingement angle are the significant factors which affects the abrasion wear rate whereas Erodent discharge rate is least significant parameter. For 0%, 5%, 10% Wt. loading has peak load of 40N, 28N, 49N respectively of flexural strength. However, the tensile strength decreased with increasing percentage of marble dust. The presence of particulate fillers in these composites improves their erosion wear resistance and this improvement depends on the type and content of the fillers. Identified essential control factors affect the rate of erosion by the successful implementation of Analysis of Variance (ANOVA). It has been shown that these models reflect the effects of various factors on erosion and loss of predictive results are consistent with the theoretical observations.
Conference Paper
div class="section abstract"> Composites made of continuous fibers generally have higher strength-to-weight ratios in fiber directions as compared to those made of discontinuous fibers. However, the latter tend to display quasi-isotropic properties which can be of advantage when directions of mechanical loading can vary. For many real-world applications such as robust design of vehicle body components for crashworthiness, impact loads are stochastic in nature both in terms of magnitude and direction. Hence, in order to realize the true potential of laminated composites with continuous fibers, instead of orthotropic laminates which are most common due to the ease of design and manufacturing, angle-ply laminates are necessary. The latter category of laminates introduce a high degree of flexibility in design options but are also simultaneously challenging in terms of mechanical characterization due to the presence of a larger number of material parameters, as compared to orthotropic laminates, with coupled normal and shear behaviors. A computationally attractive approach in large-scale numerical analysis of laminated composite structures lies in the modeling of laminates with equivalent homogeneous anisotropic properties. A practical strategy in published literature appears to be lacking on the determination of various elastic stiffness or compliance coupling coefficients for a generally-anisotropic angle-ply laminate. In the current exposition, a novel approach is demonstrated whereby using coupon specimens for tensile tests extracted at different directions with respect to a global x-direction for a square angle-ply laminate and one simple in-plane shear test, the membrane normal-shear coupling compliance parameters are determined by solving a set of linear simultaneous equations derived based on transformation of stresses and strains from global to off-axis directions. The current methodology obviates the necessity of using lateral strain gages traditionally deployed in uniaxial tests thereby making material characterization of anisotropic angle-ply laminates less time-consuming and more cost-effective. </div
Chapter
Now a days, industrial processes and domestic activities contribute towards the increase of pollutants in our environment. These anthropogenic activities lead towards increase of contaminants in our water bodies. Many researchers and technologists are involved in finding novel efficient ways for treatment of waste waters and water purification. Various methods have been deployed for removing contaminants like dyes, heavy metals, oils etc. Polymers and its composites form a group of material which have diverse applications. Polymeric composites in form of hydrogel/aerogels/membranes have shown efficient removal of different types of contaminants from water and also for purification and desalination of water. Polymer composites use different types of nanofillers like nanoclays, silica, carbon nanomaterials, nanocellulose etc. for achieving water remediation applications. This chapter provides an insight into the fabrication and characteristics of polymer nano-composites especially membranes and the role of different filler incorporations in them for the treatment of waste waters.
Article
Ablative composites have been in use for thermal protection of space vehicles for decades. Carbon-phenolic composites have proven to perform exceptionally well in these applications. However with development in aerospace industry their performance needs improvement. In this field, different carbon-based and ceramic additives have been introduced into ablative composite systems. This review article gives a comparative analysis of researches done in this field in the recent past. Density, ablative, thermal and mechanical properties of ablative composites with different ultra-high temperature ceramic particles i.e. ZrSi2, Cenosphere, nano-SiO2, BN etc. and carbon-based nanoparticles i.e. CNTs, nano-Diamonds, Graphene oxide etc. used as additives, have been compared and discussed. Emphasis is put on carbon-phenolic composite systems although some epoxy matrix systems have also been discussed for comparison.
Article
In operation, pipelines are increasingly used worldwide to meet higher transportation demands for strategic fluids while optimizing costs. Fiber-reinforced polymer (FRP) composite is used in many pipeline industries of water, oil, and gas transportation since they could resist corrosion and have lighter weight to high strength ratios. FRP composite pipes are getting cheaper in installation costs, favored in various applications. FRP has good mechanical properties, which improve stiffness, strength, pressure capacity, durability, cost-benefit, and environmental footprint when combined with other materials. This paper covers different FRP composites' applications in pipelines and their development over time. It reviews the pipelines based on the transported fluid and reviews the developed composite pipelines and the repaired conventional pipelines against corrosion or leaking using different FRP types. The review includes the most common way to employ FRP composites in different pipeline applications. Different designs and applications of FRP composites in pipelines towards achieving enhanced properties are summarized and compared. Future challenges and research directions are discussed in detail.
Article
In the present work, an attempt has been made to study the three body abrasive wear behaviour of laminate hybrid composites consisting jute fiber/glass fiber reinforced with epoxy resin. The composites samples with different stacking sequence like S1 (GJJJJJJJJG), S2 (GJGJGGJGJG) and S3 (JGJGJJGJGJ) with glass (G) and jute (J) were fabricated using hand layup technique. The three-body abrasive wear behaviour of the designed hybrid composite was carried out as per ASTM G65 standards. The results indicate that the S2 composite exhibits higher wear resistance among the developed composites. The scanning electron microscope (SEM) study was performed for the worn surfaces of the composites to find out the wear mechanism.
Article
A two-dimensional theory of flexural motions of isotropic, elastic plates is deduced from the three-dimensional equations of elasticity. The theory includes the effects of rotatory inertia and shear in the same manner as Timoshenko’s one-dimensional theory of bars. Velocities of straight-crested waves are computed and found to agree with those obtained from the three-dimensional theory. A uniqueness theorem reveals that three edge conditions are required.
Book
First published in 1991. CRC Press is an imprint of Taylor & Francis.
Book
Das Buch führt in möglichst einfacher Weise in die Grundlagen der Kontinuumsmechanik ein, wobei der Schwerpunkt bei festen deformierbaren Körpern liegt. Zahlreiche Beispiele mit vollständigen Lösungen illustrieren den theoretischen Teil und erleichtern so das Verständnis. In der 4. Auflage wurden zahlreiche Abschnitte überarbeitet und präzisiert, wobei auch die unterschiedlichen Konzepte der Kontinuumsmechanik noch deutlicher gemacht werden. Zahlreiche Fehler wurden beseitigt. Gleichzeitig wurde die Referenzliteratur erweitert sowie die Liste der weiterführenden Literatur ergänzt und aktualisiert. Der Inhalt Teil I Grundbegriffe und mathematische Grundlagen.- Einführung.- Mathematische Grundlagen der Tensoralgebra und Tensoranalysis.- Teil II Materialunabhängige Gleichungen.- Kinematik des Kontinuums.- Kinetische Größen und Gleichungen.- Bilanzgleichungen.- Teil III Materialabhängige Gleichungen.- Materialverhalten und Konstitutivgleichungen.- Deduktiv abgeleitete Konstitutivgleichungen.- Induktiv abgeleitete Konstitutivgleichungen.- Methode der rheologischen Modelle.- Teil IV Anfangs-Randwertprobleme der Kontinuumsmechanik.- Grundgleichungen der linearen Elastizitätstheorie.- Grundgleichungen linearer viskoser Fluide.- Teil V Anhang.- Elastizitäts- und Nachgiebigkeitsmatrizen. Die Zielgruppen Das Buch richtet sich an Studierende im Bereich Maschinenbau und Bauingenieurwesen, Physik und Technomathematik sowie an Wissenschaftler und Praktiker in der Industrie. Der Autor Holm Altenbach studierte am Leningrader Polytechnischen Institut (heute St. Petersburger Staatliche Polytechnische Universität „Peter der Große“), wo er auch promoviert wurde und sich habilitierte. Sein beruflicher Weg begann an der TH Magdeburg, ab 1996 war er Professor für Technische Mechanik an der Martin‐Luther‐Universität Halle‐Wittenberg, seit 2011 vertritt er dieses Gebiet an der Otto‐von‐Guericke‐Universität Magdeburg.
Article
Shell structures are used in all phases of structures, from space vehicles to deep submergence hulls, from nuclear reactors to domes on sport arenas and civic buildings. With new materials and manufacturing methods, curved thin walled structures are being used increasingly. This text is a graduate course in the theory of shells. It covers shells of isotropic materials, such as metal alloys and plastics, and shells of composite materials, such as fibre reinforced polymer, metal or ceramic matrix materials. It provides the essential information for an understanding of the underlying theory, and solution of some of the basic problems. It also provides a basis to study the voluminous shell literature. Beyond being primarily a textbook, it is intended also for self study by practising engineers who would like to learn more about the behaviour of shells. The book has two parts: Part I deals with shells of isotropic materials. In this part the mathematical formulations are introduced involving curvilinear coordinates. The techniques of solutions and resulting behavior is compared to planar thin walled isotropic structures such as plates and beams. Part II then treats the behavior of shells, involving anisotropic composite materials, so widely used today. The analysis involves the complications due to the many elastic constants, effects of transverse shear deformation, thermal thickening and offer effects arising from the properties of composite materials.