179306253 lectures in mechanics of materials ion simulescu

213
- III - TECHNICAL UNIVERSITY OF CIVIL ENGINEERING Lectures in Mechanics of Materials Eng. Ion S. Simulescu, MPh, PhD, PE Associate Professor Bucharest 2004

Upload: miyotorin

Post on 13-Dec-2015

82 views

Category:

Documents


14 download

DESCRIPTION

Lectures in Mechanics of Materials Ion Simulescu

TRANSCRIPT

Page 1: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- III -

TECHNICAL UNIVERSITY OF CIVIL ENGINEERING

Lectures in

Mechanics of Materials

Eng. Ion S. Simulescu, MPh, PhD, PE

Associate Professor

Bucharest 2004

Page 2: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- IV -

PREFACE

Twenty-five years ago, in the late summer of 1979, I left Romania with the vigor and

determination of a thirty year old to conquer the West. Doing so, I left behind family,

friends and especially, what I loved the most, my teaching position at the

Construction Institute of Bucharest, the former name of today’s Technical University

for Civil Engineering. It took the persuasion, the encouragements and the friendship

of my colleague and friend Dr. Eng. Dan Cretu, Chairmen of the Strength of Materials

and Theory of Elasticity Department, to convince me to dramatically alter my life

again and return to Romania. After much hesitation, hours of long distance

conversations and sleepless nights, what appeared for many years to be just an

impossible reverie has become reality. Starting with the fall semester of 2004-2005

academic year, I will be once more part of the faculty in his department.

Since 1980, living in United States, through my own education and professional life, I

have gain a deep understanding and first hand knowledge of the American educational

system and the way structural engineering is practiced in the most advanced country

in the world. I had the opportunity, during my graduate years at Columbia University,

an ivy-league school located in the hart of New York City, to know a number of

renowned Professors including: Masanobu Shinozuka (Probabilistic Theory of

Structural Dynamics and Reliability), Majiek Bienek (Theory of Elasticity,

Viscoelasticity and Plasticity) and Frank Dimagio (Analytical Mechanics). Taking

courses, in preparation for my PhD dissertation, I was completely captivated by the

deeply theoretical character of the lectures and also by the reduced number of

practical applications solved during theses lectures. The successful understanding of

the course theoretical domain rested exclusively with the students required, through

an intensive self-study of the course notes and additional reading, to solve an

extensive number of problems. The students were willing participants in research, and

the best of them spent their summer vacations in the school laboratories or computer

rooms. The exposure to commercial computer programs in solving practical

applications came very early during the instruction years. In contrast, the structural

engineering practice, called generically design, is extremely standardized and in

general limited to a number of simple procedures that every student passing through

an undergraduate program will learn.

It is my intention to adapt and use this accumulated experience into our Romanian

educational system. I strongly believe that the Romanian education system, which

remains almost unchanged since my departure, is in need of improvement and

modernization, especially, by injecting into it the usage of computer applications, as

early as possible in the instruction.

Page 3: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- V -

In the spring of 2004, in preparation for my new academic activity, I began writing a

textbook, entitled Lectures in Mechanics of Materials. This textbook is an

introductory course to the general topic of Mechanics of Materials and is intended to

gradually expose the students, enrolled in the Civil Engineering Section of the

Technical University of Bucharest, to this fundamental subject during the first

semester of their sophomore year. The subject is known in the Romanian technical

literature as Strength of Materials, but the more general title of Mechanics of

Materials is widely employed for undergraduate studies at the engineering colleges in

the United States. Although I followed the departmental approved syllabus, the

lectures are presented in a modern manner by following the basic approach of the

modern theory of continua, namely, the usage of the three basic groups of equations:

the equilibrium, the strain-displacement and the constitutive material equations. I

considered as essential, the early implementation of these three aspects in the

theoretical developments of this extremely important engineering course. This

presentation provides a clear perspective for the students and gives them the

opportunity to compare and follow the methods used in the more sophisticated

theoretical courses of the following study years.

Even though I tried very hard to make the text as perfect and intelligible as possible

and despite the tedious and relentless effort of my young colleague John Creighton, a

structural engineer at Washington Group International, in reading and correcting it, I

am almost certain that many improvements will become necessary as the result of the

interaction with the students.

In this important stage time of my professional carrier a special gratitude is due to

Emeritus Professor Eng. Panait Mazilu, the past Chairman of the Strength of

Materials and Theory of Elasticity Department, for bringing me on in 1975 as an

Assistant Professor and providing support for my professional activity and

development until I left the country. He represents, without a doubt, the brightest

mind I have had the pleasure knowing in my entire professional carrier. A special

thanks to my first teacher in strength of materials, and later colleague and friend, Dr.

Eng. Mihaela Kunst-Germanescu, who through her total dedication made an essential

difference in my technical education. From the American side, I wish to express a

deep appreciation to Professor Masanobu Shinozuka for his support during my

graduate study at Columbia University and his continue guidance during my PhD

preparation.

I also express my appreciation to my wife Catrinel, who sacrificed greatly by

abandoning a life long successful drama carrier to follow me to the United States and

had the audacity to endure hours of technical discussions, many moments of success

as well as difficult times during the past ten years.

Finally, the author thanks Dr. Eng. Dan Cretu without this endeavor would not be

possible.

Dr. Ion S. Simulescu

Princeton, NJ, USA.

May, 2004

Page 4: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- VI -

TABLE OF CONTENTS

Lecture 1 Introduction to Mechanics of Materials 1

1.1 What is a Deformable Body?

1.2 Geometrical Classification of the Deformable Body

1.3 Exterior Action Classification

1.4 Stress and Strain

1.5 Engineering Aspects: Analysis, Verification, Optimization

and Design

1.6 Applications of the Mechanics of Materials

Lecture 2 Fundamentals of Mechanics of Materials 10

2.1 Static Equilibrium of Deformable Solid Body

2.2 Support Reaction Forces and Beam Connections

2.3 Generalized Stress Tensor and Components

2.4 Saint Venant’s Principle

2.5 Beam Stresses and Cross-Section Resultants

2.6 Extensional and Shear Strains

2.7 Constitutive Properties of Materials 2.7.1 Tension Tests

2.7.2 Mechanical Properties of Materials

2.7.3 Elasticity, Plasticity and Creep

2.7.4 Linear Elasticity, Hook’s Law and Poisson’s Ratio

2.7.5 Generalized Hook’s Law for Isotropic Materials

2.8 Fundamental Hypotheses and Equations 2.9 Examples

2.9.1 Stress Distribution and Stress Resultants in a

Rectangular Cross-Section

2.9.2 Extensional and Shear Strain

2.9.3 Volumetric Strain

Page 5: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 7 -

Lecture 3 Geometrical Characteristics of the Beam

Cross-Section 49

3.1 Definitions

3.2 Parallel-Axis Theorems for Moment of Inertia

3.3 Moment of Inertia about Inclined Axes

3.4 Principal Moments of Inertia

3.5 Maximum Product Moment of Inertia

3.6 Mohr’s Circle Representation of the Moments of Inertia

3.7 Radii of Gyration

3.8 Examples 3.8.1 Rectangle Cross-Section

3.8.2 Composite Cross-Section

Lecture 4 Equilibrium of the Plane Linear Members 68

4.1 Type of Beams, Loads and Reactions

4.2 Equilibrium of Beams Using the Free-Body Diagrams

4.3 Differential Relations between Loads and Cross-Section

Internal Resultants

4.4 Shear Force and Bending Moment Diagrams

4.5 Examples of Shear and Moment Diagrams for Statically

Determined Beams 4.5.1. Simple-Supported Beam Loaded with a Concentrated

Vertical Force

4.5.2 Simple-Supported Beam Loaded with a Concentrated Moment

4.5.3 Beam with Overhang

Lecture 5 Axial Deformation 86

5.1 Basic Theory of the Axial Deformation

5.1.1 Strain-Displacement Equation

5.1.2 Constitutive Equation 5.1.3 Cross-Section Stress Resultants

5.1.4 Equilibrium Equation 5.1.5 Thermal Effects on Axial Deformation

5.2 Uniform-Axial Deformation 5.2.1 Member Subjected to Uniform-Axial Deformation

Page 6: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 8 -

5.2.2 Statically Determinate Structure

5.2.3 Statically Indeterminate Structure

5.3 Nonuniform-Axial Deformation 5.3.1 Non-homogeneous Cross-Section Members

5.3.2 Non-homogeneous Cross-Section Members Subjected to Thermal

Changes

5.3.3 Heated Member with a Linear Temperature Variation

5.4 Special Aspects 5.4.1 Normal Stress in the Vicinity of the Load Application

5.4.2 Stress Concentrations

5.4.3 Limits of Poisson’s Ratio

5.5 Design of Members Subjected to Axial Deformation

5.6 Examples 5.6.1 Statically Determinate Structure

5.6.2 Statically Indeterminate Structure

Lecture 6 Pure Shear 112

6.1 Basic Theory

6.2 Discrete Connectors Calculation

6.3 Weld Calculation

Lecture 7 Bending 123

7.1 Definitions

7.2 Pure Bending 7.2.1 Strain-Displacement Equation

7.2.2 Constitutive Equation

7.2.3 Cross-Section Stress Resultants

7.2.4 Normal Stress Distribution

7.3 Nonuniform Bending

7.3.1 Basic Assumptions

7.3.2 Shear Stress Distribution in Rectangular Cross-Section

7.3.3 Limitations in the Usage of the Shear Stress Formula for

Compact Cross-Sections

7.3.3.1 Cross-Section Shape

7.3.3.2 Beam Length

7.3.4 Shear Stress Distribution in Thin-Wall Cross-Sections

7.3.4.1 Assumptions

7.3.4.2 Wide-Flange Cross-Section

7.3.4.3 Closed Thin-Wall Cross-Section

Page 7: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 9 -

7.4 Build-Up Beams Connectors 7.4.1 Linear Shear Connectors

7.4.2 Discrete Shear Connectors

7.4.3 Examples

7.4.3.1 Welded Connection

7.4.3.2 Nail Connection

Lecture 8 Beam Deflection 153

8.1 Introduction

8.2 Qualitative Interpretation of the Deflection Curve

8.3 Differential Equations of the Deflection Curve 8.3.1 Integration of the Moment-Curvature Differential Equation

8.3.2 Integration of the Load-Deflection Differential Equation

8.3.3 Boundary and Continuity Integration Conditions

8.4 Examples

8.4.1 Application of the Moment-Curvature Equation

8.4.2 Application of the Load-Deflection Equation

Lecture 9 Torsion 167 9.1 Introduction

9.2 Torsional Deformation of a Member with Circular Cross-

Section 9.2.1 Strain-Displacement Equation

9.2.2 Constitutive Equation

9.2.3 Cross-Section Stress Resultants

9.3 Torsional Deformation of a Member with Closed Thin-Wall

Cross-Section

9.4 Torsional Deformation of a Member with Rectangular Cross-

Section

9.5 Torsional Deformation of a Member with Elliptical Cross-

Section

9.6 Examples 9.6.1 Rectangular Thin-Wall Cross-Section

9.6.2 Capacity Comparison of Different Shaped Solid Cross-Sections

Lecture 10 Plane Stress Transformations 181 10.1 Introduction

10.2 Plane Stress Transformation Equations

Page 8: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 10 -

10.3 Principal Stresses

10.4 Maximum Shear Stress

10.5 Mohr’s Circle for Plane Stresses

10.6 Principal Stresses Distribution in Beams

10.7 Example

Page 9: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 11 -

Page 10: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 12 -

LECTURE 1

Introduction to Mechanics of Materials

This textbook is an introductory course to the general topic of Mechanics of Materials

and is intended to gradually expose the students, enrolled in the Civil Engineering

Section of the Technical University of Bucharest, to this fundamental subject during the

first semester of their sophomore year instruction. The subject is known in the Romanian

technical literature as Strength of Materials, but the more general title of Mechanics of

Materials is widely employed for undergraduate studies at the engineering colleges in

the United States of America. The American title has been retained for two main reasons:

(a) English is the language employed for academic instruction of this course and (b)

familiarity with the advanced American educational system is considered essential for the

students of this engineering school.

The theory contained in this textbook represents the base knowledge for the professional

life of any structural engineer. Without a doubt, the theoretical subjects covered in this

textbook represent the foundation for thorough understanding of more advanced subjects

that will confront the student during the undergraduate and graduate study years.

The Classical Mechanics, discipline studied in the previous year, had introduced the

fundamental principals of Statics and Dynamics and applied them in the investigation of

the particle and rigid body behavior, cases representing a drastic idealization of the real

physical systems. The Mechanics of Materials concentrates on the study of the

deformable solid body. The principal objective of this textbook is to familiarize the

student with the basic concepts of deformation, stress and strain induced in a special

category of deformable bodies, namely the beam, by external environmental actions. The

validation of the theoretical derivations is achieved by confronting them with

experimental results obtained during laboratory tests. The understanding of the

mechanical behavior of the beams subjected to exterior actions is the most essential

factor required for successful design of all types of structures such as: bridges, high-rise

buildings, industrial buildings, power plants, aircrafts, spacecrafts, etc.

Page 11: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 10 -

Note: Before elaborating about the scope and the main directions of the course,

several conventions used in the text are described. Throughout the entire body

of this textbook, the first instance usage of important terminology necessary to

establish a base technical vocabulary is shown in boldface. Additionally, a

selection of the most important definitions are enclosed in a box and presented

using the italic font.

1.1 What is a Deformable Body?

The entire domain of Mechanics of Materials is concerned with developing the

mathematical methodology necessary to completely define the behavior of the

deformable solid body.

The mathematical description of the physical changes that take place, the modification

of volume and shape, when the three-dimensional body is subjected to exterior

actions, requires that two Cartesian orthogonal coordinate systems be employed:

(1) a global coordinate system OXYZ, being considered as fixed in three-dimensional

space, and (2) a local coordinate system oxyz, rigidly attached to the three-

dimensional body. Consequently, a material point pertinent to the three-dimensional

body can then be defined by two position vectors, each one relative the two

coordinate systems. A schematic representation of the three-dimensional deformable

body subjected to various actions and the Cartesian orthogonal coordinate systems

described above are shown in Figure 1.1. The exterior actions are shown as

concentrated vectors acting in some particular points (points 1 and 2), but in general

the exterior actions have a more complex nature. The symbols attached to some

material points (points 4, 5 and 6) of the three-dimensional body are the schematic

representation of the supports or, more appropriately, constraints.

Figure1.1 Schematic Illustration of a Three-Dimensional Solid Deformable Body

Changes in the volume or shape are generically referred to as deformations.

Consequently, any point of the three-dimensional body moves after the application of

the exterior action from its initial position to a new position in space. For example the

Definition 1.1

A deformable body is a solid three-dimensional body that changes volume and

shape in response to the application of external actions

Page 12: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 11 -

material point located in the position marked with the numeral 3 deforms into the

new position *3 . The deformation of the solid body can be small or large, but it plays

a vital role in the analysis of the behavior the three-dimensional solid under the given

exterior actions.

1.2 Geometrical Classification of the Deformable Body

In terms of geometry, any solid body located in three-dimensional space is

represented by its volume and exterior surface. For this study the exterior surface is

considered continuous without any holes or interruptions. The volume is characterized

by three dimensions (length ,l width w and high h ) or by one dimension and two

ratios. For example, if the length l , typically the dimension of greatest magnitude, is

retained, than the ratios width to length lw and height to length lh fully define the

body. The following categories of bodies can be defined as:

(a) Member or Beam is a three-dimensional solid body which has the length

l much larger than the other two dimensions. To be more précised the major

dimension l can be described by a curve called the member longitudinal axis. By

intersecting the three-dimensional volume with a normal plane to the longitudinal

axis a surface is obtained. This surface is the beam cross-section. Usually the beam is

defined:

25.0),max(

l

wh

(1.1)

The cross-section represented by a single surface is called compact, while the surface

which has a central hole is called tubular. A special type of tubular cross-section is

the thin-walls cross-section characterized by very thin wall thicknesses.

The analytical description of the beam longitudinal axis as a curve in space or plane

curve can further differentiate this category into spatial and plane curved members.

The plane members are linear (beam) or curved (arch) members.

Due to the simplicity of manufacturing, the linear member, known in the engineering

practice under the generic name of the beam, is the most commonly employed

structural element. The behavior of the linear beam is the main subject of this course.

Examples of members are shown in Figure 1.2.

Page 13: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 12 -

Figure 1.2 Examples of Member Types

(a) Rectangular-Linear, (b) Circular-Linear and (c) Circular-Curved

(b) Plate or Slab is a three-dimensional solid body which has two of its three

dimensions much larger than the third one. The smallest dimension, usually called the

thickness t , is replacing the height in the original definition. The geometry of the plate

median plane or neutral plane, the plane separating the thickness of the plate into

two equal parts, can be represented as a plane or a curved surface. The plate having a

curved median plane is called shell. By definition a plate (slab) has the following

ratios:

1.0l

t and 1.0

w

t

(1.2)

where the dimensions l and w are the length and width of the median plane,

respectively.

If the thickness t is very small comparison with the other two dimensions, the plate is

called membrane.

The plate is also a widely used structural component, but its behavior is the subject of

another course. Examples of plates are depicted in Figure 1.3

Figure 1.3 Plate Types

(a) Slab (Plane), (b) Cylindrical Shell and (c) Parabolic Shell

(c) Block is a three-dimensional solid body which has all of its three dimensions

comparable.

Page 14: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 13 -

Figure 1.4 Isolated Foundation

In structural engineering this type of solids are encountered with predilection in all

kind of buildings and machine foundations. An isolated foundation supporting a

structural column is illustrated in Figure 1.4.

Note: It is important to emphasize that the geometrical characterization of the solid

deformable body is also reflected in the mathematical model used in the

description of each body type.

1.3 Exterior Action Classification

The exterior action contributing to the deformation of the three-dimensional solid can

of mechanical or a thermal nature. In the engineering practice, these actions are

labeled under the generic name of loads. The mechanical load is the direct result of

the interaction of the three-dimensional deformable solid under investigation with

other solids, liquids or gases. For example the wind action is mechanical load induced

by the hydrodynamic pressure of the air movement. Similarly, the liquid contained in

the reservoir impinges on the walls and a hydrostatic pressure results.

In general a mechanical load is considered as continuous function of two

variables ),( trp

, where r

the position vector of the material is point where the

function is described and t is the time.

For the case of linear beam type solid, the case of interest for this course, the spatial

variable is described only by the one-dimensional position vector and the function can

be written as ),( txp . The mechanical load is a vectorial function, which is

characterized by direction and intensity.

The classification of the mechanical loads is conducted base on two criteria, both

related with the intensity: (a) the time dependency and (b) the spatial variation.

Definition 1.2

If the intensity of the applied load changes with time the load is called a

dynamically applied load or a dynamic load. As a result the inertial forces,

conforming to Newton’s law, are induced and must be considered in the equations

of equilibrium.

If the intensity of the applied load does not change in time the load is called a

statically applied loads or a static load. In this case there are no inertial forces to

be included in the equations of equilibrium.

Page 15: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 14 -

The classification of the mechanical and thermal action as dynamic or static has only

theoretical merit, because, any action is to some extent of dynamic nature. In fact, in

some new textbooks the static loads are called quasi-dynamic loads in order to

incorporate them in the same theory as the dynamic loads.

The spatial variability of the function ),( trp

can be theoretically defined by any

continuous mathematical function. In engineering practice, these functions are

typically limited for simplicity to the following: constant function, linearly varying

function and parabolically varying function. For the rare case of more complicated

spatial variability the concept of piecewise–constant (stepped) or piecewise-linear

functions may be employed.

Examples of mechanical loads commonly used in the analysis of the linear plane

beams are shown in Figure 1.5.

Figure 1.5 Line Mechanical Loads

(a) Concentrated, (b) Uniformly Distributed, (c) Linearly Distributed, and (d)

Parabolically Distributed

In the reality all mechanical loads are applied on the surface of the solid body, which

can be small or large. The application surface degenerates into a line segment in the

theoretical case of the linear solid (beam). The situation when the area of the load

application surface or the length of load application segment is small relative to the

overall area or length of the three-dimensional body suggests the theoretical definition

of the concentrated load.

Page 16: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 15 -

Obviously the concentrated load can be static or dynamic in nature. This engineering

simplification can be easily accommodated when dealing with the beam case, but

creates some theoretical difficulties when the other two categories of deformable

bodies (plates and blocks) are investigated.

A special category of loads are the body forces.

A special type of the body force frequently used in structural engineering is the self-

weight of the structural element. Because the gravitational acceleration is considered

constant, this load looses its dynamic nature and can be represented as a static

mechanical load with constant magnitude.

Note: Theoretically, it is very important to differentiate between the pure mechanical

loads and body forces, even from a practical point of view, some types of body

forces are treated as mechanical loads. The self-weight is a typical example.

Thermal loads can be categorized in a manner identical to that of the mechanical

loads, as time dependent or time independent. In engineering practice, the thermal

loads are defined as expansion-contraction thermal loads and gradient thermal

loads.

For the case of the linear member (beam) the expansion-contraction thermal load

manifests when the temperature in the beam varies only along the length of the beam

(mathematically, this means that the intensity of the load depends only on the

variable x ). In contrast, the thermal gradient load manifests only when the

temperature varies only trough the thickness of the beam.

Definition 1.3

The concentrated load is a load which has the intensity obtained by the

following limiting process:

S

trptP S

),(lim)( 0

where S is the application surface

Definition 1.4

A body force is a force representing the existence of the buoyancy, magnetic,

gravitational and inertia forces. It is obtained by multiplying the unit mass of the

material contained in the solid by the acceleration of the solid.

Page 17: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 16 -

In general the thermal loading is time dependent (transient), but is reasonably

represented as steady-state where the intensity is presumed to be constant.

All of these categories of loads will be used and studied in-depth in the lectures

concerned with the development of the linearly plane beam theory.

1.4 Stress and Strain

The deformation of the solid body subjected to various types of loads modifies the

internal equilibrium of the body by forcing the three-dimensional solid to move from

its position at rest to a new equilibrium position. The analysis of the local effect of

these deformations conducts to the theoretical concepts of the stress and strain

distribution in the deformable body, the two essential concepts of the Mechanics of

Materials.

This textbook is primarily orientated towards the study of stress and strain

distributions proper for the linearly plane beam type solids. These concepts will be

introduced in the following lectures together with the constitutive law, the functional

relation between stress and strain.

1.5 Engineering Aspects: Analysis, Verification,

Optimization and Design

The problems solved by the theoretical approaches which will be developed during

the instruction time span of this course can be organized in three categories: (a)

analysis, (b) design and (c) optimization.

The engineering activities conducted to determine the deformations, and the stress and

strain distribution of a deformable solid body, when the loads and the geometry of the

body are known, is generically called analysis. This activity will be continuously

emphasized throughout the entire length of the course.

The analysis activities precede the verification activities. The object of the

verification activity is to check the maximum stress distribution and deformations,

calculated during the analysis, against some established limiting values called

allowable stress and deformation, respectively. This aspect is also emphasized in the

following chapters.

If the loads acting on the deformable body and the allowable related to the stress

distribution and deformation have been established, various types of solids can be

analyzed and found proper to the allowable limits. Therefore the described problem is

indeterminate and additional mathematical conditions must be imposed for the

complete solution to be obtained. These additional criteria are called optimization

criteria. The minimum weight criterion is an example. The optimization is a difficult

mathematical problem which is beyond the scope of this text book. However, for a

great many engineering applications, optimization can be achieved without

Page 18: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 17 -

mathematical complexity by a simple trial-and-error analysis-verification iterative

process, drawing on the judgment, experience and creativity of the structural engineer.

The three previously described engineering activities encompass the complete

engineering effort from initial concept to final design. In engineering practice, the

loads are established according to the functionality of the structural element, while the

geometrical characteristics are established through an iterative scheme involving

analysis and verification activities. The allowable limits (criteria) are established and

standardized for the material used to manufacture the structural element. Design is the

generic name of the activity for which a structural engineer is trained and tested.

1.6 Applications of the Mechanics of Materials

Applications of the theory of deformable solid bodies, especially beams and plates,

can be found in our daily life. In the past, the more sophisticated structures were

designed by a careful assemblage of simple structural elements which could be

analyzed and verified. Since 1960, with the development and increased usage of the

digital computer, the Mechanics of Materials methods evolved and were transcribed

employing scientific programming languages (FORTRAN, etc.). This way the modern

computer programs or computer codes were born. In today engineering practice a

number of commercial computer codes, such as NASTRAN, ANSYS, ABAQUS

and GT-STRUDL, are extensively employed. These computer codes, sometimes

referred to as Computer-Aided-Engineering (CAE) codes, have the capability of

conducting the analysis as well as the verification activities for the investigated

structure.

Creation of three-dimensional models and drawings of the complex structures are

conducted today utilizing the Computer-Aided-Design (CAD) codes. Many of these

codes have the capability to idealize the constructed model and submit this model

directly or with minimal analyst intervention for analysis and verification.

The analytical methods used in the development of the commercial computer codes

are beyond the scope of this introductory course in Mechanics of Materials, but will

be an integral part of the educational process of the following years.

Page 19: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 18 -

LECTURE 2

Fundamentals of Mechanics of Materials

In this lecture the fundamental definitions and concepts used in the Mechanics of

Materials are presented.

2.1 Static Equilibrium of Deformable Solid Body

In this textbook only the time independent loads, the static loads, are considered to act

on the three-dimensional solid. Consequently, the inertial forces are zero and the

necessary conditions for equilibrium are expressed as:

0***1

kRjRiRFR ZYX

n

i

i

(2.1)

kMjMiMrFM ZYX

n

i

iiO

***)(

1

(2.2)

where the notation stands for:

n - is the number of concentrated forces;

R

- is the force resultant vector and its components;

ZYX RRR

,, are the magnitude of the components of R

along the coordinate

axes;

iF

- is a particular force;

iZiYiX FFF

,, - are the magnitude of the components of iF

along the coordinate

axes;

kji

,, - are the Cartesian unit vectors of the OXYZ system;

OM

- is the moment resultant vector and its components calculated at the

arbitrary point O , the origin of the OXYZ coordinate system;

OzOyOx MMM ,, - are the magnitude of the components of OM

along the

coordinate axes;

Page 20: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 49 -

ir

- is the position vector extending from the coordinate origin to the point

where the concentrated force iF

is applied.

The vectorial equations (2.1) and (2.2) can be re-written in a scalar form of six

algebraic equations:

01

n

i

iXX FR

(2.3)

n

i

iYY FR1

0

(2.4)

01

n

i

iZZ FR

(2.5)

0OXM (2.6)

0OYM (2.7)

0OZM (2.8)

The equations of equilibrium (2.3) through (2.8) were previously developed and

extensively used in the Classical Mechanics course, where the three-dimensional rigid

solid was studied.

The mathematical statements, (2.1) and (2.2), imply that the three-dimensional solid

body is free of constraints, a situation contradictory to the norm where the constraints

are present. It should be noted that the equilibrium equations do not refer in any way

to the nature of the material present in the solid or to any type of deformation.

Consequently, those equations also apply fully in the general case of a deformable

three-dimensional solid.

To utilize the equations (2.1) and (2.2) under the usual conditions, the three-

dimensional deformable solid is released from its constraints, and these are replaced

with the corresponding reaction forces. As described in section 1.3, the drawing that

illustrates the geometry of the three-dimensional solid, the exterior actions, and the

Definition 2.1

The necessary and sufficient conditions for an unconstrained three-dimensional solid

body to be in static equilibrium (in the absence of the inertia forces) are expressed by

the vectorial equations (2.1) and (2.2)

Page 21: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 50 -

unknown reaction forces is called a free-body diagram. A generic example of a free-

body diagram is shown in Figure 2.1.

Figure 2.1 Generic Free-Body Diagrams

(a) Constrained Body, (b) Free Body

The attempt to write the equations of equilibrium, (2.3) through (2.8), using the active

forces and the reaction forces, results in a system of six algebraic equations with

constant coefficients, containing the reaction forces as unknown quantities. This

system of equations may be solved using any linear algebra method of solution for

algebraic equation systems with constant coefficients. If the equations contain only

six unknown reactions it is called a statically determinate system and, consequently,

can be solved. When the number of unknown reactions is higher than six, the equation

system is called a statically indeterminate system and additional equations are

required for the reactions to be calculated. If a number of supports can be removed

without destroying the static equilibrium of the body, those supports may be identified

as redundant supports and the corresponding reaction forces are called redundants.

In the particular case of the plane linear beam when the exterior acting loads and the

reaction forces are contained in the vertical plane OXY only three equations of

equilibrium from the initial six are required for solution as follows:

01

n

i

iXX FR

(2.9)

n

i

iYY FR1

0

(2.10)

0OZM (2.11)

Page 22: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 51 -

Note: It can be shown that only the number of equilibrium equation is fixed at six,

but the option of writing these equations vary. For the plane case the equations

(2.9) through (2.11) may be replaced, for example, by one force equation and

two moment equations.

2.2 Support Reaction Forces and Beam Connections

The external loads applied on the three-dimensional deformable body are carried to

the supports, which are the points where the interaction with the environment (other

bodies) is considered to take place. At the supporting points the displacements,

represented by a change in the position of the point, are known quantities with

prescribed value. In general, the supported point is constrained against movement in

some direction and thus, the displacement in that direction is zero. There are also

cases where a displacement of known magnitude and/or direction is imposed a

particular support point. Due to the Newton’s first law, reaction forces are developed

at the supporting points. These reaction forces are vector quantities with known

direction, but unknown intensity.

If the studied system is composed of more than one single body, the common points

between each adjoining body are called connections. To write the equilibrium

equations for a multi- body system the connections should also be replaced by

reaction forces. A free-body diagram and a set of equilibrium equations may then be

constructed for each body allowing for solution of all external support reaction as well

as for all internal connection reactions.

Tables 2.1 through 2.3 contain the most common idealized types of supports and their

corresponding reaction forces. The two-dimensional cases are illustrated in Tables 2.1

and 2.2 which are directly related with the behavior of the linear plane beam, the case

of interest for this course. The supports and reactions pertinent to the three-

dimensional case, only occasionally considered in this course, are depicted in Table

2.3 for completeness.

Table 2.1 Plane Supports and Reactions

Page 23: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 52 -

Table 2.2 Plane Connections and Reactions

Table 2.3 Three-Dimensional Supports and Reactions

2.3 Generalized Stress Tensor and Components

When subjected to external actions, the three-dimensional solid deforms and the

internal original equilibrium is disturbed. To determine the internal forces located in

the volume of the deformed body the method of sections is employed. The

application of the methodology is depicted in Figures 2.2 and 2.3.

Figure 2.2 Three-Dimensional Free-Body Sectioned by a Cutting Plane

Page 24: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 53 -

Practically, the original body is divided into two parts, each one carrying its portion of

exterior and reaction forces, by an arbitrarily orientated plane. In a manner of

speaking, the cutting plane plays, the role of the member connection. As in the

connection case, the broken continuity of the body is replaced by internal forces

acting at every point of the common surface.

Figure 2.3 Lower Half of the Sectioned Original Body

The internal force )(PR

is called the stress vector or traction and acts on the

elementary surface A located in the vicinity of point P (see Figure 2.3). Using a

local coordinate system Pnts attached to the current point P , where n

represents the

unit vector normal to the cutting plane, the stress vector )(PR

is decomposed into

three components, )(PF

, )(PVs

and )(PVt

. The decomposition process is

illustrated in Figure 2.4. The first component of the stress vector, )(PF , is orientated

parallel to the surface outside normal unit vector n

, while the other two components,

)(PVs

and )(PVs

, are located in the cutting plane.

Note: The normal n

is univocally defined by the choice of cutting plane, while the

other two directions, s

and t

are arbitrarily chosen under the condition that

the three are mutually perpendicular.

Figure 2.4 Stress Vector and Components

The components of the stress tensor located in a three-dimensional solid body are

defined through a limiting process.

Page 25: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 54 -

The position of the point P is uniquely defined by its position vector r

. The stress

tensor )(rn

is called normal stress and acts normal to the cutting plane. The in-

plane stress tensors )(rns

and )(rnt

are called shear stresses.

Note: The normal and shear stresses belong to a special mathematical category of

algebra named tensors. It is beyond the scope of this course to elaborate on

the mathematical properties of tensors, but it must be emphasized that they

may not be manipulated as vectors. The stress tensor is transformed into a

corresponding stress vector through multiplication with an area and, only then,

manipulated through the familiar vectorial algebra.

Some clarifications regarding the sign convention of the stress tensors are necessary.

The subscript n indicates that the stresses are pertinent to the surface which has the

outward normal vector n

. It is understood that the normal stress )(rn

is positive

when parallel to n

, while the in-plane shear stresses )(rns

and )(rnt

are positive

when they are parallel to the s

and t

axes, respectively.

If a new stress vector )(rpn

pertinent to the surface characterized by the outward

normal n

is defined the following vectorial equation can be written:

tArsArnArArp ntnsnn

**)(**)(**)(*)( (2.15)

After the algebraic simplification the fundamental relation between the surface stress

tensor )(rpn

and the normal and shear tensors can be expressed as:

trsrnrrp ntnsnn

*)(*)(*)()( (2.16)

If a set of three mutually orthogonal planes, defined by a Cartesian general system

and passing through the arbitrarily chosen point P are used, as shown in Figure 2.5,

Definition 2.2

The stress tensor components are defined:

A

rFr

An

)(lim)(

0

(2.12)

A

rVr s

Ans

)(lim)(

0

(2.13)

A

rVr t

Ant

)(lim)(

0

(2.14)

Page 26: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 55 -

then three corresponding sets of stress tensors are obtained by successively

considering the unit vectors i

, j

and k

as the normal vector of the cutting plane.

Figure 2.5 Set of Three Mutually Orthogonal Planes

Consequently, the equation (2.16) is written for each one of the three orthogonal

planes as follows:

krjrirrp xzxyxx

*)(*)(*)()( (2.17)

krjrirrp yzyyxy

*)(*)(*)()( (2.18)

krjrirrp zzyzxz

*)(*)(*)()( (2.19)

The resulting stress tensors are depicted in Figure 2.6.

The following close examination and discussion of the orientation of the three sets of

tensors shown in Figure 2.6 defines the sign convention used.

Figure 2.6 Three-Dimensional Stress Tensors (Cartesian Axes)

The normal stresses are positive when orientated parallel to the outward normal of the

surface. They tend to pull the plane and consequently are said to be positive in

Page 27: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 56 -

tension. The shear stress tensor is considered to be positive when orientated parallel

to and in the positive direction of the corresponding in-plane axis. These shear stress

tensors are identified by two subscripts: the first subscript indicates the plane on

which the shear stress acts by the axis defining its normal; the second subscript

indicates the shear stress direction by identifying the coordinate system axis to which

the shear tensor is aligned.

In Figure 2.6, for clarity reasons, only the stress tensors pertinent to the visible faces

are shown. For the parallel planes having negative axes as normal the stresses have

opposite signs. Why? The infinitesimal volume has to be in equilibrium.

Figure 2.7 Three-Dimensional Free-Body Diagram of an Infinitesimal Volume

Employing the notation convention described above, it is observed that the shear

stress tensors converging towards any particular edge have the subscript order

reversed. Consider the situation depicted in Figure 2.7, where for clarity only the

stress tensors which can be used in writing a moment equation around the oz axis are

retained.

Note: The Figure 2.7 is simply a free-body diagram for an elementary volume

located around a point P.

The sign convention previously explained is used. The usage of the superscript )( or

(-) indicates the variation of the same stress tensor in-between two parallel faces. The

following equations relating identical stress tensors located on parallel faces, where

the position vector r

has been drop in the notation, are written as:

xxx (2.20)

yyy (2.21)

xyxyxy

(2.22)

Page 28: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 57 -

yxyxyx (2.23)

zyzyzy (2.24)

zxzxzx (2.25)

Note: The variation of the stresses between the parallel faces is due to the existence

of the exterior actions.

For the infinitesimal volume to be in equilibrium the total moment induced by the

stress vectors must be zero.

0oxM (2.26)

0oyM (2.27)

0ozM (2.28)

The moment component ozM written about the oz axis considering the right-hand

rule is expressed as:

2***

2***

2***

2***

******2

***

2***

2***

2***

xyx

xyx

yyx

yyx

yzxxyzx

zx

xzx

yyz

yyzM

zy

zyzxzx

yxxyy

yxxoz

(2.29)

Substituting the relations (2.20) through (2.25) and (2.29) into the equilibrium

equation (2.28), equation (2.30) is obtained:

0***)(

xyzM yxxyoz (2.30)

Note: The final expression of the equation (2.30) is obtained by neglecting the

product of the change in stress ( or ) and the geometrical elementary

quantities ( x , y and z ) as being very small quantity as x , y and

z tends to zero.

The equation (2.30) indicates that for the equilibrium to be enforced, the edge

converging shear stress tensors are equal. The equality of the edge converging shear

tensors is called the duality principle of the shear stress tensors and is expressed as:

Page 29: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 58 -

yxxy (2.31)

It can be concluded that even in the presence of the normal stresses the shear stresses

must satisfy the equation (2.31)

In a similarly manner, the other two equilibrium equations (2.26) and (2.27) involving

moments around ox and oy axes are written and the duality of the other shear stresses

is obtained:

zxxz (2.32)

zyyz (2.33)

The stress tensors can be collected in a tabular form called the generalized stress

tensor:

)()()(

)()()(

)()()(

)(

rrr

rrr

rrr

rT

zzyzx

yzyyx

xzxyx

(2.34)

The generalized stress tensor )(rT

, which plays an important role in the deformable

solid theory, has nine stress tensors components, but due to the duality principle only

six are independent.

Note: It has to be emphasized that the general stress tensor )(rT

is point dependent.

2.4 Saint Venant’s Principle

The Saint Venant’s Principle is not a rigorous law of Mechanics of Materials, but has

a great practical significance in the analysis of beams and other deformable bodies.

Definition 2.3

Saint Venant’s Principle is stated as follows:

The stresses in a deformable solid body at a point sufficiently remote from points of

application of load depend only on the static resultant of the loads and not on the

local distribution of the loads.

Page 30: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 59 -

The theoretical explanation of the Saint Venant’s Principle is depicted in the Figure

2.8.

Figure 2.8 Saint Venant’s Principle

The original free-body diagram of the three-dimensional body in equilibrium is shown

in Figure 2.8.a. The three-dimensional solid is loaded with a force q distributed over a

small surface of area A . Suppose that the generalized stress tensor )(rT

is known for

any current point )(rP

of the solid body, identified by its position vector r

. Two

concentrated loads of magnitude Q opposing each other and acting collinearly with

the resultant of the distributed force q are superimposed on the three-dimensional

body as shown in Figure 2.8.b. Their magnitude Q is obtained by integrating the

variation of the distributed force q over area A :

A

dAqQ * (2.35)

Proceeding in this manner, the equilibrium of the three-dimensional solid body

remained unchanged and consequently, the generalized stress tensor )(rT

remains

unaffected. The new loading situation shown in Figure 2.28.b may then be divided

into two load cases as indicated in Figures 2.8.c and 2.8.d. The original generalized

tensor )(rT

is written as the sum of two new generalized stress tensors )()1( rT

and )()2( rT

, each one corresponding to the loading situation illustrated in Figure 2.8.c

and 2.8.d, respectively.

)()()( )2()1( rTrTrT

(2.36)

Note: This relation is valid only under the restrictions of the Principle of Linear

Superposition discussed in section 2.8.

Because the loads shown in Figure 2.8.c are self-equilibrated (force and moment

resultants are zero) the generalized stress tensor )()1( rT

calculated in points away

from the loading is expected to tend toward a zero value.

0)()1( rT

(2.37)

Page 31: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 60 -

Consequently, the original generalized tensor )(rT

is approximated by the

generalized stress tensor )()2( rT

pertinent to the loading situation contained in Figure

2.28.d:

)()( )2( rTrT

(2.38)

Equation (2.34) provides validation for Saint Venant’s Principle.

2.5 Beam Stresses and Cross-Section Resultants

The geometrical description of the linear member or beam was given in Section 1.3.

Here, it is re-emphasized that the length of the member is much larger than the other

two dimensions. Of principal importance in characterizing the behavior of the beam is

a geometrical description of its cross-section. The geometrical properties of the cross-

section and its theoretical role in the development of the beam theory will be

elaborated on during the entire length of these lectures.

Consider the case of a beam with a general, undefined, geometrical cross-section as

shown in Figure 2.9. The volume of the beam was sectioned by a normal cutting plane

with its outward normal parallel to the longitudinal local axis ox . The general

coordinate system Pnts , previously used, is identical now with the local coordinate

system oxyz attached to the beam.

Figure 2.9 Cross-Section Beam Stresses

Definition 2.4

The surface obtained by intersecting the linear member (beam) volume with a cutting

plane having as outward normal n

coincident with the longitudinal local axis of the

member is called the beam cross-section.

Page 32: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 61 -

In this section the general formulas, expressed in equations (2.12) through (2.14), will

be applied to the case in point of this textbook, the linear beam.

The normal stress )(rn

is renamed as )(rx

, while the shear stresses )(rns

and

)(rnt

became )(rxy

and )(rxz

. The local position vector is written as:

kzjyixr

*** (2.39)

where i

, j

and k

are the unit vectors of the local Cartesian coordinate system.

The stress tensor components )(rx

, )(rxy

and )(rxz

can be recomposed in the

corresponding stress vectors F

, yV

, and zV

, respectively, by multiplying them

with the elementary surface A . By integrating over the entire surface of the cross-

section the components of the resultant force and moment particular to the cross-

section are obtained. They are named cross-sectional resultants and are expressed as:

A

x dArxF *)()(

(2.40)

A

xyy dArxV *)()(

(2.41)

A

xzz dArxV *)()(

(2.42)

A

xy

A

xz dArzdAryxT *)(**)(*)(

(2.43)

A

xy dArzxM *)(*)(

(2.44)

A

xz dAryxM *)(*)(

(2.45)

The cross-section resultants are depicted in Figure 2.10.

Fig 2.10 Cross-Sectional Resultants

Page 33: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 62 -

The cross-section forces expressed in the equations (2.40) through (2.42) are called

force resultants. Force )(xF is named cross-section normal force or cross-section

axial force, while forces )(xVy and )(xVz are called cross-section shear forces. The

cross-section normal force is positive if orientated parallel to the outward normal of

the cross-section, while the cross-section shear forces are positive when orientated

parallel to the oy and oz positive axes. The normal force is called the tension or

compression force if the action is in the positive or negative direction of the cross-

section normal, respectively.

The equations (2.43) through (2.45) are written considering that the local coordinate

system follows a right-hand rule and they are called the moment resultants. The

moment vector )(xT parallel to ox axis is named cross-sectional torsion moment or

torque. It has a twisting action on the cross-section. The other two cross-sectional

moments, )(xM y and )(xM z , acting parallel with oy and oz positive axes are

named cross-sectional bending moments.

2.6 Extensional and Shear Strains

When subjected to exterior actions the deformable solid body exhibits volume and

shape changes. Geometrically speaking, this means that the geometry of the body

before the action starts differs from the final geometry. During the deformation

process all of the geometrical characteristics (line segments, angles, surface and

volume) of the solid body are altered. A particular point P moves at the end of the

deformation process to a new position *P . Similarly, the point Q , located in the

vicinity of point P , is relocated in the new position *Q . The initial undeformed and

final deformed conditions of the three-dimensional solid body are depicted in Figures

2.11.a and 2.11.b, respectively.

Figure 2.11 Deformation of the Infinitesimal Line Segment

(a) Undeformed Body and (b) Deformed Body

The infinitesimal arc PQ , which has arclength s , is deformed into a new

infinitesimal arc **QP with arclength *s . The definition of the extensional strain is

given below.

Page 34: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 63 -

The modification of the original right angle defined by the intersection of two line

segments as a result of the deformation process is called shear strain and is

schematically shown in Figure 2.12.

Figure 2.12 Modification of the Right Angle

(a) Undeformed Body and (b) Deformed Body

The original right angle 2

changes after the deformation into a new angle * .

Definition 2.5

The extensional strain in the direction n

at a point )(rP

, identified by its position

vector r

, is defined as:

)*

(lim)(s

ssr

nalong

PQn

(2.46)

Definition 2.6

The shear strain at point )(rP

, identified by its position vector r

, represents the

change of the right angle between two line segments attached to the point and

orientated in directions n

and t

, is defined as:

)2

(lim)( *

talong

PRnalong

PQnt r

(2.47)

Page 35: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 64 -

The extensional and shear strain definitions are generalized for three mutually

orthogonal planes in a similar manner to the development given for the generalized

stress tensor in Section 2.3. By choosing the Cartesian unit vectors successively as

normal to the cutting planes, the generalized strain tensor )(rT

is obtained:

)()()(

)()()(

)()()(

)(

rrr

rrr

rrr

rT

zzyzx

yzyyx

xzxyx

(2.48)

where:

rx

( ), )(ry

and )(rz

– are the elongation strain measured in the ox , oy

and oz , respectively;

)(rxy

, )(rxz

and )(ryz

- are the shear strains.

Situations involving specific the use of specific components of the generalized strain

tensor )(rT

are commonly encountered and thus, the subject will be revisited in the

following lectures.

The rational used in the development of the Saint Venant’s principal considered only

the generalized stress tensor, but a similar argument can be made for the case of the

generalized strain tensor.

2.7 Constitutive Properties of Materials

The functional expressions relating the generalized stress tensor components to the

generalized strain tensor components are called constitutive equations. They reflect

the nature of the material contained in the deformable body. Mathematically, a

general expression such as equation (2.45) may be written to define the

aforementioned functional relationship:

),...),(),,((),( trdt

dtrftr nmnmijij

(2.49)

where ),( trij

- is the component of the generalized stress tensor;

),( trnm

- is the component of the generalized strain tensor;

ijf - is a given function;

dt

d - is the time derivative

Page 36: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 65 -

The functions ijf are chosen such that they satisfy the Laws of Thermodynamics and

with parameters established by laboratory testing to give reliable results in physical

usage. The change in the subscripts, from i and j to m and n , indicates that in

general a stress tensor component may be a function of all the strain tensor

components. Only in a few simple cases does a one-to-one relation exist.

2.7.1 Tension Tests

In a laboratory control environment, using a hydraulically actuated testing machine, a

material specimen is subjected to a tension test. The schematics of the theoretical

tension test are shown in Figure 2.13.

The tension test is conducted in the following steps: (a) the original (undeformed)

gage length 0L and cross-sectional area 0A of the specimen are measured before the

test; (b) the ends of the specimen shown in Figure 2.13.a are gripped in the testing

machine; (c) force is slowly applied to the ends of the specimen until the rupture of

the material is obtained; (d) during the load application measurements of the gage

length *L , cross-sectional area *A and the value of the applied forced *P are

tabulated. The steps (c) and (d) are repeated at regular time intervals and the loading

of the specimen is conducting slowly in order to avoid the rise in temperature and

dynamic effects. When the rupture of the specimen occurs the area minA is measured

once again. This kind of laboratory experiments will be carried out in the Department

Laboratory during the course.

Figure 2.13 Tension Test of Structural Steel Specimen

(a) Undeformed Specimen, (b) Deformed Specimen

Figure 2.14 shows the test results obtained from the testing of a structural steel

specimen, characterized by low carbon content. To emphasize the test results obtained

at lower strain values the beginning end of the test was plotted at a magnified scale.

This type of steel is commonly used in structural applications (buildings, bridges,

etc.).

Page 37: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 66 -

Figure 2.14 Stress-Strain Diagram

The values of the normal stress and elongation strain can be calculated for each

loading step as follows:

0

* *

A

Px (2.50)

0

0* *

L

LLx

(2.51)

where (*) indicates a particular measurement.

The normal stress defined by the equation (2.50) is called engineering stress, while

the extension strain expressed by equation (2.51) is known as the engineering strain.

If the normal stress is calculated using the value of the step measured area *A , the

value obtained is called the true stress and the corresponding strain is called the true

strain. They are calculated as follows:

*

**

A

P

truex (2.52)

)ln()*

ln(0

*

0*

L

L

A

A

truex (2.53)

Note: The definition (2.53) is based on the fact that the volume remained

unchanged **

00 ** LALA .

The engineering strain and the true strain can are related as expressed by equation

(2.54).

Page 38: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 67 -

)1ln( **

xtruex (2.54)

Note: Analyzing the expressions of the engineering and true stress and strain it has

to be noted that the true values are larger than the engineering values.

2.7.2 Mechanical Properties of Materials

Several important material properties are obtained from the analysis of the stress-

strain diagram. The theoretical behavior of a typical structural steel is illustrated by

the stress-strain diagram shown in Figure 2.15.

The loading and deformation process begins at point A and continues until point B is

reached. This is the elastic region and is characterized by a proportionality of the

ratio between stress and strain. The point B is named proportional limit occurring

when the stress reaches PL . The ratio of stress to strain in the linear range (elastic

region shown) is called the modulus of elasticity or Young’s modulus and is

designated by the letter E .

x

xE

when PLx (2.55)

The units for the modulus of elasticity are stress units (F/L2). Typical units are ksi or

GPa.

Figure 2.15 Theoretical Description of the Stress-Strain Diagram

Continuing beyond the stress and strain characterizing the proportional limit at point

B, the specimen begins to yield. Smaller incremental loading steps are necessary to

induce additional elongation. Two new points, C and D, named the upper yielding

point, uYL )( , and the lower yielding point, lYL )( , are obtained. For practical

Page 39: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 68 -

purposes only the lower yielding point D is retained and will be called the yielding

point YL . From point D to point E increased elongation of the specimen occurs with

no increase in stress and consequently, the stress at point E has an identical value with

the stress at point D. As shown in Figure 2.15, the diagram is horizontal and the

region is called the perfectly plastic region.

The stress begins again to increase with increased in elongation after passing point E.

The increase continues until point F, where the ultimate stress or the ultimate

strength, is reached. The phenomenon and the zone are called strain hardening and

strain hardening region, respectively.

If the elongation of the specimen continues a decrease in the stress is recorded and the

so-called “necking” appears. The “necking” represents the visual reduction of the

cross-section and continues until the fracture of the specimen is attended at point G.

The stress calculated at point F, is the maximum stress characterizing the entire stress-

strain diagram. This value is called the ultimate stress or ultimate strength and is

designated by U .

If the material stress and strain remain in the elastic region where the stress is

proportional the strain, the material exhibits elastic behavior. If the material is loaded

over the proportionality limit PL the material exhibits plastic behavior.

The material behavior described above is based on the engineering characterization of

the stress and strain. For comparison the true stress and strain behavior is also plotted

in Figure 2.15. The true stress-strain diagram is plotted using a dashed line, while the

engineering stress-strain diagram is plotted with a heavy continuous line.

The area contained under the stress-strain diagram represents the energy necessary to

deform the specimen. It can be noted, from Figures 2.14 and 2.15, that the energy

characterizing the elastic deformation is considerable smaller than the energy

characterizing the plastic deformation. This is an important characteristic typical of

ductile materials.

The behavior of the structural steel, as it was described above, can be explained

without going into the theoretical details by behavior of the metal micro-structure.

Obviously, different metals exhibit different type of stress-strain behavior when

subjected to a tension test. To emphasize this observation, the stress-strain diagrams

for low carbon steel, three high-strength alloy steels and three aluminum alloys are

presented in Figures 2.16.a and 2.16.b, respectively.

Page 40: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 69 -

Figure 2.16 Examples of Stress-Strain Diagrams

(a) High-Strength Alloy Steels, (b) Aluminum Alloys

Materials, such as structural steel, which undergo large strain before fracture are

called ductile materials. The ductility factor is represented by the ratio of the strain

measured at rupture (point G) and that corresponding to the point B in Figure 2.15.

For metals used in structural engineering the ductility factor can reach values between

10 and 20.

The materials which fail at small strain value are called brittle materials. A

qualitative difference between a ductile and brittle material are shown in Figure 2.17.

Typical examples of brittle behavior are the following materials: ceramics, glass, cast

iron. The welding material, when the weld is improperly made, also exhibits brittle

behavior which in many cases compromises the quality of the entire structural

ensemble. Drastic change in temperature, pressure and load application can very

much influence the material behavior.

Figure 2.17 Ductile and Brittle Stress-Strain Diagrams

The temperature is an important factor in the material behavior. Same materials can

change from ductile to brittle behavior as a function of the temperature.

Page 41: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 70 -

Figure 2.18 Variation of the Stress-Strain Diagram with Temperature

(a) Low Temperature, (b) High Temperature

In most common applications the materials used in structural engineering are not

subjected to temperature extremes great enough to adversely affect the ductile

behavior. For the special situations when the material is subjected to high or low

temperature corresponding stress-strain diagrams must be constructed. Figures 2.18.a

and 2.18.b illustrate the variation of the stress-strain diagram with temperature for a

type of stainless steel. Note that the strain increases for a given stress with increase in

temperature and vice versa.

The American Society of Mechanical Engineers (ASME) publishes a handbook

containing a large range of stress-strain diagrams for different metal composition and

temperatures.

2.7.3 Elasticity, Plasticity and Creep

The stress-strain diagram described above has been obtained experimentally by

continuous loading of the specimen. If at some time during the loading the load

applied to the specimen is reversed, theoretically following the same increments as

were used during loading, the process is called un-loading.

If the un-loading stress-strain diagram follows the same path as the loading branch the

material is said to be an elastic material. Obviously, the elasticity of the material is

manifested only if the un-loading occurs before the stress and strain reach the elastic

limit, represented by the stress-strain point C shown in Figure 2.19.

As described previously for a typical structural steel, the stress is proportional to

strain (the modulus of elasticity has a constant value) and the behavior is called

linear-elastic. There are materials characterized by a nonlinear-elastic behavior. An

example of non-linear elastic material is shown in Figure 2.19.

Page 42: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 71 -

Figure 2.19 Non-Linear Elastic Behavior

If the stress-strain values corresponding to point C are exceeded the material exhibits

a plastic behavior and is said to undergo a plastic flow. A typical loading and un-

loading cycle from a plastic range is schematically shown in Figure 2.20. Noted that

during un-loading the material behave as a linear-elastic material.

Figure 2.20 Plastic Behavior

For theoretical applications, the stress-strain diagram shown in Figure 2.15 is replaced

by an artificial but useful stress-strain diagram named Prandtl’s diagram. This type

of diagram is shown in Figure 2.21 and represents a material with a linear-elastic and

perfectly plastic behavior.

Page 43: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 72 -

Figure 2.21 Prandtl’s Stress-Strain Diagram

The experimental loading and un-loading of the specimen being discussed here are

conducted in a relatively short time interval (a few minutes). An interesting

phenomenon occurs if the specimen is loaded and then left under a constant load for a

long period of time. If the initial strain in the specimen is calculated and the

calculation is repeated at intervals of a few days, an increase in the strain is observed.

This phenomenon is called creep. The result of a creep experiment is shown in Figure

2.22.

Figure 2.22 Creep

Similarly, if the specimen is kept under constant strain for a long period of time a

reduction in the stress value can be observed. This phenomenon is known as

relaxation. A relaxation diagram is depicted in Figure 2.23.

Page 44: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 73 -

Figure 2.23 Relaxation

2.7.4 Linear Elasticity, Hook’s Law and Poisson’s Ratio

The elastic behavior of materials is important for the structural engineering point of

view. From the explanations regarding the stress-strain diagram it is clear that if the

material is loaded beyond the elastic limit permanent deformations will be present in

the solid body. In general, structural engineers design their structures to behave

elastically.

Analyzing the stress-strain diagram interval between points A and B illustrated in

Figure 2.15 the proportionality between the stress and strain is evident.

Mathematically, the relationship is expressed as:

xx E * for YPx 0 (2.56)

Equation (2.56) is known as Hook’s law. As noted earlier, the constant E is named

modulus of elasticity or Young’s modulus. Structural steels typically have modulus of

elasticity varying around the value of 30000 ksi or 210 GPa.

During the elongation of the tensile specimen, a reduction of the dimensions in the

other two directions normal to the deformation direction is observed. This transversal

contraction, mathematically represented by equation (2.57), is illustrated in Figure

2.24 at an exaggerated scale:

xzy * (2.57)

The constant is called Poisson’s ratio. This constant is non-dimensional and varies

for structural steel around the value of 1/3 value. Theoretically, the Poisson’s ratio is

limited to 0.5. The limits of the Poisson’s ratio are treated in Section 5.4.3.

Page 45: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 74 -

Figure 2.24 Elongation and Transversal Contraction

2.7.5 Generalized Hook’s Law for Isotropic Materials

The relationship between stress and strain described for the simple case of uni-axial

elongation will be extended to the general case involving generalized stress and strain

generalized tensors. The material constants E and , together with the thermal

coefficient of expansion are the most relevant material characteristics.

Figure 2.25 depicts the general case of the three-axial elongation. Under the influence

of applied load and temperature change, the total extensional deformation can be

decomposed into four independent effect components: (a) the deformation resulting

from the change in temperature, (b), (c) and (d) deformations due to individual

application of the three-directional normal stresses. This approach is called the

Principle of Linear Superposition.

Note: The Principle of Linear Superposition is discussed in Section 2.8. The viability

of this principle is restricted to the case of small deformation and linear elastic

materials.

Definition 2.7

An elastic solid is isotropic if its material characteristics are invariants under any

transformation of the coordinate system.

Page 46: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 75 -

Figure 2.25 Deformation Induced by a Change in Temperature and

Normal Stress Components

Mathematically, the phenomena illustrated in Figure 2.25 are described with the

following equation:

zyx

xxx

T

xx

(2.57)

zyx

yyy

T

yy

(2.58)

zyx

zzz

T

zz

(2.59)

where the notation stands for:

x - is the total elongation strain in ox direction; T

x - is the elongation strain induced by the thermal expansion;

zyx

yxx

,, - are the elongation strains induced by the action of the normal

strains zyx ,, , respectively, in the ox direction

The total strains y and z are similarly defined by substitution of the appropriate

subscripts.

The existence of the normal stress x acting along the ox direction (see Figure 2.25.c)

introduces, as described in the previous section, the following elongation strains:

E

x

xx

(2.60)

Page 47: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 76 -

E

x

yx

(2.61)

E

x

zx

(2.62)

Analogously, the formulae (2.60) through (2.62) are written for the case when the

normal stress y is considered (see Figure 2.25.d) as:

E

y

xy

(2.63)

E

y

yy

(2.64)

E

y

zy

(2.65)

Considering the application of the normal stress z depicted in Figure 2.25.e the

following equations are written:

E

zx

z

(2.66)

E

zy

z

(2.67)

E

zz

z

(2.68)

The thermal strains, which are only elongations, are expressed as:

TT

z

T

y

T

x * (2.69)

where:

- is the coefficient of thermal expansion;

T - is the change in temperature.

Introducing the formulae (2.60) through (2.69) into (2.57) through (2.59) equations

(2.70) through (2.72) are obtained as:

TE

zyxx *)]**[*1

(2.70)

Page 48: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 77 -

TE

zyxy *)]**[*1

(2.71)

TE

zyxz *)]**[*1

(2.72)

For the case of the isotropic linear material the shear strains are related to the shear

stresses by the following relationships:

xyxyG

*1

(2.73)

xzxzG

*1

(2.74)

yzyzG

*1

(2.75)

Figure 2.26 Deformation induced by the Shear Stress Components

Equations (2.73) through (2.75) are illustrated in Figure 2.26. The constant G is the

shear modulus and is obtained from E and using the following relation:

)1(2

EG (2.76)

Equations (2.70) through (2.75) are named the Generalized Hook’s Law and are

widely used in structural engineering applications. By algebraic manipulations

equation (2.70) through (2.75) may be re-written with the stress tensor components

expressed as functions of the strain tensor components. The resulting equations are:

TEE

zyxx

**)*21(

]***)1[()*21(*)1(

(2.77)

Page 49: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 78 -

TEE

zyxy

**)*21(

]**)1(*[)*21(*)1(

(2.78)

TEE

zyxz

**)*21(

]*)1(**[)*21(*)1(

(2.79)

xyxy G * (2.80)

xzxz G * (2.81)

yzyz G * (2.82)

Equations (2.77) through (2.82) represent the second form of Hook’s law.

2.8 Fundamental Hypotheses and Equations

The methods employed in Mechanics of Materials are based on three basic

hypotheses. They are as follows:

(H1) The solid body is continuous and remains continuous when subjected to exterior

actions or changes in temperature;

(H2) Hook’s Law;

(H3) There exists a unique unstressed state of body to which the body returns

whenever all the exterior actions are removed.

The body satisfying those thee hypotheses is defined as a linear elastic body. Under

the hypotheses (H1) the real microscopic atomic structure of the solids is ignored and

the solid body is idealized as a geometrical copy in the Euclidian space whose points

are identical with the material points of the body. This way the continuity is simulated

and no cracks or holes may open in the interior of the solid during the exterior actions.

A material satisfying the hypotheses (H1) is called a continuum. The hypotheses

(H2) represent the mathematical base of the definition of material elasticity. The

direct implication of the Hook’s Law is the Principle of Linear Superposition, a

principle frequently used in the Mechanics of Materials.

Page 50: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 79 -

The fundamental concepts presented in the preceding sections are grouped in three

fundamental equations extensively used in Mechanics of Materials: (1) the

equilibrium equations, (2) the geometry of deformation and (3) the material behavior.

2.9 Examples

2.9.1 Stress Distribution and Stress Resultants in a Rectangular Cross-Section

The cross-section characteristics and the stresses acting on it are depicted in Figure

2.27. The origin of the coordinate system oxyz passes though the center of the

rectangular cross-section.

Figure 2.27 Example 2.9.1

The stresses existing in the cross-section are expressed by the following functions:

zCyCCx ** 210 (2.83)

2

3 * yCxy (2.84)

Definition 2.8

Principle of Linear Superposition is stated as follows

The displacement in a material point of the solid body resulting from the consecutive

action of exterior loads is equal to the sum of the displacements in the same material

point if the loads were to act in independently.

Page 51: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 80 -

2

54 * zCCxz (2.85)

Substituting the relations (2.83) through (2.85) into equations (2.40) through (2.45)

the stress resultant forces and moment are calculated as:

ACdAzCdAyCdAC

dAzCyCCdAzyxxF

AAA

AA

x

******

*]**[*),,()(

0210

210

(2.86)

12

****

*]*[*),,()(

3

3

2

3

2

3

hbCdAyC

dAyCdAzyxxV

A

AA

xyy

(2.87)

A A

AA

xzz

hbCACdAzCdAC

dAzCCdAzyxxV

12

******

*]*[*),,()(

3

54

2

54

2

54

(2.88)

0********

*]*[**]*[*

*),,(**),,(*)(

2

3

2

54

2

3

2

54

AAA

AA

A

xy

A

xz

dAyzCdAzyCdAyC

dAyCzdAzCCy

dAzyxzdAzyxyxT

(2.89)

12

**

*******

*]**[**),,(*)(

3

2

2

210

210

hbC

dAzCdAyzCdAzC

dAzCyCCzdAzyxzxM

AAA

AA

xy

(2.90)

12

**

*******

*]**[**),,(*)(

3

1

2

2

10

210

hbC

dAzyCdAyCdAyC

dAzCyCCydAzyxyxM

AAA

AA

xz

(2.91)

For numeric application the following cross-section dimensions and constant

coefficients are assigned:

Page 52: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 81 -

10.0b m 20.0h m

6

0 10*120C Pa 6

1 10*80C3m

N 6

2 10*40C3m

N

6

3 10*100C 4m

N

6

4 10*270C Pa 6

5 10*30C4m

N

The area and the stress resultants obtained by replacing the above values into the

equations (2.86) through (2.91) are:

02.0* hbA 2m

6

0 10*4.2*)( ACxF N

33

3 10*67.612

**)(

hbCxVy N

63

54 10*4.512

***)(

hbCACxVz N

0)( xT mN *

23

2 10*67.612

**)(

hbCxM y mN *

33

1 10*33.512

**)(

hbCxM z mN *

2.9.2 Extensional and Shear Strain

The deformation of a thin square plate with an edge length a is illustrated in Figure

2.28. The corner B moved horizontally by the length b into a new position *B , while

the other three corners kept their original location. It is supposed that the edge **BA

remains linear through the deformation.

The extensional strain in the ox direction, accordingly to the definition (2.46) is:

x

xxyx xx

*

0lim),( (2.92)

Page 53: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 82 -

To determine the lengths x and *x two points P and Q as shown in Figure 2.28

are selected. After the deformation of the plate they are located in the new positions *P and *Q .

The original area of the plate is theoretically divided in both directions in n equally

spaced divisions and the mesh shown in the Figure 2.28 is obtained. For clarity, in

Figure 2.28 the number of division was limited only to five ( 5n ). The point P is

located at node ),( ji of the mesh, while the point Q is located at node ),1( ji . The

original length x is obtained as:

n

ai

n

ai

n

axxPQx PQ *)1(* (2.93)

Figure 2.28 Elongation Strain in Square Plate

(a) Undeformed and (b) Deformed Shapes

After the deformation the segment **QP is shorter and is calculated as:

yna

b

n

a

n

ya

ba

in

ya

ba

ya

b

in

ya

ba

ya

bxxQPx

PQ

**

*

)](

*

*[-

)]1(

*

*[**

***

(2.94)

The extensional strain is obtained as:

Page 54: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 83 -

ya

b

n

an

ay

na

b

n

a

x

xxyx nxx *

**limlim),(

2

*

0

(2.95)

It can be concluded that the elongation strain ),( yxx is a function only of variable

y and varies between values of zero and a

b .

The shear strain ),( yxxy which reflects the right angle change is calculated following

a similar rationale. The original right angle is defined by the segments PM and PQ

as shown in Figure 2.29.a. After the deformation, as illustrated in Figure 2.29.b, the

angle changes to angle * defined between the segments **MP and **QP .

Figure 2.29 Elongation Strain in Square Plate

(a) Undeformed and (b) Deformed Shapes

From definition 2.6 the shear strain is given by:

]2

[lim),( *

00

yxxy yx (2.96)

The following geometric relations can be written:

)arctan(2 **

***

PM

PM

yy

xx

(2.97)

in

ya

ba

ya

bx

P*

*

**

(2.98)

Page 55: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 84 -

in

b

n

b

in

ya

ba

ya

bi

n

n

ay

a

ba

n

ay

a

bx

M

*

*

*

**

)(*

)(*

2

*

(2.99)

jn

ay

P** (2.100)

)1(** jn

ay

M (2.101)

The angle * is calculated using the above formulae:

)]1(*arctan[2

]

)1(*

arctan[2

*

a

x

a

b

n

an

i

n

b

(2.102)

The shear strain is obtained as:

)]1(*arctan[]2

[lim),( *

00

a

x

a

byx

yxxy

(2.103)

The shear strain ),( yxxy is dependent only on variable x and varies from )arctan(a

b

to zero value.

2.9.3 Volumetric Strain

The rectangular parallelepiped of isotropic linear elastic material is subjected to three

normal stresses x , y and z as shown in Figure 2.30.

The absence of the shear stress indicates that only volumetric changes result without

shape change. Consequently, the undeformed volume V is modified and after the

deformation the solid body has a new volume *V . The volumetric deformation is

called dilatation. The volumetric strain is defined by the following ratio:

V

VV

V

VV

*

(2.104)

Using the notation shown in Figure 2.30 the following relations are written:

zyx LLLV ** (2.105)

Page 56: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 85 -

**** ** zyx LLLV (2.106)

Figure 2.30 Volumetric Strain

The change in the dimensions is expressed as:

xxx LL *)1(*

(2.107)

yyy LL *)1(*

(2.108)

zzz LL *)1(*

(2.109)

Substituting the above expressions into the expression of the volumetric strain the

following equation is obtained:

zyxzyzx

yxzyxzyx

zyx

zyxzzyyxx

VLLL

LLLLLL

1)1(*)1(*)1(

**

***)1(**)1(**)1(

(2.110)

If the strains are small the products can be neglected and the volumetric strain

becomes:

zyxV (2.111)

Substituting the strain equations (2.70) through (2.72) into equation (2.111) and for

the condition 0T the volumetric strain is written as a function of stresses:

)(*21

zyxVE

(2.112)

Page 57: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 86 -

Equation (2.112) indicates that the volumetric strain is always a positive value in the

presence of three dimensional tensions, which geometrically represents a volume

increase.

Page 58: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 87 -

LECTURE 3

Geometrical Characteristics of the Beam

Cross-Section

The definitions of the beam and cross-section were specified in the previous lectures.

Some geometrical characteristics of the cross-section, such as the area and moments

of inertia, have a central role in the theoretical development of Mechanics of

Materials and are the main subject of this lecture.

3.1 Definitions

The beam cross-section is a plane area bounded by a closed curve B . For

mathematical convenience the Cartesian plane coordinate system oyz , as illustrated in

Figure 3.1, is defined.

Figure 3.1 Cross-Section Area

The total area of the cross-section is calculated as:

A

dAA (3.1)

The integral contained in equation (3.1) defines the summation of the differential

areas dA over the two defining variables y and z . The area is characterized by units of

length squared [L2], with the symbol [L] representing length.

Page 59: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 68 -

The first moments of the area A about the coordinate system axes oy and

oz are called the static moments. These are defined as:

A

y dAzS * (3.2)

A

z dAyS * (3.3)

The units of the static moments are [L3].

The geometric center of the cross-section is called the centroid. Equations (3.4) and

(3.5) are used to calculate the plane position ),( zy of the centroid C . The notation is

shown in Figure 3.2.

A

dAyAy ** (3.4)

A

dAzAz ** (3.5)

Figure 3.2 Centroid Location

Note: Analyzing the integrals contained in the equations (3.4) and (3.5), the

following important conclusions regarding the position of the centroid C may

be drawn:

(a) if the cross-section area possesses one axis of symmetry, the centroid C

lies on that axis;

(b) if the cross-section area possesses two axes of symmetry, the centroid C is

located at their intersection;

(c) if the cross-section area is symmetric about a point, the centroid C is

located at the location of that point.

These cases are illustrated in Figure 3.3.

Page 60: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 69 -

Figure 3.3 Types of Symmetry for Plane Area

(a) One Symmetry Axis, (b) Two Symmetry Axes and (c) Point Symmetry

The second moments of the cross-section area A about the coordinate system

axes oy and oz are called the moments of inertia and are defined as (see

Figure 3.4 for notation) :

dAzIA

y *2

(3.6)

dAyIA

z *2

(3.7)

Figure 3.4 Second Moments of Inertia Notation

The summation of the moment of inertia yI and zI is called the polar

moment of inertia and represents the second moment of inertia about the axis

ox normal to the cross-section. The polar moment of inertia is defined as:

zy

A

po IIdAzydAI *)(* 222 (3.8)

The unit of the second moments of inertia is [L4].

Note: The second moments of inertia are always positive values.

Another important geometric property is the product of inertia of the area A

also called in the Romanian technical literature the centrifugal moment of

inertia. The definition of the centrifugal moment of inertia is given in

equation (3.9):

Page 61: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 70 -

A

yz dAzyI ** (3.9)

The unit of the product of inertia is [L4].

Note: Contrary to the moments of inertia which are always positive values, the

product of inertia moment of inertia may have either a positive or negative

value. If the area A has an axis of symmetry the product moment of inertia

yzI .calculated for a coordinate system including that axis is zero.

3.2 Parallel-Axis Theorems for Moment of Inertia

The above described moments of inertia are usually calculated relative to a coordinate

system CC zCy anchored at the cross-section centroid C . A new translated coordinate

system Oyz , with axes oy and oz parallel to the centroidal axes CCy and CCz ,

respectively, is defined in Figure 3.5. The correspondence between the moments of

inertia relative to this new coordinate system Oyz and those calculated with respect to

the centroidal coordinate system CC zCy is studied in this section.

Figure 3.5 Parallel-Axis Theorems Notation

The position ),( zy of an arbitrary point located on the cross-section area A

expressed relative to the Oyz coordinate system is written as:

Cyyy (3.10)

Czzz (3.11)

where the distances y and z are the horizontal and vertical distances, respectively,

between the two coordinate systems considered.

The moments of inertia about axes Oy and Oz , yI and zI , are expressed in the

equations (3.6) and (3.7), respectively. Substitution of equations (3.10) and (3.11) into

Page 62: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 71 -

equations (3.6) and (3.7) yields the following new expressions for the moments of

inertia:

A A

CC

A

C

A

y dAzdAzzAzdAzzdAzI ****2**)(*2222 (3.12)

A A

CC

A

C

A

z dAydAyyAydAyydAyI ****2**)(*2222 (3.13)

Note that the integrals A

C dAz * and A

C dAy * become zero when the static moments

are calculated relatively to the centroidal axes.

The moments of inertia calculated about centroid axes are expressed as:

dAzIA

CyC*

2

(3.14)

dAyIA

CzC*

2

(3.15)

Substituting equations (3.14) and (3.15) into equations (3.12) and (3.13) final

expressions of the moments of inertial calculated about the axes of the translate

system Oyz are obtained as:

Cyy IAzI *2 (3.16)

czz IAyI *2 (3.17)

Equations (3.16) and (3.17) are called the parallel-axis theorem for moments of

inertia.

Note: Examination of equations (3.16) and (3.17) shows that the minimum values

for the moments of inertia are obtained when the axes Oy and Oz are

coincident with the centroidal axes Cy and Cz , respectively.

A similar approach can be used for the case of the polar moment of inertia poI defined

by equation (3.8). Substituting equations (3.10) and (3.11) into equation (3.8), the

polar moment of inertia about point O is obtained as:

CC zyzypo IAyIAzIII ** 22 (3.18)

Grouping the terms in the equation (3.18), the final expression may be written as:

CO pp IAI *2 (3.19)

Page 63: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 72 -

where 222 zy and dAzyI C

A

CpC*)(

22

Equation (3.19) represents the parallel-axis theorem for the polar moment of

inertia.

The parallel-axis theorem for the product of inertia is derived in a similar manner to

that for the moments of inertia. By substitution of equations (3.10) and (3.11) into

equation (3.9), the following expression is obtained:

A

zyc

A

C

A

CC

A

yz

CCIdAyzdAzyAzy

dAzzyydAzyI

******

*)(*)(**

(3.20)

Since the coordinate system CC yCz passes through the centroid of the cross-section

the integrals representing the static moments are zero and consequently, the equation

(3.20) reduces to:

CC zyyz IAzyI ** (3.21)

Equation (3.21) represents the parallel-axis theorem for the product of inertia.

3.3 Moment of Inertia about Inclined Axes

Consider Figure 3.6 where a new coordinate system '' zOxy is shown rotated by an

angle from the position of the original coordinate system oyz . The rotation angle

is positive when increasing in the trigonometric positive sense (counter-clockwise).

This convention corresponds to the right-hand rule previously used.

Figure 3.6 Axes Rotation Notation

Page 64: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 73 -

The position ),( '' zy of a current point located in the cross-section A relative to the

coordinate system '' zoxy can be written as:

sin*cos*' zyy (3.22)

cos*sin*' zyz (3.23)

Accordingly, with the definition equations (3.6) and (3.7) the moments of inertia in

the rotated coordinate system '' zOxy follow as:

dAzydAzIAA

y*)cos*sin*(*)( 22'

' (3.24)

dAzydAyIAA

z*)sin*cos*(*)( 22'

' (3.25)

After algebraic manipulations the moments of inertia 'yI and

'zI are obtained as:

22 sin*cos*sin**2cos*' zyzyyIIII (3.26)

22 cos*cos*sin**2sin*' zyzyzIIII (3.27)

Further, the equations (3.26) and (3.27) are expressed in an alternative form by

substitution of the double angle )*2( formulae:

)*2sin(*)*2cos(*22

' yz

zyzy

yI

IIIII

(3.28)

)*2sin(*)*2cos(*22

' yz

zyzy

zI

IIIII

(3.29)

The product of inertia ''zyI is defined as:

A

A

zy

dAzyzy

dAzyI

*)cos*sin*(*)sin*cos*(

*)(*)( ''''

(3.30)

After the algebraic manipulations, the equation (3.30) becomes:

)sin(cos*cos*sin*)( 22'' yzzyzy

IIII (3.31)

Further more, the equation (3.31) is re-written using the double angle )*2( as:

Page 65: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 74 -

)*2cos(*)*2sin(*2

'' yz

zy

zyI

III

(3.32)

If the derivatives of the moments of inertia shown in equations (3.28) and (3.29) are

taken relative to the double angle )*2( an interesting result is obtained:

''

'

)*2cos(*)*2sin(*2)2( zyyz

zyyII

III

(3.33)

''

'

)*2cos(*)*2sin(*2)2( zyyz

zyz IIIII

(3.34)

Note: The first derivative of the moment of inertia relative to the double

angle )*2( is the product of inertia.

The sum of equations (3.28) and (3.29) reveal the following important relationship:

zyzyIIII '' (3.35)

Note: Equation (3.35) indicates the invariance of the sum of the moments of inertia

with the rotation of the axes.

3.4 Principal Moments of Inertia

The moments of inertia 'yI and

'zI , expressed by equations (3.28) and (3.29), are

functions of the angle of the rotated coordinate system '' zOxy . The extreme values

(the maximum and minimum) of the moments of inertia 'yI and

'zI are called

principal moments of inertia. The corresponding values of the rotation angle

describe the principal axes of inertia. The principal axes of inertia passing through

the centroid of the cross-section area are called centroidal principal axes of inertia.

As known from Calculus, the condition for a real function to have an extreme point (a

maximum or minimum) is that the first derivative of the function be equal to zero at

that point. For the case of the moments of inertia 'yI and 'z

I , using equation (3.33) and

(3.34), the condition of extreme, is written as:

0)*2cos(*)*2sin(*2

)( 000''

yz

zy

zyI

III (3.36)

From equation (3.36) the value of the angle corresponding to the principal

directions is obtained as:

Page 66: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 75 -

zy

yz

II

I

*2)*2tan( 0 (3.37)

From the trigonometry, it is known that equation (3.37) has two solutions, )*2( 01

and )*2( 02 , related as shown in the equation (3.38):

0102 *2*2 (3.38)

Consequently, it is concluded that the principle directions are perpendicular to each

other:

20102

(3.39)

Using the following trigonometric identities

)*2(tan1

)*2tan()*2sin(

0

2

00

(3.40)

)*2(tan1

1)*2cos(

0

20

(3.41

and substituting them into equations (3.33) and (3.34) the final expressions for the

principle moments of inertia are obtained as:

2

2

max14

)(

2yz

zyzy

p IIIII

II

(3.42)

2

2

min24

)(

2yz

zyzy

p IIIII

II

(3.43)

Note: The invariance of the sum of the moments of inertia is also preserved for the

case of the principal moments of inertia. By addition of equations (3.42) and

(3.43) the invariance is proven:

zyppIIII '' 21

(3.44)

To identify which of the two angles, 01 or 02 , corresponds to the maximum moment

of inertia max1 II p the second derivative of the function 'yI , shown in equation

(3.28), is used. The condition for the point to be a maximum is:

Page 67: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 76 -

0)]*2sin(*)*2cos(*2

[)()2(

0002

2'

yz

zyyI

III (3.45)

The expression (3.45) is re-written as:

0)*2sin(*])*2tan(

1*

2[ 0

0

yz

zyI

II (3.46)

After the trigonometric manipulations and the usage of the equation (3.37) the

inequality (3.46) became:

0tan

*cos*]4

)([*2 0

0

22

2

yz

yz

zy

II

II (3.47)

The condition for the inequality (3.47) to hold is:

0tan 0

yzI

(3.48)

Note: Here, in order for inequality (3.48) to hold true, the sign of product of inertia

yzI must be opposite to that of the tangent of the angle 0 .

Practically, the angle corresponding to the direction of the maximum moment of

inertia is obtained by successively assigning to angle 0 the values 01 and 02 and

identifying which angle verifies the inequality (3.48).

3.5 Maximum Product Moment of Inertia

Consider the angle of the principal directions 0 established and the original

coordinate system oyz shown in Figure 3.6 rotated such that the y and z axes align

with the principal directions. Then, the following expressions hold:

1py II (3.49)

2pz II (3.50)

0yzI (3.51)

Substituting equations (3.49) through (3.51) into the equations (3.28), (3.29) and

(3.32) the expression for the moments of inertia as functions of the principal moments

of inertia are obtained:

Page 68: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 77 -

)*2cos(*22

2121'

pppp

y

IIIII

(3.52)

)*2cos(*22

2121'

pppp

z

IIIII

(3.53)

)*2sin(*2

21''

pp

zy

III

(3.54)

From equation (3.54), it is easy to see that the maximum value for the product of

inertia is obtained when:

1)*2sin( (3.55)

Then,

4

(3.56)

The maximum value of the product of inertia is obtained for an angle of rotation

4

measured in the counter-clockwise direction from the position of the principal

axes is expressed in equation (3.57).

2

21

max''

pp

zy

III

(3.57)

Substituting the principal moments of inertia given by equations (3.43) and (3.44) and

(3.42) into equation (3.57) a new expression for the max''zy

I is obtained:

2

2

max 4

)('' yz

zy

zyI

III

(3.58)

The corresponding moments of inertia are obtained by replacing 4

in equations

(3.52) and (3.53):

22

21''

zypp

zy

IIIIII

(3.59)

Page 69: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 78 -

3.6 Mohr’s Circle Representation of the Moments of

Inertia

A very interesting and useful relationship, shown in equation (3.60), is obtained by

manipulating the equations (3.28) and (3.32) in the following manner: (a) the equation

(3.28) is rearranged by moving in the left hand side the term 2

zy II and then

squaring both sided of the equation, (b) the equation (3.332) is squared and (c) adding

together the previous obtained equations

2

222

4

)(]

2[ ''' yz

zy

zy

zy

yI

III

III

(3.60)

The following notation is employed in the implementation of the equation (3.60):

'yI (3.61)

''zyI (3.62)

2

2

4

)(yz

zyI

IIR

(3.63)

2

zy

C

II (3.64)

Substitution of equations (3.61) through (3.64) into the equation (3.60) yields a new

form for equation (3.60)

222)( RC (3.65)

Geometrically, equation (3.65) represents the equation of a circle located in the

Oyz plane. The circle has center C located at )0,( C and radius R .

The coordinates of the intersection points, 1P and 2P , between the circle and the

horizontal axis O , are obtained by solving the algebraic system composed of

equation (3.65) and the equation of the axis )0( :

RCP 1 (3.66)

RCP 2 (3.67)

Page 70: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 79 -

Substituting equations (3.63) and (3.64) into equations (3.66) and (3.67) the position

of the intersection points 1P and 2P are expressed as shown in equations (3.68) and

(3.69) and are identified as the principal moments of inertia.

1

2

2

14

)(

2pyz

zyzy

P IIIIII

(3.68)

2

2

2

24

)(

2pyz

zyzy

P IIIIII

(3.69)

The graphical representation of the Mohr’s circle is depicted in Figure 3.7.

Figure 3.7 Morh’s Circle Representation

Note: Practically the Mohr’s circle is constructed using the following steps:

(a) The coordinates system O is drawn as shown in Figure 2.7. The horizontal

axis O represents the moments of inertia, while the vertical axis O

represents the product of inertia (note that the positive axis is drawn

upwards). The drawing should be done roughly to scale. The representation

considers that the following conditions are met: zy II , 0yI , 0zI and

0yzI ;

(b) Using the calculated values of the moments of inertia yI and zI and the product

of inertia yzI two points noted as Y and Z are placed on the drawing. The line

YZ intersects the horizontal axis in point C which represents the center of the

Mohr’s circle;

Page 71: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 80 -

(c) The distance CY is the radius of the circle. Using the radius CY and the

position of the center C the Mohr’s circle is constructed. The intersection

points, 1P and 2P , between the circle and the horizontal axis represent the

maximum and the minimum moments of inertia;

(d) The absolute value of the )*2tan( 0 can be calculated from the graph;

(e) The angle of the principal direction 1 is the angle measured in the counter-

clockwise direction between lines CY and CP1. To obtain the position of the

two principal directions corresponding to the cross-section system Oyz an

additional point Z’, the reflection of the point Z in reference to axis , has to

be constructed. The lines Z’P1 and Z’P2 represent the principal direction 1

(associated with the maximum moment of inertia) and 2 (associated with the

minimum moment of inertia), respectively. The two directions can then be

transcribed on the cross-section sketch.

3.7 Radii of Gyration

The square root of the ratio of the moment of inertia to the area is called the radius of

gyration and has the unit of [L].

The radii of gyration relative to the original coordinate system oyz are calculated as:

A

Ir

y

y (3.70)

A

Ir z

z (3.71)

For the case of the principal moments of inertia, the corresponding radii of gyration

are:

A

Ir

p

p

1

1 (3.72)

A

Ir

p

p

2

2 (3.73)

3.8 Examples

To clarify the theoretical aspects and the formulae derived in this lecture, two

examples are presented.

Page 72: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 81 -

3.8.1 Rectangle Cross-Section

A rectangular cross-section is shown in Figure 3.8. The rectangle is characterized by

two symmetry axes and consequently, the centroid C is located at their intersection.

The coordinate system used is the centroidal coordinate system shown in the Figure

3.8.

Figure 3.8 Rectangular Cross-Section

The following cross-sectional characteristics are calculated using the formulae

previously developed:

hbA *

A

y dAzSC

* =0

A

z dAySC

* =0

12

*)*(**

32/

2/

22 hbdzbzdAzI

h

hA

yC

12

*)*(**

32/

2/

22 hbdyhydAyI

b

bA

zC

0CC zyI

32

h

A

Ir C

C

y

y

32

b

A

Ir C

C

z

z

Page 73: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 82 -

It is shown thus, that for a rectangular cross-section the centroidal coordinate system

represents the principal axes of inertia )0( CC zyI .

If the moment of inertia about the axis coinciding with the lower edge of the rectangle

is required, using the notation shown in Figure 3.8(b), the parallel-axis theorem for

moments of inertia is employed:

3

*

12

*)*(*)

4(*)(

3322'

'

hbhbhb

hIAzI

Cyy

3.8.2 Composite Cross-Section

The L-shaped cross-section illustrated in Figure 3.9(a) is proposed for investigation.

The vertical and horizontal legs have a height of t6 and t9 , respectively, while

thickness t is uniform for the entire figure. The L-shaped cross-section can be

decomposed into two rectangular areas, 1A and 2A , representing the areasof the

individual legs of the cross-section. The distances of the centroids, 1C and 2C , of the

two rectangular areas are positioned relative to the coordinates system Oyz without

any difficulty as depicted in Figure 3.9(b).

Figure 3.9 L-Shaped Cross-Section

The individual area of each leg and total area of the L-shaped cross-section are

calculated as:

2

1 *6)(*)*6( tttA

2

2 *8)(*)*8( tttA

2

21 *14 tAAA

The position of the cross-section centroid C is obtained:

322

2211 *43)*8(*)*5()*6(*)2

(*** ttttt

AyAyAy

Page 74: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 83 -

tt

ty *07.3

*14

*432

3

322

2211 *22)*8(*)2

()*6(*)*3(*** ttt

ttAzAzAz

tt

tz *57.1

*14

*222

3

A new coordinate system aligned with the system Oyz and with the origin at the

centroid C of the entire cross-section is established as CC zCy . The following

calculations are performed with reference to this centroidal coordinate system. The

moments of inertia about the centroidal coordinate axes are calculated as:

44

4444223

223

2

222

2

111

*10.40*588

23576

*196

1800*

12

8*

196

2400*

12

216])*

14

22

2(*)*8(

12

)(*)*8([

])*14

22*3(*)*6(

12

)*6(*)([]*[]*[(

tt

tttttt

ttt

ttttt

zAIzAII yyyC

44

4444223

223

2

222

2

111

*60.112*588

66208

*196

5832*

12

512*

196

7776*

12

6])*

14

43*5(*)*8(

12

)(*)*8([

])*14

43

2(*)*6(

12

)*6(*)([]*[]*[(

tt

ttttttttt

tt

ttt

yAIyAII zzzC

4

442

22

2222211111

*57.38196

7560

*196

3240*

196

4320)]*

14

22

2(*)*

14

43*5(*)*8(0[

)]*14

22*3(*)*

14

43

2(*)*6(0[]**[]**[(

t

tttt

ttt

tttt

tzyAIzyAII zyzyzy CC

The principal moments of inertia are obtained as:

424244

442

2

1

*28.129)*57.38(4

)*60.112*10.40(

2

*60.112*10.40

4

)(

2

tttt

ttI

IIIII yz

zyzy

p

Page 75: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 84 -

424244

442

2

2

*42.23)*57.38(4

)*60.112*10.40(

2

*60.112*10.40

4

)(

2

tttt

ttI

IIIII yz

zyzy

p

The angle of the principal direction of inertia is calculated as:

064.1

2

)*60.112*10.40(

)*57.38(

2

)()*2tan(

44

4

0

tt

t

II

I

zy

yz

78.46*2 0

39.230

Using the test contained in the equation (3.48) to determine if the rotation

angle 39.230 is the angle of the principal direction results in the following:

0)*57.38(

)4325.0(

)*57.38(

)39.23tan()tan(44

0

ttI yz

Consequently, the angle 39.230 is the angle of the direction of the minimum

moment of inertia and 39.2302 , while the complementary angle 61.669039.2301 represents the direction of the maximum moment of

inertia.

The angles 01 and 02 are illustrated in Figure 3.10.

Figure 3.10 L-Shaped Cross-Section Principal Directions

The radii of gyrations are obtained as:

tt

t

A

Ir C

C

y

y *69.1*14

*10.402

4

Page 76: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 85 -

tt

t

A

Ir C

C

z

z *84.2*14

*6.1122

4

tt

t

A

Ir

p

p *04.3*14

*28.1292

41

1

tt

t

A

Ir

p

p *29.1*14

*42.232

42

2

The construction of the Mohr’s circle is conducted as explained in Section 3.6. Using

the moments and the product of inertia calculated above the following values are

determined:

4

24244

2

2

*93.52

)*57.38(4

)*60.112*10.40(

4

)(

t

ttt

III

R yz

zy

444

*35.762

*60.112*10.40

2t

ttIIy

zy

C

4

1 *3.129 tRyI Cp

4

2 *4.23 tRyI Cp

Figure 3.11 L-Shaped Cross-Section Mohr’s Circle

Page 77: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 86 -

LECTURE 4

Equilibrium of the Plane Linear Members

In this lecture the variations of the internal resultants acting on the cross-section of a

linear member, are examined.

The definition of the plane linear member and its cross-section geometrical

characteristics has been discussed in the previous lectures.

For this discussion, the local coordinate system attached to the plane linear beam is a

right-hand coordinate system having the ox axis aligned along the length of the beam

with origin at the leftmost located point.

4.1 Type of Beams, Loads and Reactions

For better understanding of the following theoretical derivations, a number of

schematic diagrams representing loaded beams commonly found in structural

engineering are depicted in Figure 4.1.

The movement of a point located on the left axis of the beam resulting from the

application of forces in the oxy plane is completely determined by three generalized

components, explicitly, by two in-plane displacements and a rotation. The two

displacements are chosen, for convenience, in the directions ox and oy of the

coordinates system attached to the beam, while the rotation is described by the

angular motion about an axis parallel to the oz axis located at the point of interest

(normal to the plane of the definition plane). The forces and moments resulting from

the constraints induced by the existence of supports, accordingly with the Newton’s

First Law, are called reactions. Replacing the supports by their corresponding

reaction forces, the free-body diagram of the beam can be drawn. For the plane

linear beam the reactions at supports and connections were illustrated in Tables 2.1

and 2.2 contained in Lecture 2. In the figures, a reaction force is distinguished from an

applied force (load) by a slash mark in the shaft of the arrow symbolically describing

the force.

In the technical language employed by structural engineers, beams are identified by

the manner in which they are supported. The beam illustrated in Figure 4.1.a is called

Page 78: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 86 -

a simply supported beam. This beam has one end, point A, constrained only in the

vertical direction by a roller support, while the other end, point B, is constrained in

both vertical and horizontal directions by a pined support or clamped support. The

reactions corresponding to the particular constraints are shown in the figure.

Figure 4.1 Examples of Beams

(a) Simple Supported Beam, (b) Cantilevered Beam, (c) Propped Cantilevered

Beam and (d) Two-span Continuous Beam

The beam depicted in Figure 4.1.b is called a cantilevered beam and has only one

end, point C, constrained against movement in both, horizontal and vertical, directions

and against the rotation around an axis parallel to the oz axis. This type of support is

called a fixed support and the corresponding reactions are is composed of two forces

and a moment.

Page 79: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 87 -

More complicated beam configurations are depicted in Figures 4.1.c and 4.1.d. The

beam shown in Figure 4.1.c is called a propped cantilever beam. The end point A

has a fixed support, while the point B has a roller support attached. In Figure 4.1.d a

simply supported two-span continuous beam is represented. At point A, a pinned

support is located, while at points B and D, roller supports are attached. The

corresponding reactions are shown.

From the above description it can be concluded that in the analysis of the single-body

plane beams, three types of supports are commonly used:

1. the roller support prevents the displacement in the direction normal to the

rolling plane and is replaced in the free-body diagram by a corresponding

reaction represented by a concentrated force as shown for case1 of Table 2.1.

A similar case is that of a supporting cable illustrated in case 2 of Table 2.1;

2. the pinned support prevents all translational displacement in the beam

description plane and is replaced in the free-body diagram by a reaction,

commonly represented by two orthogonal concentrated force components as

depicted in case 3 of Table 2.1;

3. the fixed support completely prevents all translational displacement and

rotation in the beam definition plane and is replaced in the free-body diagram

by a set of reactions commonly represented by two orthogonal concentrated

force components and a concentrated moment as depicted in case 4 of Table

2.1.

If the beam is composed of a multi-body system, the decomposition of the system into

multiple single-body systems is required. At the connections between the individual

single-body systems the reactions, as shown in Table 2.2, are introduced in the free-

body diagram of each system .

The external forces (loads) applied to the plane beams are classified as:

1. distributed forces;

2. concentrated forces;

3. concentrated moments (couples).

The most commonly used distributed transverse forces are uniform and linearly

distributed. Examples of distributed transversal forces are shown in Figures 4.1.a,

4.1.c and 4.1.d. Concentrated force and moment are illustrated in Figures 4.1.b and

4.1.b.

Page 80: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 88 -

4.2 Equilibrium of Beams Using the Free-Body Diagrams

The beam depicted in Figure 4.1.a is used to explain the application of the free-body

diagram concept in the case of the plane linear member.

The beam is first conceptually freed of constraints and then, the corresponding

reactions, as described in the previous section, are introduced. As indicated before, the

roller support at point A of Figure 4.1.a is replaced by a vertical concentrated force,

while the pinned support at point B is replaced by horizontal and vertical forces. At

this point, the schematic diagram representing contains only the externally applied

forces and reaction forces and is identified as the free-body diagram. Using this

diagram, the next step is to write the equations of equilibrium and solve them for the

unknown reaction forces. From the six equilibrium equations, (2.3) though (2.8), only

three can be used, the other three are identically zero. If the beam definition plane

oxy also represents the plane of action for exterior loads, and thus, the equilibrium

equations are represented by equations (2.9) through (2.11).

If the number of unknown reactions is equal to the number of independent

equilibrium equations the beam is a statically determinate beam. If the number of

unknown reactions is larger than three the beam is called statically indeterminate

beam. Only the statically determinate beams are treated in this textbook. The beams

shown in Figures 4.1.a and 4.1.b are statically determinate beams, while Figures 4.1.c

and 4.1.d illustrate statically indeterminate beams.

After the reactions have been calculated, the beam is conceptually sectioned, as

illustrated in Figure 4.2, by a cutting-plane coincidental with a cross-section located at

distance Cx from the origin point A. In Lecture 2, this separation operation was called

the method of sections. To maintain intact the equilibrium of the two separated parts,

the corresponding internal resultants are introduced and two free-body diagrams, AC

and CB, are obtained as illustrated in Figure 4.2. In the absence of the externally

applied axial forces, only two cross-sectional internal resultants, the transverse shear

force CV and the bending moment CM , are necessary to maintain the equilibrium of

the two parts. The indices y and z , used in the previous lectures to indicate the oy

and oz axes along which the shear force and moment are acting have been omitted

for clarity.

Figure 4.2 Application of the Method of Sections

Page 81: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 89 -

If an elementary element of the beam is separated around a particular cross-section, as

shown in Figure 4.3, the sign convention established has a physical meaning, directly

related with the deformation of the element.

Note: Although it would be more consistent to define the positive shear force V in

the positive direction of the axis oy , however, the sign convention adopted

above is universally employed and is directly connected with the physical

significance explained.

Figure 4.3 Sign Convention for Internal Resultants

(a) Shear Force, (b) Bending Moment and (c) Shear Force and Bending Moment

4.3 Differential Relations between Loads and Cross-

Section Internal Resultants

The free-body diagram concept will be extended to write the equilibrium of the

elementary volume around a point of the beam. The free-body diagram of the

elementary volume is shown in Figure 4.4.

Definition 4.1

The sign convention for the cross-section internal resultants is stated as:

1. the positive shear force,V , acts in the negative direction of the oy axis at

the face x of the cross-section;

2. the positive bending moment, M , makes the face y of the beam

concave.

Page 82: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 90 -

Figure 4.4 Equilibrium of Elementary Beam Volume

The external forces acting on the elementary volume are considered to be uniformly

distributed, their variation being small enough to be neglected over the short length

x quantity. These are represented by two distributed forces, )(xpn and )(xpt , and a

distributed moment )(xm , all acting in the positive sense relative to the respective

axes.

The equilibrium equations are:

0 xF 0* xpFFF t (4.1)

0 yF 0*)( xpVVV n (4.2)

0C

zM

0)*()(

)2

**()*(

xmMM

xxpMxV n (4.3)

Definition 4.2

The sign convention for the external loads is:

1. the positive tangential distributed and concentrated loads on the beam

longitudinal axis ox act in the positive direction of ox axis;

2. the positive normal distributed and concentrated loads on the beam

longitudinal axis ox act in the positive direction of oy axis;

3. the positive distributed and concentrated bending moments act in the

positive direction of the oz axis according to the right-hand rule.

Page 83: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 91 -

After algebraic manipulations and neglecting the second-order terms equations (4.1)

through (4.3) yield a new format as follows:

)(xpx

Ft

(4.4)

)(xpx

Vn

(4.5)

)()( xmxVx

M

(4.6)

where x is the position of the cross-section.

If the length of the elementary volume x tends to zero, the volume becomes

infinitesimal and the following differential equations are obtained:

)(lim 0 xpdx

dF

x

Ftx

(4.7)

)(lim 0 xpdx

dV

x

Vnx

(4.8)

)()(lim 0 xmxVdx

dM

x

Mx

(4.9)

Equations (4.7) through (4.9) are called the differential relation between the loads

and the cross-sectional internal resultants. If the differential equations (4.7), (4.8)

and (4.9) are integrated the following expressions are obtained:

1*)()( CdxxpxF t (4.10)

2*)()( CdxxpxV n (4.11)

3*)(*)()( CdxxmdxxVxM (4.12)

The integration constants 1C , 2C and 3C are identified using the presumed known

values, boundary values of the cross-section internal resultants at the beginning of the

integration interval located at 0x :

01 FC (4.13)

Page 84: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 92 -

(4.14)

03 MC (4.15)

Introducing equations (4.13) through (4.15) into equations (4.10) through (4.1) the

following expressions are obtained:

dxxpFxF t *)()( 0 (4.16)

dxxpVxV n *)()( 0 (4.17)

dxxmdxxVMxM *)(*)()( 0 (4.18)

These equations are important and are extensively used in determination of the

variation of the cross-sectional internal resultants for beams subjected to various

loading conditions.

Note: From equations (4.16) through (4.18) important practical conclusions are

derived:

(a) the loading functions )(xpn , )(xpt and )(xm must be continuous functions on

finite-intervals for the integration to be possible;

(b) the cross-sections where concentrated forces and moments are acting represent

discontinuity points. Those cases are treated as limiting cases of the

expressions obtained above;

(c) the integrals in equations (4.16) and (4.17) represent the area contained under

the loading curve over the length of the integration interval. The shear force

related integral in equation (4.18) represents the area contained under the shear

force curve over the length of the integration interval;

(d) values representing the boundary conditions of the cross-section internal

resultants at the beginning of the integration interval must be known in order

to evaluate the cross-sectional internal resultants on that interval.

4.4 Shear Force and Bending Moment Diagrams

The plot of the shear force )(xV and bending moment )(xM variation along the

longitudinal axis ox of the entire beam are called cross-section resultants diagrams

Page 85: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 93 -

or shear force and bending moment diagrams. They play a very important role in

design and verification of beams. The shear and bending moment diagrams can be

constructed and plotted using a combination of the two methods described in the

previous section: (a) the equilibrium method (method of sections) and (b) the

differential relations.

The application of the two methods is preceded by creation of the free-body diagram

of the beam and calculation of the reaction forces.

The first method, the equilibrium method, requires a relatively large number of

sections to be created and the variation of the resultants between them to be

considered a priori linear. This is an approximation, but if the sections are relatively

closely spaced the induced error is negligible for practical purposes.

Employing the second method, the integration of the differential equations, requires

the beam loading to be decomposed into several intervals of continuity. The variation

of the shear force and bending moment obtained are exact.

Both methods have advantages and disadvantages and for complicated systems of

beams or loading, they are used together in a complementary manner. To become

proficient in the usage of the methods requires practice. A few examples are shown in

the following section.

4.5 Examples of Shear and Moment Diagrams for

Statically Determined Beams

The following examples illustrate the practical application of methods, the method of

sections and the differential relations method, for constructing shear and moment

diagrams.

4.5.1. Simply-Supported Beam Loaded with a Concentrated Vertical Force

The notation, geometry and loading are shown in Figure 4.5.a, while the

corresponding free-body diagram is depicted in Figure 4.5.b.

The roller support located at point A is replaced by a vertical reaction, while the

pinned support attached at point C is replaced by vertical and horizontal reactions. In

the absence of any horizontal load, the horizontal reaction at point B is null and, for

clarity, is not shown. Consequently, only two unknown vertical reactions remain.

Page 86: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 94 -

Figure 4.5 Simply-Supported Beam Loaded with a Concentrated Vertical Force

(a) Geometry and Load and (b) Free-Body Diagram

Because there is no horizontal load, only two equations of equilibrium can be written.

0 yF 0 CA VPV (4.19)

0C

zM 0)(** aLPLVA (4.20)

Solving equations (4.19) and (4.20) the reaction forces are obtained:

L

aLPVA

)(* (4.21)

L

aP

L

aLPPVPV AC

*)(*

(4.22)

From examination of the free-body diagram represented in Figure 4.5(b) it is seen that

the loading has two intervals of continuity, AB and BC, where the load is zero

(mathematically it may be considered as constant with a zero value). Point B

represents the point where the existence of the concentrated load P creates a

discontinuity. The free-body diagram, with the calculated reactions, is used to

calculate the shear-force and bending moment diagrams.

The shear diagram starts at point A, where the concentrated reaction AV acts. Being a

concentrated force, the point A is a discontinuity point. A similar argument is made

for point C where the other concentrated reaction exists. The concentrated force is

treated using the sections at the right and at the left. Taking sections in different

positions of x and writing the equilibrium on the vertical axis oy the shear diagram is

calculated as follows:

at point A

for 0x 0)( xV (4.23)

Page 87: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 95 -

for 0x L

aLPVxV A

)(*)(

(4.24)

at point B

for ax L

aLPVxV A

)(*)(

(4.25)

for ax

L

aP

PL

aLPPVxV A

*

)(*)(

(4.26)

at point C

for Lx

L

aP

PL

aLPPVxV A

*

)(*)(

(4.27)

for Lx 0

*)(*

)(

L

aPP

L

aLP

VPVxV CA

(4.28)

The differential relations previously obtained are used to define the variation of the

shear force in-between the points describing the continuity intervals A+B

- and B

+C

- of

the loading. From equation (4.17) it is observed that the shear force, in the absence of

transverse load, has a constant value equal to that calculated at the beginning of the

interval.

Note: Concentrated forces may be treated as “continuity” interval under the limiting

condition where the interval length tends to zero.

The moment diagram is calculated in a similar manner to the shear diagram only

replacing the vertical equation of equilibrium with the moment equation of

equilibrium. The section method is applied as follows:

at point A

for 0x 0)( xM (4.29)

at point B

for ax aL

aLPaVxM A *

)(**)(

(4.30)

Page 88: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 96 -

at point C

for Lx 0)(**

)(*

)(**)(

aLPLL

aLP

aLPLVxM A

(4.31)

Using the differential relation (4.9) leads to the conclusion that the variation of the

moment is linear in both intervals. The shear and moment diagrams are plotted in

Figure 4.6.

Figure 4.6 Shear and Moment Diagrams

(a) Free-body diagram, (b) Shear diagram and (c) Moment diagram

4.5.2 Simply-Supported Beam Loaded with a Concentrated Moment

The geometry and the notation employed are depicted in Figure 4.7. The free-body

diagram is depicted in Figure 4.7.b.

Page 89: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 97 -

Figure 4.7 Simply-Supported Beam Loaded with a Concentrated Moment

(a) Geometry and (b) Free-Body Diagram

The roller support located at point A is replaced by a vertical reaction, while the

pinned support attached at point C is replaced by vertical and horizontal reactions. In

the absence of any horizontal load the horizontal reaction at point B is null and, for

clarity, is not shown. Consequently, only two unknown vertical reactions remained

As before, with no horizontal load, the following equations of equilibrium are written

using the free-body diagram as a guide:

0 yF 0 CA VV (4.32)

0C

zM - 0* 0 MLVA (4.33)

Solving equations (4.32) and (4.33) the reaction forces are obtained:

L

MVA

0 (4.34)

L

MVV AC

0 (4.35)

Analyzing the free-body diagram represented in Figure 4.7.b it becomes obvious that

the loading has two intervals of continuity, AB and BC, where the load is nil. The

point B represents the point where the existence of the concentrated moment

0M creates a point of discontinuity. The free-body diagram shown in Figure 4.7.b,

along with the calculated reactions, is used to construct the shear-force and bending

moment diagrams.

The shear diagram starts at point A, where the concentrated reaction AV acts. Because

the existence of the concentrated reaction force acting at point A, the point becomes a

discontinuity point. A similar argument can be made for the point C where the other

concentrated force reaction acts. The concentrated reaction force is treated using

sections to the left and the right of the point of application. Taking sections at

different positions of the ox axis and writing the vertical equilibrium equations, the

shear diagram is calculated as follows:

Page 90: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 98 -

at point A

for 0x 0)( xV (4.36)

for 0x L

MVxV A

0)( (4.37)

at point B

for ax L

MVxV A

0)( (4.38)

for ax L

MVxV A

0)( (4.39)

at point C

for Lx L

MVxV A

0)( (4.40)

for Lx 0)( CA VVxV (4.41)

The differential relations previously obtained are used to define the variation of the

shear force in-between the points describing the continuity intervals A+B

- and B

+C

- of

the loading. From equation (4.17) it is evident that the shear force, in the absence of

distributed vertical load )(xpn , has a constant value equal with that calculated at the

beginning of the interval.

Note: The concentrated moment may be treated as a “continuity” interval where in

the limit the interval length tends to zero. The concentrated moment does not

create a discontinuity in the shear diagram.

The moment diagram is calculated in a similar manner as the shear diagram except

that the vertical equation of equilibrium is replaced with the moment equation of

equilibrium. The method of section is applied as follows:

at point A

for 0x 0)( xM (4.42)

at point B

for ax aL

MaVxM A **)( 0 (4.43)

Page 91: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 99 -

for ax

L

aLMMa

L

M

MaVxM A

)(*

*)(

0

0

0

0

(4.44)

at point C

for Lx 0**)( 0

0

0 MLL

MMLVxM A (4.45)

From examining of the differential relation (4.9) it is seen that the variation of the

moment is linear in both intervals. The resulting shear and moment diagrams are

plotted in Figure 4.8.

Figure 4.8 Shear Force and Moment Diagrams

(a) Free-body diagram, (b) Shear diagram and (c) Moment diagram

4.5.3 Beam with Overhang

The overhang beam geometry is shown in Figure 4.9.a. A uniformly distributed load

of 8kN/m is acting between the supports A and B and a concentrated force of 16 kN

acts at the tip of the cantilever as shown.

The pinned support at point A and a roller support at point B are replaced by the

corresponding reaction forces. In the absence of any horizontal force, the horizontal

Page 92: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 100 -

reaction at the pinned support A is zero. The free-body diagram is shown in Figure

4.9.b.

Figure 4.9 Beam with Overhang

(a) Geometry and (b) Free-Body Diagram

As previously explained the first step in solving the overhang beam is the

determination of the reactions. The two remaining equilibrium equations are:

0 yF 0164*8 CA VV (4.46)

0B

zM 0)2*16()2*4*8()4*( AV (4.47)

Solving the algebraic system of equations (4.44) and (4.45) the reactions are found as:

8AV kN (4.48)

4048 AB VV kN (4.49)

Examination of the loading function reveals the following two continuity intervals: (a)

from support A to support B where the uniformly distributed load acts and (b) from

support B to the tip of the cantilever C where the load is null. The discontinuity points

are easily identified as the points A, B and C. The calculation follows the general

methodology employed in the previous two examples. The free-body diagram and the

shear and moment diagrams are contained in Figure 4.10. Using the method of

sections, the values of the shear force are calculated as:

at point A

for 0x 0)( xV (4.50)

for 0x 8)( AVxV kN (4.51)

at point B

for 4x 24)4*8()( AVxV kN (4.52)

Page 93: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 101 -

for 4x 1640)4*8()( AVxV kN (4.53)

at point C

for 6x 1640)4*8()( AVxV kN (4.54)

for 6x 01640)4*8()( AVxV kN (4.55)

Using the differential relations (4.8) for the continuity interval AB it is seen that the

shear diagram has a linear variation, while in interval BC the shear is constant.

The moment diagram is calculated in a similar manner by applying the method of

sections as follows:

at point A

for 0x 0)( xM (4.56)

at point B

for 4x 32)2*4*8()4*()( AVxM mkN * (4.57)

at point C

for 6x mkN

VxM A

* 0)2*40(

)4*4*8()6*()(

(4.58)

From the differential relation (4.9) and the previous observations concerning the

variation of shear, it became obvious that the variation of the moment is parabolic in

the interval AB and linear in the interval BC. The maximum local moment in the

interval AB is located where the corresponding shear force is zero. An equation

similar to (4.52) is written for a general location on the interval A to B and then, used

to calculate the position of the local maximum moment:

0)*8( max xVA (4.59)

18

max AVx m (4.60)

The maximum local moment is:

4)2

**8(* max

maxmaxmax xxxVM A mkN * (4.61)

Page 94: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 102 -

Using an equation similar to (4.57) a calculation is performed to find the location

where the moment on the interval AB becomes zero:

0)2

**8(*)( x

xxVxM A (4.62)

28

*2 AV

x m (4.63)

Figure 4.10 Shear and Moment Diagrams for Beam with Overhang

(a) Free-Body Diagram, (b) Shear Force and (c) Moment Diagram

Page 95: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 103 -

LECTURE 5

Axial Deformation

Axial deformation has been experimentally introduced in Lecture 2, when the tension

test of a structural steel specimen was described. The case in point of this lecture is a

systematic description of all aspects pertinent to axial deformation.

5.1 Basic Theory of Axial Deformation

The three aspects discussed during Lecture 2, equilibrium equations, strain-

displacement equations and the constitutive relation, are applied to the case of axial

deformation. A local coordinate system oxyz , having ox axis coincident with the

longitudinal axis of the member, is employed.

5.1.1 Strain-Displacement Equation

A direct consequence of part (b) of the definition 5.1 is the fact that axial deformation

depends only on the variable x , which represents the position of a particular cross-

section on the longitudinal axis of the member.

The physical phenomenon described by definition 5.1 is illustrated in Figure 5.1,

where the undeformed and deformed conditions of the linear member are presented.

Definition 5.1

A plane linear member, when subjected to exterior loads and/or change of

temperature, undergoes an axial deformation if after the deformation:

(a) the axis of the member remains straight;

(b) the cross-sections remain plane, perpendicular to the longitudinal axis

of the beam and do not rotate about the same longitudinal axis after the

deformation.

Page 96: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 112 -

Adjacent cross-sections A and B originally located at distance x from each other, as

shown in Figure 5.1.a, are found after the deformation to be located at distance *x .

The change in the position of the two cross-sections are described by their

displacements )(xu and )( xxu , respectively.

Figure 5.1 Geometrical Aspects of the Axial Deformation

(a) Undeformed Member and (b) Deformed Member

The extensional strain )(xx is expressed as:

dx

du

x

xuxxu

x

xxx xxx

]

)()([lim)(lim)( 0

*

0 (5.1)

Equation (5.1) is called the strain-displacement equation. The cross-section

distribution of the elongation strain )(xx is shown in Figure 5.2. The elongation

strain is a function only of the cross-section position described by the variable x .

Figure 5.2 Extensional Strain Distribution

The rest of the generalized strain tensor components are:

)(*)( xx xy (5.2)

)(*)( xx xz (5.3)

0 yzxzxy (5.4)

Page 97: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 113 -

From equations (5.2) through (5.4) is evident that a transversal reduction of the cross-

section takes place concomitant with the axial deformation.

The displacement )(xu pertinent to a particular cross-section is obtained by

integration from the equation (5.1):

dxxuxux

x *)()(0

0 (5.5)

where )0(0 xuu is the displacement at the beginning at the integration interval.

Consequently, the total elongation of the member is calculated from equation (5.5) as:

L

x dxxuLue0

*)()0()( (5.6)

where L is the total length of the bar.

5.1.2 Constitutive Equation

The constitutive equation reflects, as was describe in Lecture 2, the relation between

the stress and the strain. If the linear elastic material behavior is considered, the

relation between the normal stress and extension strain )(xx for the case of the axial

deformation is written:

)(*),,(),,( xzyxEzyx xx (5.7)

Equation (5.7) represents the application of Hook’s Law for the case of axial

deformation. The material constant E , the modulus of elasticity, has a value unique to

each specific material and is obtained from tensile tests. The cross-section of the bar

is a small surface and the variation of the modulus of elasticity is negligible on this

surface. In this case the constitutive equation (5.7) is expressed as:

)(*)()( xxEx xx (5.8)

Equation (5.8) implies that the normal stress )(xx varies only along the length of the

member, but has a constant value on the entire cross-section. The representation of the

normal stress )(xx is shown in Figure 5.3.

The rest of the stress tensor components are zero:

0)()( xx zy (5.9)

0 xyxyxy (5.10)

Page 98: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 114 -

Figure 5.3 Cross-Section Normal Stress Distribution

5.1.3 Cross-Section Stress Resultants

Considering the stress distribution represented by equation (5.8) through (5.10) the

cross-section stress resultants are obtained as:

)(*)( *)( *)()( xAxdAxdAxxF x

A

x

A

x (5.11)

yx

A

x

A

xy SxdAzxdAxzxM *)( **)( *)(*)( (5.12)

zx

A

x

A

xz SxdAyxdAxyxM *)( **)(*)(*)( (5.13)

The relation between the normal stress )(xx and the cross-section resultants )(xF ,

)(xM y and )(xM z is derived using the notation shown in Figure 5.4.

Figure 5.4 Normal Stress and Stress Resultants

(a) Stress Resultants and (b) Normal Stress

Page 99: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 115 -

If the axes oy and oz of the coordinate system intersect such that the x axis passes

through the cross-section centroid, the static moments yS and zS are zero and the

axial force )(xF remains the only non-zero stress resultant. Then, from equation

(5.11) the normal stress x is calculated as:

)(

)()(

xA

xFxx (5.14)

Consequently, it is to be concluded that a beam made from a linear elastic material

undergoes an axial deformation if the axial force passes through the cross-section

centroid.

5.1.4 Equilibrium Equation

The equilibrium equation pertinent to the case of axial deformation was derived in

Section 4.3 of Lecture 4. The detailed derivation is not repeated and only the final

result expressed as the differential relation (4.7) is employed. Considering the exterior

loading and stress resultants shown in Figure 5.5, acting on an infinitesimal volume of

length x separated from the body of the beam, the following differential equation is

written:

)()(

xpdx

xdFt (5.13)

where )(xpt is the distributed loading parallel to the beam longitudinal axis.

Figure 5.5 Infinitesimal Volume Equilibrium

If the axial stress resultant )(xF is orientated as shown in Figure 5.5 and tends to

lengthen the segment the internal force is called tension. The corresponding normal

stress x is called tension stress. In contrast, if the stress resultant force )(xF is

orientated to shorten or compress the segment the corresponding internal force and

normal stress are called compression and compression stress, respectively.

Integrating equation (5.15) the stress resultant force )(xF is calculated:

Page 100: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 116 -

x

t dpFxF0

0 *)()( (5.16)

where )0(0 xFF is the value of the axial force at the origin of the integration

interval.

5.1.5 Thermal Effects on Axial Deformation

The equations obtained in the previous sections are derived considering only the

exterior load action and neglecting the change in temperature. In this section the

effects of the thermal change are introduced.

During Lecture 2, the thermal strain along the axis Ox was introduced as:

TT

x * (5.17)

where is the thermal expansion coefficient and T is the change in the member

temperature.

The total elongation strain is the sum of the elongation strain induced by the exterior

load action and thermal effects and is expressed as:

)(*)()(

)()( xTx

xE

xx xT

xxxx

(5.18)

By substitution of equation (5.14) into (5.18) the elongation strain is obtained as:

)(*)()(*)(

)()( xTx

xAxE

xFxx (5.19)

Then, using equation (5.19) in equation (5.6) the total elongation of the member is

written as:

LLL

x dxxTxdxxAxE

xFdxxe

000*)(*)(*

)(*)(

)(*)( (5.20)

5.2 Uniform-Axial Deformation

A special case of axial deformation frequently encountered in structural engineering is

the case of uniform-axial deformation shown in Figure 5.6. Systems composed of

many members subjected to uniform-axial deformation are also commonly used in

structural practice. A typical example is the plane truss, where each individual

member is subjected to uniform-axial deformation under the action of to tension or

compression forces.

Page 101: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 117 -

The formulation related with the definition of the uniform-axial member and its

application in the investigation of the statically determinate and indeterminate

structures is presented in this section.

5.2.1 Members Subjected to Uniform-Axial Deformation

Definition 5.2 is mathematically transcribed as:

0)( AxA (5.21)

0)( ExE (5.22)

0)( FxF (5.23)

.

Figure 5.6 Member Exhibiting Uniform Axial-Deformation

Note: Equation (5.23) implies the absence of the distributed load )(xpt in equation

(5.15)

Rewriting equations (5.6), (5.8) and (5.14) obtained in the previous section for the

case in point of a member with uniform axial-deformation the following equations are

obtained:

Definition 5.2

The uniform axial-deformation element is a linear member characterized by:

(a) a constant area along the entire length of the member;

(b) is made of a homogeneous elastic material;

(c) is subjected to a constant axial force .F

Page 102: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 118 -

0

0

0

)(

)()(

A

F

xA

xFxx (5.24)

0

00

0

*)(

)()(

AE

F

xE

xx x

x (5.25)

00

0

00 *

***)(

AE

LFLdxxe

L

x (5.26)

In the computer applications equation (5.26) takes the following form:

ekF * (5.27)

where L

AEk 00 * is called the axial stiffness coefficient.

The axial stiffness coefficient represents the force applied to the member ends when

the elongation is equal with the unit length. The unit is [F/L]. The reciprocal value of

the axial stiffness is called the axial flexibility coefficient. The flexibility coefficient

is calculated:

00 *

1

AE

L

kf (5.28)

Using the flexibility coefficient expression equation (5.28) is cast in a new format:

Ffe * (5.29)

If the change in temperature is also constant along the entire length of the member the

formula (5.20) are amended as follows:

LTAE

LFe **

*

* (5.30)

Using the flexibility coefficient f the elongation is calculated as:

LTFfe *** (5.31)

Rewriting the equation (5.31) the force F is obtained:

]**[* LTekF (5.32)

5.2.2 Statically Determinate Structure

An example of a statically determinate structure is shown in Figure 5.7. A rigid,

weightless beam AC is supported at end A by a column and at end C by a vertical rod

Page 103: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 119 -

CD. The rod is attached at point D to a “leveling jack”, a component which permits a

limited vertical movement and has the mission to keep the beam AC leveled. The

beam AC is loaded with a vertical force P located at distance aL from the left end A.

To calculate the axial forces acting on the column and rod the system is decomposed

into three components: the beam, the column and the rod.

Figure 5.7 Statically Determinate Structure

The free-body diagram of the beam is depicted in Figure 5.8 and is used to calculate

the reaction forces 1F and 2F . The connection between the column and the beam

requires in general, two reaction forces. In the absence of any horizontal force the

horizontal component is null and is not shown

Figure 5.8 Free-Body Diagram

The reaction forces can be calculated using the following equilibrium equations:

0 yF 021 FPF (5.33)

0C

zM 0)1(***1 aLPLF (5.34)

Solving equations (5.33) and (5.34) the reaction forces are obtained:

)1(*)1(**

1 aPL

aLPF

Compression in column (5.35)

Page 104: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 120 -

PaPaPPFF *)1(*12 Tension in rod (5.36)

Note: The column and the rod are loaded with forces having opposite directions than

the reaction forces calculated in equations (5.35) and (5.36). This became

obvious if one attempt to draw the free-body diagram for the rod and column.

From equations (5.35) and (5.36) it is evident that the column and the rod are loaded

with compression force 1F and tension force 2F , respectively. These two elements are

characterized, using definition 5.2, as members with uniform axial-deformation.

Using the formulae (5.24) through (5.26) the following related column values are

calculated:

col

col

xA

Fx 1)( (5.37)

colcolcol

col

xcol

xAE

F

E

xx

*

)()( 1

(5.38)

colcol

colAE

LFe

*

* 11 (5.39)

Note: The force 1F is compression and the column is shortened by the loading

action.

Related quantities for the rod CD are similarly calculated:

rod

rod

xA

Fx 2)( (5.40)

rodrodrod

rod

xrod

xAE

F

E

xx

*

)()( 2

(5.41)

rodrod

rodAE

LFe

*

* 22 (5.42)

Note: The force 2F is tension and the rod is elongated by the loading action.

For the beam to stay in perfect horizontal balance the displacement at point A and C

should be equal and manifest in the same direction:

CA uu (5.41)

Page 105: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 121 -

The expressions of the displacements at points A and C are calculated using the

formula (5.5).The displacement at point A is equal with the column elongation,

because the initial displacement at ground connecting point is considered zero:

colcolcolcol

colgroundAAE

LF

AE

LFeuu

*

*

*

*0 1111 (5.42)

The displacement at point C is calculated as:

rodrod

DrodDCAE

LFueuu

*

* 22 (5.43)

where Du is the displacement allowed by the “leveling jack”.

The necessary displacement allowed by the leveling device is calculated from

equation (5.41):

0*

*

*

* 2211 rodrodcolcol

rodcolDAE

LF

AE

LFeeu (5.44)

Note: The displacement Du has to be positive for the device to work properly (the

device works only in tension). This means that the deformation of the column

has to be larger than the rod deformation.

The numerical application for this example is presented in section 5.6.1.

5.2.3 Statically Indeterminate Structure

A typical example of a statically indeterminate structure is shown in Figure 5.9. This

type of structures requires a more involved methodology for in order to calculate the

stress and strain distributions.

Figure 5.9 Statically Indeterminate Structure

The system shown in Figure 5.9 is composed of a rigid member AD, pinned into the

wall at point A, and two unequal linear elastic rods, BE and CF. The rods, BE and CF,

Page 106: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 122 -

are attached to the ceiling at points E and F, respectively. The system is loaded with a

concentrated vertical force P at point D. The free-body diagram used to write the

equilibrium equations is shown in Figure 5.10.

Figure 5.10 Free-Body Diagram

The following equilibrium equations are written:

0 xF 0AH (5.47)

0 yF 021 PFFVA (5.48)

0A

zM 0*** 21 cPbFaF (5.49)

The equilibrium equations (5.48) and (5.49) contain three unknown reaction forces 1F ,

2F and AV . In order to solve these unknown quantities one additional equation is

necessary. This equation is obtained from the deformation compatibility condition

schematically described in Figure 5.11. Because the beam AD is rigid, purely

geometric relations between the rod elongations, 2e and 2e , and the rotation angle

are written as:

*1 ae (5.50)

*2 be (5.51)

Figure 5.11 Deformation Notation

Using equation (5.29) and the relations (5.50) and (5.51) the forces in the rods are

expressed as:

****1

1

1

1 akaf

F (5.52)

Page 107: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 123 -

****1

2

2

2 bkbf

F (5.53)

The stiffness coefficients, 1k and 2k , are calculated from the geometrical and material

properties characteristics of the rods:

1

111

*

L

AEk (5.54)

2

222

*

L

AEk (5.55)

Substituting equations (5.52) and (5.53) into the equilibrium equations (5.48) and

(5.49), 1F and 2F are eliminated from the system leaving only two unknowns, AV

and :

0**** 21 PbkakVA (5.56)

0***** 2

2

2

1 cPbkak (5.57)

Solving the algebraic system, the two unknowns are found as:

PPbkak

c**

**12

2

2

1

(5.58)

where 1 is the rotation angle corresponding to 1P

Pbkak

cbbkcaak

bkakPVA

***

)(**)(**

*)**(

2

2

2

1

21

21

(5.59)

Introducing the rotation angle into equations (5.52) and (5.53) the rod forces are

calculated:

PFPbkak

cakF **

**

**1_12

2

2

1

11

(5.60)

PFPbkak

cbkF **

**

**1_22

2

2

1

22

(5.61)

where 1_1F and 1_2F are the forces in the rods in 1P

The rod stresses are also calculated:

Page 108: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 124 -

1

11

A

Fx (5.62)

2

22

A

Fx (5.63)

The rod elongations are obtained using equations (5.50) and (5.51) as:

Pbkak

cae *

**

*2

2

2

1

1

(5.64)

Pbkak

cbe *

**

*2

2

2

1

2

(5.65)

If the allowable vertical force P is required, the stress in the rods must be compared

against the allowable stress all :

allxA

F

1

11 (5.66)

allxA

F

2

22 (5.67)

At limit, the relations (5.66) and (5.67) are rewritten as:

allAPFF ** 11_11 (5.68)

allAPFF ** 21_22 (5.69)

Introducing the rod forces, equations (5.60) and (5.61), into equations (5.68) and

(5.69) the allowable vertical force admitted by the system is calculated as:

]**

)**(**,

,**

)**(**min[

]*

,*

min[

2

2

2

2

12

1

2

2

2

11

1_2

2

1_1

1

cbk

bkakA

cak

bkakA

F

A

F

AP

all

all

allallall

(5.70)

The numerical application for this example is presented in section 5.6.2.

Page 109: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 125 -

5.3 Nonuniform-Axial Deformation

The definition of a member with uniform-axial deformation is specified in Section

5.3. If any one of the assumptions contained in definition 5.2 is violated the axial

deformation is called nonuniform-axial deformation. The most common cases of

nonuniform-axial deformation are treated in the following subsections.

5.3.1 Non-homogeneous Cross-Section Members

The theory developed in the previous sections assumed that the cross-section is made

from a homogeneous material described by its modulus of elasticity. In structural

engineering practice it is not uncommon to have a case in which a member

constructed from two different materials bounded together at their interface is forced

to undergo an axial deformation. The members made from different materials but

behaving together as a single member are called composite sections.

For the purpose of analysis it is assumed that the member is made from two materials,

each being characterized by a specific modulus of elasticity ( 1E and 2E ) and area ( 1A

and 2A ). The strain distribution is, as before, assumed to be a function of only the

variable x , the position of the particular cross-section of interest:

)(),,( xzyx xx (5.71)

Considering that both materials are homogeneous linear elastic materials the Hook’s

law may be written for each material as:

)(*)( 1

1 xEx xx (5.72)

)(*)( 2

2 xEx xx (5.73)

The total axial force )(xF , as expressed in equations (5.11), is divided into two axial

forces, each acting at the centroid of the corresponding bar cross-section.

)()(

*)(*)(*)()(

21

2

2

1

1

21

xFxF

dAxdAxdAxxFA

x

A

x

A

x

(5.74)

where

)(*)(*)()( 1

1

1

1

1

1

xAxdAxxF x

A

x (5.75)

)(*)(*)()( 2

2

2

2

2

2

xAxdAxxF x

A

x (5.76)

Page 110: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 126 -

Substituting the axial stresses expressed in equations (5.72) and (5.73) into equations

(5.75) and (5.76) the equilibrium equation (5.74) is written as:

)(*)](*)(*[

)(*)(*)(*)(*

)()()(

2211

2211

21

xxAExAE

xAxExAxE

xFxFxF

x

xx

(5.77)

Consequently, the elongation strain is obtained:

)(*)(*

)()(

2211 xAExAE

xFxx

(5.78)

Accordingly, the normal stresses are calculated employing the equations (5.72) and

(5.73) as:

)(*)(*)(*

)(2211

11 xFxAExAE

Exx

(5.79)

)(*)(*)(*

)(2222

22 xFxAExAE

Exx

(5.80)

In general, the stress resultant moments, as expressed in equations (5.12) and (5.13),

do not vanish and additional restrictions regarding the geometrical characteristics of

the cross-section must be imposed. Nullification of the stress resultant moments is

obtained by using symmetric cross-section about one or both centroidal axes.

The frequently encountered practical case when the areas )(1 xA and )(2 xA are

constants along the entire length and the member is subjected to a constant force P is

considered below. It is additionally assumed that the cross-section is symmetrical only

about the vertical axis oy . Then, the normal stresses developed in each material zone

are calculated, following the equations (5.79) and (5.80), as:

constPAEAE

Exx

*

**)(

2211

11 (5.81)

constPAEAE

Exx

*

**)(

2222

22 (5.82)

The assumption regarding the symmetrical aspect of the cross-section about the

vertical axis nullify the stress resultant moment yM when the bar local coordinate

system is centroidal. The second stress resultant moment zM can be made zero by

manipulating the position of the application of the force P . It was previously argued

that the individual cross-section resultants )(1 xF and )(2 xF must be located at the

Page 111: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 127 -

centroids of the cross-sections 1A and 2A . This raises the question about the location of

the applied force P in order that only axial forces are induced in each of the

individual bars composing the cross-section. This situation is solved by replacing the

axial forces )(1 xF and )(2 xF with a resultant force and a null moment written about a

horizontal axis. The stress resultant moment about the vertical axis has a zero value

due to the imposed symmetry of the cross-section about that axis. The moment

equation, written about an horizontal axis passing through the application point of the

force P , is:

0*)(* 21 PP dFddF (5.83)

where d is the distance between the centroids of cross-sections 1A and 2A , while Pd

is the distance between the application point of the force P and the centroid of the

area 2A .

The distance Pd is then calculated:

dFF

FdP *

21

1

(5.84)

Note: In general the application point of the force P does not correspond with the

centroid of the composite cross-section.

5.3.2 Non-homogeneous Cross-Section Members Subjected to Thermal Changes

Consider the composite section described in Section 5.3.1 subjected to change in

temperature. For generality, in this discussion, let be assumed that each individual bar

undergoes a different change in temperature 1T and 2T , respectively. The notation

employed in the Section 5.3.1 is maintained. In the absence of the exterior forces, the

equation of equilibrium is:

0)()( 21 xFxF (5.85)

The axial forces pertinent to each one of the bars is constant along their respective

lengths and thus, the equation (5.85) becomes:

021 FF (5.86)

The equilibrium equation contains two unknown cross-sectional forces, )(1 xF and

)(2 xF , and consequently the system is statically indeterminate. An additional

equation is necessary to obtain these forces. This equation is derived from the

condition of equality of the elongations for the two individual bars imposed by the

existence of the rigid members attached at their ends. This equation is written as:

Page 112: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 128 -

21 ee (5.87)

The elongations may be express using the formulation of equation (5.31):

LTFfe *** 11111 (5.88)

LTFfe *** 22222 (5.89)

Substituting equations (5.88) and (5.89) into equation (5.87) the second necessary

equation is obtained:

0****** 22221111 LTFfLTFf (5.90)

Solving the algebraic equation system (5.86) and (5.90) the cross-section forces are

calculated:

21

112221

*)**(

ff

LTTFF

(5.91)

The axial stress in the bars is:

1

11

A

F (5.92)

2

22

A

F (5.93)

5.3.3 Heated Member with a Linear Temperature Variation

The slender beam shown in Figure 5.12.a is heated by a heating coil capable of

producing a linearly varying temperature as shown in Figure 5.12.b. The beam has a

constant cross-section area A and is made from a linear elastic material. The beam is

attached to rigid supports at ends A and B.

Figure 5.12 Heated Uniform Member

(a) Geometry and (b) Temperature Variation

Page 113: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 129 -

In the absence of any applied forces between A and B, the equilibrium equation is

written as:

0 BA FF (5.94)

The system is statically indeterminate containing two unknown reaction forces

AF and BF . An additional equation is necessary. This equation is obtained by

observing that the total elongation of the beam is null:

0e (5.95)

Following the equations (5.20) and (5.31) the total elongation is calculated:

L

dxxTFfe0

*)(** (5.96)

where AE

Lf

*

Accordingly, with the temperature variation shown in Figure 5.12.b the thermal

variation in a particular cross-section is:

L

xTTTxT ABA *)()( (5.97)

The integral contained in the equation (5.96) is calculated using the expression (5.97):

2*)(

2*)(*

*]*)([*)(00

LTT

LTTLT

dxL

xTTTdxxT

AB

ABA

L

ABA

L

(5.98)

Substituting (5.98) into (5.96) the total elongation is expressed as:

2*)(**

LTTFfe AB (5.99)

Imposing the condition (5.95) the force F is found:

2

)(***2*)(*

ABAB TTEA

f

LTT

F

(5.100)

Using the equation (5.94) results that:

Page 114: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 130 -

FFF BA (5.101)

The normal stress is then calculated as:

2

)(**

)(

)()( AB

x

TTE

A

F

xA

xFx

(5.102)

5.4 Special Aspects

5.4.1 Normal Stress in the Vicinity of the Load Application

A vertical prismatic bar characterized by a constant cross-section along its entire

length L is loaded at one end by a concentrated force Q and supported at the other end

as illustrated in Figure 5.13. Based on Newton’s Law the constraint at the support

generates a uniform distributed reaction q opposed to the action. The free-body

diagram is shown in Figure 5.13.a. The example considered falls in the category of

members with uniform axial-deformation studied in Section 5.2. The exact

determination of the distribution of normal stress along the length of the beam

requires advanced methodologies employed in the Theory of Elasticity and for this

reason only the results are presented. Analyzing the results illustrated in Figure 5.13 it

is found that the formulae obtained in Section 5.2 are valid in the majority of the

cross-sections except of those located in the vicinity of the ends. The perturbation

zone has a length b roughly equal to the width of the cross-section.

Figure 5.13 Normal Stress Distribution

Note: It can be concluded that for practical purposes the formulae obtained in

Section 5.2 based on the assumptions contained in definition 5.1 are valid

especially if the ratio 25.0L

b. Special attention has to be given to the areas

located near the point of load application or near abrupt changes in the cross-

section. The application of Saint Venant’s Principle is valid for the case of the

beam.

Page 115: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 131 -

5.4.2 Stress Concentrations

In the theoretical development pertinent to the axial deformation of linear members,

the area )(xA of the cross-section is considered as a smoothly varying function of the

position x . If discontinuities appear in the definition of the cross-sectional area, the

formulae obtained in the preceding sections are invalid and the concept of stress

concentration must be introduced.

Figure 5.14 Concentration of Stress

A typical case is shown in Figure 5.14 where a prismatic type linear member having a

circular hole of diameter d is subjected to a constant tension forceQ . The member

cross-section is described by the height b and thickness t .

The normal stress max around the hole can be significantly greater than tdb

Q

*)(

and varies as a function of the ratiob

d. For practical applications the coefficient K ,

called the stress-concentration factor is introduced. This factor is defined as the

ratio of the maximum normal stress max around the hole to the normal stress

calculated in the absence of the hole nom , called nominal stress.

nom

K

max (5.103)

tc

Q

tdb

Qnom

**)(

. (5.104)

The average normal stress calculated with the formula (5.14) is:

tb

Pavg

* . (5.105)

The variation of the stress-concentration factor K , calculated using the Theory of

Elasticity methods is illustrated in Figure 5.15. The stress concentration factor for the

configuration under consideration varies from 2.3 to 3.0.

Page 116: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 132 -

Figure 5.15 Variation of the Stress-Concentration Factor

Analyzing the variation, it is concluded that if the diameter d decreases the

concentration factor K also increases. This is somewhat misleading and is due to the

way the chart in Figure 5.15 is constructed. Using the average normal stress avg in the

definition of the stress concentration factor K , equation (5.103) the following formula

is obtained:

)1(** maxmax

b

d

b

cK

avgavg

(5.106)

Considering the extreme cases shown in Figure 5.15 and applying these within

equation (5.106) it can be concluded that the maximum normal stress max increases

with the increase in the diameter of the hole and varies as:

avgavg *6.4*0.3 max (5.107)

For small diameter holes the stress concentration disappears at relatively small

distance from the hole. This is an example of the application of Saint Venant’s

Principle.

5.4.3 Limits of Poisson’s Ratio

The volumetric strain V was calculated in Section 2.9.3 for the general case when the

normal stresses x , y and z are applied simultaneously. For the case of axial

deformation only the normal stress x has a non-zero value and, consequently, the

equation (2.111) is written as:

xxVEV

V

*)*21()(

*21

(5.108)

Page 117: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 133 -

Because the volume can not decrease during the tensioning of the axially deformed

member the volumetric strain is a positive value. Mathematically this condition is

enforced as:

0

V

VV (5.109)

Consequently, from physical reality and equation (5.108) it is seen that:

5.00 (5.110)

The limits established for Poisson’s ratio by the expression (5.110) are generally

valid for all materials used in structural engineering. The cases representing the

limiting values, 0 and 5.0 , are pertinent to cork and water, respectively.

Structural steel has a Poisson’s ratio of 0.33. There are some cases when the material

has a negative Poisson’s ratio. These materials are called swollen solids and this

unusual behavior is characteristic of certain materials subjected to radiation.

The transversal contraction of the cross-section is similarly obtained as:

xxAEA

A

**2)(

*2

(5.111)

5.5 Design of Members Subjected to Axial Deformation

In the design of the members subjected to axial deformation two important factors, the

load L and the resistance R , are considered. The load L represents the maximum

axial force that occurs in the specific member when subjected to the action of exterior

forces or change in temperature. The most common definition for the resistance R is

the force which is developed in the member when the normal stress reaches the

yielding value Y .

The design of the member subjected to axial deformation is conducted under the

condition that the capacity of the member, represented by its resistance force R , must

always be greater or equal to the demand force L . Mathematically, this assumption is

expressed as:

RL (5.112)

In the calculation of the load L and resistance R forces there are typically a number

of factors (load magnitudes and directions, material characterization, manufacturing

tolerance, etc.) which are not known with absolute certainty. The degree of

uncertainty may be treated using the concepts proper of Probability Theory. In order

to circumvent the difficulties inherent in the rigorous application of probabilistic

methods, a global factor SF , called the safety factor, encompassing all possible

uncertainties is introduced as:

Page 118: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 134 -

allR

RSF (5.113)

The allowable resistance allR is then used in place of the resistance R in equation

(5.112).

allRL (5.114)

The safety factor is always greater than or equal to unity:

1SF (5.115)

This is the approach used by the method called ultimate strength design method and

was adopted by American Concrete Institute (ACI) and American Institute of Steel

Construction (AISC).

If the relation between the stress and strain is linear, than a similar safety factor may

be defined by limiting the value of the normal stress in the axially deformed member.

all

YSF

(5.116)

The design formula (5.114) is modified using the relationship between maximum

normal stress max and the allowable normal stress all :

all max (5.117)

The formula (5.117) was used for a long period of time in a procedure known as the

allowable-stress design. Due to the simplicity of application, this method is still

commonly used in United States for the design of steel structures.

5.6 Examples

The application of the theoretical formulae developed in this lecture is illustrated in

the following examples.

5.6.1 Statically Determinate Structure

The structure illustrated in Figure 5.7 was investigated in Section 5.5.2. The following

numerical values are considered in the numerical application of the generic

formulation:

10P kN

Page 119: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 135 -

0.5L m 0.31 L m 5.12 L m 4.0a

0.12colA 2cm 0.5rodA 2cm

1110*2colE Pa 1110*2rodE Pa

The reaction forces are:

61 F kN Compression in column

42 F kN Tension in rod

The column related values are:

610*5col

x Pa

510*5.2 col

x

510*5.7 cole m

The rod related values are:

610*8rod

x Pa

510*4 rod

x

510*6 rode m

The displacement Du of the leveling device is calculated as:

555 10*5.110*610*5.7 rodcolD eeu m

5.6.2 Statically Indeterminate Structure

The statically indeterminate structure illustrated in Figure 5.9 is solved in Section

5.2.3. For the numerical application of the formulation developed the following data

are employed:

5.1a m 5.2b m 0.4c m

0.21 L m 3.12 L m

Page 120: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 136 -

11

21 10*2 EE Pa

0.61 A 2cm 0.32 A 2cm

810*1.2all Pa

The stiffness coefficients are:

7

1 10*0.6km

N

7

2 10*615.4km

N

The rotation angle 1 is:

9

1 10*446.9 radians

The rod forces 1_1F and 1_2F are calculated as:

85.01_1 F N

09.11_2 F N

The allowable vertical force allP for the system is:

410*780.5allP N

Page 121: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 137 -

LECTURE 6

Pure Shear

In the previous lecture normal stress and elongation strain concepts were presented

and discussed. The concepts of shear stress and strain are the subject of this lecture.

6.1 Basic Theory

In accordance with the shear stress definitions (2.13) and (2.14), if on a particular

cross-section there exists only a stress resultant V , orientated parallel to any axis

describing the cross-section plane, the shear stress is defined as:

A

VA

0lim (6.1)

where V is the tangential stress vector acting on the infinitesimal area A .

The resultant shear forceV is obtained by integrating the shear stress vector over the

entire area of the cross-section A :

A

dAV * (6.2)

The shear force V and the shear stress are illustrated in Figure 6.1.

Figure 6.1 Pure Shear

(a) Shear Force and (b) shear stress

Page 122: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 123 -

The average shear stress avg is calculated by dividing the shear force by the total area

of the cross-section:

A

Vavg (6.3)

The distribution of shear stress is not uniform on the cross-section and in general, can

not be determined from the integral equation (6.2). In some cases the average shear

stress avg may be used in place of the true shear stress without introducing

significant an error. This is commonly done for the analysis of parts used for

connection of axially deformed members to other members or supports. These parts,

called connectors, are typically bolts, nails, rivets pins, welds and glue. The bolts,

rivets, pins and nails are discrete connectors, while the welds and glue are

continuous connectors.

When the exterior force acts parallel to the particular surface of the part the

phenomenon is called direct shearing. A typical example of direct shearing is the

punching shear of a metal sheet illustrated in Figure 6.2.a. A metal sheet is placed

between two rigid blocks. Through holes existing in the rigid blocks a punching rod is

pushed with a force P . The resulting small circular metal disk “sheared” from the

sheet is called a slug. The free-body diagram is shown in Figure 6.2.b.

Figure 6.3 Sheet-Metal Punch

(a) Punching Machinery and (b) Free-Body Diagram

Using the notation shown in the free-body diagram, where the exterior force and the

shear stress acting around the metal slug are depicted, the average shear stress is

calculated as:

td

Pavg

** (6.4)

Page 123: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 124 -

Another example of direct shearing is the hinge pin in an ordinary pair of pliers

shown in Figure 6.3. When force P is exerted on the arms of the pliers, the pin cross-

section is subjected to shear stress as depicted in Figure 6.3.b.

Figure 6.3 Pair of Pliers

The direct shear is characterized as single shear or multiple shear as a function of

the number of planes (cross-sections) on which the shear stress acts. Examples of

single shear and double shear are illustrated in Figures 6.5 and 6.7, respectively,

where two steel parts in tension are connected by a bolt.

The existence of shear stress implies the existence of shear strain, which represents

the change of the right angle. The original right angle shown in Figure 6.4.a is

modified into an acute angle * as depicted in Figure 6.4.b. The angular

change representing the shear strain is calculated as:

*

2

(6.5)

Figure 6.4 Shear Strain

(a) Undeformed angle and (b) Deformed angle

Page 124: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 125 -

The shear strain , measured in radians, is a small angle and for the sake of simplicity,

is approximated by its tangent:

s

s

L

)

2tan( * (6.6)

If the material is homogeneous and linear elastic, the constitutive equation is

described by Hook’s law:

*G (6.7)

where G is the shear modulus characterizing the material.

6.2 Discrete Connectors Calculation

In Figure 6.5.a, under the action of force P the two metal parts press against the bolt

in bearing and consequently, contact stresses called bearing stresses, will develop.

The tendency of the two connected pieces to pull apart induce induce a direct shearing

in the bolt. To clarify the phenomenon which takes place the connecting bolt is

isolated and shown in Figure 6.5.c.

Figure 6.5 One-Bolt Single Shear Connection

The bearing stress b exerted on the bolt is shown on the bolt free-body diagram and

is calculated as:

bear

bear

bearA

F (6.8)

where bearF and bearA are the bearing force and the bearing area, respectively.

Page 125: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 126 -

In the case of single shear the bearing stress bearF is equal the axial applied force P .

The actual distribution of the bearing stress around the contact area is a complex

distribution, which, for practical purposes, is simplified by considering that the stress

is uniformly distributed over the area obtained by multiplying the bolt diameter by the

thickness of the part bearing on the bolt. For the case shown in Figure 6.5 the

maximum bearing stress is obtained as:

brb Ftd

P

min

max_*

(6.9)

where d and mint are the bolt diameter and the minimum thickness of the two parts

connected by the bolt.

The limiting bearing stress brF is a value obtained from laboratory tests and is

uniquely determined for each type of connector.

The shearing force is transferred through the bolt section mn and by approximating

the shear stress as uniformly distributed over the bolt cross-section, the average shear

stress avgb _ is obtained as:

vavgb Fd

P

2_*

*4

(6.10)

where vF is the limiting value for the connector shear stress obtained also from

laboratory tests. The failure of a bolt in a single shear connection is illustrated in

Figure 6.6.

Figure 6.6 Failure of Bolt in Single Shear

Equations (6.9) and (6.10) may be combined into one giving the allowable force in the

connector as:

PFd

FtdP vbrallb )*4

*,**min(

2

min_

(6.11)

Page 126: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 127 -

Equation (6.11) represents the design formula for one-connector single shear

connections.

The double shear type connection is shown in Figure 6.7.a. The bearing action is

transferred in this case over three surfaces. Considering the same simplified

distribution explained above the bearing stress is calculated by application of equation

(6.9) to each of the three contact surfaces. The difference appears when analyzing the

transfer of the shearing force through the body of the connector. For double shear the

bearing force is equal to P for the middle layer contact and 2/P for each of the outer

two layer contact surfaces. Also, as illustrated in Figure 6.7.b, a resultant shear force

of 2/P occurs on each of two cross-sections of the bolt mn and pq .

Considering that the shear stress is uniformly distributed over the shearing areas the

average shear stress is calculated as:

vavgb Fd

P

2_*

*2

(6.12)

Figure 6.7 One-Bolt Double Shear Connection

The design formula is written as:

PFd

FtdP vbrallb )*2

*,**min(

2

min_

(6.13)

where ),min( 231min tttt .

Multiple shearing surfaces may occur if there are more than two parts held together by

one connector. Considering that there are 1n contact parts pulling in one direction and

Page 127: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 128 -

2n contact parts pulling in the opposite direction, the maximum bearing stress

max_b is calculated as:

bb Ftd

n

P

td

n

P

]*

,*

max[min_2

2

min_1

1max_ (6.14)

where min_1t and min_2t are the minimum thicknesses of the parts acting in opposite

directions.

The average shear stress in the bolt or pin of the case considered above is:

v

t

avgb Fdn

P

2

sec

_**

*4

(6.15)

where tnsec is the number of shearing sections.

In structural engineering practice, many connections require more than one connector

to transfer the force between the attached parts. Examples of multiple connector

connections are illustrated in Figures 6.8 and 6.9. These figures illustrate connections

with only one line of connectors.

Figure 6.8 Two-Connector Single-Shear Connection

Page 128: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 129 -

Figure 6.9 Three-Connector Single-Shear Connection

It is commonly assumed when the geometry of the connection is symmetrical and the

load passes through the connection centroid, that the load is equally sheared by all

connectors.

The most frequently encountered type of axially loaded connection is the single lap

connection. An example of two plate single lap with multiple fasteners connection is

shown in Figure 6.10. To calculate the axial load pertinent to the cross-sections 1-1

through 4 -4 the free-body diagrams are constructed and shown in Figures 6.10.b

through 6.10.e. The verification of the reduced cross-sections of the plate is required

and consequently, the evaluation of the reduced area in all perforated cross-sections is

necessary. The effective normal stress is then calculated in the usual manner by

dividing the corresponding axial force acting on the cross-section by the reduced area

of the cross-section.

Figure 6.10 Axial Load Distributions in the Single Lap Connection

Page 129: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 130 -

Suppose that plate A has a thickness t and width b and the connector diameter is d .

Accordingly, with the assumption made above, each one of the ten connectors carries

an equal load of 10

P. The reduced areas are calculated as:

dttbA **3*11 (6.16)

dttbA **2*22 (6.17)

dttbA **3*33 (6.18)

dttbA **2*44 (6.19)

Consequently, the maximum normal stress is obtained as:

]*7.0

,max[2211

A

P

A

Pplate (6.20)

Additional details pertinent to the design and verification of connections using

discrete connectors, such as the verification of the zigzagging cross-section going

through connectors 1-2-3-2-1, are discussed in specialty courses.

6.3 Weld Calculation

Welds are categorized in four basic groups: groove, fillet, slot and plug welds. These

categories are illustrated in Figure 6.11.

Figure 6.11 Basic Weld Types

Page 130: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 131 -

In the widely used Manual of Steel Construction Allowable Stress Design published

by the American Institute of Steel Construction (AISC) groove welds are further

classified in many subcategories. In engineering practice, the fillet weld is the most

commonly used type of weld, because of it is the easiest to fabricate.

Figure 6.12 Effective Throat Dimension for Fillet Welds

(a) Equal Legs and (b) Unequal Legs

The allowable stress on various types of welds is dependent upon the effective area

of the weld. The effective area of the groove or fillet weld is calculated as the product

of the effective throat dimension et and the length of the weld.

For fillet welds the effective throat dimension is defined as the shortest dimension

measured from the weld root to the face of the weld. Assuming that the weld has

equal legs of length a the effective throat is equal to 2

a. The calculation of the

effective throat accordingly to AISC is shown in Figure 6.12. For full penetration

groove welds, the AISC defines the effective throat as a function of the thickness of

the parts joined in the connection as pictured in Figures 6.13.a and 6.13.b. In the case

of partial penetration welds, the effective throat is a function of the depth of the

preparation as shown in Figures 6.13.c and 6.13.d.

Figure 6.13 Groove Weld Effective Throat

Page 131: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 132 -

The design formula for the fillet weld is:

PFtl weldalleffi

n

weldi )**( ___ (6.21)

where: n is the number of welds parallel to the axial force P ;

weldil _ , effit _ are the effective length and throat of the weld i ;

weldallF _ is the allowable shear stress in the weld.

The weld allowable shear stress weldallF _ is calculated as:

electrodeyweldall FF __ *3.0 (6.22)

where electrodeyF _ is the ultimate tensile strength of the electrode material.

For groove weld the design formula is:

PFtl weldalleffweld _** (6.23)

The allowable stress weldallF _ in the weld is:

yweldall FF *6.0_ (6.24)

where yF is the base metal yielding stress.

Page 132: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 133 -

LECTURE 7

Bending

7.1 Definitions

A linear beam subjected to transversal loading (forces and moments) deflects

laterally, a phenomenon known in structural practice as bending. In order to simplify

the study of plane linear beam bending, it has to be assumed that the beam has a

geometric longitudinal plane of symmetry as illustrated in Figure 7.1. The local

coordinate system oxyz , similar with those used in the previous lectures, has the ox

axis aligned along the length of the beam. The supports and the loading are also

considered to be symmetric in relation to the longitudinal symmetry plane of the

beam. As a direct consequence of the existence of the vertical plane of symmetry

oxy , the beam deformation manifests in this plane, and thus, it is also called the

plane of bending.

Figure 7.1 Beam with Vertical Plane of Symmetry

Page 133: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 153 -

After application of the load, the beam deforms as schematically depicted in Figure

7.2. One imagines that the beam consists of a number of parallel longitudinal fibers as

pictured in Figure 7.2.a. Then, after the deformation it is evident from the figure that

some of the fibers are stretched and some are shortened. The horizontal plane

containing the fibers which remained unchanged is called the neutral surface. The

curve resulting from the intersection between the neutral surface and the longitudinal

plane of symmetry is called the deflection curve. The above discussed terminology is

depicted in Figure 7.2.b.

Figure 7.2 Beam Deflected Shape from Bending and Descriptive Terminology

7.2 Pure Bending

Two examples of beams subjected to pure bending are shown in Figures 7.3 and 7.4.

The corresponding bending moment diagrams are also included.

The simple supported beam pictured in Figure 7.3 is loaded with two equal

concentrated moments acting at the ends A and B. Consequently, the shear force is

Definition 7.1

The segment of a plane linear beam is in pure bending if along its entire length:

(a) the cross-section is constant and symmetrical about a longitudinal plane ;

(b) the material properties are constant;

(c) it is subjected only to a constant bending moment;

(d) after the load application cross-sections remain plane and perpendicular to

the deflection curve.

Page 134: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 154 -

zero for the entire length of the beam, while the bending moment is a constant

function along the length of the member.

Figure 7.3 Simply-Supported Beam in Pure Bending

The cantilever shown in Figure 7.4 is also in pure bending, because the shear force is

zero and the bending moment, as indicated by the moment diagram, is constant for the

entire length of the beam.

Figure 7.4 Cantilever Beam in Pure Bending

An interesting situation is depicted in Figure 7.5. The beam is simply-supported and

loaded symmetrically by two concentrated forces P located at equal distance a from

the supports A and B .

Figure 7.5 Simply-Supported Beam with Central Region in Pure Bending and the End

Regions in Non-uniform Bending

Page 135: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 155 -

The shear diagram is constant in the segments AC and DB, and zero only on the

central segment CD. The bending moment, as shown in the moment diagram, has a

linear variation within the end segments AC and DB, and is constant within the

central segment CD. In accordance with definition 7.1, only the central segment CD is

in pure bending. In contrast to the situation of pure bending when only the constant

moment exists, the presence of the shear force results in a variation of the bending

moment and is referred to as non-uniform bending. The beam pictured in Figure 7.5

has its end regions in non-uniform bending and only the central region in pure

bending.

7.2.1 Strain-Displacement Equation

Assumption (d) listed in Definition 7.1 is known as the Bernoulli-Euler hypothesis.

This name was later extended to encompass the entire theory of the beam in bending

known today as the Bernoulli-Euler beam theory. This assumption, which

practically is seldom realized, introduces an important kinematic assumption very

similar in nature with that postulated in the axial deformation of the beam. The main

implication of assumptions (a) to (d) is that the fibers contained in any vertical plane

behave identically to those in the longitudinal plane of symmetry oxy .

Mathematically, this implies the independence of the deformation on the variable z

and, consequently, the beam deformation can be represented by the deformation of the

fibers located in the longitudinal plane of symmetry.

Figure 7.6 illustrates a beam segment in undeformed and deformed conditions. The

radius and the center C are called radius and center of curvature, respectively.

The radius of curvature is a function only of the variable x as a direct consequence

of assumption (d).

Figure 7.6 Beam Segment Deformation

(a) Undeformed and (b) Deformed

The undeformed segment PQ transforms during the loading application into the

curved segment P*Q*. Similarly, the undeformed segment AB located on the

deflection curve becames A*B* after the deformation, with the difference being that

Page 136: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 156 -

its original length remains unchanged. Using the general definition (2.42) the

elongation strain x is expressed as:

][lim

][lim),(),,(

*

**

x

xx

PQ

PQQPyxzyx

QP

QPxx

(7.1)

By studying Figure 7.6, the following geometrical relations can be written:

xABPQ (7.2)

*** *)( xABBA (7.3)

**** *])([ yxxQP (7.4)

Introducing the geometric relations (7.2) through (7.4) into the elongation strain

equation (7.1):

)(

]*)(

*)(*))(([lim),(

*

**

x

y

x

xyxyx QPx

(7.5)

For the case of small displacements the radius of curvature )(x is related to the

vertical displacement )(xv by the following relation:

2

2 )(

)(

1

dx

xvd

x

(7.6)

Note: Detailed explanation pertinent to equation (7.6) is provided in Section 8.3.

Substituting equation (7.6) into equation (7.5), the following ordinary differential

equation is obtained:

ydx

xvdyxx *

)(),(

2

2

(7.7)

Equation (7.7) represents the strain-displacement equation of the linear plane

member in pure bending.

Examination of equation (7.6) reveals that the elongation strain ),( yxx is inversely

proportional to the radius of curvature )(x . The deformation of the beam as a

Page 137: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 157 -

function of the curvature is illustrated in Figure 7.7. The inverse of the radius of

curvature )(x is called curvature and is notated by the letter )(/1 xk . For a

positive radius of curvature the center C is located above the beam deflection curve

and the deformation is concave as depicted in Figure 7.7.a. In contrast, the

deformation is convex if the curvature is negative and the center C is located below

the beam deflection curve as shown in Figure 7.7.b.

Figure 7.7 Beam Deformation and Radius of Curvature

(a) Positive Curvature and (b) Negative Curvature

Variation of the elongation strain ),( yxx , illustrated in Figure 7.8.a, for a particular

cross-section corresponding to a positive radius of curvature is a linear function. The

fibers located above the deflection curve are shortened, while these below the

deflection curve are extended. Consequently, the areas above and below the deflection

curve are in compression and tension, respectively. For the case of the negative

curvature, shown in Figure 7.8.b, the distribution is similar, but the compression and

tension zones are reversed. Considering the relations (7.5) and (7.8) for a particular

cross-section (x=constant) the strain ),( yxz is positive above the neutral plane and

negative below the neutral plane.

Figure 7.8 Elongation Strain Distribution

The transversal strains are expressed as:

Page 138: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 158 -

),(*),(),( yxyxyx xzy (7.8)

Note: In the classical theory of the beam transversal deformation of the cross-section

is neglected.

7.2.2 Constitutive Equation

The constitutive equation represents, as described in Lecture 2, the relation between

stress and strain. For the case of isotropic elastic material behavior under axial

deformation, the relation between the stress and strain is written as follows for the

beam in bending:

)(*)(),(*)(),(),,(

x

yxEyxxEyxzyx xxx

(7.9)

where )(xE is the modulus of elasticity, which has a value proper to each specific

material.

The normal stress ),( 0 yxxx pertinent to a particular cross-section is a function of

the variable y and varies linearly as it is directly proportional to the elongation

strain ),( yxx .

The rest of the stress tensor components are zero:

0 zy (7.10)

0 xyxyxy (7.11)

7.2.3 Cross-Section Stress Resultants

Considering the normal stress ),( yxx distribution represented by the equation (7.9)

the cross-section stress resultants )(xF , )(xM y and )(xM z illustrated in Figure 7.9

are calculated as:

z

A

AA

x

Sx

xEdAy

x

xE

dAx

yxEdAyxxF

*)(

)(*

)(

)(

*)(

*)(*),()(

(7.12)

Page 139: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 159 -

yz

A

AA

xy

Ix

xEdAzy

x

xE

dAx

zyxEdAyxzxM

*)(

)(**

)(

)(

*)(

**)(*),(*)(

(7.13)

z

A

AA

xz

Ix

xEdAy

x

xE

dAx

yxEdAyxyxM

*)(

)(*

)(

)(

*)(

*)(*),(*)(

2

2

(7.14)

where zS , yzI and zI are the static moment, the product of inertia and the

moment of inertia, respectively.

Figure 7.9 Normal Stress and Stress Resultants

(b) Stress Resultants and (b) Normal Stress

If the coordinate system oxyz is considered passing through the cross-section centroid

and the cross-section is symmetric about y axis, the cross-section stress

resultants )(xFx , )(xM z and )(xM z became:

0)( xF (7.15)

0)( xM y (7.16)

'*)(

)(zz I

x

ExM

(7.17)

where 'z is the axis passing through the cross-section centroid and parallel to z .

7.2.4 Normal Stress Distribution

Combining equation (7.17) with equation (7.9), the normal stress ),( yxx is

expressed as:

Page 140: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 160 -

yI

xMyx

z

zx *

)(),(

'

(7.18)

Equation (7.18), called Navier’s formula, expresses the linear distribution of the

normal stress ),( 0 yxxx in the beam cross-section. The variation of the normal

stress for positive and negative bending moments is illustrated in Figure 7.10.

Figure 7.10 Normal Stress Distribution

By definition the neutral axis of the cross-section is the axis where the normal stress

),( 0 yxxx is zero.

Note: In the case of pure bending, the neutral axis passes through the centroid of the

cross-section and is horizontal.

The maximum normal stresses, tension or compression, are obtained at extreme

located fibers. The ratio obtained by dividing the moment of inertia by the distance

measured from the centroid to the extreme fiber is called the-section modulus.

Obviously if the cross-section has only one axis of symmetry two section moduli can

be defined:

)(

)()(

'

xy

xIxW

bottom

zbottom (7.19)

top

ztop

y

xIxW

)()(

'

(7.20)

Consequently, if the bending moment is assumed positive, 0)( xM z , as illustrated in

Figure 7.10.a, the maximum tensile and compressive stresses are obtained as:

)(

)()(max__

xW

xMx

bottom

ztensionx (7.21)

)(

)()(max__

xW

xMx

top

zncompressiox (7.22)

Page 141: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 161 -

If the bending moment assumes a negative value 0)( xM z , the maximum stresses are

as shown in Figure 7.10.b:

)(

)()(max__

xW

xMx

top

ztensionx (7.23)

)(

)()(max__

xW

xMx

bottom

zncompressiox (7.24)

Note: If the cross-section has two axes of symmetry the section moduli are equal

and, consequently, the magnitudes of the maximum normal stresses are equal.

7.3 Nonuniform Bending

The main limitation of the pure bending theory is the requirement that only the

bending moment )(xM z is allowed to be present in the cross-section of the beam. In

general, the beams are acted on by transversal loads and, consequently, the existence

in the cross-section of the shear force )(xVy is expected. The bending of a beam which

manifests in the presence of the bending moment and shear force is called

nonuniform bending. The problem of the nonuniform bending was first studied by

D. J. Jurawski, a Russian engineer, in connection with the behavior of rectangular

timber beams. In the case of nonuniform bending, both, the normal stress and shear

stress, are present in the beam cross-section. Because the analysis of the shear stress

distribution on the cross-section is a complicated task requiring mathematical and

theoretical knowledge beyond the level of this textbook, the complexity of the

theoretical development presented herein will be gradually increased. First, the

simplest case of a rectangular cross-section will be investigated. Later, the

formulations obtained for the case of the rectangular cross-section are extended to

other geometrical type of cross-sections.

7.3.1 Basic Assumptions

As illustrated in Figure 7.11 the concurrent existence of bending moment )(xM z and

shear force )(xQy induces normal stress ),( yxx and shear stress ),,( zyxxy ,

respectively.

According to the duality principle the shear stress ),,( zyxyx acting in a longitudinal

plane exists and has a non-zero value.

),,( zyxxy = ),,( zyxyx (7.25)

Page 142: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 162 -

Figure 7.11 Nonuniform Bending of Rectangular Cross-Section

The duality of the shear stresses expressed in equation (7.25) is illustrated in Figure

7.12. In the absence of tangential forces from the exterior surface of the beam it is

expected that ),,( zyxxy to be zero at the upper and lower fibers, marked in the Figure

7.12 by points a and c .

Figure 7.12 Shear Stresses Duality

The shear stress ),,( zyxxy is related with the angular strain and ultimately involves

change in shape of the deformable body. Figure 7.13 depicts a qualitative comparison

considering the same cantilever beam in pure and nonunniform bending.

Figure 7.13 Cantilever in Pure and Nonuniform Bending

(a) Pure Bending and (b) Nonuniform Bending

Page 143: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 163 -

In accordance with the theoretical developments derived in Section 7.1, every plane

cross-section of the cantilever beam in pure bending remains planar through the

deformation as shown in Figure 7.13.a. In contrast, the cantilever shown in Figure

7.13.b, which is in nonuniform bending due to the existence of the shear force along

its entire length, results in warping of the planar cross-sections after deformation.

Note: The finding that in nonuniform bending the cross-section does not remain

plane after the deformation contradicts the basic kinematic assumption of pure

bending described by definition 7.1. However, detailed investigation using

advanced methods of the Theory of Elasticity indicate that the warping of the

cross-section due to the shear stress is insignificant and does not affect the

longitudinal strain when the ratio of cross-section height to length of beam is

small. This is the case for most ordinary structural members.

To use the distribution of ),( yxx derived for the case of pure bending in the case of

the beam subjected to nonuniform bending a supplementary assumption must be

made: the distribution of the normal stress in a given cross-section is independent of

the shear induced deformation.

Note: As a consequence of imposing the above stated assumption, the pure bending

strain-displacement relation remains valid and the influence of the angular

deformations on the longitudinal strain is neglected.

7.3.2 Shear Stress Distribution in a Rectangular Cross-Section

In the absence of a strain-displacement relation for the shear stress, the distribution of

the shear stress is obtained from equilibrium considerations instead of the approach

used for normal stress in the previous section. In the American technical literature the

shear distribution is called the shear flow, by analogy to the flow of the liquid.

A thin slice isolated from the rectangular cross-section of a linear elastic beam is

illustrated in Figure 7.14.a. To simplify the figure only the cross-section stress

resultants, the shear force and the bending moment are shown. The corresponding

stresses are depicted in Figure 7.14.b.

By further subdividing the thin element with a horizontal plane ab the free- diagram

of the remaining upper portion is obtained. Using the notation shown in Figure 7.14.c

the horizontal equilibrium of this remaining upper portion can be written as:

12 FFH (7.26)

The horizontal force H represents the resultant force of the longitudinal shear

stresses ),,(),,( zyxzyx xyyx acting on the plane ab , while the horizontal forces 1F

and 2F are the resultant forces of the normal stresses ),( yxx acting on the vertical

areas. To determine the shear stress distribution in the cross-section the duality

principle is applied.

Page 144: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 164 -

Figure 7.14 Stress Resultants and Stresses

(a) Cross-section Resultants, (b) Stresses and (c) Free-body Diagram

Using the notation shown in Figure 7.15 and the Navier’s normal stress formula

(7.18) the forces 1F and 2F are expressed as:

)(*)(

**)(

*),(

'

''

''

1

ySI

xM

dAI

xMdAxF

A

z

z

z

Az

z

A

x

(7.27)

)(*)(

**)(

*),(

'

''

''

2

ySI

xxM

dAI

xxMdAxxF

A

z

z

z

Az

z

A

x

(7.28)

where )('

yS A

z and zI is the static moment of the section 'A and the moment of inertia

with respect to the neutral axis, respectively.

Note: The Figure 7.15 illustrates the rectangular cross-section after the deformation

emphasizing the lateral deformation of the rectangular cross-section.

Figure 7.15 Notation

The static moment )('

yS A

z of the section 'A is expressed as the product of the area 'A and the distance from the centroid of 'A centroid to the neutral axis.

Page 145: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 165 -

'

0

' *)('

yAyS A

z (7.29)

where '

0y is the distance from the centroid of area 'A to the neutral axis.

Substituting equations (7.27) and (7.28) into equation (7.26) yields the following

equation:

)(*])()(

[

)(*)(

)(*)(

'

''

ySI

xM

I

xxM

ySI

xMyS

I

xxMH

A

z

z

z

z

z

A

z

z

zA

z

z

z

(7.30)

The shear flow q which is the shear force per unit length is found by taking the limit

of the expression (7.30):

z

A

zy

z

A

zzzxx

I

ySxV

I

yS

x

xMxxM

x

Hyxq

)(*)(

)(*]

)()([limlim),(

'

'

00

(7.31)

Note: The relationship between bending moment and shear force was established in

Lecture 4 (equation (4.6)).

The average shear stress is calculated in a similar manner with the calculation of the

shear flow if instead of the length x the elementary area tx * is used:

tI

ySxV

tI

yS

x

xMxxM

tx

Hyx

z

A

zy

z

A

zzzx

xavryx

*

)(*)(

*

)(*]

)()([lim

*lim),(

'

'

0

0_

(7.32)

Equation (7.32) is called Jurawski’s formula for shear stress. In Figure 7.16 a typical

rectangular cross-section is shown. The distribution of the average shear stress

),(_ yxavryx calculated according to Jurawski’s formula (7.32) is illustrated in Figure

7.16.b.

Page 146: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 166 -

Figure 7.16 Shear Stress Distribution for a Rectangular Cross-Section

Using the notation shown in Figure 7.16.a, the static moment )('

yS A

z of the area above

the cutting plane cd is calculated:

]4

[*2

]2

)2

(

[*)2

(*)( 22

'

yhb

y

yh

yh

byS A

z

(7.33)

The moment of inertia about the horizontal axis passing through the cross-section

centroid is:

12

* 3hbI z (7.34)

Introducing equations (7.33) and (7.34) into equation (7.32) the variation of the

average shear stress is obtained:

]*4

1[**2

)(*3

*12

*

]4

[*2

*)(

),(),(

2

2

3

22

__

h

y

A

xV

bhb

yhb

xV

yxyx

y

y

avryxavryx

(7.35)

The distribution of the average shear stress in the rectangular cross-section is

parabolic and has the maximum value at the neutral axis location, where the normal

stress is zero. The maximum average shear stress is:

A

xVyx

y

avrxy

)(*5.1)0,(max

_ (7.36)

Page 147: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 167 -

Note: The formulae (7.31) and (7.32) derived for the case of the rectangular cross-

section can be extended without any theoretical difficulty to the general case

as follows:

)(

),(*)(),(

'

xI

yxSxVyxq

z

A

zy (7.37)

),(*)(

),(*)(),(

'

_yxtxI

yxSxVyx

z

A

zy

avryx (7.38)

The general formulae (7.37) and (7.38) are mainly restricted by the assumption that

the shear stress ),( yxxy is parallel to the shear force acting on the cross-section, an

assumption which is not always valid.

7.3.3 Limitations in the Usage of the Shear Stress Formula for Compact Cross-

Sections

Cross-sections without holes such as solid rectangular or circular cross-sections are

called compact cross-sections in contrast to the “thin-wall” or “built-up” cross-

sections. Jurawski’s formula for shear stress expressed by equation (7.32) is derived

under the assumption that the cross-section is a rectangular shape, but in engineering

practice this formula is applied to a large variety of compact cross-section shapes.

Additional limitations in the usage of Jurawski’s formula for shear stress are

discussed below.

7.3.3.1 Cross-Section Shape

The correct distribution of shear stress can be calculated using advanced methods of

the Theory of Elasticity. Some of these results are presented and compared with the

average shear stress calculated using equation (7.36).

The exact shear stress distribution at the neutral axis (NA) obtained from analyzing

the case of the cantilever beam subjected to a concentrated load applied at the tip

point and having different cross-sectional ratios bh is illustrated in Figure 7.17. The

cross-sections of the cantilevers shown in Figure 7.17.a and 7.17.b have an aspect

ratio of 25.0/ bh and 2/ bh , respectively.

The distribution of the shear stress calculated at the neutral axis of the wider cross-

section ( 25.0/ bh ) indicates a substantial departure from the average shear stress

calculated from equation (7.36). Contrary, the maximum shear stress calculated for

the narrow rectangular cross-section ( 2/ bh ) shows a very good agreement with the

value obtained from the application of the same formula (7.36). The above findings

limit the application of the formula (7.36) to the case of narrow rectangular cross-

sections. The following limit in the cross-section aspect ratio is imposed:

Page 148: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 168 -

0.2b

h (7.39)

Figure 7.17 Rectangular Cross-Section Beams

(a) Wide and (b) Narrow

The case of the cantilever with a circular cross-section is shown in Figure 7.18.a. The

difference between the value of the shear stress from the exact solution and the

average shear stress from Jurawski’s formula at the neutral axis location is negligible.

Figure 7.18 Circular Cross-Section

7.3.3.2 Beam Length

The verification of the average shear stress value calculated with equation (7.38) by

comparison with the exact solution from the Theory of Elasticity it can be shown that

Page 149: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 169 -

the error in using Jurawski’s formula for shear stress is negligible when the ratio of

cross-section height h to beam length l is limited as follows:

0.4h

l (7.40)

7.3.4 Shear Stress Distribution in Thin-Wall Cross-Sections

The compact sections are seldom used in the structural engineering. The majority of

beams used in steel structures are classified as beams with “thin-wall” cross-sections.

The thin-wall cross-section is categorized in (a) open and (b) closed cross-sections as

illustrated in Figures 7.19.a and 7.19.b, respectively.

Figure 7.19 Thin-Wall Cross-Sections

The two most common types of thin-wall cross-sections used in structural

engineering, tubular and wide-flange cross-sections, are illustrated in Figures 7.21.

The thin-wall cross-sections are characterized by a reduced aspect ratio between the

wall thickness and the overall dimensions of the cross-section. The characteristic ratio

0/ dt and ht f / of the tubular and wide-flange cross-section, respectively, are

extremely small.

Figure 7.20 Thin-Wall Beams

(a) Tubular and (b) Wide-Flange

7.3.4.1 Assumptions

The reduced thickness of the wall of the cross-section facilitates the introduction of

two important assumptions:

(a) the shear flow is always tangent to the local centerline of the cross-section;

Page 150: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 170 -

(b) the shear stress is constant in the thickness of the cross-section.

These two assumptions mimics the situation described in the case of the rectangular

cross-section.

The shear flow ),( sxq and the shear stress ),( sx characteristic to a thin-wall cross-

section are shown in Figure 7.21.

Figure 7.21 Shear Flow and Shear Stress in Thin-Wall Cross-Section

(a) Shear Flow and (b) Shear Stress

They are related through the following formula:

),(

),(),(

sxt

sxqsx (7.41)

7.3.4.2 Wide-Flange Cross-Section

The calculation of shear stress for wide-flange cross-section will closely follow the

methodology employed for the shear stress determination in the case of the

rectangular cross-section. A thin slice with length x is isolated from the body of the

beam. The normal stress and the shear flow are illustrated in Figure 7.22.a and 7.22.b,

respectively.

Figure 7.22 Isolated Slice of a Wide-Flange Beam

Page 151: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 171 -

The shear flow in the web is calculated using the equilibrium of the upper remaining

portion of the slice after further subdividing it with a horizontal plane cutting at

y above the NA as shown in Figure 7.23. The free-body diagram of the upper portion

is pictured in Figure 7.23.b

Figure 7.23 Free-Body Diagram and Notation

(a) Free-Body Diagram Perspective, (b) Free-Body Diagram Longitudinal View and

(c) Cross-Section View and Notation

The horizontal equilibrium equation of the isolated upper body is written as:

021 wFFF (7.42)

Using a rational similar to that used for the rectangular cross-section, equation (7.42)

is solved to obtain the distribution of the shear flow in the web:

z

A

zy

webxyI

SxVyxq

w'

*)(),(_ (7.43)

Using the notation from in Figure 7.23.c, the static moment 'wA

zS and the moment of

inertia zI calculated about the neutral axis are:

)4

(*24

)(*

2

]2

[*2

1*]

2[*]

22[*

2

1*]

22[*

2

222

'

yhthhb

yh

yh

thhhh

bS

www

www

wwA

zw

(7.44)

)***(*12

1 333

wwwz hthbhbI (7.45)

Incorporating equations (7.44) and (7.45) in equation (7.43) the following expression

for the web shear flow is obtained:

Page 152: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 172 -

]****

*4

***

***[*

2

)(*3),(

2

333

333

222

_

yhthbhb

t

hthbhb

hthbhbxVyxq

www

w

www

wwwy

webxy

(7.46)

Consequently, the shear stress in the web is obtained:

]****

*4

***

***[*

*2

)(*3),(

2

333

333

222

_

yhthbhb

t

hthbhb

hthbhb

t

xVyx

www

w

www

www

w

y

webxy

(7.47)

Equations (7.47) can be written as:

]*[**

)(*5.1),( 2

21_ yCCht

xVyx ww

w

y

webxy (7.48)

where the coefficients 1wC and 2wC are constants and depend only on the cross-section

dimensions.

Formula (7.47) indicates a parabolic variation of the shear stress xy valid in the web

for the following range of y values.

22

ww hy

h (7.49)

If the horizontal cutting plane passes into the flange the static moment 'fA

zS is

calculated as:

]4

[**2

1 22

'

yh

bS fA

z (7.50)

The resulting shear flow and shear stress equations follow:

]****

*4

***

*[*

2

)(*3),(

2

333

333

2

_

yhthbhb

b

hthbhb

hbxVyxq

www

www

y

flangexy

(7.51)

]****

*4

***

*[*

*2

)(*3),(

2

333

333

2

_

yhthbhb

b

hthbhb

hb

b

xVyx

www

www

y

flangexy

(7.52)

Page 153: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 173 -

Equation (7.52) shows that the shear stress flangexy _ is zero as expected for 2

hy

and reaches its maximum value at2

why . The shear stress calculated with equations

(7.47) and (7.52) exhibit a discontinuity at 2

why , the border between web and

flange. Equation (7.52) has been applied to a wide rectangular shape with ratio of

height to width in violation of the limits established in section 7.3.3.1. Consequently,

an erroneous linear variation between the zero value and the value calculated from the

equation (7.47) is suggested for the shear stress ),(_ yxflangexy in the flange.

The shear flow in the flange is obtained by sectioning the thin slice of beam with a

vertical plane as shown in Figures 7.22.a and 7.22.b. The shear horizontal force

fF necessary to equilibrate the forces 1F and 2F induced by the normal stress

variation is shown in Figure 7.24.a.

Figure 7.24 Free-Body Diagram and Notation

(a) Free-Body Diagram Perspective, (b) Free-Body Diagram Longitudinal View and

(c) Cross-Section View and Notation

The horizontal equilibrium equation is written:

021 fFFF (7.53)

The shear flow ),( yxq f in the flange calculated using the notation shown in Figure

7.24c is:

z

A

zy

fI

SxVyxq

f'

*)(),( (7.54)

The geometrical quantities used above are expressed functions of the cross-section

dimensions:

)***(*12

1 333

wwwz hthbhbI (7.55)

Page 154: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 174 -

)(*8

)22

(*** 22'''

w

f

fff

A

z hhsth

tsyAS f (7.56)

Inserting equations (7.55) and (7.56) into equation (7.54) the shear flow in the flange

is obtained:

sxVhthbhb

hhsxq y

www

w

flangezx *)(*]***

[*2

3),(

333

22

_

(7.57)

The shear stress calculated in accordance with the relation (7.41) is:

f

y

www

w

flangezxt

sxV

hthbhb

hhsx

*)(*]

***[*

2

3),(

333

22

_

(7.58)

The shear stress calculated in equation (7.58) acts on the vertical plane parallel to the

longitudinal axis of the beam. The duality of the shear stresses indicates that an equal

shear stress acts horizontally on the flange but in the plane of the cross-section. This

is a flangexz _ shear stress.

The variation of the shear stress xy in a particular cross-section is shown in Figure

7.25 and is obtained by giving particular values to y and s in equations (7.47) and

(7.58), respectively.

Figure 7.25 Shear Stress Distribution in Wide-Flange Cross-Section

From equations (7.47) and (7.58) it is evident that the shear stress varies linearly on

the flange and parabolicly on the web.

The resultant of the shear stress acting on the web is equal to:

)](*3

2[** minmaxmin wwwwwweb thT (7.59)

Page 155: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 175 -

Note: In the case of the structural steel shapes (W, S, etc.) the shear force taken by

the web accounts for more than 90% of the total shear force acting on the

entire cross-section. To obtain conservative results steel design standards

require the entire shear force to be sustained by the web alone.

7.3.4.3 Closed Thin-Wall Cross-Section

The most commonly used closed thin-wall cross-sections are the box and tubular

cross-sections depicted in Figure 7.19.b. The analysis of the distribution of shear

stress for a closed thin-wall cross-section is conducted in a similar manner with the

theoretical development employed for open cross-sections. A thin slice of length x is

isolated from the body of the closed thin-wall beam. As illustrated in Figure 7.26.b

the beam has the longitudinal vertical plane as a plane of symmetry. The shear flow

shown in Figure 7.26.b must to be directed in such a manner that the resultant is equal

to the vertical shear force acting on the cross-section.

Figure 7.26 Closed Thin-Wall Cross-Section

The main problem arising in evaluation of shear stress in closed thin-wall members is

how to position the cutting plane. If the vertical cutting plane passing through the axis

of symmetry oy is used the beam is separated in two symmetric halves. According to

assumption (a) stated in Section 7.3.4.1 the shear flow is tangent to the centerline of

the cross-section. Thus, for example, in Figure 7.26.c it is obviously that the shear

flow is horizontal at ends a andb . The shear flow must be zero in the plane of

symmetry and as a consequence of the duality principle the shear flow in the cross-

section at these locations is also zero. Since the shear flow is zero at the cross-section

points intersecting the axis of symmetry oy , the general distribution is obtained by

using cutting planes symmetrical to this axis and writing the horizontal equilibrium

for the beam segment contained between these planes.

The tubular circular cross-section illustrated in Figure 7.27 is used to exemplify the

methodology. Two symmetric cutting planes located by the angle are shown in

Figure 7.27.a. The free-body diagram obtained is pictured in Figure 7.27.b. The

equilibrium equation is written as:

12*2 FFH (7.60)

Page 156: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 176 -

Figure 7.27 Circular Thin-Wall Cross-Section

The unbalanced force H is expressed as:

xqH * (7.61)

Following the same steps as in the previous sections the shear flow is obtained as:

)(*2

),(*)(),(

xI

xSxVxq

z

Ay (7.62)

Consequently the shear stress is:

txI

xSxVx

z

Ay

yz*)(*2

),(*)(),(

(7.63)

Using the notation shown in Figure 7.27.c, the static moment ),( xS A and the

moment of inertia )(xI z relative to the neutral axis oz are calculated as:

sin***2),( 2 trxSA (7.64)

trxI z **)( 3 (7.65)

Substituting equations (7.64) and (7.65) into equations (7.62) and (7.63) the shear

flow and stress are calculated as:

r

xVxq

y

*

sin*)(),(

(7.66)

tr

xVx

y

yz**

sin*)(),(

(7.67)

Page 157: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 177 -

The angle varies in the interval from zero to 2

and, consequently, the shear flow

and stress attain their maximum values at the neutral axis. The maximum shear stress

is:

tr

xVx

y

yzyz**

)()

2,(max_

(7.68)

If the expression of the area is employed equation (7.68) becomes:

)(

)(*2)

2,(max_

xA

xVx

y

yzyz

(7.69)

where trxA ***2)( is the area of the cross-section.

The distribution of the shear flow is illustrated in Figure 7.28.

Figure 7.28 Shear Flow Distribution

7.4 Built-Up Beam Connectors

The cross-sections of the beams discussed in the previous sections are considered to

be made from continuous material without any interruptions. They are called

homogeneous cross-sections. In structural engineering practice the compact cross-

sections, which are homogeneous, are seldom used in steel structures. For shorter

span beams or lighter loads the wide-flange homogeneous beams are commonly used.

In general, the industrial steel buildings are subjected to heavy loads and the most

commonly used cross-sections, such as the wide-flange and closed thin-wall, are

fabricated from parts joined together by connectors. These types of beams are called

build-up members. The connectors may be glue, welds, rivets, bolts or nails. Glue

(adhesives) and nails are connectors proper to be used in the wood built-up members,

Page 158: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 178 -

while welds, rivets and bolts are commonly used for the steel built-up members.

Examples of build-up members are illustrated in Figure 7.29.

Figure 7.29 Built-Up Beams

(a) Glued Wood Beam, (b) Box Wood Beam, (c) Welded Steel Beam and

(d) Reinforced Steel Beam

The first two cross-sections shown in Figures 7.29.a and 7.29.b are wood fabricated

beams constructed using glue or nails, respectively. Figures 7.29.c and 7.29.d are

illustrating steel beams constructed using wells or bolts.

The shear flow induced by nonuniform bending must be transferred in-between the

adjoining parts comprising the beam cross-section. This shear transfer is realized in

three ways: (a) by shear distributed over the interface areas, (b) by shear distributed

along a connector line and (c) by discrete shear connectors. The shear transfer type (a)

is common for the laminated wood beams composed of multiple pieces as shown in

Figure 7.29.a. The linear transfer, type (b), is proper to the welded beams as pictured

in Figure 7.29.c. The discrete type shear transfer, shown in Figures 7.29 .b and 7.29.d,

are pertinent to bolt or nail connectors. There are special cases where welds are

discontinuously applied, referred to as stitch welds.

The shear flow determined using equation (7.31) is employed to calculate the required

shear flow recq , which represents the shear force per unit length that must be

transferred from one part of the cross-section to the adjacent part under the given

loading condition. The required shear is calculated as:

z

y

z

A

zy

recI

yAxV

I

SxVq

**)(*)( (7.70)

where A is the area on which the unbalanced flexural stresses are exerted.

7.4.1 Linear Shear Connectors

The required shear flow calculated with equation (7.69) determines the actual

necessary force per unit length which must be transmitted by the weld. The shear

capacity of the weld weldq must be larger than the required shear flow:

Page 159: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 179 -

weldreq qq (7.71)

The capacity of the weld is determined from laboratory test data.

7.4.2 Discrete Shear Connectors

The discrete connectors have a shear capacity connectorQ , also established by laboratory

testing, and are assumed spaced at a equal distance s . For the discrete connector to

transfer the shear the following relation must be maintained:

connectorreq Qsq * (7.72)

Equation (7.72) is applied in different zones of the beams, considering in the

calculation the maximum value of the vertical shear force in the particular zone. This

way, the distance between the connectors may be increased or decreased within the

various zones of the beam in accordance with the total shear flow requirement for the

zone.

7.4.3 Examples

Two examples of shear connectors are investigated in the following sections. The

first, contained in section 7.4.3.1 exemplifies the requirements for welds within a

built-up steel beam. The second, demonstrates the calculation of discrete connectors,

as applied in the design of a nail connected wooden box beam.

7.4.3.1 Welded Connection

The build-up beam is assembled from three rectangular steel plates connected by

welds. The dimensions of the steel plates and the notation are shown in Figure 7.30.

The cross-section is subjected to a vertical shear force 45yV .kN

Figure 7.30 Steel Build-Up Beam

Page 160: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 180 -

The required shear flow is calculated using equation (7.64):

z

y

z

A

zy

reqI

yAxV

I

SxVq

**)(*)(

Note: The area considered in the calculation is the flange which during the bending

tries to slip relatively to the web.

Using the dimensions shown in Figure 7.30 the static moment and the moment of

inertia are calculated:

3

1

5.377437)2

15

2

290(*)165*15(

)22

(*)*(*

mm

thbtyAS

f

fflange

flange

z

483

23

3

121

3

10*3045.112

5.7*290])

2

15

2

290(*)165*15(

12

165*15[*2

12

*])

22(*)*(

12

*[*2

mm

ththbt

btI

f

f

f

z

The required shear flow is:

130.010*3045.1

5.377437*458reqq mmkN /

Two welds are located at the interface between the flange and the web and,

consequently, they share the required shear flow. The shear flow in the weld is:

065.02

req

weld

qq mmkN /

7.4.3.2 Nail Connection

The cross-section of a built-up wooden box type beam is illustrated in Figure 7.31.

The beam is assembled from four planks of wood having the dimensions indicated in

the figure. The connection between planks is realized by employing nails along the

four interfaces. The capacity of each nail is NQnail 800 . A vertical shear force

kNVy 15 acts on the cross-section.

Page 161: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 181 -

Figure 7.31 Wood Box Beam

The shear flow is transferred between the upper flange and the two webs and is

calculated as:

z

y

z

A

zy

reqI

yAxV

I

SxVq

**)(*)(

The static moment and the moment of inertia are obtained as:

3 864000)2

40

2

280(*)40*180( mmS A

z

4833

10*642.212

180*200

12

210*280mmI z

The required shear flow is calculated as:

mmkNqreq / 0491.010*642.2

10*64.8*158

5

There are two rows of nails at this location of the cross-section. The nails are spaced

at the unknown distance s along the length of the beam.

mmkNq rownailreq / 0246.02

0491.0__

Using equation (7.72) the distance between the nails is calculated as:

mmq

Qs

rownailreq

nail 52.326.24

800

__

Page 162: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 182 -

LECTURE 8 Beam Deflection

8.1 Introduction

A beam with a straight longitudinal axis subjected to transversal loads acting in its

longitudinal plane of symmetry deforms and the resulting curve is called the

deflection curve. In the previous lecture, the curvature of the deflection curve was

used to determine the stress and strain distribution in the cross-section of a beam

under the restrictions established for pure and nonuniform bending. In this lecture, the

equation of the defection curve will be derived and consequently, the displacements at

any point of the beam may be calculated. The calculation of beam deflections is an

important part of structural analysis and design. The deflections are limited to

prescribed tolerances imposed by the functionality of the particular structural element.

The cantilever beam pictured in Figure 8.1 deforms in its vertical plane of symmetry

under the action of the exterior transversal loading.

Figure 8.1 Example of Beam Deflection

The deflection curve and its slope are mathematically represented by the real

functions )(xv and )(x , respectively. The vertical distance measured from a given

point located on the undeformed axis of the beam to the corresponding point located

on the deflection curve is called the transverse displacement.

Page 163: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 167 -

8.2 Qualitative Interpretation of the Deflection Curve

For the general case of nonuniform bending the equation relating the bending moment

)(xM z with the radius of curvature )(x is:

)(*)(

)()( xI

x

xExM zz

(8.1)

where )(xE - modulus of elasticity;

)(xI z - moment of inertia about the horizontal centroidal axis of the cross-

section;

)(xM z - cross-section bending moment;

)(x - radius of curvature of the deflection curve.

Note: If the material is homogeneous linear elastic and the beam is of uniform cross-

section the product zIE * is called the flexural rigidity and is constant.

From the relation given by equation (8.1), the deflection )(xv can be anticipated

using the sign convention established in Lecture 7 and re-plotted in Figure 8.2. The

moment diagram is always plotted at the beam side where the fibers are subjected to

tension, while the curvature center is placed on the opposite side. When the bending

moment is positive the deflection curvature is concave, while for the negative bending

moment the deflection curvature is convex. At the supports the deflection must

correspond to the prescribed support constraint condition. Locations where the

deflection curve changes from concave to convex, or vice versa, are called inflection

points.

Figure 8.2 Sign Convention for Beams in Bending

An example of qualitative interpretation of the deflection curve is shown in Figure

8.3. The moment diagram changes from positive to negative in the interval AB at

Page 164: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 168 -

point D. In between points A and D the moment is positive and the curvature center is

located above the deflection curve. Between points D and B the moment diagram is

negative and, thus, the curvature center is located below the deflection curve. The

change in curvature takes place at point D, which is an inflection point. On the

overhanging end BC, the moment is negative and the curvature is convex.

Considering that at the supporting points the beam is constrained not to move

vertically, the qualitative deflection diagram, as shown in Figure 8.3, can be sketched.

Figure 8.3 Example of Qualitative Interpretation of Deflection Curve

8.3 Differential Equations of the Deflection Curve

From Calculus it is known that the slope of a real function at any particular point of

its defined continuity interval is the first derivative of the function. For the case in

point, the real function is represented by the beam displacement curve )(xv . Using the

notation shown in Figure 8.1 the following relation is written:

dx

xdvx

)()(tan (8.2)

Under the assumption of small displacements, the first derivative is a very small

value:

1)(

dx

xdv (8.3)

consequently, the angle can approximate by its tangent:

dx

xdvxx

)()(tan)( (8.4)

In Calculus the relation between the radius of curvature )(x and the deflection )(xv is

established as:

Page 165: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 169 -

2

3

2

2

2

]))(

(1[

)(

)(

1

dx

xdv

dx

xvd

x

(8.5)

Under the small displacement assumption (8.3), equation (8.5) can be simplified and

written as:

2

2 )(

)(

1

dx

xvd

x

(8.6)

Substituting equation (8.6) into equation (8.1) yields the following equation:

)(*)(

)()(2

2

xIxE

xM

dx

xvd

z

(8.7)

The moment-curvature equation, a second order differential equation, is obtained

from equation (8.7) and expressed in the following standard form:

"*)(*)()( vxIxExM z (8.8)

where the notation 2

2" )(

dx

xvdv is employed.

Consider the differential relation between the transverse loading and the cross-section

stress resultants, )(xV and )(xM , previously obtained from the equilibrium of the

beam infinitesimal volume element:

)()(

xpdx

xdVn (8.9)

)()(

xVdx

xdM (8.10)

where )(xV and )(xM are the vertical shear force and the bending moment,

respectively.

Differentiating the moment-curvature equation (8.8) and using equations (8.9) and

(8.10) the shear-deflection and load-deflection equations are obtained:

]'"*)(*)([)( vxIxExV z (8.11)

]""*)(*)([)( vxIxExp zn (8.12)

Page 166: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 170 -

The load-deflection equation (8.12) is a forth-order differential equation.

If the beam is made from a homogeneous linear elastic material and has a constant

cross-section in the continuity interval of the bending moment or transversal loading

equations (8.8) and (8.12) became:

"**)( vIExM z (8.13)

IV

zn vIExp **)( (8.14)

The general theory of differential equations with constant coefficients can be

employed to obtain the solution of the second-order differential equation (8.8) and

forth-order differential equation (8.12). The continuity intervals pertinent to the

functions involved in these differential equations, the bending moment and transverse

load, must be recognized in order for the integration process to be properly

conducted.

8.3.1 Integration of the Moment- Curvature Differential Equation

Integration of the second-order differential equation (8.8) on an interval of continuity

for the bending moment )(xM z yields the following relations:

1

' *)(*)(

)()()( Cdx

xIxE

xMxvx

z

z

(8.13)

21 ***)(*)(

)()( CxCdxdx

xIxE

xMxv

z

z (8.14)

For the case of a beam with constant bending stiffness constIE z )*( , the integrals

expressed in equations (8.13) and (8.14) may be simplified as follows:

1

' *)(*

1)()( CdxxM

IExvx z

z

(8.15)

21 **)(*

1)( CxCdxxM

IExv z

z

(8.16)

The integration constants 1C and 2C are calculated by imposing the boundary

conditions of the specific problem at hand.

Page 167: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 171 -

8.3.2 Integration of the Load-Deflection Differential Equation

Successive integration of the forth-order differential equation (8.12) in a continuity

interval of the transverse loading function yields the following relations:

1

'" *)(])(*)(*)([)( CdxxpxvxIxExV nz (8.17)

21

" ***)()(*)(*)()( CxCdxdxxpxvxIxExM nz (8.18)

321 *)(*)(

1**

)(*)(*

*]**)([)(*)(

1)()(

CdxxIxE

CdxxIxE

xC

dxdxdxxpxIxEdx

xdvx n

z

(8.19)

43

21

***)(*)(

1

**)(*)(

1***

)(*)(*

**]**)([)(*)(

1)(

CxCdxdxxIxE

dxdxxIxE

CdxdxxIxE

xC

dxdxdxdxxpxIxE

xv n

z

(8.20)

The integration constants 1C , 2C , 3C and 4C are calculated by imposing the boundary

conditions applicable to the specific problem.

Note: The moment-curvature equation (8.8) can be used only if the bending moment

variation is known, the case for statically determinate structures. The load-deflection

curve equation (8.14) requires only that the variation of transverse load be known

and, consequently, can be used for either statically determinate or indeterminate

beams.

For the case of a beam with constant bending stiffness constIE z )*( , the integrals

expressed in equations (8.17) through (8.20) may be considerably simplified as

follows:

1

''' *)()(**)( CdxxpxvIExT nz (8.21)

21

" ***)()(**)( CxCdxdxxpxvIExM nz (8.22)

32

2

1 ]*2

*

***)([*

1)()(

CxCx

C

dxdxdxxpIEdx

xdvx n

z

(8.23)

Page 168: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 172 -

43

2

2

3

1 *]2

*6

*

****)([*

1)(

CxCx

Cx

C

dxdxdxdxxpIE

xv n

z

(8.24)

8.3.3 Boundary and Continuity Integration Conditions

As previously stated, the functions (8.15) though (8.21) obtained above, represent the

general solutions. Only after the boundary conditions are imposed and the integration

constants determined the solutions became representative for a specific case study.

The boundary conditions commonly encountered in the application of the equations

(8.15) and (8.16) or equations (8.21) through (8.24) are presented in Table 8.1.

In general, the transverse load )(xpn and the bending moment )(xM functions are

described for a particular case by a number of continuity intervals. Therefore, the

continuity conditions at the common ends of the intervals must be described in order

for the constants to be calculated. Each continuity interval is treated as an independent

interval and the boundary and continuity conditions are applied. The most common

continuity conditions are summarized in Table 8.2.

Table 8.1 Boundary Conditions

Page 169: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 173 -

Table 8.2 Continuity Conditions

8.4 Examples

The methodology used in the application and integration of the moment-curvature

(8.8) and load-deflection (8.14) differential equations is illustrated in examples 8.4.1

and 8.4.2, respectively.

8.4.1 Application of the Moment-Curvature Equation

The deflection curve is required for a cantilever beam subjected to a concentrated

force BP and a concentrated bending moment BM both acting at the tip of the beam,

point B. The beam is characterized by a constant cross-section along its entire length,

with geometry and loading as shown in Figure 8.4.

zIE * constant (8.25)

The corresponding reactions, AP and AM , are found using the equilibrium equations:

BA PP (8.26)

LPMM BBA * (8.27)

Page 170: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 174 -

Figure 8.4 Cantilever Beam

The moment-curvature equation (8.8) requires knowledge of the bending moment

function and identification of the continuity intervals. For the case in point, the

bending moment )(xM diagram is continuous on the entire length of the beam and is

expressed as:

)(**)( xLPMxPMxM BBAA (8.28)

Using equation (8.28) in the moment-curvature equation (8.13), the problem specific

differential equation is obtained:

xIE

P

IE

Mx

IE

P

IE

M

IE

xM

dx

xvdv

z

B

z

A

z

A

z

A

z

***

****

)()("

2

2

(8.29)

Integrating the differential equation (8.29) twice yields the following expressions for

deflection and slope:

1

2***2

**

)()( Cx

IE

Px

IE

M

dx

xdvx

z

B

z

A (8.30)

21

32 ****6

***2

)( CxCxIE

Px

IE

Mxv

z

B

z

A (8.31)

The integration constants 1C and 2C are identified using the boundary conditions at

point A:

for 0x 0)0( x (8.32)

0)0( xv (8.33)

Page 171: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 175 -

Solving the algebraic equations (8.32) and (8.33) by substitution of equations (8.30)

and (8.31) the integration constants 1C and 2C are found as:

01 C (8.34)

02 C (8.35)

Substituting equations (8.34) and (8.35) into equations (8.30) and (8.31), the final

deflection and slope expressions are obtained:

)2

(***

**

)(x

LxIE

Px

IE

Mx

z

B

z

B (8.36)

)3

(****2

***2

)( 22 xLx

IE

Px

IE

Mxv

z

B

z

B (8.37)

The maximum deflection value is obtained at the tip of the cantilever:

]3

*2*[*

**2

2 LPM

IE

Lv BB

z

B (8.38)

and the corresponding rotation

]2

*[**

LPM

IE

LBB

z

B (8.39)

The deflection curve is a cubic (third-order) polynomial and is schematically plotted

in Figure 8.5

Figure 8.5 Deflection curve

Equations (8.38) and (8.39) may be written for the case when only the concentrated

force BP is considered:

B

z

B PIE

Lv *

**3

3

(8.40)

Page 172: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 176 -

B

z

B PIE

L*

**2

2

(8.41)

In the absence of the concentrated force BP equations (8.38) and (8.39) become:

B

z

B MIE

Lv *

**2

2

(8.42)

B

z

B MIE

L*

* (8.43)

8.4.2 Application of the Load-Deflection Equation

The simply supported beam shown in Figure 8.6 is subjected to a concentrated

vertical force BP acting at distance a from the fixed support A. The beam has a

constant flexural rigidity along its entire length.

zIE * constant (8.44)

The reaction force and moment at point A are obtained by solution of the equilibrium

equations:

BA PP (8.45)

aPM BA * (8.46)

Figure 8.6 Cantilever Beam

The transverse load has two continuity intervals, AB and BC, and, consequently, the

forth-order differential equation must be integrated for each one of them as:

Page 173: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 177 -

For interval AB (the interval origin is point A) the transversal load is zero:

0)( xp AB (8.47)

Using equation (8.47) in equations (8.21) through (8.24) yields the following:

AB

ABzAB CxvIExV 1

''')(**)( (8.48)

ABAB

ABzAB CxCxvIExM 21

" *)(**)( (8.49)

ABABAB

z

AB CxCx

CIE

x 32

2

1 ]*2

*[**

1)( (8.50)

ABABABAB

z

AB CxCx

Cx

CIE

xv 43

2

2

3

1 *]2

*6

*[**

1)( (8.51)

For interval BC (the interval origin is point B) the transversal load is also zero:

0)( xpBC (8.52)

Equations (8.21) through (8.24) are written considering equation (8.52) as:

BC

BCzBC CxvIExV 1

"' )(**)( (8.53)

BCBC

BCzBC CxCxvIExM 21

" *)(**)( (8.54)

BCBCBC

z

BC CxCx

CIE

x 32

2

1 ]*2

*[**

1)( (8.55)

BCBCBCBC

z

BC CxCx

Cx

CIE

xv 43

2

2

3

1 *]2

*6

*[**

1)( (8.56)

Eight (8) integration constants must be calculated. Four (4) boundary conditions (at

points A and C) and four (4) continuity conditions (at point B) are employed. The

boundary conditions are expressed as:

Point A at x=0 (interval AB)

0)0( xAB 03 ABC (8.57)

0)0( xvAB 04 ABC (8.58)

Page 174: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 178 -

Point C at x=b (interval BC)

0)( bxVBC 01 BCC (8.59)

0)( bxM BC 02 BCC (8.60)

The continuity conditions at point B (both intervals) are:

)0()( xvaxv BCAB

BCABAB

z

Ca

Ca

CIE

4

2

2

3

1 ]2

*6

*[**

1 (8.61)

)0()( xax BCAB

BCABAB

z

CaCa

CIE

32

2

1 ]*2

*[**

1 (8.62)

)0()( xMaxM BCAB 0* 221 BCABAB CCaC (8.63)

)0()( xVPaxV BCBAB 011 BC

B

AB CPC (8.64)

Solving the system of algebraic equations (8.57) through (8.64) the integration

constants are calculated:

B

AB PC 1 aPC B

AB *2 03 ABC 04 ABC (8.65)

01 BCC 02 BCC z

BBC

IE

aPC

**2

* 2

3 z

BBC

IE

aPC

**3

* 3

4

(8.66)

Substituting the integration constants (8.65) and (8.66) into equations (8.48) through

(8.51) and (8.53) through (8.56), respectively, the variation of the shear force, bending

moment, slope and deflection for each interval are obtained:

For the interval AB (the interval origin is point A)

BAB PxV )( (8.67)

)(*)( xaPxM BAB (8.68)

)2

(**

*)(

xa

IE

xPx

z

BAB (8.69)

Page 175: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 179 -

)3

(***2

*)(

2x

aIE

xPxv

z

BAB (8.70)

For the interval BC (the interval origin is point B)

0)( xVBC (8.71)

0)( xM BC (8.72)

z

BBC

IE

aPx

**2

*)(

2

(8.73)

)3

*2(*

**2

*)(

2a

xIE

aPxv

z

BBC (8.74)

The deflection and slope at points B and C are:

z

BB

IE

aPv

**3

* 3

(8.75)

z

BB

IE

aP

**2

* 2

(8.76)

)3

(**2

* 2a

LIE

aPv

z

BC (8.77)

z

BC

IE

aP

**2

* 2

(8.78)

Page 176: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 180 -

LECTURE 9

Torsion

9.1 Introduction

In the physical sense, torsion of a linear member refers to the action of twisting

(rotation) of a structural member about its longitudinal axis when subjected to a

twisting moment called torque. The twisting moment is a vector collinear with the

member longitudinal axis. Examples of torque are illustrated in Figure 9.1.

Figure 9.1 Torque Examples

In engineering practice, a linear member subjected to torsion is named torsional

member or shaft. The geometrical shape of the member subjected to torsion plays an

important role in the resulting deformation. To substantiate this assertion, two

examples of shafts in torsion, one with circular cross-section and the other with

Page 177: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 181 -

rectangular cross-section, are illustrated in Figures 9.2 and 9.3. To better identify the

deformation, a two-directional mesh is marked on the lateral surfaces of both

members. The mesh is photographed before and after the application of the torque at

the left end of the member, while the right end is kept fixed.

Figure 9.2 Torsional Deformation of a Bar with Circular Cross-Section

Figure 9.3 Torsional Deformation of a Bar with Rectangular Cross-Section

Examination of Figure 9.2 reveals that adjacent cross-sections rotate about the

longitudinal axis relative to one another without undergoing other type of visible

deformation. In contrast, the member with rectangular cross-section, illustrated in

Figure 9.3, suffers a visible warping of the cross-section in addition to the relative

rotation of adjacent cross-section.

This is an indication that only the torsional-deformation of the members with circular

cross-sections can be treated exactly using simple kinematic assumptions. Members

having more complicated cross-sections solved using the advanced methods of the

Theory of Elasticity.

9.2 Torsional Deformation of a Member with

Circular Cross-Section

9.2.1 Strain-Displacement Equation

Following the observations derived from Figure 9.2 the torsional-deformation of a

member with circular cross-section is schematically illustrated in Figure 9.4.

Page 178: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 182 -

Figure 9.4 Torsional-Deformation of a Member with Circular Cross-Section

The radial plane ABEF is located in the body of the circular cross-section beam

before the application of torque. After the application of a torsional load T and the

associated deformation, at the fixed end the radius AB remains in its original position,

while the radius EF at the free end rotates to a new position *EF . Similarly, in a

particular cross-section identified by the coordinate x the radius CD rotates, after the

deformation, to the position *CD .The rotation angle measured between the original

position and the deformed position is notated )(x . Consequently, the torsional

deformation of a member with circular cross-section is described by three

fundamental kinematic assumptions.

It is necessary to establish a consistent sign convention for which will be used

throughout the theoretical development. The torque )(xT and rotation angle )(x are

positive on the cross-section if they rotate in the right-hand rule sense of the outer

normal to the cross-section. The sign convention established for the torque )(xT and

angle of rotation )(x is shown in Figure 9.5.

Definition 9.1

A plane linear member when subjected to torsional moment undergoes a torsional

deformation if after the deformation:

(c) the axis of the member remains straight and without longitudinal

extension;

(d) the cross-sections remain plan and perpendicular to the longitudinal

axis of the beam;

(e) radial lines remain straight and radial as the cross-section rotates about

the longitudinal axis of the member.

Page 179: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 183 -

Figure 9.5 Torque and Rotation Angle Sign Convention

To study the relation between the rotation angle and the shear strain the notation

shown in Figure 9.6.a is employed. Consider the elementary volume of length x and

radius isolated from the body of the beam as shown in Figure 9.6.b.

Figure 9.6 Torsional-Deformation of the Elementary Element

After the rotation of the cross-sections, the original right angle QRS deforms into

angle Q*R*S*, which is no longer a right angle. The shear strain is expressed as:

**'***

2),( SRSSRQx

(9.1)

where x is the position of the cross-section and is varies between zero and the

cross-section radius.

Since ),( x is a small angle its value can be approximated by its tangent:

dx

d

xSR

SSx xx

*

*limlim),( 0'*

'*

0

(9.2)

Equation (9.2) is called the strain-displacement equation for torsional deformation

of a circular cross-section. The shear strain ),( x is a function of the cross-section

position x and the distance measured radially from the center of the circular cross-

section. In a particular cross-section, the shear strain ),( constx varies linearly

with and reaches its maximum value on the cross-section periphery.

Page 180: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 184 -

Figure 9.7 Shear Strain Distribution for circular Cross-Sections

(a) Solid Cross-Section and (b) Tubular Cross-Sections

The shear strain distribution pertinent to circular and tubular circular cross-sections is

depicted in Figure 9.7.

9.2.2 Constitutive Equation

The constitutive equation reflects, as was describe in Lecture 2, the functional

relationship between stress and strain. If the homogeneous linear elastic material

behavior is considered, the relationship between shear stress and strain is written

according to Hook’s Law as:

),(*),( xGx (9.3)

The material constant G is called shear modulus. Equation (9.3) indicates that the

shear stress follows the linear distribution of the shear strain when the material is

elastic and homogeneous. The distribution of the shear stress in solid and tubular

cross-sections is shown in Figure 9.8.

Figure 9.8 Shear Stress Distribution for circular Cross-Sections

(b) Solid Cross-Section and (b) Tubular Cross-Section

Substituting equation (9.2) into equation (9.3), the shear stress is expressed as a

function of the angle of rotation as:

Page 181: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 185 -

dx

dGx

**),( (9.4)

Accordingly with the consideration of the duality of shear stress principle, the

torsional deformation induced cross-section shear stress demands the existence of

shear stress acting in the radial-longitudinal planes as illustrated in Figure 9.9. The

shear stress and strain shown in the cross-sectional and longitudinal planes are

pictured in Figures 9.9.a and 9.9.b, respectively.

Note: The orientation of the in-plane cross-section shear stress results from the

positive direction of the torque acting on the cross-section. The orientation of the

longitudinal shear stress on radial planes shown in Figure 9.9.b is a direct result of the

duality principle.

Figure 9.9 Shear Stress and Strain Distribution in Radial Planes

9.2.3 Cross-Section Stress Resultants

The torque )(xT is related with the shear stress ),( x by the following integral

equation:

A

dAxxT *),(*)( (9.5)

Introducing equation (9.4) into equation (9.5):

p

AA

Idx

dGdA

dx

dGdA

dx

dGxT ******]**[*)( 2

(9.6)

where AdIA

p *2 is the polar moment of inertia of the cross-section.

Equation (9.6) can be rewritten as:

Page 182: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 186 -

pIG

xT

dx

d

*

)(

(9.7)

Equation (9.7) is called the torque-twist equation. Considering the torque )(xT as a

continuous real function on a given interval of continuity, equation (9.7) can be

integrated and the angle of rotation at a particular cross-section is calculated as:

)0(**

)()( xdx

IG

xTx

p

(9.8)

The total angle of twist between the ends of the continuity interval is obtained as the

difference between the angles of rotation at both ends:

L

p

total dxIG

xT

0

**

)( (9.9)

where L is the length of the member.

By combining equations (9.4) and (9.7), a new expression for the shear stress is

obtained:

*)(

),(pI

xTx (9.10)

Equation (9.10) is widely used in the analyses of beams with circular cross-sections.

The maximum value for the shear stress is obtained as:

RI

xTRxx

p

*)(

),()(max (9.11)

where R is the exterior radius of the cross-section.

9.3 Torsional Deformation of a Member with

Closed Thin-Wall Cross-Section

In the previous section, the primary focus of the theoretical development was on

members with solid or thick walled hollow cross-sections. In structural engineering

and especially in the aerospace engineering the usage of members with a closed thin-

wall cross-section is prevalent. The torsional deformation of thin-wall members can

be solved if some simplifying assumptions are imposed. These simplifying

assumptions are:

(a) the member is cylindrical and the cross-section does not vary along the

length;

Page 183: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 187 -

(b) the cross-section is closed and is comprised of a single-cell with a

small thickness relative to the general dimensions of the cross-section;

(c) the shear stress is constant through the thickness and parallel to the

median curve defining the cross-section;

(d) the member is subjected to end torque only;

(e) the warping of the cross-section is unrestrained at both ends;

The first three assumptions, (a), (b) and (c) are common for the definition of the thin-

wall cross-section and were used before, while the other two have a specific

application for torsion investigation. An example of a single-cell thin-wall torsional

member is illustrated in Figure 9.10.

Note: An important aspect of this discussion is that the cross-section is not required

to be circular.

Figure 9.10 Single Cell Thin-Wall Cross-Section

(a) Geometry and Notation, (b) Shear Flow Distribution and

(c) Shear Stress Distribution

The shear flow q is defined as:

tq * (9.12)

where is the shear stress.

Figure 9.11 Free-Body Diagram of the Thin-Wall Element

Consider the elementary volume element ABCD, shown in Figure 9.10.a and

magnified for clarity in Figure 9.11, isolated from the body of a cylindrical thin-wall

Page 184: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 188 -

member subjected to torques at both ends. As a result of the duality of shear stress

principle, the existence of shear flow in the transverse cross-sections induces shear

flows in the longitudinal direction. The equilibrium of the shear forces in the

longitudinal direction, in the absence of any forces acting on the exterior surface

planes, is written as:

024 VV (9.13)

Using the notation shown in Figure 9.11, the equilibrium equation (9.13) shows that:

BA qq (9.14)

Equality (9.14) implies that the shear flow q is constant on the entire cross-section

and, consequently, the shear stress is expressed as:

constt

q (9.15)

Equation (9.15) reflects the underlying assumption c that the shear stress is constant

across the thickness of the thin-wall.

Using the notation shown in Figure 9.12, the integral relation between the torque

T and the shear flow q is written as:

mmm CC

sC

dsqdsqFdT ***** (9.16)

It is evident from Figure 9.12.b that the following geometrical relation holds:

dsdA **2

1 (9.17)

Figure 9.12 Thin-Wall Cross-Section Notation

Substituting equation (9.17) into the equilibrium equation (9.17) the torque T is

expressed as:

Page 185: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 189 -

mAqT **2 (9.18)

where mA is the area enclosed by the median curve.

From equation (9.18) the shear flow is obtained as:

mA

Tq

*2 (9.19)

The shear stress is calculated by substitution of equation (9.19) into equation

(9.15):

tA

T

m **2 (9.20)

9.4 Torsional Deformation of a Member with Rectangular

Cross-Section

As illustrated in Figure 9.3 the prismatic member suffers not only relative the rotation

of adjacent cross-sections, but also warping of the cross-sections when subjected to

torque application. This type of torsional problem is exactly solved using methods of

the Theory of Elasticity. Without going into theoretical details, only some practical

formulae are presented in this section. The distribution of shear stress for a

rectangular cross-section is illustrated in Figure 9.13.

Figure 9.13 Shear Stress Distribution In a Rectangular Cross-Section

It is notable that the shear stress attains a the maximum value at the middle of the

longer edge, while becoming zero at the corners. The maximum shear stress

corresponding to a rectangular cross-section is expressed as:

2max

** td

T

(9.21)

Page 186: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 190 -

where is the ratio of the dimension d to t satisfying the relation 1td .

The rotation angle is calculated:

JG

LT

*

* (9.22)

where the torsional moment of inertia J is obtained as:

3** tdJ (9.23)

The constants and are listed in Table 9.1.

Table 9.1 Torsional Constants for Rectangular Cross-Section

9.5 Torsional Deformation of a Member With Elliptical

Cross-Section

The shear stress distribution in an elliptical cross-section is illustrated in Figure 9.14.

The maximum value occurs on the periphery of the cross-section at the two ends of

the ellipse minor axis and is calculated using the following formula:

2max

**

*2

ba

T

(9.24)

Figure 9.14 Shear Stress Distribution in the Elliptical Cross-Section

The rotation angle for a member with elliptical cross-section is calculated as:

JG

LT

*

* (9.25)

Page 187: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 191 -

where the torsional moment of inertia J is obtained as:

22

33 **

ba

baJ

(9.26)

9.6 Examples

9.6.1 Rectangular Thin-Wall Cross-Section

An unknown torque T is applied to an aluminum member with a rectangular thin-

wall cross-section. The geometry of the cross-section is illustrated in Figure 9.15. The

allowable shear stress for the aluminum is 110all MPa .

Figure 9.15 Thin-Wall Rectangular Cross-Section

The unknown torque T is calculated using equation (9.20):

allm tAT ***2

The median curve encloses the area mA as shown in Figure 9.15.b:

232 10*016.2201636*56)2

4

2

432(*)

2

4

2

452( mmmAm

The thickness t of the tube is:

mmmt 004.04

Consequently, the allowable torque for the cross-section is calculated as:

mNT *08.1774)10*110(*)004.0(*)10*016.2(*2 63

If the tube was manufactured such that the thicknesses of the walls varies as depicted

in Figure 9.16, the torque T is calculated as:

Page 188: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 192 -

qAT m **2

Figure 9.16 Imperfect Thin-Wall Cross-Section

The shear flow q , which is accordingly to equation (9.15) constant, is determined

using the minimum thickness as follows:

mNtq allall /10*85.3)10*110(*0035.0* 56

min

The allowable torque is then calculated as:

mNqAT allm *32.1552)10*110(*)0035.0(*)10*016.2(*2**2 63

9.6.2 Capacity Comparison of Different Shaped Solid Cross-Sections

Three types of solid cross-section shapes are illustrated in Figure 9.17. The allowable

shear stress is the same for all the cross-sections considered.

Figure 9.17 Solid Cross-Sections

(a) Circular, (b) Square and (c) Rectangular

The torsional capacity allT of the circular cross-section is calculated as:

c

IT

allp

all

*

The polar moment of inertia is obtained as:

Page 189: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 193 -

2

*

32

)2(* 44 ccI p

Consequently, the torsional capacity is:

allall

allp

all cc

c

IT

**57.1*

2

**3

3

The torsional capacity bT of the squared cross-section is calculated using the formula

(9.19) and Table 9.1 as:

allallb aaaT **208.0***208.0 32

The torsional capacity cT of the rectangular cross-section is:

allallc aaaT **141.0*)*5.0(*)*2(*282.0 32

If the condition of equal areas is imposed, the radius of the circular cross-section is

found as:

ac *5642.0

and the capacity aT of the circular cross-section becomes:

allalla acT **282.0**57.1 33

To compare these three cross-sections, the following ratios are calculated considering

the torsional capacity of the circular cross-section as the base for comparison:

7376.0**282.0

**208.01

3

3

all

all

a

b

a

a

T

Tratio

500.0**282.0

**141.02

3

3

all

all

a

b

a

a

T

Tratio

Thus, it is shown that the most effective cross-section for torsional-deformation is the

circular cross-section.

Page 190: 179306253 Lectures in Mechanics of Materials Ion Simulescu

- 194 -

LECTURE 10

Plane Stress Transformation

10.1 Introduction

As discussed in Lecture 2, the general three-dimensional state of stress is obtained, as

was discussed in Lecture 2, by passing a set of three orthogonal planes though a point

of the solid body and isolating the infinitesimal volume located around the point. The

three-dimensional state of stress is illustrated in Figure 10.1 where, for clarity, only

the stresses drawn on the faces with positive normal are represented.

Figure 10.1 Three-Dimensional State of Stress

Mathematically the three-dimensional state of stress is represented by a generalized

stress tensor ),,( zyxT defined by nine distinct stress components:

zzyzx

yzyyx

xzxyx

zyxT

),,( (10.1)

Applying the shear stress duality principle the number of independent tensorial

components is reduced from nine to six and, consequently, the symmetric shear stress

components are related as:

yxxy (10.2)

Page 191: 179306253 Lectures in Mechanics of Materials Ion Simulescu

195

zxxz (10.3)

zyyz (10.4)

There are instances when some of the stress tensor components vanish and the general

three-dimensional stress tensor ),,( zyxT degenerates into a tensor characterized by

only three independent components. This condition is called plane state of stress. For

the present theoretical development, it is assumed that all stress components pertinent

to the planes having normal vector parallel to the z axis are zero:

0z (10.5)

0 xzzx (10.6)

0 yzzy (10.7)

This type of plane stress tensor is illustrated in Figure 10.2.

Figure 10.2 Plane Stress Tensor

Additionally, the remaining non-zero stress tensor components contained in equation

(10.1) are assumed to be independent of the variable z .

),(),,( yxTzyxT (10.8)

Consequently, the plane stress tensor ),( yxT can be represented in plane oxy as

depicted in Figure 10.3:

Figure 10.3 Plane State of Stress

Page 192: 179306253 Lectures in Mechanics of Materials Ion Simulescu

196

The plane stress tensor ),( yxT is defined by three non-zero components:

yyx

xyxyxT

),( (10.9)

Several of the conditions encountered in previous lectures, including the study of the

axial and torsional deformation and pure and non-uniform bending, are characterized

by different states of plane stress. In reality, the plane stress tensor is a direct result of

the assumptions imposed on the deformation.

Examples of plane stress tensors, such as uni-axial, pure shear and bi-axial, are

illustrated in Figure 10.4.

Figure 10.4 Examples of Plane State of Stress

(a) Uni-Axial, (b) Pure Shear and (c) Bi-Axial

10.2 Plane Stress Transformation Equations

Suppose that the components of the plane stress tensor ),( yxT , as expressed by

equation (10.9), are defined at any point ),( yxP of the vertical plane oxy . The

variation of stress components when the reference system attached to point ),( yxP is

rotated with a counterclockwise angle , as illustrated in Figure 10.5, is the subject of

this section.

Page 193: 179306253 Lectures in Mechanics of Materials Ion Simulescu

197

Figure 10.5 Representation of the Plane State of Stress

(a) Normal Planes and (b) Rotated Planes

The transformation relations between stresses n , t and nt pertinent to a rotated

plane and the stresses x , y and xy are obtained by writing the equilibrium

equations for the infinitesimal triangular element depicted in Figure 10.6. The

inclined plane is defined by its positive normal n which is rotated counterclockwise

with an angle from the horizontal direction x .

Figure 10.6 Equilibrium of the Infinitesimal Triangular Element

(a) Stresses and (b) Forces

Using the notation shown in Figure 10.6.b the equilibrium equations are written as:

0 nF

0cos*)*(sin*)*(

sin*)*(cos*)*()*(

yyxxxy

yyxxn

AA

AAA (10.10)

0 tF

0sin*)*(cos*)*(

cos*)*(sin*)*()*(

yyxxxy

yyxxnt

AA

AAA (10.11)

Page 194: 179306253 Lectures in Mechanics of Materials Ion Simulescu

198

From Figure 10.6 the following geometrical relations can be derived:

cos*AAx (10.12)

sin*AAy (10.13)

Substituting equations (10.12) and (10.13) into equations (10.10) and (10.11) and

using the shear stress duality principle the following expressions are obtained for

normal n and shear nt stresses:

cos*sin**2sin*cos* 22

xyyxn (10.14)

)sin(cos*cos*sin*)( 22 xyyxnt (10.15)

Equations (10.14) and (10.15) can be re-written using the trigonometric relations

between the angle and the double angle )*2( :

)*2sin(*)*2cos(*22

xy

yxyx

n

(10.16)

)*2cos(*)*2sin(*2

xy

yx

nt

(10.17)

Equation (10.16) and (10.17) are called the plane stress transformation equations.

In general, two faces are needed to express the plane stress tensor around a point

),( yxP . Consequently, the formulae (10.16) and (10.17) are applied twice: first,

considering the rotation angle 'xx and, secondly, for the complementary

angle 'xy . The notation is illustrated in Figure 10.7. The rotation angles 'xx

and

'xy are related as:

2

''

xxxy (10.18)

The relation between the double angles necessary in equations (10.16) and (10.17) is

obtained as:

'' *2*2xxxy

(10.19)

Consequently, the following trigonometric relations can be established:

)*2sin()*2sin( '' xyxx (10.20)

Page 195: 179306253 Lectures in Mechanics of Materials Ion Simulescu

199

)*2cos()*2cos( '' xyxx (10.21)

Figure 10.7 Stresses on Orthogonal Rotated Faces

The stresses on two orthogonal rotated faces are expressed as:

)*2sin(*

)*2cos(*22

)(

'

'''

xxxy

xx

yxyx

xxnx

(10.22)

)*2cos(*

)*2sin(*2

)(

'

''''

xxxy

xx

yx

xxntyx

(10.23)

)*2sin(*

)*2cos(*22

)90(

'

'''

xxxy

xx

yxyx

xxny

(10.24)

)*2cos(*

)*2sin(*2

)90(

'

''''

xxxy

xx

yx

xxntxy

(10.25)

If equations (10.22) and (10.24) are summed, the invariance of the summation of

normal stresses is established:

yxyx '' (10.26)

Page 196: 179306253 Lectures in Mechanics of Materials Ion Simulescu

200

10.3 Principal Stresses

The maximum and minimum normal stresses are called principal stresses and

mathematically represent the extreme values of the normal stress function )( n . The

extreme values are obtained by imposing the condition that the first derivative of the

normal stress )( n relative to the rotation angle is zero:

0)(

d

d n (10.27)

The explicit expression for equation (10.27) is obtained by differentiating equation

(10.16):

0)*2cos(**2)*2sin(*)( xyyx (10.28)

Dividing by 2* )*2cos( the trigonometric equation (10.28) is transformed into

equation (10.29) relating the tangent of twice the principal directions angle, p , to the

stresses ion the orthogonal planes x and y :

2

)*2tan(yx

xy

p

(10.29)

The angle p represents the angle for which the normal stress )( n reaches its

extreme value. The geometrical illustration of the equation (10.29) is presented in

Figure 10.8.a where the distance R is calculated as:

2

2

4

)(xy

yxR

(10.30)

Figure 10.8 Geometrical Representation of Equation (10.29)

Solution of the trigonometric equation (10.29) yields two solutions, )*2( 1p

and )*2( 2p , where the two angles are related as:

Page 197: 179306253 Lectures in Mechanics of Materials Ion Simulescu

201

12 *2*2 pp (10.31)

From equation (10.31), the orthogonality of the two principal directions is established:

212

pp (10.32)

To calculate the values of the normal stress )( n corresponding to the angles 1p and

2p it is necessary to evaluate the trigonometric functions )*2sin( p and )*2cos( p

contained in equation (10.16). Using the notation shown in Figure 10.8.a and the

trigonometric relations (10.20) and (10.21) these functions can be expressed as:

R

xy

p

)*2sin( 1 (10.33)

R

yx

p2)*2cos( 1

(10.34)

R

xy

pp

)*2sin()*2sin( 12 (10.35)

R

yx

pp2)*2cos()*2cos( 12

(10.36)

Substituting first the trigonometric expressions (10.33) and (10.34) into equation

(10.16) and then (10.35) and (10.36) the principal stresses 1 and 2 are obtained:

Ravxy

yxyx

pn

2

2

114

)(

2)( (10.37)

Ravxy

yxyx

pn

2

2

224

)(

2)( (10.38)

The average normal stress is calculated as:

2

yx

av

(10.39)

The principal stresses are schematically depicted in Figure 10.9.

Page 198: 179306253 Lectures in Mechanics of Materials Ion Simulescu

202

Figure 10.9 Principal Stresses and Directions

The right-hand expression of equation (10.28) being identical with the expression

(10.17), representing the shear stress nt , allows equation (10.28) to be re-written as:

0)(

nt

n

d

d

(10.40)

Equation (10.40) indicates that the principal normal stresses are obtained for a

rotated plane where the shear stress is zero.

The invariance of the sum of the normal stresses is again shown to be valid for the

case of the principal stresses. Summation of equations (10.37) and (10.38) yields:

yxav *221 (10.41)

To identify which of the two angles, 1p or 2p , corresponds to the maximum

principal stress 1 the second derivative of the function )( n relative to the rotation

angle is employed. The condition for the point to be a maximum is:

0)()(

2

2

p

n

d

d

(10.42)

The condition (10.42) is explicitly written as:

0)*2sin(*)*2cos(*2

pxyp

yx

(10.43)

The inequality (10.43) can be manipulated and cast in a new form:

0)*2sin(*])*2tan(

1*

2[

pxy

p

yx

(10.44)

If the )*2sin( p is expanded the following trigonometric expression is established:

Page 199: 179306253 Lectures in Mechanics of Materials Ion Simulescu

203

2)(cos*tan*2cos*sin*2)*2sin( ppppp (10.45)

Substituting equations (10.29) and (10.45), representing the )*2tan( p and

)*2sin( p , into the inequality (10.44) the following expression is obtained:

0tan

*cos*]4

)([*2 22

2

xy

p

pxy

yx

(10.46)

The condition for the inequality (10.46) to hold true is:

0tan

xy

p

(10.47)

Note: It is important to note that for the inequality (10.47) to hold true, the signs of

the tangent of the angle p and shear stress xy must be identical.

The angle corresponding to the direction of the maximum normal stress can also be

obtained by successively assigning to the angle in equation (10.16) the values 1p

and 2p and observing which angle produces the maximum principal stress.

10.4 Maximum Shear Stresses

The maximum shear stresses are determined in a similar manner as the principal

stresses. The extreme condition for the shear stress function )( nt contained in

equation (10.17) is written as:

0)(

)(

d

d nt (10.48)

The explicit format of equation (10.48) is obtained by differentiating the expression

(10.17):

0)*2sin(**2)*2cos(*)( xyyx (10.49)

Dividing by )*2cos(*2 the trigonometric equation (10.49), the tangent of the

principal directions angle s is obtained:

)*2tan(

12)*2tan(pxy

yx

s

(10.50)

Page 200: 179306253 Lectures in Mechanics of Materials Ion Simulescu

204

The angle s represents the angle for which the shear stress )( nt reaches its extreme

value. Equation (10.50) is illustrated in Figure 10.10.

Figure 10.10 Angular Relation between )*2( s and )*2( p

Solving the trigonometric equation (10.50), two solutions )*2( 1s and )*2( 2s are

obtained. They are related as:

12 *2*2 ss (10.51)

Dividing equation (10.50) by two (2), the orthogonality of the two angles 1s and 2s is

obtained:

212

ss (10.52)

Equation (10.50) indicates that a relation between the angles )*2( p and )*2( s can

be established. With this intent, equation (10.50) is recast into a new format as

follows:

0)*2sin(

)*2cos(

)*2cos(

)*2sin(

p

p

s

s

(10.53)

First, multiplying by the terms in the denominator, equation (10.53) becomes:

0)*2cos(*)*2cos()*2sin(*)*2sin( spps (10.54)

Then, equation (10.54) is simplified as:

0)]*2()*2cos[( ps (10.55)

Therefore,

2

)]*2()*2(

ps (10.56)

Page 201: 179306253 Lectures in Mechanics of Materials Ion Simulescu

205

The relationship between angles s and p is calculated from equation (10.56):

4

ps (10.57)

Still, equation (10.57) does not indicate how to identify the direction of the maximum

shear stress. By examination of Figure 10.10, the following angular relations can be

established:

2

*22

*3*2

2*2*2 1111

ppps (10.58)

The relationship between the angles of the maxim principal and shear stress

directions, s and p , is obtained from (10.58) as:

4

11

ps (10.59)

Again using the notation shown in Figure 10.10, the following trigonometric relations

are obtained:

R

yx

s2)*2sin( 1

(10.60)

R

xy

s

)*2cos( 1 (10.61)

R

yx

s2)*2sin( 2

(10.62)

R

xy

s

)*2cos( 2 (10.63)

Successively substituting the two groups of expressions, (10.60) and (10.61), and,

(10.62) and (10.63), into equation (10.17) the maximum and minimum shear stresses

are calculated as:

max

2

2

114

)()(

Rxy

yx

snts (10.64)

min

2

2

224

)()(

Rxy

yx

snts (10.65)

Page 202: 179306253 Lectures in Mechanics of Materials Ion Simulescu

206

The normal stresses corresponding to the maximum and minimum shear stresses are

calculated by substituting the expressions (10.60) through (10.63) into equation

(10.16):

2)( 11

yx

sns

(10.66)

2)( 22

yx

sns

(10.67)

The maximum and minimum shear stresses and the corresponding normal stresses are

illustrated in Figure 10.11.

Figure 10.11 Relationships between Principal Planes and Maximum Shear Stress

Planes

Note: From Figure 10.12.b it can be concluded that, in contrast, to the principal

planes which are free of shear stress, the planes on which the shear stress

achieves extreme values are not necessarily free of normal stresses.

10.5 Mohr’s Circle for Plane Stresses

Mohr’s circle is a graphical construction reflecting the variation of the plane state of

stress around a particular point, including information pertinent to the principal and

maximum shear stresses.

From equations (10.16) and (10.17), first squared and then summed, the following

relation is obtained:

222)( Rntavn (10.68)

Page 203: 179306253 Lectures in Mechanics of Materials Ion Simulescu

207

Equation (10.68) represents the equation of a circle of radius R defined in the ntn ,

plane. The center of the circle is located at point )0,( avC . The values of circle

radius R and average normal stress av are calculated employing equations (10.30) and

(10.39), respectively.

Intersecting the equation of the circle (10.68) with the horizontal axis 0nt the

intersection points )0,( 11 nP and )0,( 22 nP are obtained:

11 Ravn (10.69)

22 Ravn (10.70)

It can be concluded that these intersection points represent the principal stresses

1 and 2 .

The Mohr’s circle for plane stress condition is drawn relative to a Cartesian system

with the abscissa and the ordinate axis representing the normal stresses and the

shear stress , respectively. The following sign convention is employed as illustrated

in Figure 10.12:

(a) the positive shear stress axis is downward;

(b) the positive angle is measured counterclockwise;

(c) the shear stress on a face plots as positive shear if tends to rotate the face

counterclockwise.

Figure 10.12 Mohr’s Circle Notation

Page 204: 179306253 Lectures in Mechanics of Materials Ion Simulescu

208

Note: The positive direction of the vertical axis, representing the shear stress ,

pointing downward (sign convention (a)) is elected in order to be able to

enforce the positive measurement of the angle (sign convention (b)).

Examining the notations shown in Figure 10.12 clarifies that all the angles are

measured from the line XY ( 0 ) in the anticlockwise direction.

Morh’s circle for plane stress is constructed in the following steps:

(f) The coordinate system is drawn as shown in Figure 10.12. The horizontal axis

represents the normal stress , while the vertical axis represents the shear

stress . To gain full advantage of the graphical benefits of the method it is

necessary that the drawing to be made on scale. However, the method is also

helpful as a conceptual tool in combination with the governing equations

wherein it may be drawn more roughly. The representation considers that the

following conditions are met: yx and 0xy ;

(g) Using the calculated values of the normal stresses x and y and the shear

stress xy two points noted as ),( xyxX and ),( xyyY are placed on the

drawing. The line XY intersects the horizontal axis at point C which

represents the center of the Mohr’s circle;

(h) The distance CX represents the radius of the circle. Using the radius CX and

the position of the center )0,( avgC the Mohr’s circle is constructed. The

intersection points, 1P and 2P , between the circle and the horizontal axis

represent the maximum and the minimum principal stresses;

(i) The value of the )*2tan( p can be calculated from the graph. The angle

)*2( p is identified on the graph by the 1XCP angle and is measured from

the 0 to the principal directions line in the counterclockwise direction;

(j) The lines 1XP and 2XP represent the principal direction1 (associated with the

maximum principal stress) and 2 (associated with the minimum principal

stress), respectively.

Every point on the Mohr’s circle corresponds to a pair of stresses and on a

particular face. To emphasize the face involved the point is labeled identically with

the face where it belongs. For example, the face x , y and n are represented on the

Mohr’s circle by the points X , Y and N . To reinforce the shear stress sign convention

(c) two icons indicating the rotation sense induced by the shear stress are shown in

Figure 10.12. The angle measured from 0 to the radius line CN in the

counterclockwise direction is equal to twice the angle of the plane rotation )*2( .

Note: The points X andY represent the case of orthogonal planes having normals

parallel to axes x and y , respectively. The line XY corresponds to angle

Page 205: 179306253 Lectures in Mechanics of Materials Ion Simulescu

209

0 . The double angle )*2( 1p of the maximum principal direction is

measured from the line CX to the line 1CP . By these conventions, the double

angle sense is established as being positive in the counterclockwise direction.

The angle 1p is the angle 12 PXP and has the same direction as the double

angle )*2( 1p . The angle associated with the minimum direction 2p is

perpendicular to the angle 1p . The stresses corresponding to a plane rotated

with an angle are obtained by placing on the circle the radius CN located by

measuring in the counterclockwise direction an angle of )*2( from the

lineCX . The corresponding stresses n and nt are a function of the location

of the point ),( ntnN position in the coordinate system and may be

obtained by scaling them from the figure or by the use of the analytical

equations (10.30), (10.39) and (10.68). The opposite point

),( ntnT represents the stresses on the orthogonal rotated face.

In comparison with the technique used to show the principal axes in the oxy

representation (Figures 10.7 through 10.11) the plot obtained from the Mohr’s circle

appears to be misleading. The cause is that the Mohr’s circle is drawn in the

coordinate system. In the oxy representation the principal directions are correctly

plotted by artificially rotating the principal directions obtained from the Mohr’s

circle around the point C with an angle )*2( 1p measured in the counterclockwise

direction.

10.6 Principal Stresses Distribution in Beams

One of the most important applications of the plane state of stress theory described

above is found in the study of variation of the stresses in beams under non-uniform

bending. Recall from Lecture 7 that under some imposed kinematic assumptions, a

beam subjected to transversal loading is in a state of plane stress. With the exception

of some areas (around the supports or the application points of concentrated loads) the

beam theory characterizes the existence of only two types of stresses: normal stress

x and shear stress xy . The normal stress x is calculated using Navier’s formula

expressed by equation (10.71), while the shear stress xy is obtained employing

Jurawski’s formula contained in equation (10.72):

yI

xMyx

z

zx *

)(),( (10.71)

tI

ySxVyx

z

A

zy

xy*

)(*)(),(

'

(10.72)

Page 206: 179306253 Lectures in Mechanics of Materials Ion Simulescu

210

The notation used in the formulae (10.71) and (10.72) is explained in Lecture 7 and is

not repeated herein.

The plane stress tensor previously expressed in equation (10.9) is written for the case

of the beam in nonuniform bending as:

0

),(yx

xyxyxT

(10.73)

The entire theoretical development described in the previous sections can be without

restriction applied to the study of the particular plane stress tensor (10.73).

Consequently, the variation of the stresses around any point in a beam subjected to

nonuniform bending can be calculated. Figure 10.13 represents an example of the

application of plane stress theory for the case of a simply supported beam.

In the example, the beam has a rectangular cross-section and is subjected to a single

concentrated force P*2 located at the mid-span. It is evident that the ratios of the

beam dimensions and the loading do not violate any of the assumptions related to the

applications of the formulae (10.71) and (10.72). The shear force diagram )(xV and

the bending diagram )(xM , where the axis identification indices were dropped for

clarity, are plotted.

Figure 10.13 Simple Supported Beam

The geometrical characteristics of the rectangular cross-section involved in the

evaluation of the formulae (10.71) and (10.72) are:

33

**3

2

12

)*2(*cb

cbI z (10.74)

Page 207: 179306253 Lectures in Mechanics of Materials Ion Simulescu

211

bt (10.75)

)1(***2

1

2*)(*)(

2

22

c

ycb

cyycbyS z

(10.76)

For the left half of the beam cx 50 the shear force )(xV and the bending moment

)(xM are expressed as:

PxV )( (10.77)

xPxM *)( (10.78)

Substituting equations (10.74) through (10.78) into equations (10.71) and (10.72), the

expressions for normal and shear stresses are obtained as:

yxcb

Py

cb

xPyxx **

**

2

3*

**3

2

*),(

33

(10.79)

)1(**

*4

3

***3

2

)1(***2

1*

*

)(*)(),(

2

2

3

2

22'

c

y

cb

P

bcb

c

ycbP

tI

ySxVyx

z

A

zy

xy

(10.80)

Note: The minus (-) sign appearing in formula (10.81) has been inserted in order to

comply with the shear sign convention (c).

To obtain an illustrative variation of the principal stresses, the rectangular domain of

the beam is divided by superimposing a rectangular mesh. For the case under study,

the mesh has five spaces in the longitudinal x direction and eight spaces in the

vertical y direction. Using Mathcad programming capabilities the principal stresses

and corresponding angular directions can be easily calculated for every point of the

mesh. The principal stresses calculated for two cross-sections cx and cx *4 and

all nine points vertically describing the cross-sections are contained in Table 10.1.

The ratios 0

max

and

0

min

are tabulated instead of the max and min ,

wherecb

P

**20 .

A review of the results presented in Table 10.1 shows that at the extreme fibers the

principal stresses correspond with the normal stresses and reach the maximum values.

At the extreme fiber locations the shear stress xy is zero. The situation is different for

Page 208: 179306253 Lectures in Mechanics of Materials Ion Simulescu

212

the case of wide-flange beams where both x and xy have significant values at the

junction between the web and the flange.

Table 10.1

Today, with the help of modern computer codes, the formulae involved in the

calculation of the principal stresses and directions can be computed using a very

refined mesh. The graph containing the curves tangent to the principal directions in

every point of the mesh is called the stress trajectory. Two sets of curves are drawn

and they are orthogonal at every point. The stress trajectory graph pertinent to the

simply supported beam investigated above is pictured in Figure 10.14. A typical

example of practical usage of the stress trajectory curves is the placement of the

reinforcement in reinforced concrete beam. Because the stress trajectory graph does

not give any indication about the magnitude of the principal stresses another type of

graph is also used. This is called a stress contour plot and contains curves of equal

principal stress magnitudes. The commercial codes employed today can provide these

plots.

Page 209: 179306253 Lectures in Mechanics of Materials Ion Simulescu

213

Figure 10.14 Stress Trajectory Plot

10.7 Example

The theoretical formulation derived above is used to investigate the following

practical case:

20x MPa

10y MPa

10xy MPa

The corresponding stress tensor is written as:

MPayxT *1010

1020),(

The state of stress for the case above is shown in Figure 10.13.

Figure 10.13 Example Plane State of Stress

The following values illustrated in Figure 10.13 are calculated as:

MPayx

avg 52

)10(20

2

Page 210: 179306253 Lectures in Mechanics of Materials Ion Simulescu

214

MPayx

152

)10(20

2

MPaR xy

yx028.18)10(15

4

)(222

2

667.015

10

2

)*2tan(

yx

xy

p

69.33*2 p

Figure 10.14 Geometrical Relations

From equation (10.47) it is established that the angle related to the maximum

principal direction must have a negative tangent. Consequently, the angles of the

principal direction are:

84.161 p

16.739084.162 p

The principal stresses, shown in Figure 10.15, are obtained as:

MPaRavg 028.23028.1851

MPaRavg 028.13028.1852

Page 211: 179306253 Lectures in Mechanics of Materials Ion Simulescu

215

Figure 10.15 Principal Stresses

The angle of the maximum shear stresses is calculated as:

84.614584.164511 ps

The maximum shear stresses are calculated as:

MPaxy

yx

snts 028.184

)()(

2

2

11

MPaxy

yx

snts 028.184

)()(

2

2

22

The normal stresses acting on the maximum shear planes are calculated as:

MPayx

sns 52

)( 11

MPayx

sns 52

)( 22

Figure 10.16 Maximum Shear Stresses

Page 212: 179306253 Lectures in Mechanics of Materials Ion Simulescu

216

The maximum shear stresses and the corresponding normal stresses are illustrated

Figure 10.16

Morh’s circle pertinent to the problem is illustrated in Figure 10.17.

Figure 10.17 Mohr’s Circle

Note: The points 'X and 'Y represent the case of orthogonal planes having the

normal directions rotated with angles of 30 and 120 , respectively, from the

x axis. Successively substituting the above angular values in equations

(10.16) and (10.18) the following stresses pertinent to points 'X and 'Y are

obtained:

for 30

MPaX

84.3'

MPaYX

99.17''

for 120

MPaY

16.6'

MPaXY

99.17'''

Page 213: 179306253 Lectures in Mechanics of Materials Ion Simulescu

217

REFERENCES

1. Mazilu, Panait, Rezistenta Materialelor, Institutul de Constructii

Bucuresti, Romania, 1974.

2. Craig, Roy R., Mechanics of Materials, Second Edition, John Willey &

Sons, New York, 2000.

3. Gere, James M., Mechanics of Materials, Fifth Edition, Brooks/Cole,

Pacific Grove, CA, 2001.

4. Higdon, A., Ohlsen, E.H., and Stiles, W.B., Mechanics of Materials, Third

Edition, John Willey & Sons, New York, 1962.

5. Popov, E.P., Introduction to Mechanics of Solids, Prentice-Hall Inc.,

Englewood Cliffs, New Jersy, 1968.

6. Timoshenko, A., and Goodier, J.N., Theory of Elasticity, Third Edition,

McGraw-Hill Inc., New York, 1970.

7. Timoshenko, S.P., Strength of Materials, Part 1, Elementary Theory and

Problems, Third Edition, CBS Publishers & Distributors, New Deli, India,

2002.

8. Hartog, J.P.D., Strength of Materials, Dover Publications Inc., New York,

1961.