Sie sind auf Seite 1von 264

Cornel Sandi BI

STRENGTH OF MATERIALS

2014
CONTENTS

1. Introduction. 2
1.1 Generalities. 2
1.2 The object and problems of strength of materials. Connections with
other engineering sciences... 2
1.3 A short review of the history of strength of materials.. 4
1.4 Bodies schematic representation in strength of materials. 9
1.5 Loads schematic representation in strength of materials calculus.. 12
1.6 Supports and reactions 16
1.7 Equilibrium equations 18

2. Internal forces in statically determinate members... 23


2.1 Internal forces computation 23
2.2 Differential relations among load, shear and bending moment... 25
2.3 Examples concerning the main types of diagrams. 28
2.3.1. Axial forces diagrams... 28
2.3.2 Shear and bending-moment diagrams... 29
2.3.3 Torque diagrams 40
2.4 Superposition method. 41
2.5 Moving loads.. 42
2.6 Internal forces diagrams for plane structures (2d structures - frames) and
spacial (3d) structures. 43
Problems to be assigned... 53

3.First moments and second moments of an area 57


3.1 First moments of an area. Centroid of an area.. 57
3.2 Second moments of an area 59
3.3 Parallel - axis theorem (Steiners relations)... 60
3.4 Moments of inertia of simple surfaces.. 62
3.5 Moments of inertia of complex surfaces (composite areas)... 65
3.6 Moments of inertia with respect to inclined axes... 66
3.7 Radius of gyration. Ellipse of inertia.. 70
Problems to be assigned... 74

4. Displacements, stress and strain.. 75


4.1 Displacements. 75
4.2 The concept of strain.. 75
4.3 The concept of stress.. 77
4.4 Relationships among internal forces and stresses within a beam (member)
cross section.. 82

5. Strength of materials basic assumptions.. 83

6. Axial loading... 87
I
6.1 Stresses and strains. 87
6.2 Poisson's ratio. 98
6.3 Stress concentrations.. 100
6.4 Own weight effect. Members of constant strength. 102
6.5 Stress-strain diagram.. 107
6.5.1 Generalities. 107
6.5.2 Safety coefficients. Allowable strengths 110
6.6 Statically indeterminate problems.. 111
6.6.1 Members with unhomogeneous cross sections... 113
6.6.2 Straight members fixed at both ends.. 116
6.6.3 Systems of parallel members.. 118
6.6.4. Systems of concurrent bars 120
6.6.5 Problems involving stresses produced by assembling imperfections. 121
6.6.6 Problems involving temperature changes... 123
6.6.7 Two more numerical examples concerning statically indeterminate
problems 125
Problems to be assigned... 128

7. Shearing stresses in members of small cross sections. 135


7.1 Shearing stress and shearing strain... 136
7.2 Strength of a simple riveted joint... 137
7.3 Bearing stress in connections. 138
7.4 Welded connections 139
Problems to be assigned... 141

8. Basic elements of the theory of elasticity............................................................ 143


8.1 Plane state of stress. 143
8.2 Spatial state of stress.. 148
8.3 Monoaxial state of stress 150
8.4 Pure shear state... 151
8.5 Generalized Hookes law. 151
8.6 Strain energy... 155
8.6.1 Elastic strain energy under axial loading. 156
8.6.2 Shearing stresses elastic strain energy. 158
8.6.3 Strain energy for a general state of stress 158
8.6.4 Elastic strain-energy density associated with a change in volume.
Elastic strain-energy density associated with a distortion (a change in
shape)... 160
8.6.5 Relation among E, G and . 161

9. Torsion. 163
9.1 Torsion of circular members.. 164
9.1.1 Stresses and strains.. 164
9.1.2 Statically indeterminate shafts. 175
9.2 Torsion of noncircular (rectangular) members... 180
II
9.3 Thin-walled hollow shafts (tubes).. 185
Problems to be assigned.. 191

10. Bending. 195


10.1 Prismatic members in pure bending. 196
10.2 Shearing stresses in beams subjected to simple bending. 203
10.3 Prevention of longitudinal sliding in case of composite sections. 210
10.4 Beams of constant strength... 213
10.5 Unsymmetric bending.. 218
10.6 General bending 221
Problems to be assigned... 222

Strength of materials typical key words and phrases.. 225


Apendix I. 247
Apendix II.... 248
Apendix III.. 249

III
PREFACE

Today, more than ever, engineering applications are often interdisciplinary,


involving the interrelationship of the basic engineering sciences (mechanics,
metallurgy, chemistry, physics, mathematics, etc.). This is the reason why the modern
engineer must have a fundamental knowledge in each of these particular areas. There
is no doubt that one of the most important matter within the preparing of the
engineering students is represented by the course of Strength of Materials, which is a
very important link of this knowledge. Furthermore, for a successful machine or
structural design, a thorough mastery of strength of materials is a must.
This book is intended to be a basic course of Strength of Materials which to
develop in the engineering student the ability to analyze a given problem in a simple
and logical manner and to apply to its solutions a few fundamental and well-
understood principles. But there is also another important reason this book has been
written for: to accustom the Romanian engineering students with the main concepts,
symbols and trends belonging to the English Engineering Life. As we know, the
world of modern science has progressed swiftly across hitherto impossible barriers.
There has been more change in the last twenty years than in the previous fifty and
more in fifty years than in the previous 200. Never was it more important to have
available information on the countless facts of modern scientific knowledge and
achievements. Every science-oriented engineer should equip himself with the most
up-to-date achievements of science and technique. But for this, he must know the
other mans language. Research people need knowledge of reports and publications
in many languages. You can not have a proper access to the impressive quantity of
Internet information if you dont master a foreign language, especially English. The
barrier that is set up by a difference in language may constitute a serious obstacle in
pure science or in the more practical world of technology.
There are three main ways in which scientific information may be acquired:
- by adequate translation;
- by persuading scientists to publish their work in the better-known
languages;
- by teaching scientists to read, to write and to speak foreign languages.
From these three solutions, the third one is, of course, much more convenient to
anyone.
Only from the above mentioned ideas one can conclude that it is so important
to master a foreign language, especially English which is spread all over the world.
Engineering English is no longer a problem nowadays. Every student can easily and
rapidly master it, provided he is set on learning it. This is the conclusion to be drawn
as far as the difficulties in the search for technical information are concerned. A great
philosopher used to say: The knowledge of a language is another weapon in life.
Strength of Materials

1. INTRODUCTION

1.1 GENERALITIES

During the whole history of the human society, the practical needs have been
those which fundamentally contributed to the development of sciences like
mathematics, physics, chemistry, astronomy, biology, etc.
In the same manner, the Strength of Materials has arisen from the necessity of
building safely, economically and aesthetically. Through the ages, the application of
materials in engineering design has posed difficult problems to mankind. In the Stone
Age the problems were mainly in the shaping of material. In the early days of the
Bronze Age and the Iron Age the difficulties were both in production and shaping. For
many centuries metal-working was laborious and extremely costly. Estimates go that
the equipment of a knight and horse in the thirteenth century was of the equivalent
price of a Centurion tank in Word War II. With the improving skill of metal working,
applications of metals in structure increased progressively. Then it was experienced
that structures built of these materials did not always behave satisfactorily and
unexpected failures often occurred. Detailed descriptions of castings and forgins
produced in the Middle Age exist. When judged with present day knowledge, this
production methods must have been liable to build important technical deficiencies
into the structures.
The vastly increasing use of metals in the nineteenth century caused the
number of accidents and casualties to reach unknown levels. For example, the
number of people killed in railway accidents in Great Britain was in the order of two
hundred per year during the decade 1860-1870. Most of the accidents were a result of
derailing caused by fractures of wheels, axles or rails. Most of these accidents were
certainly due to a poor design, which mainly means a poor understanding and use of
Strength of Materials, when design a certain mechanical structure.
Within the context of modern engineering design, Strength of Materials has
continued to occupy a very important place, providing a useful and necessary tool in
understanding the behaviour of certain mechanical structures subjected to external
loads.

1.2 THE OBJECT AND PROBLEMS OF STRENGTH OF


MATERIALS. CONNECTIONS WITH OTHER
ENGINEERING SCIENCES

Three fundamental areas of engineering mechanics are statics, dynamics and


strength of materials. While statics and dynamics are devoted primarily to the study
of the external effects of forces on rigid bodies (i.e. bodies which do not deform

2
Introduction

under the action of the external loads), strength of materials deals with the relations
between externally applied loads and their internal effects on bodies. Moreover, the
bodies are no longer assumed to be rigid; the deformations, however small, are of
major interest. In mechanical design, the engineer must consider both dimensions and
material properties to satisfy requirements of strength and rigidity. When loaded, a
machine component or structure should neither break nor deform excessively.
Let us now consider for example a body subjected to several external loads
(forces and couples) Fig.1.1.
Before the action of the loads,
one can find a certain state inside the
body. But, due to the action of the
external loads, the mechanical state
inside the body will change. Each
particular point of the involved body
feels in a certain way the action of the
external loads, depending upon the
position of the point inside the body and
the physical nature of the material at that
level.
Within the above mentioned
context Strength of Materials tries to
offer reasonable answers to some
important questions such as: Fig. 1.1
- what does a certain point of the body feel if the external loads act?
- how does the involved body deform under the action of the external loads?
- which are the mathematical connections between the mechanical effects
occured at the level of the points of the body and the values of the external
loads?
- which are the critical values of the external loads for which the body fails or
deforms excessively ?
These are only some matters Strength of Materials is concerning with.
In its obvious intention to give valuable answers to the above mentioned
questions, Strength of Materials has ordered its problems in three important classes:
Dimensioning problems:
- input data : values and physical nature of the external loads; geometry
of the body; mechanical properties of the material the body is made of;
- output data : the required dimensions so that the body would not fail or
deform excessively.
Determination of the allowable external loads:
- input data: body dimensions and geometry; mechanical properties of
the material;
- output data: the maximum allowable values of the external loads so
that the body would not fail or deform excessively.

3
Strength of Materials

Checking problems:
- input data : body dimensions and geometry; mechanical properties of
the material; external loads values and orientation;
- output data : a checking calculus which to conclude if the involved
body would fail or not within such conditions.

Strength of Materials provides the engineers with the mathematical means of


getting inside the bodies, of understanding the intimate behavior of matter, in order to
give valuable answers to the questions mentioned above.
As a branch of science, Strength of Materials belongs to the Mechanics of
Solid Deformable Bodies together with: The Theory of Elasticity, Theory of Plasticity
and Statics, Dynamics and Stability of Constructions. Although the Theories of
Elasticity and Plasticity concern with the same problems as Strength of Materials,
they use only a limited number hypothesis (assumptions), are much more exact but,
many times, exceedingly complicated. The Theories of Elasticity and Plasticity use
complex mathematical tools, in many cases less adequate to the engineering
applications. On the contrary, Strength of Materials uses a large number of
theoretical simplifying hypothesis, approximations and experimental analysis
methods, providing a useful tool for immediate engineering applications.
Through its practical and theoretical investigations Strength of Materials
interferes with a lot of fundamental sciences such as: mathematics, physics,
technology, chemistry, etc. even with biology, proving again if it is still the case
the absence of once called hard lines between the sciences.

1.3 A SHORT REVIEW OF THE HISTORY OF STRENGTH OF


MATERIALS

From the earliest times when people started to build, it was found necessary to
have information regarding the strength of materials so that rules for determining
safe dimensions of different bodies or structural elements could be experienced.
No doubt the Egyptians had some empirical rules of this kind, for without
them it would have been impossible to erect their monuments, temples, pyramids and
obelisks, some of which still exist.
The Greeks further advanced the art of building. They developed statics, which
underlies the mechanics of materials. Archimedes (287-212 B.C.) gave a rigorous
proof of the conditions of equilibrium of a lever and outlined methods of determining
centers of gravity of bodies. He used his theory in the construction of various hoisting
devices.
The Romans were great builders. Not only some of their monuments and
temples remain, but also roads, bridges and fortifications. We know something of
their building methods from the book of Vitruvius, a famous Roman architect and
engineer of the time of Emperor Augustus.
Most of the knowledge that the Greeks and Romans accumulated in the way of

4
Introduction

structural engineering was lost during the Middle Ages and only since the
Renaissance has it been recovered. Thus when the famous Italian architect Fontana
(1543-1607) erected the Vatican obelisk at the order of Pope Sixtus V, this work
attracted wide attention from European engineers. But we know that the Egyptians
had raised several such obelisks thousands of years previously, after cutting stone
from the quarries of Syene and transporting it on the Nile. Indeed, the Romans had
carried some of the Egyptians obelisks from their original sites and erected them in
Rome; thus it seems that the engineers of the sixteenth century were not as well
equipped for such difficult tasks as their predecessors.
During the Renaissance there was a revival of
interest in science and art leaders appeared in the
field of architecture and engineering. Leonardo da
Vinci (1452-1519) was a most outstanding man of
that period. He was not only the leading artist of his
time but also a great scientist and engineer.
Leonardo da Vinci was greatly interested in
mechanics and, in one of his notes, he states:
Mechanics is the paradise of mathematical science
because here we come to the fruits of mathematics.
The first attempts to find the safe dimensions
of structural elements analytically were made in the
seventeenth century. Galileo Galileis (1564-1642) Leonardo da Vinci
famous book Two New Sciences shows the writers
efforts to put the methods applicable in stress analysis into a logical sequence. It
represents the beginning of the science of Strength of Materials.
In 1678, the paper Of Spring was published, a paper whose author was Robert
Hooke (1635-1703). It contains the results of Hookes experiments with elastic
bodies. This is the first published paper in which the
elastic properties of materials are discussed. The
linear relation between the force and the
deformation is the so-called Hookes law, which
later on was used as the foundation upon which
further development of the Strength of Materials of
elastic bodies was built.
Bernoulli family produced outstanding
mathematicians for more than a hundred years:
Jacob, Nicholas, John, Daniel Bernoulli. Besides
mathematics these outstanding mathematicians were
also attracted by Mechanics and Strength of
Materials. Galileo Galilei
For example, Daniel Bernoulli was the first
to derive the differential equations governing lateral
vibrations of prismatic bars, using it to study particular modes of this motion and to
John Bernoulli belongs the well-known hypothesis of plane sections for beams in
bending.
5
Strength of Materials

Leonard Euler (1707-1783), Daniel Bernoullis pupil, was one of the greatest
mathematicians of all times. In the field of Strength of Materials he was principally
interested in the geometrical forms of elastics curves. His contributions in the
buckling phenomenon, for example, is essential.

Daniel Bernoulli
Leonard Euler

Navier (1785-1836) published in 1826 the


first real book on Strength of Materials where his
main achievements in this field were incorporated. If
we compare this book with those of the eighteenth
century, we clearly see the great progress made in
mechanics of materials during the first quarter of the
nineteenth. Engineers of the eighteenth century used
experiments and theory to establish formulas for the
calculation of ultimate loads. Navier, from the very
beginning, states that it is very important to know
the limit up to which structures behave perfectly
elastically and suffer no permanent deformations.
Navier Within the elastic range, deformation can be
assumed proportional to force and comparatively
simple formulas can be established for calculating these quantities. Beyond the elastic
limit the relation between forces and deformations becomes very complicated and no
simple formulas can be derived for estimating ultimate loads. Navier suggests that the
formulas derived for the elastic conditions should be applied to existing structures,
which had proved to be sufficiently strong, so that safe stresses for various materials
could be determined which later could be used in selecting proper dimensions for
new structures.
S. D. Poisson (1781-1840) was born in a small town near Paris in a poor
family and, until he was fifteen years old, he had no chance to learn more than to read
and write. In 1796 he was sent to his uncle at Fountainebleau, and there he was able
to visit mathematics classes. His excellent progress made him became in 1812 a
member of the French Academy. In the field of Strength of Materials, the principal
results obtained by Poisson are incorporated in two memoirs which he published in
6
Introduction

1829 and 1831, and in his course of mechanics (Trait de Mcanique) published in
1833.
The two outstanding engineers, G. Lam
(1795-1870) and B.P.E. Clapeyron (1799-1864),
graduated from the cole Polytechnique of Paris in
1818. They had important contributions to the
Strength of Materials within their help offered to the
new Russian engineering school of that time the
Institute of Engineers of Ways of Communication in
St. Petersburg. This new Russian school was later to
have a powerful influence upon the development of
engineering science in Russia. Lam and Clapeyron
had to teach applied mathematics and physics in this
school and they had also to help with the design of
various important structures in which the Russian S. D. Poisson
government was interested: for example, several
suspension bridges were designed and erected in St.
Petersburg at that time. These bridges (constructed
between 1824 and 1826) were the first suspension
bridges build on the European Continent.
After their return to Paris, Lam and
Clapeyron continued their work and, as a
recognition of their fundamental contributions
within the engineering world, they were elected
members of the French Academy of Sciences.
Barr de Saint-Venant (1797-1886) was born
in the castle de Fortoiseau (Seine et Marne). His
talent in mathematics was noticed very early and he
was given a careful coaching by his father who was
G. Lam
a well-known expert in economy. Later he studied at
the Lyce of Bruges and in 1813, when sixteen years
of age, he entered the cole Polytechnique after
taking the competitive examinations. Here he
showed his outstanding ability and became the first
in his class. Unfortunately, the political events of
1814 (when referring to Napoleon, Saint-Venant
said : My conscience forbids me to fight for an
usurper.) forced him to interrupt his academic
studies being proclaimed a deserter and never
again allowed to resume his study at the cole
Polytechnique. After nine years, the government
permitted him to enter the cole des Ponts et
Chausses without examination. After graduating,
he devoted all his life to engineering (especially to Barr de Saint-Venant

7
Strength of Materials

the Theory of Elasticity and Strength of Materials), in 1868 being elected a member
of the Academy of Sciences.
D.J. Juravski (1821-1891) graduated in
1842 from the Institute of Engineers of Ways of
Communication in St. Pertersburg. His career
was closely related with the development of
railroad construction in Russia.
A. Whler (1819-1914) was born in the
family of a schoolmaster in the province of
Hannover and he received his engineering
education at the Hannover Polytechnical
Institute. Being an outstanding student, he won
a scholarship after his graduation, which
enabled him to get a practical training, both at
D.J. Juravski the Borsing locomotive works in Berlin and in
the construction of the Berlin-Anhalter and
Berlin-Hannover railways. In this way he had to solve many problems concerning the
mechanical properties of materials and started his famous investigations upon the
fatigue strength of metals. It can justly be said that scientific investigation in the field
of fatigue of materials began with Whlers work. For each kind of fatigue test,
Whler designed and built all the necessary
machines and measuring instruments. In
designing these machines, he imposed stringent
requirements upon the accuracy with which
forces and deformations were to be measured
so that his machines represent an important
advance in the technique of structural materials.
The above mentioned scientists were
only a few of those who fundamentally
contributed to the development of Strength of
Materials. Of course, there are much more
names involved in this problem and it would be
unfair if we did not even mention them here: C.
A. Whler A. Coulomb (1736 1806), Augustin Cauchy
(1789 18557), Thomas Young (1773 1829),
J. C. Maxwell (1831 1879), Otto Mohr (1835 1918), Alberto Castigliano (1837
1884) etc.
The Romanian school has also offered excellent researchers and scientists to
the Strength of Materials engineering world: C. C. Teodorescu, Gh. Buzdugan,
tefan Ndan, Radu Voinea, D. R. Mocanu, Petre Augustin, etc.
Nowadays, the courses of Strength of Materials occupy a fundamental place
within the Romanian academic engineering education, with a strong tradition and
hope for the future.

8
Introduction

1.4 BODIES SCHEMATIC REPRESENTATION IN STRENGTH


OF MATERIALS

It is to be recalled from Section 1.2 that the main task of Strength of Materials
concerns the mechanical behaviour of different bodies or structural elements
subjected to external loads. Within its theoretical and practical investigations,
Strength of Materials operates with three important types of bodies:

a) Bars: represent those bodies


whose length is much greater
than the other two dimensions
(Fig.1.2). Examples of this class:
beams, members, rods, columns,
pins, rivets, shafts, etc.

Fig. 1.2

square cross section rectangular cross circular cross section annular cross section
section

I - Shapes U - Shapes T Shapes Angles

Fig. 1.3

The cross section of a bar is defined as the plane section of minimum area
obtained through the intersect between the involved bar and a plane passed through
the bar at the some arbitrary point. The cross section of a bar has a certain area
generally denoted by A [mm2] and may have different geometric shapes (Fig.1.3).
The cross section of a bar may be constant or variable, as shown in Fig.1.4.

9
Strength of Materials

a. b.
Fig. 1.4
The bar axis is defined as the geometric locus of the cross sections centers of
gravity (centroids). It is also called the medium curve. Along its different portions the
bar axis may be straight (Fig.1.5a) or curved (Fig.1.5b).

a. b.
Fig. 1.5
In Strength of Materials calculus schemes the bars are usually represented
through their longitudinal axes (Fig 1.6).
Simplified representation

Fig. 1.6

10
Introduction

It is important to be noted that, in most cases, a bar is geometrically


represented within an Oxyz coordinate system, the bar cross section being located in
the zOy plane while Ox is the longitudinal axis, i.e. the bar axis or the medium curve.

b) Plates: represent those bodies whose two dimensions (length and width) are
much greater than the third dimension (thickness), Fig. 1.7.

Fig. 1.7 Fig. 1.8

The midsurface of a plate is defined as the geometric locus of the plate


thickness midpoints.
Depending upon the shape of the midsurface, the plates may be classified in
two important types: plane plates (Fig. 1.7) and curved plates (Fig 1.8).

Fig. 1.9 Fig. 1.10

Depending upon the thickness variation there are two main distinct types of
plates: with constant thickness (Fig. 1.9) and with variable thickness (Fig. 1.9).

c) Blocks: represent those bodies whose dimensions are of the same size order
(Fig 1.10)

Although the above classification has been done on geometrical criteria, it does
also remain available for Strength of Materials calculus. On the other hand anyone
can observe the high variety of the surrounding bodies, objects or structural elements.
It is the engineering designers' practical experience which has to choose, for each
particular body or structural element, the appropriate class from those presented
above. The fundamental Strength of Materials problems, concepts and theories have
been constructed on the bars class. Many of these theoretical and practical
investigations results have been then extended to the other two classes: plates and
blocks.

11
Strength of Materials

1.5 LOADS SCHEMATIC REPRESENTATION IN STRENGTH


OF MATERIALS CALCULUS
Since Strength of Materials deals with the relations between externally applied
loads and their internal effects on bodies, the clear understanding of the external
loads action represents a first and necessary step in the analysis of a given mechanical
structure.
In fact, a load represents the action of a certain body on another body. It
models therefore the interaction between two or more bodies or structural elements.
There are two main classes the loads may be classified in: forces and couples
(moments of forces).
When International System metric units are being used, forces (usually denoted
by P or F) are expressed in newtons (N) and couples (moments of forces - usually
denoted by M) are expressed in newton-meters (N m).
However, when one finds that these units are exceedingly small or big
quantities, multiples or submultiples of these units may be also used.
Within the framework of different criteria, several specific types of loads have
to be differentiated:

a) Depending on the size of the interacting bodies contact surface area, we may
have:
concentrated forces: when the contact surface area may be theoretically
reduced to a point, Fig 1.11.

Fig. 1.11

distributed forces: when the interaction of the bodies in contact is


transmitted through a surface of a certain area, Fig 1.12.

Fig. 1.12

Furthermore, the forces may be distributed on surfaces (Fig 1.13) or, if a


dimension of such surfaces is too small, one can talk about forces distributed on lines,
Fig.1.14.
12
Introduction

Fig. 1.13 Fig. 1.14

In Fig. 1.15 some examples of forces distributed on lines in case of a simply


supported beam have been represented.

Fig. 1.15

13
Strength of Materials

The above mentioned concentrated and distributed forces may be encountered


in a wide range of practical situations or engineering applications, Fig.1.16.

a.

A liquid in a vessel. The lateral vessel walls are


subjected to a linearly distributed force (pressure
p g x , where is the liquid density, g is the
gravitational acceleration and x is the depth at the b.
level where the pressure is measured. A pressurized gas in a thin
walled vessel. The force p acting
on the vessel walls is uniformly
distributed.

A uniform snow layer acts on a


house root with a uniformly
distributed force.

A railway vagon acts on the


railways with mobile
c.
concentrated forces.

d.

Fig. 1.16

14
Introduction

It is to be noted that the moments of forces may be classified in the same manner.

b) Depending on the loads variation in time we may have:

constant loads: the loads (P or


M) remain constant in time (Fig.
1.17).

Fig. 1.17

loads with a periodical variation


in time (Fig. 1.18).
In such cases the loads do
periodically vary in time, between a
maximum and a minimum value.
Fig. 1.18

loads with a random variation in


time (Fig. 1.19).

Fig. 1.19

c) Depending on the time in which a load is applied to a certain body we may


have:

static loads: which do slowly vary from zero up to the nominal value, remaining
then constant in time (Fig. 1.20a).

a. b.
Fig. 1.20

15
Strength of Materials

dynamic loads: which vary in a very short time from zero up to the maximum
value (Fig. 1.20b). An example of this type has been represented in Fig.1.21.
The rod BD of uniform cross
section is hit at its end B by a body
of mass m, moving with a velocity
v0. The rod deforms under the
impact with . After vibrating for
a while, the rod will come to rest
and all its internal stresses will
disappear. Such a sequence of
events is referred to as an impact
loading (this load is an example of
dynamic loads class).

Fig. 1.21

1.6 SUPPORTS AND REACTIONS

The mechanical connections between bodies subjected to different loads -


and the surrounding environment are provided by the mechanical supports.
From our knowledge of statics we have to recall three basic types of
mechanical supports:
simple supports: which prevent the supported element from moving linearly
along a direction perpendicular to the supporting base ( Fig. 1.22).

Fig. 1.22

In Fig. 1.23 an example of a simply supported beam has been represented.


Usually the supporting points are denoted by capital letters A , B , C

Fig. 1.23 Fig. 1.24

16
Introduction

pin-connections: which prevent the supported element from moving linearly along
two perpendicular directions at the supporting point (Fig. 1.24).

In Fig. 1.25 an example of a


beam, simply supported at A and pin
connected at B has been shown.
Fig. 1.25

fixed connections: which prevent the


supported element from moving
linearly or rotating at the supporting
point. In Fig.1.26 a column with one
fixed end B and one pin-connected end
A has been represented. If a structural
element is fixed at a certain point, it is
usually said that the element is
embedded at that point.
For any of the above presented
types of connections, if the support
prevents the supported element from
moving or rotating along a certain
direction then, along that direction a
Fig. 1.26 reaction occurs.

Depending upon the number of prevented linear movements or rotations, in


case of plane problems, the following reactions will develop (Fig. 1.27).

simple support pinconnection fixed connection:

Fig. 1.27

As shown in Fig.1.27, the horizontal reactions are denoted by X while the


vertical reactions by Y, with subscripts representing the involved supporting point.
The reactions represent in fact the action of the surrounding environment on bodies or
structural elements subjected to external loads. These reactions can be then calculated
using different equations representing the bodies or structural elements conditions of
mechanical equilibrium.

17
Strength of Materials

Let us consider now, for example,


a beam, simply supported at end A
and pinconnected at end B,
subjected to a uniformly distributed
force p and a concentrated force P
(Fig. 1.28). Due to the action of the
external loads p and P, at the level
of the supporting points A and B
the reactions YA, XB and YB will
develop.
Fig. 1.28

These reactions may be computed from the beam conditions of mechanical


equilibrium. After the reactions have been computed, there will be no difference
between the external loads p and P and these reactions, all representing a global
system of loading for a body in mechanical equilibrium (Fig.1.29).

Fig. 1.29

1.7 EQUILIBRIUM EQUATIONS

The equilibrium equations represent the mathematical expression of


equilibrium for a body or a structural element subjected to external loads (original
loads and reactions).
In case of plane problems (2D problems) the conditions of equilibrium consist
in two equations representing the projections of all forces about two arbitrary
perpendicular directions in the plane and one equation representing a summation of
moments about an arbitrary point of the plane.
Let us consider for example a beam, simply supported at A and pinconnected
at B (Fig. 1.30), supporting a load P.

18
Introduction

Fig. 1.30

Due to the action of force P, reactions YA, XB, YB will develop. In such cases any
strength of materials problem has to start with the computation of the unknown
reactions YA, XB, YB . In writing the mathematical conditions of equilibrium a
coordinate system must be firstly chosen (Oxy), Fig. 1.30. After that, three necessary
equilibrium equations have to be written as follows:
summation of forces about Ox axis equals zero :
X 0 X B
0,
summation of forces about Oy axis equals zero :
Y 0 YA YB, P 0 YA YB P
summation of moments about the pin support B equals zero:
P(a b)
M B
0 P(a b) - Y A b 0 YA .
b
Substituting for YA into Y =0, we write:
P a b Pb Pa Pb Pa
YB P YA P .
b b b
We may write therefore:
P a b Pa
X B
0;Y A ;YB .
b b

The negative sign of YB expresses that the real physical sense of reaction YB is
opposite than that shown in Fig. 1.30.
Once the reactions have been computed, strength of materials doesnt make
any differentiation between the original loading (represented by force P) and these
reactions, all being treated as a global system of forces acting upon a body in
mechanical equilibrium (Fig. 1.31).
Following the same
reasoning, in case of 3D problems
six equilibrium equations are
required: three equations of forces
and three equations of moments.

Fig. 1.31

19
Strength of Materials

SAMPLE PROBLEMS
For the 2D mechanical structural elements presented below, supported and loaded as
shown, determine the values of reactions using the required mathematical equilibrium equations:

X 0 X B
0;

Y 0 YA YB P 0;

M B
0 YA P b 0.

We write therefore:
P b P a
YA ; YB ; X B
0.

Fig. 1.32

X 0 X B
0;

Y 0 YA YB 3P 0;

M B
0 Y A 800 600 P 400 600 2 P 300 0;

We may finally write:


YA 114 , 28 N ;

YB 185 , 72 N ;

X B
0.

2
X 0 X A
10 2 0;
2
X A
10 kN .

Fig. 1.33 2
Y 0 YA 10 2 5 2 0;
2
YA 20 kN .

2
M A
0 20 10 2 3 5 2 1 M A
0;
2
M A
20 kN m.

Fig. 1.34 2
X 0 X A
20 0 X A
14 ,142 kN ;
2
2
Y 0 YA YB 10 20 0 ;
2
YA YB 24 ,142 kN .

2
M A
0 Y B 1500 10 1000 20 500 0;
2
YB 11 , 38 kN .

Finally we may write:


Fig. 1.35 X A
14 ,142 kN ; Y A 12 , 762 kN ; Y B 11 , 38 kN ;

20
2. INTERNAL FORCES IN STATICALLY DETERMINATE
MEMBERS
2.1 INTERNAL FORCES COMPUTATION

Consider a body of arbitrary shape acted upon by several external loads (Fig.
2.1). In statics, we would start by determining the resultant of the applied loads to
determine whether or not the body remains at rest. If the resultant is zero, we have
static equilibrium a condition generally
prevailing in structures. If the resultant is
not zero we may apply inertia forces to
bring about dynamic equilibrium. Such
cases will be discussed later under dynamic
loading. For the present, we consider only
cases involving static equilibrium.
Generally speaking, when loads are
applied to a certain mechanical structure or
machine, each component of such a
Fig. 2.1 structure or machine is subjected to
external loads of different values (Fig. 2.1). Under the action of the external loads,
internal forces occur inside the involved component (assimilated to the arbitrary
body represented in Fig. 2.1). If these internal forces reach critical values the body
(component) will fail.
One of the methods commonly used for the determination of internal forces in
strength of materials is known as the method of sections. In fact the problem remains
the same like that presented in the previous chapter: what does every point of the
body (generically represented in Fig. 2.1) ,,feel when the body is subjected to
external loads in mechanical equilibrium?

Fig. 2.2 Fig. 2.3


The method of sections consists in passing an exploratory plane through an
arbitrary point of the body and with an arbitrary orientation (Fig. 2.2). In this way
two distinct segments of the body will occur (Fig.2.3), the left surface (SL) and the
Strength of Materials

right surface (SR) representing the internal plane surfaces of the body, originally in
contact. Since the body represented in Fig. 2.2 is in equilibrium, neither of the two
segments of Fig. 2.3 can be in equilibrium. If we want to bring the segment II for
example in the same state it is in Fig. 2.2, the action of segment I on segment II
(which actually exists inside the body represented in Fig. 2.2) has to be considered.
This action may be reduced at the
centroid O of surface SR to a resultant
force R and a resultant moment M (Fig.
2.4).
In other words R and M represent
the action of segment I on segment II, as
a global mechanical effect occured at the
level of the entire section SR. In fact this
mechanical effect develops inside the
body represented in Fig. 2.2.
Fig. 2.4 Furthermore it is to be noted that M and
R represent the effect of the external
loads acting on the segment I (i.e. P1, Pn, M1 Fig. 2.2) which develops inside the
body at the level of SR.
The resultant force R and the resultant moment M are called internal forces.
Under the action of M, R, P2, Pk, Mk the segment II of the body is now in mechanical
equilibrium (as it really is in the actual state of Fig. 2.2). Now using the adequate
equilibrium equations for segment II, the values of internal forces R and M can be
derived.
A similar reasoning may be also applied to the segment I of the body, on which
internal forces R and M develop (Fig. 2.5). From action and reaction mechanical
law we may write:
R = - R,
M = - M.
Let us now apply the above
reasoning to a loaded statically
determinate member (Fig. 2.6). It is to
be mentioned that a statically
determinate member is a member for
which all reactions can be completely
computed from statics alone.
Fig. 2.5 After computing the specific
reactions (YA, YB, ZBetc) corresponding to the supporting points A and B, the
member represented in Fig. 2.6 is in fact a body subjected to several external loads in
mechanical equilibrium. Passing an exploratory plane at some arbitrary point of the
member, perpendicular to the axis of the member, and considering only a segment of

22
Internal forces in statically determinate members

the member (just like in the preceding discussion) the internal forces R and M are
revealed (Fig. 2.7).
We do also attach to the segment
considered in Fig. 2.7 a coordinate
system whose origin is taken at the
centroid O of the exploratory cross
section (SR). Ox is the axis of the
member while Oz and Oy represent the
axes to which the exploratory cross
section of the member is reported.
For convenience, the internal
forces M and R are resolved into Fig. 2.6
components that are normal and tangent to the cross section considered, within the
chosen coordinate system (Fig. 2.8) - R resolved into components N, Ty and Tz, while
M into components Mx, Miy , Miz.

Fig. 2.7 Fig. 2.8


Each component reflects a certain effect of the applied loads on the member
and is given a special name as follows:

N: axial force (R component along Ox axis, or, more briefly, x axis). This
component measures the pulling (or pushing) action perpendicular to the section
considered. A pull represents a tensile force that tends to elongate the member,
whereas a push is a compressive force that tends to shorten it.

Ty, Tz: shearing forces (R components along y axis and z axis respectively). These are
components of the total resistance to sliding the portion to one side of the
exploratory section past the other. The resultant shearing force (acting on zOy plane)
is usually denoted by T, and its components by Ty and Tz to identify their directions.

Mx: twisting couple (twisting moment or torque) (M component along x axis). This
component measures the resistance to twisting the member and is commonly given
the symbol Mt.

Miy, Miz: bending moments. These components measure the resistance to bending the
member about the y or z axes and are often denoted merely by Miy and Miz.

23
Strength of Materials

The quantities N, Ty, Tz, Mx, Miy, Miz are also called internal forces. Each of
them produces a certain type of mechanical effect on the involved member:

N : axial loading ;
T y , T z : shearing loading ;

M t : torsion ;

M iy
,M iz
: bending .

The simultaneous presence on the current member cross section of two or


more types of internal forces determines a combined loading.
Although the type of coordinate system used within such analysis is, in a way,
controversial, we shall use the following sign convention:

N, Ty and Tz should be considered positive if orientated to the opposite sense


of the axes;

Mt, Miy and Miz should be considered positive if orientated to the sense of
the axes.

From the preceding discussion, it is obvious that the internal effect of a given
loading depends upon the selection and orientation of the exploratory section. In
particular, if the loads act in a single plane, say the xy plane as is frequently the case,
the six components of Fig. 2.8 reduce to only three namely, the axial force (N), the
shearing force (T) and the bending moment Miz. This in why, in case of plane
problems (when plane members are subjected to loads contained in the same plane)
the internal forces refer to only three components whose positive sign convention
should be taken as follows:

Fig. 2.9

The positive sign convention represented in Fig. 2.9 should be used for
plotting the axial forces, shearing forces and bending moments diagrams. As it will
be explained later, the positive sign convention corresponding to the face SR is used
when the member is covered from the left to the right while the positive sign
convention corresponding to the face SL is used when the member is covered from
the right to the left.

24
Internal forces in statically determinate members

2.2 DIFFERENTIAL RELATIONS AMONG LOAD, SHEAR AND


BENDING MOMENT

Let us now consider a simply supported beam AB of span , carrying a


distributed load p per unit length (Fig. 2.10a), and let C and C be two points of the

a. b.
Fig. 2.10
beam at an infinitely small distance dx from each other.
We shall detach the portion CC of the beam and draw its free body diagram
(Fig. 2.10b). The forces exerted on the free body include a load of magnitude pdx
and the internal forces at C and C as shown. The shear and bending moment at C
will be denoted by T and M respectively, and will be assumed positive while the
shear and bending moment at C will be denoted by T + dT and M + dM respectively.
Since the shear and bending moment are assumed to be positive, the internal forces
will be directed as shown in Fig. 2.10 b. It is also to be mentioned that since the
distance dx between C and C is considered infinitely small, the load p may be
assumed uniformly distributed per length dx and may be replaced by a resultant pdx,
(Fig. 2.11).

Fig.2.11.

From the mechanical equilibrium of the detached segment CC we write:


the summation of forces about the vertical direction is zero:
Fy 0 T dT pdx T 0 .

25
Strength of Materials

We write
. pdx dT
Dividing by dx the two members of the equation, we have:
dT
p . (2.1)
dx

writing now that the summation of moments about C is zero, we have:


dx
M C '
0 M dM pdx M Tdx 0 .
2
2
p dx
The third member of this equation being an infinitely small quantity
2
compared with the others, it may be neglected and we may write:
dM Tdx .
Dividing now the two members of the equation by dx we obtain:
dM
T . (2.2)
dx

Relations (2.1) and (2.2) may be written in a single one as follows:


2
d M dT
2
p . (2.3)
dx dx

The above presented relations may be successfully used for plotting the shear
and the bending moment diagrams. Generally speaking, internal forces diagrams (i.e.
diagrams of axial forces, shearing forces, torsion and bending moments) are a
graphical representation of the successive values of axial force N, shearing force T,
torque Mt and bending moment Mi in the various sections against the distance
measured from one end of the involved member.
In particular, relations (2.1), (2.2) and (2.3) bring us several important rules
concerning the shear and bending moment diagrams:

The distributed force p measures the tangent slope of the shear curve (shear
diagram). If p = 0, the shearing force will be constant;

It should be observed that Eq. (2.1) is not valid at a point where a concentrated
force is applied. At such a point the shear curve is discontinues and a sudden
change occurs in the diagram. The value of the sudden change in the shear
diagram, when a concentrated force is applied, equals the value of that
concentrated force;
dM
Equation (2.2) indicates that the slope i
of the bending moment diagram is
dx
equal to the value of the shearing force. This is true at any point where the
26
Internal forces in statically determinate members

shearing force has a well defined value, i.e. at any point where no concentrated
load is applied;

Equation (2.2) does also show that T = 0 at points where M is maximum. This
property facilitates the determination of the points where the beam (a member in
bending is often referred to a beam) is likely to fail under bending;

If a concentrated couple is applied at an arbitrary point of the beam, a sudden


change in the bending moment diagram occurs, the change value being equal to
the applied concentrated moment (couple);

Equation (2.3) shows that the shear and the bending moment curves will always
be, respectively, one or two degrees higher than the load curve. For example if the
load curve is a horizontal straight line (the case of an uniformly distributed load
p), the shear curve is an oblique straight line and the bending moment curve is a
parabola. If the load curve is an oblique straight line (first degree), the shear curve
is a parabola (second degree) and the bending moment curve is a cubic (third
degree).

With the above rules in mind, we should be able to sketch the shear and the
bending moment diagrams without actually determining the function T(x) and M(x)
along the member, once a few values of the shear and the bending moment have been
computed. The sketches obtained will be more accurate if we make use of the fact
that, at any points where the diagrams are continuous, the slope of the shear curve is
equal to (- p) and the slope of the bending moment curve is equal to T.

For plotting the internal forces diagrams, the following steps have to be
covered:

a) Denoting of the important points. An important point of a member is a point


where a certain change (geometrical, loading, etc) occurs. The supporting points
are usually denoted by capital letters A, B, C, etc. and the other important points
by figures 1, 2, 3 etc.;
b) Two successive important points define a portion of the member;
c) Determination (when necessary) the magnitude of the reactions at the supports;
d) A covering sense of the member has to be chosen (from the left to the right, from
the right to the left or both);
e) For each distinct portion of the member a current cross section at distance x from
one end of the involved portion has to be considered;
f) For the current cross section considered, each distinct internal force (N, T, Mt, Mi)
has to be mathematically expressed as a function of x: N(x), T(x), Mt(x), Mi(x);
g) Plotting the functions N(x), T(x), Mt(x), Mi(x) along the entire member, the
internal forces diagrams are finally obtained.

27
Strength of Materials

2.3 EXAMPLES CONCERNING THE MAIN TYPES OF


DIAGRAMS

2.3.1 AXIAL FORCES DIAGRAMS

Example 1
Draw the axial force diagram for the horizontal member with one fixed end
and uniform cross section, shown in Fig. 2.12.

a. b.
Fig. 2.12

Step 1 important (main) points: 1, 2 and A;


Step 2 main portions of the member: 1 - 2; 2 - A;
Step 3 the magnitude of reactions may be determinated using the condition of
mechanical equilibrium:
Fx = 0 P + 2P - XA = 0 XA = 3P ;
Step 4 the covering sense of the member: let us say it is from the left to the
right;
Step 5 we first consider the first portion of the member (1 -2) and an
exploratory current cross section located at distance x from end 1 of the
portion. Looking to the left one can conclude that the single axial force
component acting upon the current cross section considered is equal to
P. For any value of x this component remains constant. This is why, for
portion 1 - 2, the axial force will be constant (N P). The corresponding
axial force diagram of portion 1 - 2 has to be hachured perpendicularly
to a reference horizontal line. Since the covering sense of the member
was chosen from the left to the right, the positive sign convention I has
been used (Fig. 2.12a).
In the same manner, the axial force for the second portion 2 - A of the member
is:
N2-A = P + 2P = 3P = ct.

28
Internal forces in statically determinate members

It is to be mentioned that portion 2 - A for example, could have been covered


from the right to the left as well. In such a case the current cross section is taken at
distance x from A and, looking to the right, we have:
NA-2 = XA = 3P (the same value as above).
When covering the member from the right to the left, the positive sign
convention II should be taken (Fig.2.12b).
The above presented algorithm for plotting the axial forces diagrams remains
unchanged even if the loading or the geometry are much more complicated or the
internal forces are not axial but shearing forces or bending moments.
The following examples will be accompanied by no supplementary
explanations.
Example 2
Draw the axial force diagram
for the member supported and
axially loaded as shown in Fig. 2.13.
Portion 1 - 2 or, more simple, 1 - 2:
N(x) = 0;
Portion 2 - A:
Fig. 2.13 N(x) = 3P = ct.

Example 3
Draw the axial force diagram
for the member shown in Fig. 2.14.
Fx = 0; XA - 20 - 10 - 5 2 = 0;
XA = 40 kN.
Portion 1 - 2:
N(x) = 20 kN;
Fig. 2.14
Portion 2-A:
x 0; N 2 30 kN ;
N(x) = 20 + 10 + 5 x = 30 +5 x ;
x 2 m; N A 40 kN .

2.3.2 SHEAR AND BENDING-MOMENT DIAGRAMS

Fig. 2.15 shows a simply supported beam that carries a concentrated load P,
being held in equilibrium by the reactions YA and YB. For the time being we neglect
the mass of the beam and consider only the effect of load P. Applying the method of
sections, let us assume that a cutting plane d - d, located at a distance x from point A,
divides the beam into two segments.
29
Strength of Materials

Fig.2.15
The free-body diagram of the left segment (Fig. 2.16) shows that the externally
applied load is YA . To maintain equilibrium in this segment of the beam the internal
forces occurring at the level of the exploratory section d - d must supply the resisting
forces necessary to satisfy the conditions of static equilibrium. In this case, the
external load is vertical, so the condition Fx= 0 (the x axis is horizontal) is
automatically satisfied.

Fig. 2.16
Since the left segment of the beam is in equilibrium, the resisting shearing
force T acting on the left segment has to be numerically equal to YA. In other words,
the shearing force in the beam may be determined from the summation of all vertical
components of the external loads acting on either side of the section. However, it is
simpler to restrict this summation to the loads that act on the segment to the left of
the section. This definitions of the shearing force (also called vertical shear or just
shear) may be expressed mathematically as:
T Fy
L
, (2.4)

the subscript L emphasizing that the vertical summation includes only the external
loads acting on the beam segment to the left of the section being considered.
In computing T, when the beam is covered from the left to the right, upward
acting forces and loads are considered as positive (see also the sign convention
presented in the preceding section). This rule of sign produces the effect shown in
Fig. 2.17, in which a positive shearing force tends to move the left segment upward
with respect to the right segment, and vice versa.

Fig. 2.17

30
Internal forces in statically determinate members

For a complete equilibrium of the free-body diagram in Fig. 2.15 and Fig. 2.16
the summation of moments must also balance. In this discussion YA and T are equal,
thereby producing a couple Mi that is equal to YA x and is called the bending moment
because it tends to bend the beam.
Analogous to the computation of T at the current cross section, the bending
moment is defined as the summation of moments about the centroidal axis of any
selected cross section of all loads acting either to the left or to the right side of the
section, being expressed mathematically as:
M i
M L
M R
, (2.5)
where the subscript L indicates that the bending moment is computed in terms of the
loads acting to the left of the section, while the subscript R referring to loads acting
to the right of the section.
Why the centroidal axis of the exploratory section must be chosen as the axis
of bending moment may not be clear at this moment; this will be explained later.
To many engineers, bending moment is positive if it produces bending of the
beam concave upward, as in Fig. 2.18.

Fig. 2.18
We prefer to use an equivalent convention, which states that the upward acting
external forces cause positive bending moments with respect to any section while
downward forces cause negative bending moments. Therefore, if the left segments of
the beam is concerned (Fig. 2.16), this is equivalent to taking clockwise moments
about the bending axis as positive, as indicated by the moment sense of YA. With
respect to the right segment of the beam (Fig. 2.16) this convention means that the
moment sense of the upward reaction YB is positive in counterclockwise direction.
This convention has the advantage of permitting a bending moment to be computed,
without any confusion in sign, in terms of the forces to either the left or the right of a
section, depending on which requires the least mathematical work. We never need
think about whether a moment is clockwise or counterclockwise; upward acting
forces always cause positive bending moments regardless of whether they act to the
left or to the right of the exploratory section.
The definition of shearing force and bending moment may be summarized
mathematically as follows:
T Fy Fy ;
L R

M i
M L
M R
,

31
Strength of Materials

in which positive effects are produced by upward forces and negative effects by
downward forces.
This rule of sign will be used exclusively hereafter. To avoid conflict with this
rule, we must compute vertical shear in terms of the forces lying to the left of the
exploratory section. If the forces acting to the right of the section were used, it would
be necessary to take downward forces as positive so as to agree with the sign
convention shown in Fig. 2.17.
Example 1
Draw the shear and bending-
moment diagrams for the cantilever beam
shown in Fig. 2.19. (A cantilever beam is
a beam with a fixed end, subjected at its
free end to a single concentrated force P).
We observe that the internal forces
exerted on a current cross section at
distance x from the free end 1 are
represented by:
a shearing force T of magnitude T = -
P (see the positive sign convention);
Fig. 2.19 a bending moment
x 0 Mi 0;
Mi = - P x: 1
x Mi P .
A

We note (Fig. 2.19) that the


negative values corresponding to the
bending-moment diagram are represented
above the reference line. In this way the
bending-moment diagram shows us how
the involved beam deforms under the
action of the external loads.
Example 2
Draw the shear and bending-
moment diagram for a simply supported
beam AB, of span subjected to a single
concentrated load P (Fig. 2.20) the case
Fig. 2.20 of Fig. 2.15.
We first determine the reactions at the supports from the free-body diagram of
the entire beam (Fig. 2.20); we find that:
P b P a
YA ; YB .

32
Internal forces in statically determinate members

For the portion A - 1, cutting the beam at distance x from end A, we have:
T = YA = constant;
x 0 Mi 0;
A
Mi = YA x: P b Pab
x a Mi YA a a .
1
Pab
While the bending moment increases linearly from M = 0 at x = 0 to M at x =

a, we note that the shear has a constant value. Even if the problem is quite simple, it
is more convenient to cover the second portion of the beam from the right to the left.
Therefore, cutting the beam at distance x from end B and using the adequate
sign convention we have:
B - 1:
P a
T YB ;

x 0 M i 0;
B
Mi = YB x: Pab
x b M i .
1
We can now complete the shear and bending-moment diagrams (Fig. 2.20).
For portion B - 1 the shear has a negative constant value while the bending moment
Pab
increases linearly from M = 0 at B to M at 1 (for x = b).

Remarks
If a concentrated traverse force acts at a section of the beam, a sudden change in
the shear diagram at that section occurs, the sudden change value being equal to
that concentrated force. In our case of Fig. 2.20, at point 1, the sudden change is:
Pb Pa P a b P
P.

If, for a certain portion of the beam, the shear is constant, the bending-moment
diagram is linear;

Covering the beam from the left to the right within the portion A - 1 and then from
the right to the left within the portion B - 1, and since at point 1 there is no
concentrated moment, there will be no sudden change in the bending-moment
diagram at point 1. This is why we have obtained the same value of the maximum
bending moment at 1;
The covering sense of the beam, when plotting such diagrams, has no importance.
It may be chosen from the left to the right, or from the right to the left or
combined, as it is convenient to us;

33
Strength of Materials

When designing a beam like that presented in Fig. 2.20, we must note that the
strength of the beam is usually controlled by the maximum absolute value Mi max
Pab
of the bending moment in the beam (in our case M i
max
).

We note from the foregoing example that, when a beam is subjected only to
concentrated forces, the shear is constant between the applied forces while the
bending-moment varies linearly between the forces. In such situations, therefore, the
shear and bending-moment diagrams may easily be drawn, once the values of T and
Mi have been obtained at sections selected just to the left and just to the right of the
points where the loads and reactions are applied.

Numerical examples
1. Draw the shear and bending-moment diagrams for a simply supported beam subjected to two
concentrated loads (forces) as shown in Fig. 2.21.
Determination of the reactions at the supports
Fy= 0 ; YA -5 - 10 + YB = 0 YA + YB = 15 kN ;
MA = 0 ; YB 4 - 10 3 - 5 1 = 0 YB = 8,75 kN ;
MB = 0 ; YA 4 - 5 3 - 10 1 = 0 YA = 6,25 kN .

Portion A - 1
T = YA = 6,25 kN ;
x 0 M iA 0;
Mi = YA x; x 1m Mi 6 , 25 kN m.
1

Portion 1 - 2
T = YA - 5 = 6,25 - 5 = 1,25 kN ;
Mi = YA (1 + x) - 5x .
This means that
x 0 Mi 6 . 25 kN m;
1

x 2m Mi 6 , 25 1 2 5 2 8 , 75 kN m.
2

For the last portion it is much more convenient to


Fig. 2.21 cover the beam from the right to the left.
Portion B - 2
T = - YB = - 8,75 kN;
x 0 M i 0;
Mi = YB x B
x 1m M i 8 , 75 1 8 , 75 k N m.
2

We obtain therefore the shear and bending-moment diagrams shown in Fig. 2.21.

34
Internal forces in statically determinate members

2. Draw the axial force, shear and bending - moment diagrams for the beam shown in Fig. 2.22.
The beam represented in Fig. 2.22 can
be drawn in a simplified manner as shown
in Fig. 2.23.
As in preceding example, the reactions
are determined by considering the entire
beam as a free body, they are:
XB = 10 kN; YA = 22,5 kN; YB = 2,5 kN.
Portion 1 - A
Fig. 2.22
N 10 2 cos 45 10 kN ;

T 10 2 sin 45 10 kN ;

2
Mi 10 2 x 10 x :
1 A 2
x 0 Mi 0;
1

x 1m Mi 10 kN m.
A

Portion A - 2

N 10 2 cos 45 10 kN ;


T 10 2 sin 45 YA

10 22 , 5 12 , 5 kN ;

2
M i 10 2 1 x YA x
A 2
2
10 1 x 22 , 5 x :
x 0 M i 10 kN m;
A
Fig. 2.23 x 1m M i 2 ,5 k N m.
2

It is more convenient to us to cover the last portion of the beam from the right to the left.
Portion B 2
N X B 10 kN ;

T YB 2 , 5 kN ;
x 0 M i 0;
B
M i
B 2
YB x 2 ,5 x :
x 1m M i 2 ,5 k N m.
2

We can now complete the axial force, shear and bending-moment diagrams of Fig. 2.23. We
note that the axial force has a constant value along the beam; the shear has also constant values
between the important points of the beam while the bending moment varies linearly. At points
(sections) where concentrated forces act, sudden changes in the shear diagram occur (whose values
equals the applied concentrated forces). Since there are no concentrated moments on the beam there
will be no sudden changes in the bending-moment diagram.
35
Strength of Materials

Example 3
Draw the shear and bending-moment diagrams for a beam, simply supported
at its ends and subjected to a uniformly distributed load p (Fig. 2.24).
Due to the symmetry of loading and geometry, the reactions are:
p
YA YB .
2

As usually, we cut the beam at distance


x from A and note that:
p
T YA px px :
2
p
x 0; TA ;
2

x ; T 0;
2
p
Fig. 2.24 x ; TB .
2
2
x p px x 0; M A 0;
M i YA x p x x :
2 2 2 x ; M B 0.

Within the calculus, the distributed load over the current portion of the beam
has been replaced by its resultant px applied at the midpoint of the involved portion.
Since at the midpoint of the beam the shear equals zero, the bending moment
reaches a maximum value at that point:
2 2
p p p
M max M i .
2 2 2 2 2 8

We do also note that the shear diagram is represented by an oblique straight line (Fig.
2.24), while the bending-moment diagram by a parabola. In the section where T = 0,
the bending-moment has a maximum value.
Example 4
Draw the shear and bending-moment diagrams for a beam, simply supported
at its ends and subjected to a linearly distributed load (Fig. 2.25).
The entire beam is taken as a free body, and, from the conditions of
equilibrium, we write:

36
Internal forces in statically determinate members

p0
Fy = 0 ; YA YB ;
2

p0
MB = 0 ; YA 0;
2 3
p0
YA ;
6
p0 2
MA = 0 ; Y B 0;
2 3
p0
YB .
Fig. 2.25 3

Using the first equation of equilibrium ( Fy = 0) we check that the values


already obtained for YA and YB are correct.
Now writing the mathematical expressions of the shear and bending-moment
at an arbitrary section at distance x from end A, we have:
px x
T YA ;
2
px x x
M i YA x .
2 3

From similar triangles we may write:


px x x
px p0 ,
p0

which substituted in the preceding expressions of T and Mi, leads to :


2
px x p0 x x p0 p0 x
T YA p0 ;
2 6 2 6 2

p0
x 0; TA = ;
6
2
p0 p0 p0
x = ; TB = = - ;
6 2 3

2 3 x 0; M = 0;
px x x p0 x x p0 x p0 x iA
M i YA x x p0 :
2 3 6 6 6 6 x = ; M iB = 0;

37
Strength of Materials

The shear curve is thus a parabola while the bending-moment curve is a third degree
function. The shear curve intersects the x axis at a distance given by equation:
2
p0 p0 x
T 0 0 x .
6 2 3

Therefore, the maximum value of the bending moment occurs at x , since T
3
dM i
(and thus ) is zero for this value of x:
dx
3 2
p0 p0 p0
M i .
3 6 3 6 3 9 3

Example 5
Draw the shear and bending-moment diagrams for a simply supported beam,
subjected to a concentrated moment M0 applied at point 1 (Fig. 2.26).

The entire beam is taken as a free body and


we have:
M 0 M 0
YA ; YB .

The negative sign of YB indicates that the
real sense of this reaction is opposite to
that represented in Fig. 2.26.
The shear at any section is constant
and equal to M0 / . Since a concentrated
moment (couple) is applied at 1, the
Fig. 2.26 bending-moment diagram is discontinuous
at 1; the bending-moment decreases
suddenly by an amount equal to M0.

Remark
The concentrated moment in Fig.
2.26 symbolizes for example the action of
two equal and opposite concentrated forces
as shown in Fig. 2.27, where M0 = P d.
Fig. 2.27

38
Internal forces in statically determinate members

A complex sample problem


Sketch the shear and bending-moment diagrams for the simply supported beam shown in
Fig. 2.28.

Considering the entire beam as a


free body, we determine the reactions as
follows:
Fy = 0 ;YA + YB + 5 - 10 1 = 0;
YA + YB = 5;
MB = 0 ; 5 3 -10 1 2,5 + YA 2 + 15 = 0;
YA = - 2,5 kN;
MA = 0 ; 5 1 -10 1 0,5 + 15 - YB 2 = 0;
YB = 7,5 kN.
Using the first equation of
equilibrium ( Fy = 0) we check that the
Fig. 2.28 two values obtained for YA and YB are
valid.
Next we draw the shear and bending-moment diagrams. The sketches obtained will be more
accurate if we make use of the fact that, at any point where the curves are continuous, the slope of
the shear curve is equal to -p while the slope of the bending-moment curve is equal to T.
Portion 1 - A
x 0; T1 5 kN ;
T 5 10 x:
x 1m; TA 5 kN .

This means that at the midpoint between 1 and A (for x = 0,5 m) the shear is zero, and, therefore, the
bending-moment reaches a maximum value. It is to be mentioned that this point of maximum for
the bending moment is valid only for the involved portion (i.e. 1 - A). Within the other portions of
the beam the bending-moment could reach grater values as well. This is why, the maximum value of
the bending reached within a certain portion of the beam is called a local maximum. There are cases
in which a local maximum does also represent a global maximum too.

x 0; M i 0;
1
x 2
M i 5x 10 x 5x 5x : x 1 m; M i 0;
A
2
x 0 ,5 m ; M i 1 , 25 k N m .
MAX

Portion A - 2
T 5 10 1 YA 5 10 2 ,5 7 , 5 kN ;

x 0; M i 0;
A
M i 5 1 x 10 1 0 ,5 x :
x 1 m; M i 7 ,5 k N m .
2

39
Strength of Materials

M i
2
= - 7,5 kN m tells us that close to the end 2 of the portion A - 2, the bending moment reaches
such a value.
It will be more convenient to us to cover the last portion of the beam from the right to the
left (i.e. from B to 2):
Portion B - 2
T YB 7 , 5 kN ;

x 0; M i 0;
B
M i YB x 7 ,5 x :
x 1 m; M i 7 ,5 k N m .
2

This time, M i
2
= - 7,5 kN m tells us that close to the end 2 of the portion B - 2, the bending
moment reaches such a value. In this way we have obtained two values for the bending-moment at
point 2: one for the portion A - 2, close to the point 2 to the left and one for the portion B - 2 close to
the same point 2 but to the right. Since at point 2 there is a concentrated moment acting on the beam
(equal to 15 kN m), the sudden change in the bending-moment diagram at point 2 is correct.

2.3.3 TORQUE DIAGRAMS

In the preceding sections we have discussed about axial forces, shear and
bending-moment diagrams. Here we shall consider members which are in torsion.
More specifically we shall learn to draw internal forces diagrams for members
subjected to twisting couples or torques.
We say that a member is subjected to torsion if at any cross section of the
member, the internal forces are represented by a torque vector directed along the axis
of the member.
To sketch the torque diagrams the method of sections may be used, as
presented in the preceding sections.

Draw the torque diagram for a member fixed at one end and subjected to
concentrated and uniform distributed torques as shows in Fig. 2.29.
Considering the entire
member as a free body we obtain
the reactions at A.
M x 0
M
M A
M 0
2M 0
0
4M 0
;

Using the method of sections
and covering the member from 1 to
A we have:
1-2
Mt M 0;
2-3
Fig. 2.29 M t M 0 2M 0 3M 0 ;

40
Internal forces in statically determinate members

3-A
M x 0; M 3
3M 0
;
M t
M 0
2M 0
m x M 0
2M 0
0
x ;
x ; M A
4M 0
.

We note that, in such a case, the sign used for torques is not so important. As
soon as a certain sign has been adopted for the first met torque, the signs for the
other torques have to be adopted consequently. Therefore, the torque diagrams may
be sketched above or below the reference line. Like in the preceding examples, if a
concentrated torque acts at a certain section of the member, at that point a sudden
change in the torque diagram occurs (the change being equal to that concentrated
torque). The torque diagrams are usually hachured as shown in Fig. 2.29.

2.4 SUPERPOSITION METHOD


The superposition principle is a consequence of the material linear-elastic
behaviour: the effect at any point of a linear-elastic mechanical structure subjected
to several loads represents the summation of the effects produced by each of these
loads acting separately.
Using the superposition method, a complicated problem may be solved
through a summation of simple problems. For example, the shear and bending-
moment diagram for the beam shown in Fig. 2.30a, may be sketched as an
algebraical summation of the three diagrams of Fig. 2.30b.

a.
Fig. 2.30

Important remark

The drawing of the internal forces


diagrams (axial forces, shear, bending-
moment or torque diagrams) may be
performed in a unique, simple and logical
manner: the involved member is cut at an
arbitrary point, the internal force of a
certain type representing the summation of
all corresponding external loads (or
b.
moments) acting to the left or to the right
of the cross section considered (and using the adequate sign convention).
41
Strength of Materials

In case of shear and bending moment diagrams, a particular case may also
arise. The presence of one, two or more intermediate pin connections between
different segments of a beam, offers one, two or more additional conditions for the
computation of the external reactions.

Draw the shear and bending-moment diagrams for the beam shown in Fig.
2.31.
Due to the presence of the intermediate pin connection at point 2, the bending
moment (as internal force) at that section is zero.
On the other hand, the
bending moment at point 2,
represents the summation of all
bending moments given by the
loads acting to the left side of
section 2. We obtain therefore:
M i2 0

YA 4a p 2 a 3a 0 YA 1, 5 ap .

From now on the shear and


bending moment diagrams may be
sketched as if the support A and
the intermediate pin connection
would have not existed, the beam
being subjected at A by an upward
Fig. 2.31 vertical concentrated external
force equal to 1,5ap. We finally
obtain the shear and bending moment diagrams shown in Fig. 2.31.

2.5 MOVING LOADS


A truck or other vehicle rolling across a beam or girder constitutes a system of
concentrated loads at fixed distance from each other. For beams carrying only
concentrated loads the maximum bending moment occurs under one of the loads.
Therefore the problem here is to determine the bending moment under each
load when each load is in a position to cause a maximum moment to occur under it.
The largest of these various values is the maximum moment that governs the design
of the beam.
In Fig. 2.32, P1 ,P2, P3 and P4 represent a system of loads at fixed distances a,
b and c from each other; the loads move as an unit across the simply supported beam
with span . Let us locate the position of P2 when the bending moment under this
load is maximum. If we denote the resultant of the loads on the span by R and its
position from P2 by e, the value of the left reaction is:

42
Internal forces in statically determinate members

R
YA ( e x) .

The bending moment under P2
is then:
M i ( M )L

R
Fig. 2.32 M 2 ( e x) x P1 a .

To compute the value of x that will give the maximum M2, we set the
derivative of M2 with respect to x equal to zero:
dM 2 R
( e 2 x) 0 ;
dx
from which:
e
x . (2.6)
2 2

This value of x is independent of the number of loads to the left of P2, since
the derivative of all terms of the form P1a with respect to x will be zero.
Equation (2.6) may be expressed in terms of the following rule: the bending
moment under a particular load is a maximum when the center of the beam is
midway between that load and the resultant of all loads then on the span. With this
rule we locate the position of each load when the moment at that load is a maximum
and compute the value of each such maximum moment.
The maximum shearing force occurs at, and is equal to, the maximum reaction.
The maximum reaction for a group of moving loads on a span occurs either at the left
reaction, when the leftmost load is over that reaction, or at the right reaction when
the rightmost load is over it. In other words, the maximum reaction is the reaction to
which the resultant load is nearest.

2.6 INTERNAL FORCES DIAGRAMS FOR PLANE STRUCTURES


(2D structures - FRAMES) AND SPATIAL (3D) STRUCTURES

The principle presented above for sketching the straight beams internal forces
diagrams may be easily extended to the plane or spacial structures. Let us consider
for example the plane beam shown in Fig. 2.33, for which we have to draw the axial
force, shear and bending-moment diagrams.
An observer "O covering the beam from 1 to A (or from A to 1, as it is easier
from the mathematical point of view) sees each straight portion of the beam as a
beam for which applies the rules presented in the preceding sections.

43
Strength of Materials

Therefore, for portion 1-2, at a


current section at distance x from 1,
we have (Fig. 2.33):

N 0;

T P;
x 0; M i1 0;
M i P x:
x a; M i2 Pa .

Fig. 2.33

Fig.2. 34
The second straight portion of the beam may be covered from 2 to A or A to 2
as well. Let us suppose that the second case is being used.

Fig.2. 35

The effect of the concentrated force P is transmitted up to the current cross


section located at distance x from point 2, where the observer is placed. Therefore, at
that section we have:

44
Internal forces in statically determinate members

N P;

T 0;

M i P a.

We note that all internal forces corresponding to the portion 2 - A are constant.
We are now in the position to draw the internal forces diagrams (N, T, Mi). It
is to be mentioned that, in such cases the diagrams are sketched with respect to a
reference line representing the x axis of the beam (the axis directed along the beam).
Analogous to the straight beams, for N and T diagrams + means above the reference
line. For the bending-moment diagrams, "+" means below and "-" means above the
reference line (from the observer's point of view). With these remarks, the internal
forces diagrams have been represented in Fig. 2.36.

Fig. 2.36
The diagrams represented in Fig 2.36 tell us what does the plane beam feel (as
a global effect) at each particular cross section, when subjected to the external load
P.
We also note that it was not necessary to compute the reactions XA, YA, MA for
sketching the internal forces diagrams.

SAMPLE PROBLEMS
a) Draw the axial force, shear and bending
moment diagrams for the frame and the
loading shown in Fig. 2.37.
Considering the frame as a free body we
first determine the reactions:
Fx 0 xA 0;

Fy 0 YA YB P;

M A 0
P
YB 2 P 0 YB
2

Fig. 2.37
45
Strength of Materials

P
YA YB .
2
Applying the above presented principle and letting an observer to cover the
beam from A to 1 and then from B to 1 we have:
A - 1:
P
N YA ;
2
T X A 0;

M i X A x 0.
B - 2:
N 0;
P
T YB ;
2
x 0; M iB 0;
P
M i YB x x; P P
2 x ; M .
i2
2 2
2 - 1:
N 0;
P P
T YB P P ;
2 2
P
P x 0; M i2 ;
M i Y B ( x) Px ( x) Px ; 2
2
x ; M i1 0.

The axial force, shear and bending moment diagrams have been represented in Fig.
2.38.

Fig. 2.38

b) Draw the axial force, shear and bending-moment diagrams for the frame shown
in Fig.2.39.

46
Internal forces in statically determinate members

In Fig. 2.39b the simplified form of the frame together with the reactions have
been represented. Considering the entire beam (frame) as a free body and using the
external reference coordinate system Oxy, we determine the reactions as follows:

Fig. 2.39
F x 0 X A 5 ap 0 X A 5 ap ;

F y 0 YA YB 2 ap 0 YA YB 2 ap ;

M A 0 2 aY B p 2a a 5 ap 2a 0 YB 6 ap ;

M B 0 YA 2a 5 ap 2a p 2a a 0 YA 4 ap .
We have found therefore that:
X A 5 ap ;

YA 4 ap ;

YB 6 ap .

The second equation of equilibrium (Fy = 0) may be used as a checking equation


when the reactions YA and YB were computed from MA = 0 and MB = 0. The N, T
and Mi diagrams are shown in Fig. 2.40.

Fig. 2.40
47
Strength of Materials

c) Draw the axial force, shear and bending-moment diagrams for the curved beam
of radius R shown in Fig. 2.41.
Although the axis of the beam is not a straight line, the principle presented
above for sketching the N, T, Mi diagrams remains valid.
The problem consists in determining the internal forces corresponding to each
particular cross section of the beam. In order to locate the current cross section an
angular parameter must be used (instead of the linear parameter x which has
been used up the now) for each particular portion of the curved beam.

Fig. 2.42 Fig. 2.43


Let us now consider the first portion 12 of the built-in arch shown in Fig.
2.41. The axial force, shear and bending moment equations for segment 1 2 are
obtained similarly by passing a cross section aa anywhere between 1 and 2. As
discussed above, the cross section is located by the parameter . When varies
between 0 and 90 the whole portion 1 - 2 of the curved beam is covered. The
concentrated load P which acts at section 1, transmits its effect through the segment
BCDE up to the current cross section D'E located by parameter (Fig. 2.42).
This means that a vertical downward load P will act at the centroid O of the
current cross section DE. It is in fact the internal force exerted on the current cross
section and may be resolved into two components: one component perpendicular to
the current cross section DE and the other one contained within the plane of the
cross section. The first component represents the current axial force (N) while the
second component represents the corresponding shearing force (T). An observer O
placed at the current cross section, using the proper positive sign convention will see
that:
0;N P;

0 ;T 0;
N P cos :

T P sin : ;N 0;
2

;T P.
2

48
Internal forces in statically determinate members

The bending moment exerted by the concentrated load P, acting at point 1,


with respect to the centroid O of the current cross section is (Fig. 2.43):

M i P(R R cos ) PR (1 cos ):

0;M i1 0;

;M i2 PR .
2

The same reasoning may be


applied when the second portion of the
curved beam is to be covered. This time
it will be more convenient to us to cover
Fig. 2.43 the beam from A to 2. But in this case
we have to compute the reactions YA and MA at first. This will be done by considering
the entire curved beam as a free body and using the corresponding equations of
equilibrium, Fig.2.42.
Fy 0 YA P 2P 0 YA 3P ;

M A 0 M A P 2R 2P R 0 M A 4 PR .

We write:
N Y A cos 3 P cos ;

0;N A 3P ;

;N 2 0.
2

T Y A sin 3 P sin ;

0 ;T A 0;

;T2 3P .
2
Fig. 2.44
M i M A YA (R R cos ) 4 PR 3 PR (1 cos );

0; M iA M A 4 PR ;

;M i2 PR .
2

The axial force, shear and bending moment diagrams have been represented in Fig.
2.45. It is to be noted that all the established properties of the N, T, Mi diagrams of
straight beams remains valid.
49
Strength of Materials

Fig. 2.45
We shall now consider a simple spatial structure (a 3D structure, i.e. a three
dimensional structure) represented by a beam, fixed at one end, and subjected to two
concentrated loads P and Q at the other end, Fig. 2.46.

Fig. 2.46

d) Draw the axial force, shear, bending moment and torque diagrams for the 3D
beam shown in Fig. 2.46.
Although the problem seems to be a little bit more completed the basic
principle for sketching the internal force diagrams remains unchanged.
Before solving the problem there are still some important remarks to be done:
As it will be discussed later, the sign of the shearing force has no physical
consequences in designing a beam or a certain mechanical structure. This is why,
in case of complicate structures, we shall give the T sign up and we shall
represent the T diagrams as they are convenient to us;
The sign of the bending moment does also depend upon the relative position of
the observer, Fig. 2.47. Although the two beams represented in Fig. 2.47 are
entirely identical from the geometrical and loading point of view, the two
corresponding bending - moment diagrams have different signs. These signs are,
after all, simple conventions. On the other hand, it is to be observed that in both
cases of Fig 2.47, the bending-moment diagrams occupy the same position with
respect to the reference line. In other words, this means that the position of the
bending-moment diagrams do not depend upon the observer's position. The
50
Internal forces in statically determinate members

location of the bending moment diagram with respect to the reference line
corresponds to the position of the beam fibres in tension.

Fig. 2.47
For this reason, in many cases, we shall not note the sign of the bending
moment diagrams and we shall represent these diagrams on the side corresponding to
the beam fibres in tension, Fig. 2.48.
Let us now return to the original
problem regarding the simple 3D
structure shown in Fig. 2.46. Covering
the beam from 1 to A and attaching a
proper coordinate system (whose Ox
axis is usually directed along the beam)
to each main portion of the beam (Fig.
2.46) we obtain the axial force, shearing
force, bending - moment and torque
diagrams shown in Fig. 2.49. It should
Fig. 2.48 be noted that, for each main portion of
the beam, two shearing forces and two bending-moments could exist simultaneously
(about Oy and Oz axes).
We shall now conclude our analysis concerning the main types of internal
forces by observing that these diagrams are in fact a graphical representation of a
global mechanical effect occured at any particular cross section of a given member
subjected to external loads.

51
Strength of Materials

Fig. 2.49
While these diagrams represent a first and necessary step in the analysis of a given
structural member, they do not tell us whether the external loads may be safely
supported. Whether or not a given structural member will break under the external
loading clearly depends upon the ability of the material to withstand the
corresponding elementary forces occurred at the level of each particular point of the
member cross sections. This is why, after a short study of the moments of inertia
within the next chapter, some other chapters of the text will be devoted to the
analysis of the stresses and of the corresponding deformations in various structural
members, considering axial loading, shearing loading, torsion and bending
successively. Each analysis will be based upon a few basic concepts, namely, the
conditions of equilibrium of the forces exerted on the member, the relations existing
between stress and strain in the material and the conditions imposed by the supports
and loading of the member. The study of each type of loading will be complemented
by examples, sample problems and problems to be assigned, all designed to
strengthen the students understanding of the subject.

52
Internal forces in statically determinate members

PROBLEMS TO BE ASSIGNED

P2

P.2.1 Draw the axial force, shear and bending moment diagrams for the members, frames and
loading shown (Fig. P.2.1).

a. b.

c. d.

e. f.

g. h.

i. j.

Fig. P.2.1

53
Strength of Materials

k.
l.

m.
n.

o.
p.

Fig. P.2.1 (continued)

54
Internal forces in statically determinate members

r. s.

t. u.
Fig. P.2.1 (continued)

P.2.2 Draw the torque diagrams for the members and loading shown (Fig. P.2.2).

Fig. P.2.2

P.2.3 Draw the axial force, shear, bending moment and torque diagrams for the 3-D structures
and loading shown (Fig. P.2.3).

55
Strength of Materials

a. b.

c. d.

e. f.

Fig. P.2.3

56
3.FIRST MOMENTS AND SECOND MOMENTS OF AN
AREA

Many engineering formulas and applications such as those relating to strength


of beams, columns, shafts, etc., involve the use of different mathematical expressions
which describe, from the mathematical point of view, the shape and dimensions of
the cross sections. These mathematical expressions are called: geometrical
characteristics.
For members under axial loading (tension or compression) the single involved
geometrical characteristic is represented by the cross-sectional area A of the member.
A higher value of the cross-sectional area A of the member means a higher strength of
the member under axial loading (Fig.3.1).
For structural elements in bending,
torsion etc. there are also other
geometrical characteristics involved
within the strength calculus: the static
moments (first moments of an area) and
the moments of inertia (second moments
of an area). Fig. 3.1

3.1 FIRST MOMENTS OF AN AREA. CENTROID OF AN


AREA

Consider an area A located in the zOy plane (Fig.3.2). Denoting by z and y the
coordinates of an element of area dA, we define the first moment of area A with
respect to z axis as the integral:

S z y dA . (3.1)
A

Similarly, the first moment of area A with respect to the y axis is defined as the
integral:
S y z dA . (3.2)
A

We note that each of these integrals may be positive, negative or zero,


depending upon the position of the coordinate axes. The first moments of area, Sz and
Sy, are expressed in mm3, cm3, m3, etc.
Strength of Materials

Since in almost all cases the area A


of Fig.3.2 is assimilated to the cross-
sectional area of a beam, a shaft etc., it is
presented as being located in the zOy
plane, the Ox axis being directed along
the beam, shaft etc.
The centroid of area A is defined as
the point G of coordinates zG and yG
(Fig.3.2), which satisfy the relations:
Fig. 3.2
z dA
S y
A
zG S y A zG ;
A A (3.3)
y dA
A Sz
yG Sz A yG .
A A

Comparing (3.1) and (3.2) with (3.3), we note that the first moments of area A may be
expressed as the products of the area and the coordinates of its centroid:

Sy A zG ; Sz A yG . (3.4)

When an area possesses an axis of


symmetry, the first moment of the area
with respect to that axis is zero. Indeed,
considering the area A of Fig.3.3, which
is symmetric with respect to the Oy axis,
we observe that to every element of area
dA of abscissa z corresponds an element
of area dA of abscissa -z. It follows that
Fig. 3.3 the integral in (3.2) is zero and, thus,
Sy=0.
It does also follow from the first of the relations (3.3) that zG = 0. Thus, if an area A
possesses an axis of symmetry, its centroid G is located on that axis. If an area
possesses two axes of symmetry (Fig.3.4) the centroid G coincides with its geometric
center.The coordinate axes passing through the centroid of a given area are called
centroidal (or central) axes.
It is to be observed that the integrals involved in relations (3.1) and (3.2) are
actually double integrals, which have to be calculated with the well known
mathematical methods (Fig. 3.5).

S z y dA y dz dy ; S y z dA z dz dy .
A (D ) A (D )

58
First moments and second moments of an area

Fig. 3.4 Fig. 3.5

3.2 SECOND MOMENTS OF AN AREA

Consider again an area A located in the zOy plane (Fig. 3.6) and an element of
area dA of coordinates z and y.
The second moment, or moment of
inertia, of area A with respect to the Oz
axis, and the second moment, or the
moment of inertia, of area A with respect
to the Oy axis are defined, respectively,
as:
2
Iz
2
y dA ; Iy z dA . (3.5)
A A

While each of the above integrals is


actually a double integral, it is possible in
Fig. 3.6
many applications to select elements of
area dA
in the shape of thin horizontal or vertical strips, and thus reduce the computation to
simple integration. This will be illustrated later.
We now define the centrifugal moment of inertia (or the product of inertia) of
area A with respect to Oz and Oy axes (Fig. 3.6) as the integral:

I zy zy d A .
(3.6)
A

Relations (3.5) show that the moments of inertia of an area are positive
quantities and are expressed in mm4, cm4, m4 etc. On the other hand, relation (3.6)
shows that the centrifugal moment of inertia may be positive, negative or zero,
depending upon the locations of the area relative to the involved axes. It is positive if
the area lies principally in the first or third quadrants and negative if the area lies
principally in the second or fourth quadrants.
We define the polar moment of inertia of area A with respect to point O (Fig.
3.6) as the integral:
59
Strength of Materials

2
I r dA,
p (3.7)
A

where r is the distance from O to the element dA. While this integral is again a double
integral, it is possible in the case of a circular area to select elements of area dA in the
shape of thin circular rings, and thus reduce the computation of Ip to a simple
integration. This will be illustrated later. It is to be noted that the polar moment of
inertia is also a positive quantity, being expressed in mm4, cm4, m4 etc.
An important relation may be established between the polar moment of inertia
Ip of a given area and the moments of inertia Iz and Iy of the same area. Noting that
r2 = z2 + y2 (Fig. 3.6) we write:

2 2 2 2 2
I p r dA z y dA z dA y dA
A A A A

or
I p Iz Iy . (3.8)

If an area has an axis of symmetry,


this axis together with any axis
perpendicular to it will form a set of axes
for which the centrifugal moment of
inertia is zero. Consideration of the
symmetrical section shown in Fig. 3.7
will disclose that, for any differential area
dA, there is a symmetrically placed equal
differential area dA. With respect to the
Oy axis of symmetry, the z coordinates of
dA and dA are equal but of opposite sign,
whereas their y coordinates are equal and
Fig. 3.7 of the same sign regardless of the
position of the Oz axis.
Hence the sum of the products zy dA for each such pair of symmetrically
placed elements dA and dA will be zero. It follows, therefore, that the value of
zy d A for the entire area will be zero if either or both reference axes are axes of
A

symmetry.

3.3 PARALLEL - AXIS THEOREM (STEINERS RELATIONS)

Consider the moments of inertia Iz and Iy and the centrifugal moment of inertia
Izy of an area A with respect to two arbitrary perpendicular axes Oz and Oy (Fig. 3.8).
We assume to know the quantities Iz, Iy and Izy, where
60
First moments and second moments of an area

2 2
Iz y dA ; I y z dA ;
A A

I zy zy d A .
A

Let us now consider another


coordinate system z1O1y1, translated with
quantities a and b with respect to the axes
Oy and Oz of the first coordinate system.
The problem which arises consists in
determining the quantities I z , I y and
1 1 Fig. 3.8
Iz y of the same area A but with respect to the axes of the new coordinate system.
1 1

We write:
2 2 2 2
Iz y1 d A y b dA y 2 by b dA
1
A A A

2 2 2
y dA 2b y dA b dA Iz 2 bS z b A;
A A A

2 2 2 2
I y z1 d A z a dA z 2 az a dA
1
A A A

2 2 2
z dA 2a z dA a dA I y 2 aS y a A;
A A A

Iz y z1 y1d A y b z a dA zy zb ay ab dA
1 1
A A A

zy d A b z dA a y dA ab dA I zy bS y aS z abA .
A A A A

Thus, the mathematical connection between the moments of inertia Iz , I y and


I zy of an area A and the same quantities Iz , I y and Iz y calculated with respect to
1 1 1 1

the translated coordinate system z1O1y1, is described by the following relations:


2
Iz Iz 2 bS z b A
1
2
I y1 I y 2 aS y a A (3.9)
Iz y1 I zy bS y aS z abA ,
1

where
a is the distance between axes Oy and O1y1,
b is the distance between axes Oz and O1z1,
61
Strength of Materials

Sz, Sy are the static moments (first moments) of area A with respect to axes Oz
and Oy.
If the point O is the centroid of area A, it follows from relations (3.4) that Sz =
Sy = 0 and we may write:

2
Iz Iz b A;
1
2
I y
1
I y a A; (3.10)
Iz y I zy abA .
1 1

Relations (3.10) are known as Steiners formulas.


For example, the first relation of (3.10) expresses that the moment of inertia
Iz of an area with respect to an arbitrary Oz1 axis is equal to the moment of inertia Iz
1

of the same area with respect to the centroidal Oz axis parallel to the Oz1, plus the
product b2A of area A and the square of the distance b between the two axes. This
result is also known as the parallel-axis theorem. It makes it possible to determine the
moment of inertia of an area with respect to a given axis, when its moment of inertia
with respect to a centroidal axis of the same direction is known. Conversely, it makes
it possible to determine the moment of inertia Iz of an area A with respect to a
centroidal axis Oz, when the moment of inertia I z of A with respect to a parallel axis
1
2
is known, by subtracting from I z the product b A. We should note that the parallel-
1

axis theorem may be used only if one of the two axes involved is a centroidal axis.

3.4 MOMENTS OF INERTIA OF SIMPLE SURFACES

a) Rectangular area
For the rectangular area A shown in
Fig. 3.9, determine the moments of
inertia Iz, Iy and Izy with respect to the
centroidal Oz and Oy axes.
As mentioned before:
2
Iz y dA .
A

We select as an element of area (dA)


a horizontal strip of length b and
thickness dy (Fig.3.9). We write:

dA = b dy.
Fig. 3.9

It follows that:

62
First moments and second moments of an area

h h
2 3 2 3 3 3 3
2 2 by 1 h h b 2h bh
Iz y dA y b dy b .
h
3 h 3 8 8 24 12
A
2 2

Thus, the moment of inertia Iz of a rectangular area with respect to the centroidal Oz
axis is:
3
bh
Iz . (3.11)
12

In the same manner, it follows that :


3
hb
Iy . (3.12)
12

Since Oz and Oy are axes of symmetry, we have:


I zy 0.

b) Circular area
For the circular area shown in Fig. 3.10 determine the polar moment of inertia I p
and the moments of inertia Iz, Iy and Izy.

Fig. 3.10 Fig. 3.11

We select as an element of area (dA) a ring of radius r and thickness dr,


(Fig.3.11). The polar moment of inertia of area A is:
d d
d
2 2 4 4 4
2 2 3 r 2 d d
I p r dA r 2 r dr 2 r dr 2 0 .
4 2 16 32
A 0 0
0

63
Strength of Materials

Thus
4
d
I p . (3.13)
32

Due to the symmetry of the circular area, we have Iz = Iy.


Recalling (3.8), we write:
4
d
I p Iz I y 2I z 2I y
32

and, thus

4
d
Iz I y . (3.14)
64

c) Triangular area

Determine the moments of inertia for a triangle of base b and altitude h with
respect to an axis coinciding with its base and a centroidal axis parallel to its base.
Select the differential element as shown in Fig. 3.12.

From similar triangles, we have


b h y
m . The moment of inertia with respect
h
to z1 axis is obtained from:
h h
2 2 2 b h y
Iz y dA y m dy y dy
1
h
A 0 0
3 4 3 3 3
bh b h bh bh bh
.
3 h 4 3 4 12

We have thus obtained:


Fig. 3.12

3
bh
Iz
1
. (3.15)
12

To determine the centroidal moment of inertia Iz , we transfer the known value of


h
Iz , from the base axis z1 to the parallel axis z. Since the transfer distance is as
1
3
shown in Fig. 3.12, we write:

64
First moments and second moments of an area

2
h bh
Iz Iz .
1 3 2

It follows that
2 3 3 3
h bh bh bh bh
Iz Iz . (3.16)
1 3 2 12 18 36

3.5 MOMENTS OF INERTIA OF COMPLEX SURFACES


(COMPOSITE AREAS)

To determine of the moments for inertia of a complex surface the following steps
have to be covered:
- the complex surface (area) A has to be divided into several component parts
of areas A1, A2 ...;
- determination of the centroidal point G of the complex area;
- since the integral representing the moment of inertia of area A may be
subdivided into integrals extending over A1, A2 ..., the moment of inertia of
A with respect to a given axis will be obtained by adding the moments of
inertia of areas A1, A2 ... , with respect to the same axis. Before adding the
moments of inertia of the component areas, however, the parallel-axis
theorem should be used to transfer each moment of inertia to the desired
axis. This is shown in the following example.

Determine the moments of inertia Iz, Iy and Izy of area A shown in Fig. 3.13, with respect to
the centroidal axes.

Fig. 3.13 Fig. 3.14

We first divide the complex area A into the two rectangular areas A1 and A2 (Fig 3.14) and
denote their centroids and their own centroidal axes by G1 ,G2 , z1 , y1 , z2, y2 respectively.
65
Strength of Materials

We may now determine the coordinates zG and yG of the centroid G of the composite area A,
using, for example, the coordinate system z1 G1 y1 as follows:

y i Ai
i 3 a 1, 5 a 4a
yG
2
1, 5 a ; (i = 1,2)
Ai 4a 4 a 1, 5 a
i

1, 5 a
z i Ai 2a 4 a 1, 5 a
i 2
zG 0 , 75 a .
2
Ai 4a 4 a 1, 5 a
i

Recalling the formulas (3.11) and (3.12) and using the parallel-axis theorem we may write
the moments of inertia of the composite area A as follows:
3 3
4a a 2 2 1, 5 a 4a 2 4
Iz 1, 5 a 4a 2 ,5 a 1, 5 a 4 a 1, 5 a 23 , 33 a ;
12 12

3 3 2
a 4a 2 4a 1, 5 a 1, 5 a
Iy 0 , 75 a 4a a 2a 0 , 75 a 4 a 1, 5 a
12 12 2
4
10 , 2 a ;

2 1, 5 a
I zy 0 1 , 565 a 0 , 652 a 4a 0 2 ,5 a 1 , 565 a 2a 0 , 652 a 4 a 1, 5 a
2
4
7 , 43 a .

3.6 MOMENTS OF INERTIA WITH RESPECT TO INCLINED


AXES

In some cases, it is necessary to


determine the moments of inertia with
respect to axes that are inclined to the
usual axes. The moments of inertia in
such cases can be obtained by formal
integration, but a general formula is
usually easier to use.
The problem may be stated as
follows: assuming the values Iz, Iy and
Izy with respect to the Oz and Oy axes
Fig. 3.15 to be known, determine the values of

66
First moments and second moments of an area

I z1 , I y1 and Iz y with respect to the Oz1 and Oy1 axes inclined at an angle with
1 1

Oz and Oy axes, as shown in Fig. 3.15.


The coordinates for a typical differential area dA are given by z and y with
respect to the y and z axes, and by y1 and z1 relative to the z1 and y1 axes. The relation
between these coordinates can be obtained by projecting the coordinates z and y on
the z1 and y1 axes. This gives (Fig. 3.15):

z1 z cos y sin ; (3.17)


y1 y cos z sin .

By definition, the values of Iz and I y are:


1 1

2 2
Iz y1 d A ; I y z1 d A ; Iz y z1 y1 d A .
1 1 1 1
A A A

Replacing the values of z1 and y1 from (3.17) we have:


2 2 2 2 2 2
Iz y1 d A y cos z sin dA y cos 2 yz cos sin z sin dA
1
A A A
2 2
I z cos I zy sin 2 I y sin ;

2 2 2 2 2 2
I y z1 d A z cos y sin dA z cos 2 zy cos sin y sin dA
1
A A A

2 2
I y cos I z sin I zy sin 2 ;

Iz y z1 y1 d A z cos y sin y cos z sin dA


1 1
A A

2 2 2 2
zy cos z sin cos y sin cos yz sin dA
A

2 2
Iz I y sin cos I zy cos sin .

We, thus, obtain:


2 2
Iz I z cos I y sin 2 I zy sin cos ;
1
2 2
I y
1
I z sin I y cos 2 I zy sin cos ; (3.18)
2 2
Iz y Iz I y sin cos I zy cos sin .
1 1

If the relations
67
Strength of Materials

2 1 cos 2 2 1 cos 2
sin , cos
2 2

are substituted in (3.18), we may write:

1 cos 2 1 cos 2
Iz Iz I y I zy sin 2 ;
1 2 2
1 cos 2 1 cos 2
I y Iz I y I zy sin 2 ;
1 2 2
Iz I y
Iz y sin 2 I zy cos 2
1 1 2

or

Iz I y Iz I y
Iz cos 2 I zy sin 2 ;
1 2 2
Iz I y Iz I y
I y cos 2 I zy sin 2 ; (3.19)
1 2 2
Iz I y
Iz y sin 2 I zy cos 2 .
1 1 2

When the values of Iz, Iy and Izy are known, relations (3.19) permit the values of
I z , I y and I z y with respect to the Oz1 and Oy1 axes, inclined at an angle to the
1 1 1 1

Oz and Oy axes, to be determined without further integration. In a sense, these


relations do for inclined axes what the Steiners formula does for parallel axes.
A simple analysis of relations (3.19) tells us that I z , I y and I z y are
1 1 1 1

functions of angle . One could ask: which are the values of angle that make these
quantities ( I z , I y and I z y ) maximum or minimum? The angles defining the
1 1 1 1

maximum and the minimum moments of inertia, also called the principal moments of
inertia, may be found by differentiating (3.19) with respect to and setting the
derivative equal to zero:

d Iz Iz I
1 y
2 sin 2 2 cos 2 I zy 2Iz y 0;
d 2 1 1
(3.20)
dI y Iz I
1 y
2 sin 2 2 cos 2 I zy 2Iz y 0.
d 2 1 1

We find that:

68
First moments and second moments of an area

2 I zy
tg 2 . (3.21)
Iz I y

Equation (3.21) gives us two values of ( 1 and 2 1 ) for which Iz and


1
2
I y have extreme values. This is why the equation (3.21) is always written as:
1

2 I zy
tg 2 1, 2 . (3.22)
Iz I y

The extreme conditions for Iz , I y (3.20 ) - mean in fact that the product of
1 1

inertia Iz y equals zero. In the same time, a second differentiation of (3.20) shows
1 1

that:
2 2
d Iz d I y
1 1
, (3.23)
2 2
d d

which means that a maximum value of Iz implies a minimum value of I y and


1 1

vice versa.
Substituting for from equation (3.21) into (3.19) we obtain the extreme
values of quantities I z and I y , called the principal moments of inertia:
1 1

Iz I y 1 2 2
I 1, 2 Iz I y 4 I zy (3.24)
2 2

with respect to the axes Oz1 and Oy1, rotated with angle 1.

In this way we found two


perpendicular directions given by 1
and 2 1 for which the moments
2
of inertia of the original area have
extreme values (a maximum value I1 with
respect to one of these directions and a
minimum value I2 with respect to the
other direction). Usually, the axis of
maximum is denoted by 1 while the axis
Fig. 3.16
of minimum by 2.
It is important to be mentioned again that the product of inertia of the original
area with respect to the coordinate system 1O2 (Fig. 3.16) is zero.

69
Strength of Materials

Axes 1 and 2 are called principal axes.


One could demonstrate that if Izy < 0 the axis of maximum is placed in the first
quadrant while, if Izy > 0, the axis of maximum is placed in the second quadrant.

3.7 RADIUS OF GYRATION. ELLIPSE OF INERTIA

The term radius of gyration is used to describe another mathematical


expression and occurs most frequently in column formulas. Radius of gyration is
usually denoted by the symbol i and is defined as:

I
i , (3.24)
A

where I is the moment of inertia and A the area.


Thus, we have:

Iz I y
iz ; iy ;
A A

(3.25)
I1 I2
i1 ; i2 .
A A

The following is a geometric interpretation of this relation. Assume the area of Fig.
3.2 to be squeezed into a long narrow strip as shown in Fig. 3.17.
Each differential element of area
dA will then have the same distance iz
from the Oz axis. The moment of inertia
is given by:
2 2
Iz y dA iz dA
A A

2 2
iz dA iz A.
Fig. 3.17 A

The strip may be placed on either side of the reference axis, since if iz is negative,
squaring it will automatically make it plus. Also, part of the strip may be at a distance
iz from one side of the reference axis and the remainder of the strip at equal distance iz
from the other side of the axis.
In view of this discussion, the radius of gyration is frequently considered to be
the uniform distance from the reference axis at which the entire area may be assumed
to be distributed. For an area whose dimensions perpendicular to a reference axis are
negligibly small compared with its distance from that axis, the radius of gyration is
practically equivalent to the centroidal location of the area.
70
First moments and second moments of an area

The ellipse of equation


2 2
z y
2 2
1 0 (3.26)
i2 i1

represents the centroidal principal ellipse of inertia with respect to a certain area A.

Sample problems
1. For the area shown in Fig. 3.18 determine: (a) the centroidal point G of the area A; (b) the
moments of inertia Iz, Iy and Izy with respect to the centroidal reference system zGy; (c) the
principal axes of inertia 1 and 2; (d) the principal moments of inertia I1 and I2; (e) the
principal radii of inertia i1 and i2 and (f) draw the ellipse of inertia.

Solution

We first divide the area A of the whole


surface into three rectangular areas with centroidal
points G1, G2 and G3 (Fig. 3.18). We observe that
the centroid of the second rectangular area G2
coincides with the centroid G of the whole area A.
Thus, axes z2 and y2 coincide with the centroidal
axes Gz and Gy of the whole area A.
Recalling (3.11) and (3.12) and using the
parallel-axis theorem we may write the moments of
inertia of the composite area A as follows:

3 3
165 30 2 30 240
Iz 135 30 165 2
12 12

8 4
2,1573 10 mm ; Fig. 3.18

3 3
30 165 2 240 30 8 4
Iy 82 , 5 165 30 2 0 , 9038 10 mm ;
12 12

8 4
I zy 135 82 , 5 30 165 2 1 ,1026 10 mm .

The principal directions of inertia are given by

2 I zy 2 1 ,1026
tg 2 1 , 2 1 30 ,19 ; 2 1 59 , 81 .
Iz Iy 2 ,1573 0 , 9038 2

Since Izy > 0, the principal axis of maximum is placed in the second quadrant (Fig. 3.18).

71
Strength of Materials

The principal moments of inertia are:

Iz Iy 1 2 2
I 1, 2 Iz Iy 4 I zy
2 2
8 8
2 ,1573 10 0 , 9038 10 1 8 8 2 8 2
2 ,1573 10 0 , 9038 10 4 1 ,1026 10 .
2 2

We finally have:
8 4
I1 2 , 799 10 mm ;
8 4
I2 0 , 262 10 mm .

The principal centroidal radii of inertia are:


8
I1 2 , 799 10
i1 127 , 94 mm ;
A 30 165 2 30 240

8
I2 0 , 262 10
i2 39 ,14 mm .
A 30 165 2 30 240

The principal centroidal ellipse of inertia has been represented in Fig. 3.18.

2. For the composite area of Fig. 3.19, composed of two U-shaped profiles, determine the same
quantities like in the previous example.

Solution
We first divide the area A of the whole surface
into two areas 1 and 2 (the two U shapes) having
the centroidal points at G1 and G2 (Fig. 3.19).
From the appropriate tables containing the
geometrical characteristics of rolled-steel shapes
(APPENDIX III) we may get all the necessary
data:

4 4
I 248 10 mm ;
z
1
4 4
1: I 3600 10 mm ;
y
1
2 2
A 42 , 3 10 mm ;
1
Fig 3.19
4 4
Iz 3600 10 mm ;
In these tables we may also find the location 2
of the centroidal point with respect to the U- 2: Iy 248 10
4
mm
4
;
shaped section (22,3 mm - Fig. 3.19). 2
2 2
Using the coordinate system z1 G1 y1 , we A2 42 , 3 10 mm .
can compute now the position of the centroidal
point G of the entire area as follows:

72
First moments and second moments of an area

2
y i Ai 120 22 , 3 42 , 3 10
yG 48 , 85 mm ;
2
Ai 2 42 , 3 10

2
z i Ai 120 22 , 3 42 , 3 10
zG 71 ,15 mm .
2
Ai 2 42 , 3 10

We have thus located the centroidal reference coordinate system of the composite area: zGy
(Fig.3.19).
Recalling again the formulas (3.11) and (3.12) and using the parallel axis theorem we may
write the moments of inertia of the composite area as follows:

4 2 2 4 2 2
Iz 248 10 48 , 85 42 , 3 10 3600 10 120 22 , 3 48 , 85 42 , 3 10

6 4
58 , 668 10 mm ;

4 2 2 4 2 2
I 3600 10 71 ,15 42 , 3 10 248 10 120 71 ,15 22 , 3 42 , 3 10
y

6 4
81 , 30 10 mm ;

2
I zy 0 48 , 85 71 ,15 42 , 3 10 0 120 22 , 3 48 , 85 120 71 ,15 22 , 3

2 6 4
42 , 3 10 29 , 4 10 mm .

The principal directions of inertia are given by

2 I zy 6
2 29 , 4 10 1 34 , 47 ;
tg 2 1 , 2
6
Iz Iy 58 , 668 81 , 30 10 2 124 , 47 .

Since Izy > 0, the principal axis of maximum is placed in the second quadrant (Fig 3.19). The
principal moments of inertia are

Iz Iy 1 2 2
I 1, 2 Iz Iy 4 I zy .
2 2

We finally have

6 4
I1 101 , 48 10 mm ;
6 4
I2 38 , 48 10 mm .

The principal centroidal radii of inertia are:

73
Strength of Materials

6 6
I1 101 , 48 10 I2 38 , 48 10
i1 109 , 52 mm ; i2 67 , 44 mm .
2 2
A 2 42 , 3 10 A 2 42 , 3 10

The principal centroidal ellipse of inertia is shown in Fig. 3.19.

PROBLEMS TO BE ASSIGNED
P.3
For the areas shown in the figures below, locate the centroids of the areas and then determine:
the second moments of the involved areas ( Iz, Iy and Izy ) with respect to the centroidal axes;
the principal axes of inertia 1 and 2 ;
the principal moments of inertia I1 and I2;
the principal radii of inertia i1 and i2, and, finally, draw the ellipse of inertia.

Fig. P.3.1 Fig. P.3.2 Fig. P.3.3

Fig. P.3.4 Fig. P.3.5 Fig. P.3.6

Fig. P.3.7 Fig. P.3.8 Fig. P.3.9 Fig. P3.10

74
4. DISPLACEMENTS, STRESSES AND STRAINS
4.1 DISPLACEMENTS

Consider a body subjected to several external loads P1, P2,, Pk, Pn in


mechanical equilibrium, (Fig. 4.1). The body has been represented in a rectangular

coordinate system Oxyz, with i , j and k the versors of the reference axes.
Let us assume that, before
the action of the external loads, a
certain point of the body occupies
the position A (Fig. 4.1). Due to the
action of the external loads the
body deforms and, after
deformation of the body, the point
A is displaced to A . The length
A A is called the total displacement
of point A.
Denoting the displacement
Fig. 4.1
components of point A by u, v and
w, we may write

AA u i v j w k (4.1)
The components u, v, w are called linear displacements.

4.2 THE CONCEPT OF STRAIN

Consider again a body subjected to several external loads in mechanical


equilibrium, P1,,Pn. Before the action of the loads the involved body occupies a
certain position in space. Due to the action of the external loads the body deforms
(Fig. 4.2) and all its points
are displaced to the other
positions.
Two points occupying
the positions A and B, within
the initial undeformed state
of the body, are displaced to
A and B respectively. In
this way the distance
between points A and B
changes due to the action of
the external loads. Denoting
by the original length of
Fig. 4.2 the segment AB and by
Strength of Materials

the length of AB we define the linear absolute deformation as


- . (4.2)
Letting approach zero, we define the deformation per unit length at point A (or the
normal strain at point A) along direction AB as

lim . (4.3)
0

Since and are expressed in the same units, the normal strain obtained by
dividing by is a dimensionless quantity.
Returning now to Fig. 4.2, let us consider three points (O, C, D) of the body in
its original, undeformed state, where segments OC and OD are perpendicular. After
deformation of the body, the points O, C, D are displaced to O , C , D the segments
O C and O D losing their perpendicularity. In the same time a change of the right
angle between OC and OD occurs ( COD C O D ). Letting the segments OC and
OD approach zero, we define the shearing strain as
lim COD C 'O ' D ' .
OC 0 (4.4)
OD 0

In fact represents the change of the original right angle COD. In case of a
rectangular element of infinitely small sides a, b and c the shearing strain
corresponding to the x and y directions is, (Fig. 4.3):

Fig. 4.3
If we attach a rectangular coordinate reference system Oxyz to the body represented
in Fig. 4.2, the knowledge of the body deformation state implies the knowledge, for
each particular point of the body, of all components contained in the table bellow.

x xy xz

T yx y yz
. (4.5)
zx zy z

T is called the deformations tensor. In other words, the deformation state of a body
subjected to certain external loads is defined by the nine components of table (4.5).
Since xy yx
, xz zx
, yz zy
, only six components of table (4.5) are distinct.
This is why the deformation state is a tensorial quantity.
76
Displacements, stresses and strains

4.3 THE CONCEPT OF STRESS

While the method of sections (i.e. internal forces diagrams) presented in


chapter II represents a first and necessary step in the analysis of a certain mechanical
structure (member, beam, etc) it does not tell us whether the loads may be safely
supported. This is why we have to introduce the concept of stress.
Let us now return for a while to the method of sections, considering a body
subjected to several external loads in mechanical equilibrium (Fig. 4.4a).

a. b.
Fig. 4.4

Passing an exploratory plane through the body at some arbitrary point we


obtain the two segments shown in Fig. 4.4b. (I and II). In this way we expose the
internal forces acting on the exploratory section that are necessary to maintain the
equilibrium of either segment. As presented in chapter II, in general, the internal
forces reduce to a force and a couple that,
for convenience are resolved into their
components that are normal and tangent to
the section, (Fig. 4.5).
The origin of the reference axes is
always taken at the centroid which is the key
reference point of the section. Although we
are not get ready to show why this is so, we
shall prove it as we progress.
Anyhow, the internal forces defined
above (R, M, R, M or their components)
represent a global effect occurred within the
Fig. 4.5 exploratory section considered. But, if the
considered body is to fail under the action of the external loads, this failure begins at
the level of a certain point of the exploratory section. This is why we have to analyze
the local effect occurred at the level of the exploratory sectional points, when the
involved body is subjected to several external loads in mechanical equilibrium. Such
an analysis requires the concept of stress.

77
Strength of Materials

Let us now consider the Ist segment of the


body shown in Fig.4.4, for which we take an
elementary area A of the section surface
considered, around an arbitrary point C,
(Fig.4.6).
We denote by P the surface force
acting on A. This force models the
mechanical interaction between the points of
A and the points of the area A of the
second segment of the body with which A
comes in contact. The vector quantity
Fig. 4.6

P
pm (4.6)
A

is called the total average stress developed on A.


Now letting A approach zero (which means that A reduces to the point C), the
vector quantity
P
p lim (4.7)
A 0 A

is called the total stress at point C or simply the stress at point C, (Fig.4.6).
Generally speaking, the stresses on different sections passing though the same
point in a body are different. Usually, the stress is resolved into a component
(called the normal stress) acting along the normal direction n to the A surface and a
tangential component (called the shearing stress) acting along a tangential
direction t , contained within the plane of surface A (Fig.4.6).
Since the force P is expressed in newtons (N) and A in square millimeters,
the stress p and the stress components and will be expressed in N/mm2. This
unit is called a megapascal (MPa). We have

N
1 MPa 1
2
.
mm

If we take a rectangular coordinate reference system xyz with its origin at point
C (with Cx axis perpendicular to A and Cz and Cy axes contained within the A
plane), the shearing stress may be resolved into other two components acting along
Cy and Cz respectively, Fig. 4.7.To indicate the acting plane and the direction of a
normal stress, we associate the stress with a coordinate subscript.

78
Displacements, stresses and strains

For instance, indicates the normal


x

stress acting on a plane perpendicular to the


x axis and also in the x direction. As to a
shearing stress, we should associate it with
two coordinate subscripts. For instance, xy

indicates the shearing stress acting on a


plane perpendicular to the x axis, but in the y
direction.
The stress p shown in Fig. 4.7 may
be thus expressed as
Fig.4.7

p x
i xy
j xz
k , (4.7)


where i ,jand k are the versors of the reference axes.
The stress components positive sign convention has been shown in Fig. 4.8.

Fig. 4.8 Fig. 4.9


Returning to the exploratory plane considered, it is to be noted that there is
an infinit number of planes (sections), which may be passed through the point C of
the body. For each plane passed through point C, we find a certain stress p .
Although there is an infinit number of stresses which may be associated to the point
C, one could demonstrate (and this will be done later) that the stress state at point C is
defined by the stresses corresponding to the three perpendicular planes at C (Fig.4.9),
where

px x
i xy
j xz
k

py yx
i y
j yz
k (4.8)

pz zx
i zy
j z
k

The quantity

79
Strength of Materials

x xy xz

T yz y yz (4.9)
zx zy z

is called the stress tensor at point C.


To facilitate the visualization of the stress condition at point C, we shall
consider a small cube of side a centered of C and the stresses exerted an each of the
six faces of the cube (Fig.4.10).

Fig. 4.10 Fig. 4.11

The stress components shown in the figure are , and , which represent x y z

the normal stresses of faces respectively perpendicular to the x, y and z axes, and the
six shearing stress components , , etc. We recall that, according to the definition
xy xz

of the shearing stress components, represents the y component of the shearing


xy

stress exerted on the face perpendicular to the x axis, while represents the x yx

component of the shearing stress exerted on the face perpendicular to the y axis. Note
that only three faces of the cube are actually visible in Fig.4.10 and that equal and
opposite stress components act on the hidden faces. While the stresses acting on the
faces of the cube differ slightly from the stresses at C, the error involved is small and
vanishes as side a of the cube approaches zero.
Important relations between the shearing stress components will now be
derived. Let us consider the free-body diagram of the small cube centered at point C
(Fig.4.11).
The normal and shearing forces acting on the various faces of the cube are
obtained by multiplying the corresponding stress components by the area A of each
face. Selecting coordinate axes centered at C, we may write the six equilibrium
equations:
Fx 0 ; Fy 0 ; Fz 0;
(4.10)
M x
0; M y
0 ; M z
0 . (4.11)
80
Displacements, stresses and strains

Since forces equal and opposite to the forces actually shown in Fig. 4.11 are
acting on the hidden faces of the cube, it is clear that equations 4.10 are satisfied.
Turning now to the equations (4.11), we first consider the last of these equations,
M z
0 . Using a projection on the xy plane (Fig.4.12), we note that the only forces
with moments about the z axis different from zero are the shearing forces.
These forces form two couples one of
counterclockwise moment A a , yx

the other of clockwise moment


A a . We write therefore
xy

M z
0 ( xy
A) a ( yx
A) a 0

from which we conclude that


xy yx
. (4.12)
The relation obtained shows that
the y component of the shearing stress
exerted on a face perpendicular to the x
axis is equal to the x component of the
Fig. 4.12 shearing stress exerted on a face
perpendicular to the y axis. From the remaining two equations (4.11) , we derive in a
similar manner the relations
yz zy
; zx xz
. (4.13)

Fig. 4.13

We conclude from equations (4.12) and (4.13) that only six stress components
are required to define the condition of stress at a given point C, instead of nine as
originally assumed in (4.9). These stress components are , , , , , .x y z xy xz yz

We also note that, at a given point, shear cannot take place in one plane only; an
equal shearing stress must be exerted on another plane perpendicular to the first one
(Fig.4.13). The two shearing stresses have the same orientation with respect to the
common edge. This property is called the shearing stresses duality law.

81
Strength of Materials

4.4 RELATIONSHIPS AMONG INTERNAL FORCES AND STRESSES


WITHIN A BEAM (MEMBER) CROSS SECTION

The method of sections tells us that, in the most general state, the global
internal force at the level of a given cross section of a beam or member, may be
expressed through six components (Fig.4.14): N , T , T , M M , M and M . y z x t iy iz

Fig. 4.14 Fig. 4.15


These components do also represent a global effect corresponding to the entire
cross section. In the same time these global internal forces determine the
development of normal and shearing stresses at the level of the cross - sectional
points, (Fig. 4.15). For example, at the level of a certain element of area dA, of
coordinates z and y, the normal stress and the shearing stresses
x
and may xy xz

occur. The internal forces N , T , T , M


y z
M , M and M represent in fact the
x t iy iz

summation of the elementary effects given by normal and shearing stresses at the
level of the cross - sectional elements of area dA. We may write therefore the
following relationships:

N x dA;
A

Ty xy dA;
A

Tz xz dA;
A
(4.14)
M x M t xy z xz y dA;
A

M iy x z dA;
A

M iz x y d A.
A

82
5. STRENGTH OF MATERIALS BASIC ASSUMPTIONS

To evaluate the stress, strains and displacements in a strength of materials


problem, we must derive a series of basic equations. During the process of derivation,
however, if we consider all the influential factors in an all-round way, the results
obtained will be so complicated that practically no solutions can be found. Therefore,
we have to make some basic assumptions about the properties of the body
considered. Under such assumptions, we can neglect some of the influential factors of
minor importance temporarily, thus simplifying the strength of materials calculus. In
this text we shall comply with the following assumptions in classical strength of
materials:

1. The body is continuous, i.e., the whole volume of the body is filled with
continuous matter, without any void. Only under this assumption, can the physical
quantities in the body, such as stresses, strains and displacements, be continuously
distributed and thereby expressed by continuous functions of coordinates in space. In
reality, all engineering materials are composed of elementary particles and do not
accord with the assumption of continuity. However, it may be conceived that this
assumption will lead to no significant errors so long as the dimensions of the body are
very large in comparison with those of the particles and with the distances between
neighboring particles.

2. The body is homogeneous: the properties are the same throughout the body.
Under this assumption we may analyse an elementary volume isolated from the body
and then apply the results of analysis to the entire body.

3. The body is isotropic so that the material properties are the same in all
directions. Thus, the strength study will be independent of the orientation of
coordinate axes.

Most engineering materials do not satisfy the above last two assumptions
completely. Structural steel, for instance, when studied with a microscope, is seen to
consist of crystals of various kinds and various orientations. It seems that the material
is far from being homogeneous and isotropic. However, since the dimensions of any
single crystal are very small in comparison with those of the entire body, and since
the crystals are orientated at random, the behavior of a piece of steel, on average,
appears to justify the assumptions of homogeneity and isotropy. This is the reason
why the strength of materials calculus based on these assumptions can be applied to
steel structures with very great accuracy so long as none of the members has been
Strength of Materials

subjected to the process of rolling which may produce a definite orientation of the
crystals. In contrast with steel, wood is definitely not isotropic, since the properties of
wood in the direction of the grain differ greatly from those in the perpendicular
directions. In assuming isotropic material, we shall of course exclude the treatment of
wooden structures.

4. The displacements and strains are small, i.e., the displacement components of all
points of the body during deformation are very small in comparison with its original
dimensions and the strain components and the rotations of all line elements are much
smaller than unity.
Thus when we formulate the
equilibrium equations relevant to the
deformed state, we may use the lengths and
angles of the body before deformation. In
addition, when we formulate the
geometrical equations involving strains and
displacements, we may neglect the squares
and products of small quantities.
For example, when writing the
Fig. 5.1 moment given by the force P at point A
(Fig. 5.1) we choose the arm instead of , the difference between the two lengths
being a very small quantity. This is why we may write all the equilibrium equations
on the undeformed state of the body.

a. b.
Fig. 5.2
Following the same idea, if the two members of Fig. 5.2a are subjected to a
force P applied at B, within the context of the above mentioned assumption the angle
made by each member with the vertical direction remains approximately equal to
(Fig. 5.2b).

84
Strength of Materials Basic Assumptions

A direct consequence of the small displacements and strains assumption is the


method (principle) of superposition.

5. The body is perfectly elastic, i.e., it wholly obeys Hookes law of elasticity, which
shows the linear relations between the stress components and the strain components
(Fig. 5.3). Under this assumption, the elastic constants will be independent of the
magnitudes of these components. The justifications for this assumption lies in the
physical behavior of nearly all materials in engineering construction.
In the other words, if the strains caused in a certain body by the application of a
given load disappear when the load is removed, the material is said to behave
elastically.

Fig. 5.3

The linear relations between stress and strain in their simplest way, may be
expressed by Hookes law as follows :

=E and = G , (5.1)
where E is called the modulus of elasticity of the material involved (or Youngs
modulus, after the English scientist Thomas Young (1773-1829) and G is called the
modulus of rigidity or shear modulus of the material. For structural steel we have:

E = 2,1105 MPa ; G = 8 104 MPa.


Fig. 5.3 does also tell us that the involved material behaves linearly.

6. Saint-Venants principle: except in the immediate vicinity of the points of


application of the loads, the stress distribution in a body may be assumed independent
of the actual mode of application of the loads. This statement practically applies to
any type of load.
While Saint-Venants principle makes it possible to replace a given loading by
a simpler one for the purpose of computing the stresses in a structural member, we
should keep in mind two important points when applying this principle:

a. The actual loading and the loading used to compute the stresses must be
statically equivalent;

85
Strength of Materials

b. Stresses cannot be computed in this manner in the immediate vicinity of the


points of application of the loads. Advanced theoretical or experimental methods
must be used to determine the distribution of stresses in these areas.

Applying the above principle at a section K of a beam, for example, (where K


is not placed in the immediate vicinity of the points of application of the loads), the
effect occured at K is the same for both types of loading (Fig.5.4a,b).

a. b.
Fig. 5.4

7. Bernoullis hypothesis

For structural members in tension


or compression, the plane and parallel
cross sections before the deformation of
the members, remain plane and parallel
after deformation (Fig.5.5).
For beams (structural elements in
bending), the plane sections,
perpendicular to the beam axis before the
deformation of the beam, remain plane
and perpendicular to the beam axis after
deformation (Fig.5.6).
Fig. 5.5

Fig. 5.6

The above presented assumptions let us simplify the strength of materials


calculus without loosing the accuracy in a fundamental manner.

86
6. AXIAL LOADING

One of the basic problems facing the engineer is to select the proper material
and proportion it to enable a structure or machine to perform its function efficiently.
For this purpose, it is essential to determine the strength, stiffness and other
properties of materials. These properties are then compared with those computed
using the strength of materials models. From this comparison the desired unknown
elements may be found.

6.1 STRESS AND STRAIN

A bar is said to be under axial loading (tension - Fig. 6.1a or compression -


Fig. 6.1b) if, at any cross section, the internal forces are represented by a single
component: the axial force N.

a. b.
Fig. 6.1

To determine the stress and strain in a bar under axial loading, the following
basic assumptions will be considered:
The material is continuous, homogeneous, isotropic and linear-elastic;
The displacements and strains are small;
Bernoullis hypothesis is available;
Hookes law is expressed through: x E x .

Static considerations

Suppose that a cutting plane , perpendicular to the bar axis (Ox), isolates the
right segment of the bar (Fig. 6.2a). Then, as shown in Fig. 6.2a, the resisting internal
force over the cut section must balance the applied load P. From the condition of
equilibrium we have:
Fx 0 N P 0 N P .
Strength of Materials

Fig. 6.2 a.
The internal axial force N represents a
global effect occurred at the level of the entire
cut cross section considered.
On the other hand, at the level of the
cross sectional points, normal stresses develop
(Fig. 6.2b). Thus, considering an element of
area dA around an arbitrary cross sectional
point Q, the elementary axial force developed
at this level is (Fig. 6.3). b.

dN dA . (6.1)
This means that the relation between the global axial force N and the normal stress
over the cross section considered may be written as follows:
N dA , (6.2)
A

where A is the bar cross-sectional area.


Geometrical considerations
Since Bernoullis hypothesis is available,
we may note that all points of a certain cross
section displace with the same quantity u, the
section remaining plane after deformation (Fig.
Fig. 6.3
6.4).

We may write therefore



x constant . (6.3)

From Hookes law we have
x E x constant . (6.4)

Since the stress distribution over the cross section considered is constant, may be
written outside the integral (6.2), obtaining

88
Axial loading

a. b.
Fig. 6.4
N dA dA A
(6.5)
A A

and therefore
N P
. (6.6)
A A
We should also note that, in formula (6.6) is obtained by dividing the magnitude P
of the resultant of the internal forces, distributed over the cross-section, by area A; it
represents, therefore, the average value of the stress over the cross section, rather than
the stress at a specific point of the cross section.
To define the stress at a given point Q of a current cross section, we should
consider a small area dA (Fig. 6.3) around Q. Dividing the magnitude of dN by dA,
we obtain the average value of the
stress over dA.
In general the value obtained
for the stress at a given point Q
of the cross section is different
from the value of the average
stress, given by formula (6.6), and
is found to vary across the
section.
In a slender rod subjected to
equal and opposite concentrated
forces P (Fig. 6.5a) this variation is
small in a section away from the
points of application of the
concentrated forces (Fig. 6.5c), but a. b. c. d.
it is quite noticeable in the Fig. 6.5.
neighborhood of these points (Fig. 6.5b,d).
In practice, we shall assume that the distribution of normal stresses in an
axially loaded member is uniform, except in the immediate vicinity of the points of
application of the loads. The value of the stress is then equal to average and may
be obtained from formula (6.6). However, we should realize that, when we assume a
uniform distribution of stresses over a given cross section, the concentrated forces P
must be applied at the centroid of the section (Fig. 6.1). In other words, an uniform

89
Strength of Materials

distribution of stresses is possible only if the line of action of the applied


concentrated loads P passes through the centroid G of the section considered. This
type of loading is called centric loading (Fig. 6.6).

a. b.
Fig. 6.6 Fig. 6.7
However, if a two-force member is axially loaded but eccentrically, as shown
in Fig. 6.7a, we find from the conditions of equilibrium of the portion of member
shown in Fig 6.7b, that the internal forces at a given section must be equivalent to a
force P applied at the centroid of the section and a couple M = aP. In such a case the
distribution of stresses cannot be uniform. This point will be discussed in detail later,
within another chapter.
Let us now return to formula (6.6). To determine whether the bar of Fig. 6.1
may support the given load P, we must compare the value obtained for under this
loading with the maximum value of the stress which may be safely applied to steel (in
case the bar is made of steel). From a table of properties of materials we find the
maximum allowable stress in the type of steel to be used ( a ). Let us suppose we
have to design a simple member (like that represented in Fig. 6.1, for example) so
that it would not fail under specified loading conditions. In such a case, on one hand,
we must compute the actual stresses in the involved member using formula (6.6):
N P
.
A A

On the other hand we have to find out the allowable value of the normal stresses a
for the involved material (using different engineering tables containing materials
properties). In the end, the strength condition of the member considered is

P
. (6.7)
a
A

The condition (6.7) may be used for all three specific types of strength of materials
problems.
90
Axial loading

Another important aspect of the analysis and design of structures relates to the
deformations caused by the loads applied to the structures. Clearly, it is important to
avoid deformations so large that they may prevent the structure for fulfilling the
purpose for which it was intended. Let us now consider a bar BC, of length and
uniform cross-sectional area A, which is suspended from B (Fig. 6.8a).
If we apply a load P to end C, the bar elongates (Fig. 6.8b), with .

Since

x ,

from Hooke's law, we have:



x E x E . (6.8)

Substituting for from (6.6) into (6.8) we may
write:
N
E ,
A
a. b.
Fig. 6.8 from which it follows that

N P
, (6.9)
EA EA

where E is the modulus of elasticity of the material involved and N is the axial
internal force which is constant along the member considered.
Equation (6.9) may be used only if the bar is homogeneous (constant E), has a
uniform cross section of area A and is loaded at its ends. If the bar is loaded at other
points, or if it consists of several portions of various cross sections and possibly of
different materials, we must divide it into component parts which satisfy individually
the required conditions for the application of formula (6.9). Denoting respectively by
Ni, i, Ai and Ei the internal force, the length, the cross sectional area and the modulus
of elasticity corresponding to the part i, we may express the deformation of the entire
bar as:

N i i
. (6.10)
i
E i Ai

For example, let us determine the total deformation of the rods shown in Fig. 6.9,
under the given loading conditions.

91
Strength of Materials

a. b.
Fig. 6.9
P 2a 3 Pa 2 Pa 7 Pa
a) 1 B
1 2
2 3
3 B
.
EA EA EA EA
7 Pa
Thus, in the case a of Fig 6.9, the total deformation of the bar shown is .
EA
b) 1 B 1 2 2 3 3 4 4 B

P 2a 2 Pa 2 Pa 5 Pa Pa
7 ,5 .
EA EA 2 EA 2 EA EA

In case of a rod of variable cross section and/or variable axial internal force
(when N=N(x) and A=A(x), x being the parameter along the rod considered), formula
(6.9) must be applied only to an element of the bar with an infinitely small length
dx (Fig. 6.10).
Since such an element has an infinitely small length dx, the variation of the
cross sectional area A and/or that of the axial internal force N may be neglected along
the element considered, formula (6.9) becoming available.
Thus, in the case of the bar shown
in Fig. 6.10, we may write the
deformation of the element of length dx
as:
N x dx
dx (6.11)
E A x

which is in fact the formula (6.9).


The total deformation of the
bar shown in Fig. 6.10 is therefore
obtained by integrating (6.11) over the
length of the bar:

N (x)
dx . (6.12)
Fig. 6.10 0 EA ( x )

92
Axial loading

Example 6.1

For the rod shown in Fig. 6.11, draw the axial force diagram, determine the required
diameter d and the total displacement of point 1, knowing that the allowable normal stress for the
involved material is a= 150 MPa and the modulus of elasticity is E = 2,1 105 MPa.
Considering the free body of rod 1,2,3,B,
we note that the reaction at B is :
YB = 250 kN.
It follows that:
N1-2 = 125 kN = constant;
N2-3 = 125 kN = constant;
N3-B = 125 + 125 =
= 250 kN = constant.
The axial force diagram has been represented in
Fig. 6.11.
To determine the required diameter d so
that the rod would not fail under the specified
loading, we must use the condition (6.7):
max a.

Fig. 6.11 The maximum value of normal stresses


along the rod considered may be established
only after the computation of normal stress for each particular portion of the rod. We write
therefore:

3 3
N1 2
125 10 4 125 10
1 2 2 2
;
A1 2
d d
4
3 3
N 2 3
125 10 4 125 10
2 3 2 2
;
A2 3
(1 , 5 d ) (1 , 5 d )
4
3 3
N3 B
250 10 4 250 10
3 B 2 2
.
A3 B
(1 , 5 d ) (1 , 5 d )
4
We note that the maximum value of the normal stresses corresponds to the first portion of the rod:
3
4 125 10
max 1 2 a 150 .
2
d

It follows that:
3
4 125 10
d 32 , 57 mm .
150

We write:
d = 32,57 mm 32,6 mm .

93
Strength of Materials

The total vertical displacement of point 1 (i.e. the deformation of the rod) is obtained from (6.10)
after the rod has been divided into four portions, as shown in Fig. 6.11. We write:
N i i
1 2 2 3 3 B
E i Ai
i

3 3 3
125 10 2000 125 10 1000 250 10 1500
2 , 69 mm .
2
5 32 , 6 5 2 5 2
2 ,1 10 2 ,1 10 (1 , 5 32 , 6 ) 2 ,1 10 (1 , 5 32 , 6 )
4 4 4

Example 6.2

The rigid bar BCD is supported by two vertical rods 1 and 2 (Fig. 6.12). The two rods are
made of steel with E = 2,1 105 MPa and a=150 MPa. For the 76,5 kN force shown, determine
the required diameters of the two rods (d1 and d2) and the vertical displacement of point C.

Fig. 6.12 Fig. 6.13


The axial forces in the two vertical rods may be determined by cutting the two rods and using the
equilibrium equations for the free body obtained (Fig. 6.13). We write:

M (D ) 0 N 1 2 , 25 76 , 5 1 , 25 0 N1 42 , 5 k N ;

Fy 0 N1 N 2 76 , 5 k N N 2 34 k N .

It is important to be mentioned that the determination of the axial forces represents a first and
necessary step within axial loading problems.
We do also note that, in our case, the two axial forces in rods 1 and 2 are positive, which
means that the two rods are in tension.
The strength conditions for the two rods are:
3 3
N1 42 , 5 10 42 , 5 10 4
1 a 150 d1 18 , 99 mm ;
2
A1 d1 150

4
3 3
N 2 34 10 34 10 4
2 a 150 d2 16 , 98 mm .
2
A2 d 2
150

94
Axial loading

We adopt d1 = 19 mm; d2 = 17 mm.


The deformations of the two rods are (Fig. 6.14.):

Fig. 6.14
3
N 1 1 42 , 5 10 1000
1 0 , 713 mm ;
2
EA 1 5 19
2 ,1 10
4

3
N 2 2 34 10 1 , 25 1000
2 0 , 891 mm .
2
EA 2 5 17
2 ,1 10
4

From similar triangles, we may finally


find that the vertical displacement of
point C is CC"= 0,792 mm.

Example 6.3
Two rods of uniform circular
cross-sections BC and BD support a 30
kN vertical force (Fig. 6.15). Knowing
that the allowable stress of the rods
material is a = 150 MPa and E = 2,1
105 MPa, determine the required values
of the diameters d1 and d2 and the
vertical displacement of point B.
From the geometry of the
structure shown in Fig. 6.15 one can find
that the lengths of the two rods are
Fig. 6.15 respectively:

1 0 , 956 mm ; 2 0 , 743 mm .

Passing a cross-section through each of the two rods, the two corresponding internal axial
forces N1 and N2 are revealed. From the mechanical equilibrium of point B we have (Fig. 6.16):

95
Strength of Materials

X B
0 N 1
cos 30 N 2
cos 40 0;


YB 0 N 1
cos 60 N 2
cos 50 30 .

We find:
N1 = 24,46 kN; N2 = 27,64 kN .

Applying the strength condition for the two rods, we


Fig. 6.16 write:
3 3
N1 24 , 46 10 24 , 46 10 4
1 a 150 d1 14 , 41 mm ;
2
A1 d1 150
4

3 3
N 2 27 , 64 10 27 , 64 10 4
2 a 150 d2 15 , 31 mm .
2
A2 d 2
150
4

The deformations of the two rods are (Fig. 6.17):

3 3
N 1 1 24 , 46 10 0 , 956 10
1 0 , 682 mm ;
2
EA 1 5 14 , 41
2 ,1 10
4
3 3
N 2 2 27 , 64 10 0 , 743 10
2 0 , 531 mm .
2
EA 2 5 15 , 31
2 ,1 10
4

It is to be observed that, although the


concentrated load P acts vertically, due to
the lack of symmetry of the system
considered, the point B displaces along
direction BB which is not vertical. In the
same time, the deformation of the two rods,
1 and 2, are revealed by sketching a
perpendicular line from the initial position of
point B to the two rods after deformations
(although the method is approximate, it
models with a sufficient accuracy the
deformations of such bars when the
hypothesis of small deformations is
available). Using the same hypothesis of
small deformations, we may also consider
that, after deformation, the two rods 1 and 2
form with the vertical direction the same
angles of 60 and 50 respectively. Assuming
that the angle between BB and the vertical Fig. 6.17
direction is , we may write (Fig. 6.17):

96
Axial loading

1 2
1 2
BB ; BB '

cos 60 cos 50 cos 60 cos 50

1
2

cos 60 cos sin 60 sin cos 50 cos sin 50 sin

1 2

cos cos 60 sin 60 tg cos cos 50 sin 50 tg


1 cos 50 1 sin 50 tg 2 cos 60 2 sin 60 tg


tg 1 sin 50 2 sin 60 2 cos 60 1 cos 50 .

We have:

1 cos 50 2 cos 60 0 , 682 cos 50 0 , 531 cos 60
tg 10 .

1 sin 50 2
sin 60 0 , 682 sin 50 0 , 531 sin 60

The vertical displacement of point B is therefore (Fig.


6.18):
1
v B
BB B B cos cos

cos 60

0 , 682
Fig. 6.18 cos 10 1, 045 mm .

cos ( 60 10 )

Example 6.4

The rigid bar BOC is supported by the rod CD, having a circular uniform cross section of
diameter d. Rod CD is made of steel with a = 150 MPa and E = 2,1 105 MPa. For the 15,7 kN
force shown determine the required value of diameter d and the vertical displacement of point B.
Free Body: the bar BOC
M (O )
0 N 1 0 ,4 P 0 ,3 0 N 1 0 ,4 15 , 7 0 , 3 0 N1 11 , 775 kN ( tension );

Strength condition:
3 3
N1 11 , 775 10 11 , 775 10 4
1 a 150 MPa d 10 mm .
2
A1 d 150
4

Vertical displacement of point B, v B (Fig. 6.20).


We denote by C and B the displaced positions of points C and B respectively. Since the bar
BOC is rigid, the segments OC and OB rotate with the same angle. From similar triangles we may
write:
CC 1 0 ,4
.
BB v B
0 ,3

97
Strength of Materials

Fig. 6.19 Fig. 6.20


Thus:
3 3
0 ,3 N 1 1 0 ,3 11 , 775 10 0 , 7 10
v B 1 0 , 4997 mm .
2
0 ,4 EA 1 0 ,4 5 10
2 ,1 10
4

6.2 POISSON'S RATIO


When a homogeneous slender bar is axially loaded, the resulting stress and
strain satisfy the Hookes law as long as the elastic limit of the material is not
exceeded.

a. b.
Fig. 6.21

Assuming that the load P is directed along the x axis (Fig. 6.21a), we have x = P A,
where A is the cross-sectional area of the bar, and from Hookes law:

x
X
, (6.13)
E

where E is the modulus of elasticity of the material.


We do also note that the normal stresses on faces respectively perpendicular to
the y and z axes are zero: y = z = 0 (Fig. 6.21b). It would be tempting to conclude
that the corresponding strains y and z are also zero. This, however, is not the case. In
all engineering materials, the elongation produced by an axial tensile force P in the
direction of the force is accompanied by a contraction in any transverse direction
(Fig. 6.22).

98
Axial loading

Since the material under


consideration is assumed to be
homogeneous and isotropic, the
strain must have the same value
for any transverse
direction: . This value is y z

referred to as the lateral strain.


The absolute value of the ratio of
the lateral strain over the axial
strain is called Poissons ratio,
denoted by the Greek letter . We
Fig. 6.22
write:

y z
. (6.14)
x x

For the loading condition represented in Fig. 6.21, an increase of dimension along the
x axis, means a decrease of transverse dimensions, so that we have:
.
y x
and z x
. (6.15)
It follows that the relations which fully describe the condition of strain under an axial
load applied along a direction parallel to the x axis are:

x
x
; y z
x
. (6.16)
E E

The above presented phenomenon is also referred to as the transverse contraction.


During such a contraction the original dimensions of the bar under consideration
change as follows (Fig. 6.22):
h h h h h y
h1 y
h 1 x
;

b b b b b z
b 1 z
b 1 x
; (6.17)
0 0 0 0 0 x
0 1 x
.

The new volume of the bar after deformation is:


V h 1 x
b 1 x
0 1 x
,

the original volume (before deformation) being:


V0 b h 0.
The variation of the volume due to the axial load applied is therefore:
2
V V V0 b h 0 (1 x
) (1 x
) b h 0

2 2
b h 0 1 2 x x
1 x
1

2 2 2 2 3
b h 0 1 x
2 x
2 x x x
1

99
Strength of Materials

2 2 2
b h l0 x [1 2 2 x ].
x x
0 0 0

Since the strains , , are much smaller than unity, the last three terms in the
x y z

above relation may be omitted. We have, therefore:

V V V0 b h 0 x
1 2 . (6.18)
On the other hand, it has been experimentally observed that a bar in tension does
always increase its volume. This means that:
V 0 b h 0 x
1 2 0 1 2 0 .
Since the very definition of Poissons ratio requires it to be a positive quantity, we
may conclude that, for any engineering material:
0 0 ,5 . (6.19)
In numerous cases (for numerous materials) = 0,3.

6.3 STRESS CONCENTRATIONS


The exact computation methods of the theory of elasticity and the experimental
photoelastic methods show that, for members in tension or compression having a
variable cross section, the corresponding normal stresses are not uniformly
distributed over the cross sections. Therefore, Bernoullis hypothesis does loose its
validity. When a structural member contains a discontinuity, such as a hole or a
sudden change of the cross section, high localized stresses may also occur near the
discontinuity. Fig. 6.23 and 6.24 show the distribution of stresses at the level of the
critical sections corresponding to two such cases.

Fig. 6.23 Fig. 6.24

100
Axial loading

The ratio between the maximum stress ( max) and the average stress ( ave)
computed at the level of the critical (narrowest) section of the involved discontinuity
is called the stress - concentration factor:
max
k
. (6.20)
ave

It is to be noted that the average stress ( ave) is computed with the basic
formula (6.6) while the maximum stress ( max) represents the real maximum stress
occured due to the stress-concentration phenomenon. To determine the maximum
stress occurring near a discontinuity in a given member in tension or compression,
the designer needs only to compute the average stress ave = P A in the critical
section and multiply the result obtained by the appropriate value of the stress -
concentration factor k. On the other hand, the stress-concentration factor is given in
form of tables or graphs, as shown in Fig. 6.25 for the two cases represented in Fig.
6.23 and Fig. 6.24.

Fig. 6.25

It is also to be noted that the above presented procedure remains valid only if
max does not exceed the proportional limit of the involved material (the concept of
proportional limit of a material will be explained later).

101
Strength of Materials

6.4 OWN WEIGHT EFFECT. MEMBERS OF CONSTANT


STRENGTH
As we saw in the preceding sections, the strength calculus for different types of
members in tension or compression does neglect the own weight of the members.
This procedure remains valid only if the length of the involved members has low
values. Otherwise, if the length of the members has important values, the own weight
cannot be neglected.
Consider for example a
homogeneous rod BC of length
and uniform cross-sectional area A,
which is suspended from B, hanging
under its own weight, (Fig. 6.26).
Let us also assume that the
rod material specific weight is .
is defined as the ratio between the
weight G and the corresponding
volume of the material, V. We may
Fig. 6.26
write therefore:
G
or G V . (6.21)
V

At an arbitrary cross section of the involved bar, located at distance x from end C we
write:
x 0; NC P;
N x P G x P A x;
x ; N B
P A.

P
x 0; C 1
;
N x P Ax P A
x x;
A A A P
x ; B 2
,
A

where G(x) represents the weight of the bar portion below the cross section
considered.
The diagrams of the axial forces and the normal stresses have been shown in
Fig. 6.26. Note that the maximum values of axial force and normal stress occur
within the upper cross section of the rod (for x = ). The strength condition is:
P
max B
a
. (6.22)
A

We may write therefore that, the required cross-sectional area of the rod is:

102
Axial loading

P
A , (6.23)
a

where all
is the allowable value of the normal stress for the material the rod is made
of.
In the absence of force P (P = 0) we have:

max . (6.24)

This means that, in such a case, if the length of the rod reaches a critical value c, the
rod breaks under its own weight. For = c, the maximum stress in the rod equals the
breaking strength of the material ( B) the rod is made of. We write:
B
max B
c c . (6.25)

The quantity B will be defined later.


The total deformation of the rod shown in Fig. 6.26 is obtained by
integration over the length of the rod:


N x 1 1
dx N x dx P A x dx
EA x EA EA
0 0 0

2 2
1 Ax 1 A
Px P .
EA 2 0
EA 2

Since:
G
A V G ,
V
we may write:
2
1 A 1 G G
P P P .
E A 2 E A 2 E A 2

Thus, it follows that under its own weight the total deformation of the rod subjected
to the concentrated force P is:
G
P

2
,
(6.26)
E A

where G is the total weight of the involved rod. We note that the rod shown in Fig.
6.26 has a constant cross-section. It follows from (6.22) that the normal stress is
maximum at section B.

103
Strength of Materials

Fig. 6.27 Fig. 6.28

This means that only at B, the rod can be dimensioned in a proper way, all other
cross-sections having bigger dimensions than necessary. This is the reason why the
engineers have tried to design a bar so that each cross-section of the bar to be equally
loaded (i.e. at every cross-section of the bar the normal stress has the same value).
Thus the concept of MEMBERS OF CONSTANT STRENGTH has arisen.
Obviously, such a member of constant strength must have a continue variable
cross section, increasing from the lower end to the upper end (Fig. 6.27). At distance
x from end 1, the cross-sectional area is A(x). The problem is to determine the
mathematical expression of the function A(x), finding in this way the cross-sectional
area variation mode along the involved member. Let us now consider again the
member shown in Fig. 6.27. We denote by G(x) the weight of the member portion of
length x. We do also consider two cross-sections (mn and m'n') at an infinite small
distance dx from each other (Fig. 6.28). The cross-sectional area increases from mn to
m'n' with an infinite small quantity dA(x).
At the level of the cross-section mn we may write:

N x P G x
x . (6.27)
A x A x

Recalling that a member of constant strength must have the same stress at each
particular cross-section (equal to a) we write:

P G x
x a ,
A x

and thus
A x a
P G x . (6.28)

Applying formula (6.28) for the cross-section m'n' (for which the area is increased
with dA(x) and the weight of the lower member portion with dG(x)) we have:

104
Axial loading

A x dA x a
P G x dG x
or (6.29)
A x dA x a
P G x A x dx ,

Substracting (6.28) from (6.29) we write:

A x dA x A x a
P G x A x dx P G x ,
or
d A x a
A x dx .
This means that
d A x
dx , (6.30)
A x a

Integrating member by member we have:

ln A x x C ,
a

or
x C
A x e a
, (6.31)

where C is an integration constant.


We have obtained therefore an exponential variation of the cross-sectional area
along the involved member.
The constant C may be obtained from (6.27) and (6.31) making x = 0 and =
a. We write:

N 0 P P
a .
C
A 0 A 0 e

Thus
C P P
e C ln . (6.32)
a a

It follows that
P P
x ln x ln x
a a a a
P a
A x e e e e
a

and thus
x
P
A x e a
. (6.33)
a

105
Strength of Materials

Using (6.28), the corresponding weight of the member portion below the cross
section mn is
x
a
G x A x a
P P e 1 . (6.34)

The total deformation of the member is obtained by integration over the length
of the member:

N x 1 N x 1 a

dx dx a
dx . (6.35)
EA x E A x E E
0 0 0

Although the relation (6.33) insures the condition of equal strength, such a
member is enough complicated from the technological point of view. A much more
practical solution is shown in Fig. 6.28.
We may observe that the member shown in Fig. 6.29 has three distinct portions
of cross-sectional areas A1, A2, A3 and lengths 1, 2, 3 respectively. For each portion
the maximum stress is to be reached at the upper end. For the first portion we may
write:
P A1 1 P
m ax 1.
1 A1 A1

Since this stress has to be equal to the allowable stress ( a) of the material the
member is made of, we have

P
m ax 1 a
.
1 A1

The necessary value of the cross-sectional area A1, so that the first portion of the
member would not fail, is
P
A1 . (6.36)
a
1

In the same manner, for the second portion we may write

P A1 1 A2 2 P A1 1
m ax 2.
2 A2 A2 A2

1
m ax a
P A1 1 2 a
2 A2

P A1 1 a
A1 P a
A2 .
a
2 a
2 a
1 a
2

106
Axial loading

For the third portion we have

Fig. 6.29

2
P a
A3 .
a
1 a
2 a
3

In general, for a member with n distinct portions, the required cross-sectional area Ai
of the ith portion is therefore:
i -1
P a
Ai . (6.37)
a
1 a
2 a
i

In particular, if all portions of the member have the same length we may write:
i -1
P a
Ai i
. (6.38)
a

6.5 STRESS-STRAIN DIAGRAM

6.5.1 GENERALITIES

Although apparently simple, the tensile tests of the material provide the
strength of materials with necessary and useful information. It is to be noted that, the
strength of a material is not the only criterion that must be considered in engineering
design. The stiffness of a material is frequently of equal importance. On the other
hand, the mechanical properties such as hardness, toughness and ductility are
generally those which determine the selection of a certain material within a certain
engineering context. These properties are determined by making tests on the
materials and comparing the results with established standards. Although a complete
description of these tests is the province of materials testing and hence will not be

107
Strength of Materials

given here, one of the tests (the tension test of steel) and its results will be considered
because it helps to develop several important basic concepts.
The tensile tests stress-strain diagram represents the graphical visualization of
the relation between stress and strain in a given material, being a very important
characteristic of the material. To obtain the stress-strain diagram of a material, one
usually conducts a tensile test on a specimen made of that material. One type of
specimen commonly used is shown in Fig. 6.30.
In a tensile test, the specimen is gripped between the jaws of a testing machine
(Fig. 6.31). Before the applying
of the tensile load P, two marks
are inscribed on the specimen
(k and k - Fig. 6.30), at a
distance 0 from each other.
After the specimen is placed in
the testing machine, as the load
P increases, the distance
between the two marks does Fig. 6.30
also increase. At a given
moment i, the axial force at any cross-section of the cylindrical portion of the
specimen is Ni (equal to P),
while the distance between k
and k is i. This means that,
at any moment, we have two
values:
- the axial force in the
involved specimen Ni = P;
- the instant distance
between the two marks k and
k.

The distance i is measured


with a dial gage and the
elongation i = i -0 is
recorded for each value of P.
We then convert each pair of
readings (Ni and i) at a
certain moment i into the
Fig. 6.31 instant normal stress and
strain of the member as
follows:

N i i i 0
i and i , (6.39)
A 0 0

108
Axial loading

2
d0
where A is the original cross-sectional area of the specimen A .
4

Plotting these data on a graph with the ordinate representing the stress i and the
abscissa representing the strain i of the specimen at different moments throughout
the test, we obtain the diagram called the stress-strain diagram. Fig. 6.32 represents
such a graph for structural steel.
The diagram obtained by plotting the stress versus the strain, reflects the
behavior of the involved
material. It is to be
mentioned that the
following conditions have
to be fulfilled within a
tensile test:
- the specimen must have a
constant cross section;
- the material must be
homogeneous;
- the load must be axial,
Fig. 6.32 i.e. produces uniform
stress.
Finally, note that since strain represents a change in length divided by the
original length, strain is a dimensionless quantity. However, it is common to use units
of meters per meter (m / m) or millimeters per millimeter (mm/mm) when referring to
strain. In engineering work, strains of the order of 1 10-3 are frequently encountered.
The main points and concepts developed from the stress-strain curve (Fig.
6.32) are:

The proportional limit (P) represents the maximum value of the stress p up
to which the stress is directly proportional to the strain . From the origin O up to
the proportional limit, the stress-strain diagram is a straight line. We may write:

=E , (6.40)

a relation known as Hookes


law, after the English
mathematician Robert Hooke.
The quantity E is called the
modulus of elasticity of the
material involved, or Youngs
modulus.

The elastic limit (E)


represents the stress e up to
which the material will return Fig. 6.33
109
Strength of Materials

to its original shape when unloaded;


The yield limit (Y) is the stress y at which there is an appreciable elongation
or yielding of the material without any corresponding increase of the load. However,
the phenomenon of yielding is peculiar to structural steel; other grades of steel and
steel alloys or other materials do not possess it, as is indicated by the typical stress-
strain curves of these materials shown in Fig. 6.33. The yield strength is closely
associated with the yield point.
The ultimate stress, or ultimate strength ( u) as it is more commonly called,
is the highest ordinate on the stress-strain curve.
The rupture strength ( r) is the stress at failure. For structural steel it is
somewhat lower that the ultimate strength because the rupture strength is computed
by dividing the rupture load by the original cross-sectional area, which, although
convenient is incorrect. The error is caused by a phenomenon known as necking.
As failure occurs, the material stretches very rapidly and simultaneously
narrows down, as shown in Fig. 6.34, so that the rupture load is actually distributed
over a smaller area.

Fig. 6.34

If the rupture area is measured after failure occurs, and divided into the rupture
load, the result is a truer value of the actual failure stress. Although this is
considerably higher than the ultimate strength, the ultimate strength is commonly
taken as the maximum stress of the material.

6.5.2 SAFETY COEFFICIENTS. ALLOWABLE STRENGTHS

The working stress, also called the allowable stress, is the maximum safe stress
a material may carry. In design the allowable stress all should be limited to values
not exceeding the proportional limit so as not to invalidate the stress-strain relation of
Hookes law on which all subsequent theory is based. However, since the
proportional limit is difficult to determine accurately, it is customary to base the
allowable stress on either the yield point or the ultimate strength, divided by a
suitable number c, called the factor of safety (or the safety coefficient):

y
a
or a
u
. (6.41)
c c

The determination of the factor of safety which should be used for various
applications is one of the most important engineering tasks. On the one hand, if a
factor of safety is chosen too small, the possibility of failure becomes unacceptably
large; on the other hand, if a factor of safety is chosen unnecessarily large, the result
110
Axial loading

is an uneconomical or nonfunctional design. The choice of the factor of safety which


is appropriate for a given design application requires engineering judgement based on
many considerations, such as the following:
- variations which occur in material properties;
- the number of loadings which may be expected during the life of the structure
or machine;
- the type of loadings which are planned for in the design, or which may occur in
the future;
- the type of failure which may occur;
- uncertainty due to the methods of analysis;
- deterioration which may occur in the future due to poor maintenance or due to
unpreventable natural causes;
- the importance of a given member to the integrity of the whole structure.
For the majority of structural and machine applications, the factors of safety
are specified by design specifications or building codes written by committees or
experienced engineers working with professional societies, with industries or
agencies.

6.6 STATICALLY INDETERMINATE PROBLEMS

In the problems considered in the


preceding sections, we could always use the
conditions of mechanical equilibrium to
determine the internal forces produced in the
various portions of a member under given loading
conditions. But there are many problems,
however, in which the internal forces cannot be
determined from statics alone (i.e. the equations
of static equilibrium are not sufficient for a
solution). Such cases are called statically
indeterminate and require the use of additional
relations that depend on the elastic deformations
Fig. 6.35 in the members. The following examples will
show how to handle this type of problems.
Consider, for example, two rods BC and DC, pin-connected at C and subjected
to a vertical concentrated force P, applied at C, Fig. 6.35. The axial forces (N1 and
N2) in the two rods may be completely computed from statics alone. We write

Fx 0 N 1 sin N 2
sin 0 ,

Fy 0 N 1 cos N 2
cos P 0 .

111
Strength of Materials

from which we obtain the values of N1 and N2. This is a typical example of statically
determinate problem. But what happens if another rod EC is attached (Fig. 6.36)? In
such a case a part of the internal forces N1 and N2 will be taken over by the
supplementary rod EC. So that, the internal forces in the three rods will be N 1' , N 2'
and N 3' . We have therefore three unknowns in the problem ( N 1' , N 2' , N 3' ). On the
other hand, from statics we can use only two equations of mechanical equilibrium as
in the preceding case:
Fx 0 and Fy 0 .
Clearly, the two above equations are not
sufficient to determine the three unknown
internal forces N 1' , N 2' and N 3' . We have
therefore, a statically indeterminate problem. The
difference between the number of unknown
quantities (3 in this case) and the number of
equations available from statics is called the
indetermination degree. In our example, the
Fig. 6.36 indetermination degree is 1. It is to be noted that
2, 3 or more supplementary rods may be attached to the system shown in Fig. 6.35. In
consequence, the indetermination degree will be 2, 3, etc.
Let us now consider another example. A
rigid bar is pin-connected at O, and suspended
from a vertical rod BC, as shown in Fig. 6.37.
Determine the internal axial force N1 in the rod
caused by the load P applied at D.
Since, from the summation of moments at
O, we can determine the value of N1, the problem
is statically determinate.
We write
P
M O 0 N1 a P 0 N1 .
Fig. 6.37 a
But a supplementary vertical rod BC
attached to the system shown in Fig. 6.37,
transforms the problem into a statically
indeterminate one, Fig. 6.38.
This time, we have two unknowns and only
a single equilibrium equation available from
statics. We write
M O 0 N 2 a' N1 a P 0 .
Thus the indetermination degree is 1. In the same
time the presence of the supplementary attached
Fig. 6.38 rod determines the decrease of the original axial
112
Axial loading

force N1 shown in Fig. 6.37. If a new supporting rod is attached, the indetermination
degree increases with 1 unit (becoming 2).
Such statically indeterminate problems (in which the reactions and the internal
forces cannot be determined from statics alone) may be resolved by using adequate
additional relations involving deformations which are obtained from the geometry of
the problem (as presented below).
Several specific types of statically indeterminate problems may be
differentiated.

6.6.1 MEMBERS WITH UNHOMOGENEOUS CROSS SECTIONS

A rod (1) of length , cross-sectional area A1, and modulus of elasticity E1, has
been placed inside a tube (2) of the same length, but of a cross sectional area A2 and
modulus of elasticity E2, as shown in Fig 6.39. Determine the internal forces in the
two members when the load P is applied through a rigid plate as shown.
We denote by N1 and N2 the unknown axial forces
in the two involved members. The single equation which
may be written from statics is:
N1 N 2
P . (6.42)

Clearly, one equation is not sufficient to determine


the two unknown internal forces N1 and N2.. Thus, the
problem is statically indeterminate. However, the
geometry of the problem shows that the two members
have the same axial deformation. We may write,
therefore
Fig. 6.39 1 2

N 1 N 2 N1 N 2
(6.43)
E 1 A1 E 2 A2 E 1 A1 E 2 A2

Equations (6.42) and (6.43) may be solved simultaneously for N1 and N2:

A1 E 1 P A2 E 2 P
N1 ; N 2
. (6.44)
A1 E 1 A2 E 2 A1 E 1 A2 E 2

The above presented example may be also extended to the general case shown
in Fig 6.40, where a member is composed of other "n" members, all being subjected to
a concentrated load, applied through a rigid plate. Denoting by N1, N2,...,Nn the
unknown axial forces in the n members we may write

N1 + N2 + N3 +...+ Nn = P (6.45)

113
Strength of Materials

But one equation (written


from statics) is not sufficient to
determine the n unknown axial
forces. The problem is (n-1) degree
statically indeterminate. However,
since the axial deformations of the n
members have the same value, we
may also write another (n-1)
supplementary equations as follows:
1 2 3 n
Fig. 6.40

N 1 1 N 2 2 N 3 3 N n n
... . (6.46)
E 1 A1 E 2 A2 E 3 A3 E n An

Equations (6.45) and (6.46) may be solved for N1, N2, N3,...,Nn.
Numerical example
A rod (1) with a variable circular cross section has been placed inside a tube (2) of the same
length . The rod is made of steel with E E 1 2 ,1 10 5 MPa while the tube is made of aluminum
with E E 2 0 , 7 10 5 MPa . The two members are compressed by a concentrated load P acting
through a rigid plate, as shown in Fig. 6.41. Draw the axial forces and normal stresses diagrams
for the two members.

Fig. 6.41
Denoting by N1 and N2, respectively, the axial forces in the rod and in the tube, we may
write from statics
N1 N 2 P . (6.47)
The geometry of the problem shows that the axial deformations of the two members are
equal. We write therefore:
1 2. (6.48)
Since the rod has a variable circular cross section, its total axial deformation may be
expressed as

114
Axial loading
6d
N1 N1
1 dx dx ; (6.49)
0
E 1 A1 ( x ) 0
E 1 A1 ( x )

where A1(x) represents the current cross-sectional area of the rod, at distance x, as shown in Fig.
6.42. We write
2
BC
A1 ( x ) ;
4

BC 2 BB ' B' C ' 2 BB ' d .


From similar triangles we may write
BB ' x
2d d 6d
2

2 BB ' x x
2 BB '
d 6d 6
x x 6d
BC d
6 6 Fig. 6.42
2
x 6d
A1 ( x ) (6 d x)
2
.
4 36 144

In other words, the rod circular cross sectional area at distance x from the rod end is

A1 ( x ) (6 d x)
2
. (6.50)
144
Now, the rod axial deformation may be expressed as follows:
6d 6d 6d 6d
N1 N1 144 144 N1 1 144 N 1 (6 d x )'
1 dx 2
dx 2
dx 2
dx
0
E 1 A1 ( x ) E1 0
(6 d x) E1 0
(6 d x) E1 0
(6 d x)

1 6d
144 N 1 (6 d x) 144 N 1 1 1 144 N 1 1 1 144 N 1 12 N 1
.
E1 ( 1) 0
E1 6d 6d 6d E1 6d 12 d E 1 12 d E1d

In the same time, the tube axial deformation is


N 2 N 6d N 6d N 24 d 24 N
2
2 2 2
2
2
.
E 2 A2 E 2 A2 2 2 E 2 5d 5 E 2d
E2 9d 4d
4

Since the two axial deformations have the same value, we may write
12 N 1 24 N 2
1 2
,
E1d 5 E2d

obtaining that
N 2
0 ,833 N1. (6.51)
Equations (6.47) and (6.51) may be solved simultaneously for N1 and N2. We shall finally have:
N1 36 k N ;

N 2
30 k N .
The axial forces and the normal stresses diagrams have been sketched in Fig. 6.41.

115
Strength of Materials

6.6.2 STRAIGHT MEMBERS FIXED AT BOTH ENDS

A horizontal member BC is fixed at its both ends before the external axial
concentrated force P is applied. The member is of length , uniform cross section of
area A, and modulus of elasticity E (Fig. 6.43). Compute the stresses corresponding
to the portions B-1 and 1-C due to
the application of load P at point
1?
Due to the action of the
concentrated load P, applied at
Fig. 6.43 point 1, two horizontal reactions at
points B and C develop (Fig. 6.44).
From statics we may write
X B
X C
P . (6.52)
Thus, the problem is statically indeterminate (the indetermination degree being
equal to 1). However, the unknown reactions XB and XC may be determined if we
observe from the geometry that the total elongation of the member must be zero
(points B and C are fixed from the beginning). Denoting by B 1 and 1 C ,
respectively, the elongations of the portions B-1 and 1-C, we write
B C
B 1
1 C

X B
a P X B
b
0
EA EA
or

X B a P b X B b 0

Pb Pb
X B
. (6.53)
a b
Carrying this value into (6.52) we
write
Fig. 6.44

P b P b P b P a
X C
P X C
P . (6.54)

We may write therefore


Pb
X B
;

Pa
X C
.

116
Axial loading

With this values in mind we are now in the position to draw the axial forces
diagram. This diagram has been sketched in Fig. 6.44. The normal stress for each
specific portion of the member is
Pb
N Pb
B 1
B 1
;
A A A

Pa
N1 Pa
1 C
C
.
A A A

The normal stresses diagrams have been also sketched in Fig. 6.44.

Numerical example
A tube of length , with the inner diameter of 50 mm and the outer diameter of 80 mm, has
been fixed at its both ends as shown in Fig. 6.45. Determine the reactions at B and C and draw the
axial forces and normal stresses diagrams for the loading shown.

Following the same procedure as presented above, we write


Fx 0 X B
180 80 X C
0 X B
X C
260 ; (6.55)
3 3 3
X B
0 , 9 10 X B
180 10 1, 2 10
B C
B 1
1 2
2 C
EA EA

3 3 3
X B
180 10 80 10 1, 2 10
0 . (6.56)
EA

Equations (6.55) and (6.56) may be solved


for XB and XC:

X B
160 k N ;

X C
100 k N .

With these values, the axial forces


diagrams have been sketched in Fig. 6.45.
The corresponding normal stresses for each
distinct portion of the member are:

3
N 160 10
B 1
B 1
72 , 75 MPa ;
A 2 2
80 60
4
3
N1 20 10
1 2
2
9 , 09 MPa ;
A 2 2
80 60
4
3
N 100 10
2 C
2 C
45 , 47 MPa .
A 2 2
80 60
Fig. 6. 45 4

117
Strength of Materials

If the tube is made of steel, for example, with a


= 160 MPa, we can conclude that it does
not fail under the action of the two concentrated forces ( max
=72,75 MPa a
160 MPa ).

6.6.3 SYSTEMS OF PARALLEL MEMBERS


A horizontal rigid bar B1...Bn+1 is suspended from "n" vertical members (bars)
as shown in Fig. 6.46. Determine the internal axial forces for the n bars and for the
loading shown. Since from
statics there are only two
equations available ( F 0 Y

and M 0 ) the problem is


statically indeterminate (the
indetermination degree being
"n-2" ).
Due to the action of the
vertical concentrated forces P1,
P2,..., Pn the rigid bar displaces
vertically and rotates with a
certain angle as shown in Fig.
6.46. We denote by
' ' '
B 1 , B 2 ,..., B n 1 the displaced
Fig. 6.46.
positions of points B1, B2,..,Bn+1.
Since the bar B1 B2 B3... Bn+1 is rigid, points B 1 , B 2 , B 3' ,..., B n' 1 lie in a straight line. In
' '

the same time the vertical displacements of the supporting points B2, B4, Bi+1,...,Bn+1
coincide with the elongations of the vertical bars 1, 2 , 3, ...n respectively. This is
why, the geometry of the problem lets us write other n-2 relations among
1 , 2 ,..., n , and, thus n-2 relations among N 1 , N 2 , N 3 ,..., N n .
In this way, n distinct equations may be written among the axial forces in the n
vertical bars: N 1 , N 2 , N 3 ,..., N n . Solving this system we reach to the values of
N 1 , N 2 , N 3 ,..., N n as functions of the external applied loads P1, P2, ...,Pn. The normal
stress corresponding to each of the n vertical bars may be then computed using the
formula:
Ni
i
, i 1 ,..., n .
Ai

Numerical example 1
The rigid bar BCD is suspended from three vertical bars 1, 2 and 3 as shown in Fig. 6.47.
Knowing that the length of the three vertical bars are 1 = 1 m, 2 = 0,75 m, 3 = 1,25 m
respectively, and the allowable stress of the material the bars are made of is a =120 MPa,
determine the required value of the diameter d if a concentrated force P = 180 kN is applied at K.

118
Axial loading

Considering the free body of the rigid bar


BCD, two equilibrium equations may be
written from statics:

Fy 0 N1 N 2
N 3
P ; (6.57)

M A
0

N1 1 P 1, 5 N3 2 0

N 2
2 N3 1, 5 P . (6.58)

On the other hand, we may also write a


geometrical relation among 1 , 2 and
3 as follows:

3 1
Fig. 6.47 2 or
2
2 2 3 1 .
We may write therefore

N 2 2 N 3 3 N 1 1
2 2N 2
0 , 75 N 3 1, 25 N1 1 N1 1, 5 N 2
1, 25 N3 0 (6.59)
EA EA EA

Equations (6.57), (6.58) and (6.59) may be solved simultaneously for N1, N 2
and N 3
:

N1 1, 9 k N ;

N 2
36 , 2 k N ;

N 3
41 , 9 k N .

Since the three vertical bars 1, 2 and 3 have identical cross-sections (i.e. identical cross sectional
areas A), the maximum normal stress develops in the bar for which the axial force is maximum (i.e.
the bar 3). We write therefore
3 3
N 41 , 9 10 4 41 , 9 10
max 3
3
2 all
120 d 21 mm .
A d 120
4

Thus, the minimum acceptable value for the diameter d, so that the system would not fail, is
21 mm.

Numerical example 2
The rigid bar OBCD is supported as shown by two steel vertical bars 1 and 2 of 10 mm
diameter ( E 2 ,1 10 5 MPa ). Determine the normal stress in each vertical bar caused by the load P
= 19 kN applied at C . (Fig. 6.48).

119
Strength of Materials

From statics we may write

M O
0

N 1 2a P 3 , 25 a N 2 5a 0

2N1 5N 2
3 , 25 P .
(6.60)
Thus, the problem is statically
indeterminate (indetermination degree is 1).
However, from similar triangles we have:
BB ' 2a 1 2a
5 1 2 2
DD ' 5a 2
5a

5 N 1 1 2 N 2 2
5 N 1 750 2N 2
1250
EA EA

Fig. 6.48
N 2 1, 5 N 1 .
(6.61)

From (6.60) and (6.61) we finally have


N1 6 ,5 k N ;

N 2
9 , 75 k N .

With these values, the corresponding normal stresses in the two vertical bars are:
3 3
N1 6 , 5 10 6 , 5 10
1 2 2
82 , 76 MPa ,
A d 10
4 4
3
N 9 , 75 10
2
2
2
124 MPa .
A 10
4

6.6.4. SYSTEMS OF CONCURRENT BARS

Numerical example

120
Axial loading

Three members made of the same material and having the same cross-sectional areas are
used to support a load P, as shown in Fig. 6.49.
Knowing that E 2 ,1 10 5 MPa and A 600 mm 2
determine the axial force and stress in each member

when 30 and compute the vertical displacement of
point B.
From statics we may write
Fx 0


N 1 sin 30 N 3 sin 30 0 N1 N3 ; (6.62)

Fy 0 N 1 cos 30 N 2
N 3 cos 30 0

Fig. 6.49 2 N 1 cos 30



N 2 120 . (6.63)
Due to the symmetry of the system, the point B displaces to B' as shown in Fig. 6.49. We have
1

BB ' 2
. (6.64)
cos 30

The lengths of the three bars are

1200
1
1385 , 6 mm 3 ;
cos 30

2 1200 mm .

With these values equation (6.64) becomes:


N 1 1 N 2 N 1 1 1385 , 6

2
N 2
N1 1,15 N1 . (6.64')
EA cos 30 EA 2
1200

Equations (6.62), (6.63) and (6.64') may be now solved for N1, N2, N3:

N1 41 , 63 k N N 3;

N 2
1 ,15 N1 47 , 87 k N .

The corresponding normal stress in each member are:


3 3
N1 41 , 63 10 N 2
47 , 87 10
1 3
69 , 38 MPa and 2
79 , 78 MPa .
A 600 A 600

The vertical displacement of point B is


3
N 2 2
47 , 87 10 1200
v B
2 5
0 , 456 mm .
EA 2 ,1 10 600
NOTE:
The statically indeterminate non-symmetrical systems of concurrent bars will be discussed later,
using energetic methods.

6.6.5 PROBLEMS INVOLVING STRESSES PRODUCED BY ASSEMBLING


IMPERFECTIONS

121
Strength of Materials

Let us consider a straight bar of uniform cross-section of area A and modulus


of elasticity E. This bar has to be fixed between two supporting points B and C as
shown in Fig. 6.50a. During the assembling process it comes out that the bar is
shorter than necessary with a small quantity .

a. b.
Fig. 6.50
To fix the bar B'C' between points B and C, one must pull with an external
axial force N until point C' reaches the position C. Only at that moment the assembly
is possibly. But, at that moment, we find an internal axial force in the bar, which is
equal to N. The necessary axial force used for the compensation of the imperfection
is:
N EA
N . (6.65)
EA
We may write that, the normal stress in the bar due to the compensation of the
assembling imperfection is
N E
. (6.66)
A
This assembling normal stress will be then added to the stress resulting from
the functional use the bar has been designed for. This is why such assembling stresses
are undesirable stresses, in many cases being unknown when starting the operation of
the involved member.

Numerical example
The rigid bar OBCD is supported as shown by three vertical steel roads 1, 2 and 3 (with the cross-
sectional areas A 1 100 mm 2 , A 2 150 mm 2 and A 3 300 mm 2 respectively and E 2 ,1 10 5 MPa )
and pin connected at point O. The vertical rods have the same length l 3 m . During the
assembling it comes out that a clearance exists between the supporting point D and the end of
the rod 3, Fig. 6.51. Knowing that 0 , 6 mm and a 1 m determine the normal stresses in the
three rods and the rotation angle of the rigid bar OBCD if the imperfection is removed.

122
Axial loading

The clearance is
compensated by pulling
downward from point D and
pulling upward from the end F
of rod 3, such as point D
displaces to D' with quantity y
while point F displaces to D'
with quantity 3 . When points
D and F are pin connected at D',
the three vertical rods 1, 2 and 3
will be in tension, the internal
axial forces in the three rods
being N1, N2 and N3 respectively.
Since the compensated clearance
is a very small quantity the
geometry of the system will not
change essentially, the points B,
Fig. 6.51 C and D displacing vertically to
B', C' and D' respectively. In the
same time, since the bar OBCD is rigid, points O, B', C' and D' will be in a straight line after
compensation of .
Considering the free body of bar OBCD, from statics we may write:
M O
0 N1 a N 2
2a N 3
3a 0 N1 2N 2
3N 3
. (6.67)
With three unknown quantities (N1, N2, N3) and only a single equilibrium equation from statics, the
problem is statically indeterminate (the indetermination degree being 2). However, two more
equations may be added from the geometry of the deformed system. From similar triangles we may
write:
BB ' 1 a N 2 2 N 1 A2 150
2
2 1 N 2
2N1 2N1
CC ' 2
2a EA 2
EA 1 A1 100

N 2
3N1 (6.68)
and
1 a
y 3 1 .
y 3a
But, on the other hand,
N 3 N 1
3 y 3 3 1 3 . (6.69)
EA 3 EA 1

Equations (6.67), (6.68) and (6.69) may be solved simultaneously for N1, N2 and N3:
N1 1058 N ;

N 2
3174 N ;

N 3
2478 N .

With these values, the corresponding normal stresses in the three vertical rods due to the
compensation of are

123
Strength of Materials

N1 1058
1
10 , 58 MPa ;
A1 100

N 2
3174
2
21 ,16 MPa ;
A2 150

N 3
2478
3
8 , 26 MPa .
A3 300

The angle of rotation of the rigid bar OBCD is

BB ' 1 N 1 1 1058 3000 1 4


tg 5
1, 58 10 rad .
a a EA 1 a 2 ,1 10 100 1000

6.6.6 PROBLEMS INVOLVING TEMPERATURE CHANGES

The mechanical structures elements which have been considered so far were
assumed to remain at the same temperature while being loaded. We shall now
consider the situation involving changes in temperature.
Let us consider for example a homogeneous member BC, of length 0 and
uniform cross section, fixed at its both ends as shown in the figure below.
If the member temperature is
raised by t the member tends to
elongate with the quantity
0 t , (6.70)
Fig. 6.52 where is a constant characteristic
of the material, called the coefficient
1
of thermal expansion (expressed in C ).
Since the member BC is fixed at its both ends, it cannot elongate, the supports
exerting equal and opposite axial forces N on the member after the temperature has
been raised, to keep it from
elongating (Fig. 6.53). It follows that
a state of stress occurs in the
member. Since N cannot be
computed from statics, the problem
is statically indeterminate (the
Fig. 6.53 indetermination degree being 1).
To compute the unknown
value of reaction N, we detach the member from its support at C (Fig. 6.54a) and let it
elongate freely as it undergoes the temperature change t (Fig. 6.54b). Equation
(6.70) tells that, in such conditions, the member elongates with the quantity
0 t .
124
Axial loading

After the elongation has been


produced we may return to the
original state of Fig. 6.53 by
a. applying the force N at point C and
making the elongation become zero
again, (Fig. 6.54c).
From Fig. 6.54b,c we may write
b.
N0
0 t
EA

0 EA t
N EA t.
c. 0
(6.71)
The normal stress in the member due
Fig. 6.54 to the temperature change is
therefore

N EA t
E t (compression). (6.72)
A A

Numerical example
Determine the normal stresses 1 and 2 corresponding to the two portions BC and CD of
the bar shown in Fig. 6.55, when the temperature is raised by t 30 C . Use the values:
E1 2 ,1 10
5
MPa ; E 2 0 , 7 10
5
MPa ; A 1
2
200 mm ; A 2 400 mm
2
; 1=500 mm; 2=600 mm;
6 1 6 1
1
12 10 C ; 2 24 10 C .
Since the temperature

increases with t C , the bar tends to
elongate but cannot because of the
restrains imposed on its ends. The
elongation of the rod is thus zero.
Equalizing the elongation
produced freely (in case the bar is
detached from its support at B) due to
Fig. 6.55 the temperature change with that
produced by the action of N, we have
N 1 N
1
1 t 2
2
t
2
. (6.73)
E 1 A1 E 2 A2
We may write therefore
1 2
N t 1
1 2
2
E 1 A1 E 2 A2

125
Strength of Materials
6 6
t 1 30 12 10 500 24 10 600
N
1 2 2
18 , 360 k N .
1 2
500 600
5 5
E 1 A1 E 2 A2 2 ,1 10 200 0 , 7 10 400

Noting that the axial forces in the two distinct portions of the bar are equal to N 18 , 360 kN , we
obtain the following values of the stress in the portions BC and CD of the bar:

N 18360
1 BC
91 , 8 MPa ;
A1 200

N 18360
2 CD
45 , 9 MPa .
A2 400

6.6.7 TWO MORE NUMERICAL EXAMPLES CONCERNING


STATICALLY INDETERMINATE PROBLEMS

Example 1
The rigid bar OBC is supported as shown by two steel rods of cross-sectional area A=200
2
mm and modulus of elasticity E 2 ,1 10 5 MPa . Determine the stresses in the two rods caused by
the load P applied at C.
Considering the free body of the bar OBC and using statics we may write

M O
0 N 1 sin 1
N 2
sin 2
2 P 2 0 , (6.74)

Fig. 6.56
where
1, 5
1 atg 56 , 30 ;
1

1, 5
2 atg 36 ,86 .
2
Since we may write only a single equation from statics, the problem is statically indeterminate (with
two unknowns: N1 and N2). However, the geometry of the problem shows that (Fig. 6.56) the points

126
Axial loading

B and C displace to B' and C' respectively, due to the action of force P. From similar triangles we
may write:
1

BB ' 1 cos 90 1 1
CC ' 2 2 2

cos 90 2

1
sin 1 1 sin 1
1 2
2
sin 1
2 1 sin 2
. (6.75)
2 2 2
sin 1
2
sin 2

Equations (6.74) and (6.75) may be solved simultaneously for N1 and N2:

N1 12 , 035 k N ;

N 2
12 , 49 k N .

With these values, the corresponding stresses in the two rods are:

3
N1 12 , 035 10
1
60 MPa ;
A 200
3
N 2
12 , 49 10
2
62 MPa .
A 200

Example 2
Draw the axial force and stress diagrams for the steel member shown in Fig. 6.57, where a
0 , 3 mm clearance exists between the end of the member and the supporting point C before the
application of the loads. Assume that E 2 ,1 10 5 MPa .

Fig. 6.57

127
Strength of Materials

Due to the action of forces P1 and P2 the end C' of the member reaches the support C and
thus a reaction XC occurs at this support. This is why we shall have two unknown reactions (XB and
XC) and only a single equation of mechanical equilibrium which may be written from statics:

FX 0 X B
X C
250 125 375 kN (6.76)
The second necessary equation involving XB and XC will be determined from the condition that the
total deformation of the member corresponding to the final state (after the loads are applied) is
equal to . We write
BC B 1 1 2 2 3 3 C 0 , 3 mm . (6.77)
Thus,
X 600 X 250 600 X 250 600 X 250 125 600
B B B B
0 , 3 mm , (6.77')
EA 1 EA 1 EA 2
EA 2

where
2
70 2
A1 3848 , 45 mm
4
and
2
50 2
A2 1963 , 49 mm
4
are the cross-sectional areas of the portions B-2 and 2-C respectively.
Equations (6.76) and (6.77') may be solved simultaneously for XB and XC :
X B
317 , 413 k N ;

X C
57 , 587 k N .
We are now in the position to draw the axial force diagram. This has been done in Fig. 6.57.
Dividing the axial force (N) by the corresponding cross-sectional areas, we obtain the normal stress
diagram also sketched in Fig. 6.57.
PROBLEMS TO BE ASSIGNED

P6

P.6.1 A solid cylindrical rod and a solid rectangular rod are welded together at C as shown (Fig.
P.6.1). Draw the axial force diagram and determine the corresponding normal stresses for each
particular portion of the member and the total displacement of point B, knowing that the modulus of
5
elasticity for the two rods is E 2 ,1 10 MPa .

128
Axial loading

Fig. P.6.1

P.6.2 Determine the total deformations of the steel rods shown (Fig. P.6.2a,b), under the given
loads, if E 2 ,1 10 5 MPa .

a. b.
Fig. P.6.2
P.6.3 For the rod shown in the figure below:
a) draw the axial force diagram;
b) determine the required value of area A;
c) compute the displacement of point 2;
d) compute the total elongation of the rod.
The rod is made of steel with the allowable normal stress a
150 MPa and the modulus of
5
elasticity E 2 ,1 10 MPa . It is also known that a 250 mm and P 100 kN .

Fig. P.6.3

P.6.4 The two rigid bars O1BC and O2 DE are pin - connected at points O1 and O2, respectively.
The bars are supported by two rods 1 and 2 as shown in Fig. P.6.4. Knowing that a 0 , 5 m , the
allowable normal stress for the two rods a
150 MPa , the modulus of elasticity
5
E 2 ,1 10 and the applied force P = 50 kN , determine:
MPa
a) the axial forces (N1 and N2) in the two rods;
b) the required value of area A;
c) the elongations of the two rods 1 and 2 .

129
Strength of Materials

Fig. P.6.4

P.6.5 Compute the total elongation of a tapered rod subjected to an external load P as shown
in Fig. P.6.5.

Fig. P.6.5

P.6.6 A rigid bar OBC is supported by a vertical rod of uniform annular cross section as shown
(Fig. P.6.6). The vertical rod 1 is made of aluminum with E 0 , 7 10 5 MPa . For the 40 kN force
shown, determine the required values of the diameters d and D and the vertical displacement of
point C if the allowable stress of the material the rod is made of is a 120 MPa .

Fig. P.6.6

130
Axial loading
5
P.6.7 Two cylindrical rods, one made of steel (with ES 2 ,1 10 MPa ) and the other one made of
5
aluminum (with E A 0 , 7 10 MPa ) support a 60 kN force as shown (Fig. P.6.7). Knowing that
the normal stress allowable values of the two materials are a
180 MPa and steel

a A
, determine the required values of the diameters d1 and d2, and the vertical
110 MPa

displacement of point B.

Fig. P.6.7

5
P.6.8 Two cylindrical rods, one made of aluminum ( E A 0 , 7 10 MPa ) and the other one made of
5
brass ( E B 1, 05 10 ) are joined at point 1. The composite rod is fixed at one end (B) while a
MPa
0 , 3 mm gap exists between the other end and the right wall (C). Determine the value of force
P P 0 required to cancel the gap between the end 3 of the composite rod and the wall. If P 2 P0 :

Determine the reactions at B and C;


Draw the axial force diagram for the composite rod;
Draw the normal stresses diagram;
Determine the horizontal displacement of point 2.

Fig. P.6.8

P.6.9 A rigid bar OBCDF is supported as shown by three vertical cylindrical rods of diameters
d 1 10 mm , d 2 12 mm and d 3 16 mm . The rods are made of steel with a
150 MPa and
5
E 2 10 . Determine the maximum value of force P which may be safely applied at F and
MPa
the vertical displacement of points C and F.

131
Strength of Materials

Fig. P.6.9

P.6.10 A rod consisting of the two portions, BC and CF is fixed at its both ends and subjected to an
axial force P as shown (Fig. P.6.10). Knowing that D 2 d 60 mm , a 180 MPa and
E 2 ,1 10 MPa , determine the maximum value of force P which may be safely applied at C and
the horizontal displacement of point C.

Fig. P.6.10

P.6.11 A rigid bar BCDF is suspended from four identical wires as shown (Fig. P.6.11). Determine
the axial forces in the wires (N1, N2, N3 and N4) when a force P is applied at D.

Fig. P.6.11

P.6.12 The rigid bar OBCD is supported by two cylindrical rods of diameters d1 10 mm and
d2 16 mm respectively (Fig. P.6.12). The two rods are made of steel with a
160 MPa and
5
E 2 ,1 10 MPa . Determine the maximum value of force P which may be safely applied at D.

132
Axial loading

Fig. P.6.12

P.6.13 Two cylindrical rods, one of steel and the other one of aluminum are joined at C. The
composite rod is fixed at its both ends and then subjected to the loading shown. Knowing that
5 5
a Steel
180 MPa , a
120 MPa , E Steel 2 ,1 10 MPa , E A 0 , 7 10 MPa , P 10 kN and
A

, determine the reactions at B and F, draw the axial force diagram and determine the
0 ,5 m
required values of the diameters d1 and d2 of the composite rod.

Fig. P.6.13

P.6.14 A steel rod of circular cross section of diameter d is placed inside a steel tube of diameters
d1 and D1. The rod is shorter with 0 ,1 mm than the tube. A rigid plate is attached to the tube
and then an external axial force P is applied as shown (Fig. P.6.14). Knowing that
5
a
190 MPa , E 2 ,1 10 MPa , d 40 mm , d 1 50 mm , D 1 70 mm and 300 mm ,
Steel

determine the maximum value of force P which may be safely applied.

Fig. P.6.14

133
Strength of Materials

P.6.15 Four identical cylindrical rods of diameter d 10 mm and length 1m , made of steel
5
(with E 2 ,1 10 MPa and a
200 MPa have to be joined, while a 0 ,15 mm gap exists between
B and B (Fig. P.6.15). Knowing that 25

and 35

, determine the normal stress in each
rod when the gap is compensated.

Fig. P.6.15
P.6.16 Four identical members are used to support the load P 40 kN (Fig.P.6.16). Determine the

normal stress in each member when 20 , d 15 mm , 1m and E 2 10 5 MPa . Compute
the vertical displacement of point O.

Fig. P.6.16
P.6.17 A rigid bar OBC is supported by three steel rods as shown (Fig. P.6.17). Knowing that
5
d1 40 mm , d 2 30 mm , d 3 10 mm and E 2 ,1 10 MPa determine the normal stress in each
rod when an external force P 50 kN is applied at B.

134
Axial loading

Fig. P.6.17
P.6.18 A rod consisting of two cylindrical portions of distinct materials is fixed at its both ends as
shown (Fig. P.6.18). Determine the axial force in the rod and the corresponding normal stresses
induced in the two portions by a temperature rise of t .

Fig. P.6.18

P.6.19 Two horizontal rigid beams are suspended from three identical steel rods as shown (Fig.
P.6.19). Knowing that E 2 ,1 10 5 MPa , A 10 mm 2 and 1 m determine the maximum value of
force P, which may be safely applied at C and the vertical deflection of points B, C and D.

Fig. P.6.19

135
7. SHEARING STRESSES IN MEMBERS OF SMALL
CROSS SECTIONS

Bolts, pins, rivets and welds used within different machine components and
technical systems are considered pieces of small dimensions, providing certain
simplifying assumptions in Strength of Materials calculus. In many cases, these
elements are subjected to transverse forces, as shown in Fig. 7.1.

Fig. 7.1

If the distance e between the two shearing forces application points 1 and 2 is
too small, the bending moment at any cross section between the two points may be
neglected, and thus the single internal force in the member remaining the shearing
force T = P. The shearing force T determines the development of the shearing
stresses at any cross section of the member considered (between points 1 and 2).
Shearing stresses are commonly found in bolts, pins, rivets and welds used to
connect various structural members and machine components. Consider for example,
the two plates B and C, connected by a rivet as shown in Fig. 7.2.
Due to the action of the opposite
forces P, the cross section FG of the rivet
is sheared by a shearing force T equal to P.
The analysis of riveted, bolted and
welded connections involve so many
indeterminate factors that exact
computation solutions are impossible.
Nevertheless, by making certain
Fig. 7.2 simplifying assumptions, we can easily
obtain practical solutions.
Strength of Materials

7.1 SHEARING STRESS AND SHEARING STRAIN

Let us consider for example, a current cross section of a member subjected to


shear. The shearing force T may have a certain orientation in the section considered,
as shown in Fig.7.3. The shearing force T may be resolved into two components Tz
and Ty, acting along the axes Oz and Oy respectively. Due to the action of Tz and Ty
at the level of an arbitrary element of area dA the shearing stresses xz and xy develop.
For the entire cross section considered, the following relations may be written:

xy d A T y ;
A

(7.1)
xz d A T z .
A

The actual distribution of the


shearing stresses in the section
considered is therefore statically
indeterminate. However, since the rivets,
bolts and welds are in general
components of small dimensions, we may
assume that the shearing stresses are
Fig. 7.3
uniformly distributed over any cross
section.
This means that we may write:
xy constant ;
(7.2)
xz constant .

Relations (7.1) become therefore:


xy Ty ; Ty

dA A Ty ; xy ;
xy A

A
(7.3)
xz dA T z ; A Tz ;

Tz
xz
xz
,
A A
or, in general,
T
. (7.4)
A
The strength condition is therefore:

T
a , (7.5)
A

136
Shearing Stresses in Members of Small Cross Sections

where A is the cross-sectional area (associated with the area of the plane surface
shown in Fig. 7.3) and a the allowable value of the involved material shearing
stress.
In general:
a= (0,5 0,8) a. (7.6)
Relation (7.5) may be used for all three types of strength of materials characteristic
problems:
- dimensioning problems;
- checking problems;
- calculus of the allowable external loads.
It should be emphasized that the value obtained by using the relation (7.4) is an
average value of the shearing stress over the entire section. As we shall see later, the
actual value of the shearing stress varies
from zero at the surface of the member to
a maximum value max which may be
much larger than the average value.
Nevertheless, the accuracy obtained using
the relation (7.4) is sufficiently high
when the cross-sectional areas of the
members considered have low values.
The shearing strain is not of a great
importance in such cases. It consists in
fact in a relative displacement v of the
sheared cross sections located at distance
Fig. 7.4 e from each other, as shown in Fig.7.4.
If the material obeys the Hookes law, we write (Fig. 7.4):
v Te
tg v e e .
e G GA

Thus, the relative displacement of the sheared cross sections 1 and 2 is:
Te
v (7.6)
GA

where T is the shearing force, e is the distance between the applied external forces, G
is the shear modulus while A represents the cross-sectional area of the involved
member.

7.2 STRENGTH OF A SIMPLE RIVETED JOINT

The rivet shear is produced at those sections where it withstands the relative
displacement tendency of the joined elements, as shown in Fig.7.5. We may observe
that the shearing force at the rivet sheared cross section T equals P. The allowable
maximum applied force P is therefore:
137
Strength of Materials

2
d
T m ax P A s a a ,
4
where a represents the allowable
shearing stress of the rivet involved
material. If the rivet does join a number
of i + 1 elements, then i sheared
sections exist, with a cumulated area:
Fig. 7.5 d
2
As i . (7.8)
4

7.3 BEARING STRESS IN CONNECTIONS

A rivet (pin, bolt etc.) is also subjected to bearing stresses.


Consider for example the three
plates of thickness t connected by a
rivet as shown in Fig. 7.6.
Since there are two sheared
cross sections in this case, the rivet
is said to be in double shear.
The contact between the rivet
and the joined elements develops on
Fig. 7.6 half cylinder surfaces as shown in
Fig. 7.7.

Fig. 7.7 Fig. 7.8

Since the actual distribution of the contact forces and of the corresponding
stresses- is quite complicated (Fig. 7.8), one uses in practice an average nominal
value b of the stress, called the bearing stress, obtained by dividing the load by the
area of the rectangle representing the projection of the rivet on the plate section (Fig.
7.9)

138
Shearing Stresses in Members of Small Cross Sections

Since this area is equal to t d in the case of element 2 (for example) of Fig. 7.7 (Fig.
7.9), where t is the plate thickness and d the diameter of the rivet, we write:

P P
b . (7.9)
A t d

Fig. 7.9

This value has to be equal to or less than b a (the allowable value of the
involved material bearing stress). We write therefore:

P
b b a . (7.10)
td

In conclusion, a rivet has to be designed considering two important matters: the


shearing stress and the bearing stress. Both computed stresses must be equal to or
less than the allowable corresponding stress values, so that the involved structure
would not fail under the action of external loads.

7.4 WELDED CONNECTIONS

The reliability of welded connections has increased to the point where they are
used extensively to supplement or replace riveted or bolted connections in structural
and machine design. It is frequently more economical to fabricate a member by
welding simple component parts together than to use a complicated casting.
Welding is a method of joining metals by fusion. With heat from either an
electric arc or an oxyacetilene torch, the metal at the joint is melted and fuses with
additional metal from a welding rod. When cool, the weld material and the base
material form a continuous and almost homogeneous joint. To protect the weld from
excessive oxidation, a heavily coated welding rod is used that releases an inert gas
that envelopes the arc stream; this technique is called the shielded arc process.
The welded connections offer several important advantages:
- the connected members strength is not diminished through additional drilling

139
Strength of Materials

like the riveted or bolted connections do;


- the welded connections require a relatively simple technological process,
leading to a low price of the manufacturing;
- the maintenance technological process require a minimum effort.
Here are the main types of welds:
a) Butt welds (Fig. 7.10)
In such a case, the weld is subjected to
tension, the normal stress being
computed with the formula:
P
. (7.11)
(b 2t ) t

We note that a length b 2t is used


instead of b. This happens due to the
technological flaws, which usually occur
at the two ends of the weld. The strength
Fig. 7.10 condition is therefore:
P
a w , (7.12)
(b 2t ) t
where a w represents the allowable stress of the weld, usually taken to be:
a w 0 ,8 a

( a = the allowable stress of the base material).


b) Transverse fillet welds (Fig. 7.11)
The strength of transverse fillet
welds is assumed to be determined by
the shearing resistance of the weld
throat regardless of the direction of
the applied load.
In the 45 fillet weld (Fig. 7.11),
with the leg equal to t, the shearing
area through the throat is the length of
weld b times the throat depth.
Considering the real length of the
weld (as specified above) equal to b -
2a, the two welds shearing stress is
(Fig. 7.11):
P
w . (7.13)
2 a b 2 a
Fig. 7.11

140
Shearing Stresses in Members of Small Cross Sections

The strength condition may be written therefore as follows:


P
w a , (7.14)
2 a b 2 a w

where a represents the allowable shearing stress of the weld, usually taken to be:
w

a 0 , 65 a .
w

( a the allowable stress of the base material).


c) Side fillet welds (Fig. 7.12)

Fig. 7.12
Using the same reasoning we may write:
P
w a . (7.15)
2 a 2 a w

PROBLEMS TO BE ASSIGNED
P.7
P.7.1 Two steel plates are joined together by four rivets of 35 mm diameter as shown (Fig. P.7.1).
Knowing that the allowable values of the shearing and bearing stresses of the material the rivets
are made of are a = 70 MPa, b = 120 MPa respectively, and the allowable normal stress of
a

the plates is a = 160 MPa, determine the maximum value of force P which may be safely applied
at C.
P.7.2 A gusset plate is riveted to a larger plate by three rivets of 30 mm diameter as shown (Fig.
P.7.2). Determine the shearing stresses developed in the rivets.
P.7.3 Two plates are welded to resist a load P as shown (Fig. P.7.3). What maximum value of P can
be applied if a weld = 70 MPa and a plate = 160 MPa ?

P.7.4. Determine the maximum value of force P which may be applied to the welded plates shown
(Fig. P.7.4), if a plates = 160 MPa and a weld 0 , 8 a plates .

141
Strength of Materials

Fig. P.7.1 Fig. P.7.2

Fig. P.7.3 Fig. P.7.4

Fig. P.7.5 Fig. P.7.6

P.7.5 A load P is applied to a steel rod supported as shown (Fig. P.7.5). Knowing that a steel
=
190 MPa; a steel = 110 MPa and b a = 250 MPa determine the maximum value of force P
which may be safely applied to the steel rod.

P.7.6 Two wooden planks are joined by the glued joint shown (Fig. P.7.7). Determine the required
value of length so that a 6 kN load to be safely supported if a glue = 1,5 MPa.

142
8. BASIC ELEMENTS OF THE THEORY OF ELASTICITY

The examples of the previous sections were limited to members under axial
loading and connections under transverse loading. Most structural members and
machine components are under more involved loading conditions. Even in the case of
members under axial or transverse loading, the normal and shearing stresses on
planes which are not perpendicular to the axis of the involved members, obey
complex and difficult mathematical and physical matters. This is why, before going
further, we have to understand several important basic elements concerning the
theory of elasticity. They will help us to go deeper inside the investigated mechanical
elements or machine components and to find reasonable answers to the problems of
Strength of Materials.

8.1 PLANE STATE OF STRESS

Every elastic body is spatial and, in general, every external loads system is
spatial. Hence, strictly speaking, any strength of materials problem (or elasticity
problem) is a spatial problem. For its solution, we have to consider all the
components of stress, strain and displacement. However, if the body has a particular
shape and the external loads are distributed in a particular manner, we may consider
the spatial problem as a plane one and neglect some of the components.
This will greatly
simplify the mathematical
aspect of solution while the
results may still be applied in
engineering design with
sufficient accuracy.
Let us now consider a
thin plate of uniform
thickness, subjected to loads
acting within the plate middle
plane as shown in Fig. 8.1. In
Fig. 8.1 such a case the thin plate is
said to be in a plane stress
condition.
From the thin plate we isolate an infinitely small element in shape of a prism
(Fig. 8.2). For convenience the plate dimension (or that of the infinitely small
element considered) in the Oz direction is taken to be unity.
The problem which arises consists in the determination of stresses at the
arbitrary point O of the plate, on planes which are perpendicular to the plate middle
Strength of Materials

plane. In other words, we have to determine the normal and shearing stresses on
arbitrary planes BCBC inclined with angle with respect to the Oy axis, as
shown in Fig. 8.2. We assume the elementary area of the rectangular surface BCBC
to be equal to dA. While the stresses acting on surface BCBC differ slightly from the
stresses at O, the error involved is small and vanishes as sides OB and OC approach
zero.

Fig. 8.2

On the other hand, due to the action of the external loads P1, P2,...,Pn, normal
and shearing stresses develop on the faces OBOB, OCOC and BCBC of the
element shown in Fig. 8.2. These stresses are in fact a direct consequence of the
interaction between the element and the surrounding material of the plate. For
convenience, the element OBCOBC has been represented in Fig. 8.3 in a simplified
manner.
In Fig. 8.3, since the plane
BC approaches point O, the stress
acting on BC will become the
stress acting on the inclined plane
considered passing through the
point O. Now assuming to know
the values of stresses acting on the
planes OB and OC (i.e. y, yx, x,
xy ), we shall determine the
stresses and on the inclined
plane BC.
From the summation of
Fig. 8.3
moments about point D (Fig. 8.3)
we write:
BC BC
M D 0 yx d A sin cos xy d A cos sin 0 .
2 2

This leads to a relation already known:

xy yx .
144
Basic elements of the theory of elasticity

From the mechanical equilibrium of all forces projected on and directions


respectively we have:
F 0
d A x d A cos cos y d A sin sin xy d A cos sin yx d A sin cos 0 ;
F 0
d A x d A cos sin y d A sin cos xy d A cos cos yx d A sin sin 0 .

We may write therefore:


x cos 2 y sin 2 2 xy sin cos ;

(8.1)


x y sin cos xy cos sin .
2 2

Using the relations
2 1 cos 2 2 1 cos 2
sin ; cos ,
2 2
(8.1) becomes
1 cos 2 1 cos 2
x y xy sin 2 ;
2 2

x y sin 2 (8.1)
xy cos 2 .
2

or
x y
x
y
cos 2 xy sin 2 ;
2 2
(8.1)
x y
sin 2 xy cos 2 .

2

A simple analysis of relations (8.1) tells us that the normal and the shearing stresses
and on surface BCBC inclined with angle (and thus the normal and the shearing
stresses at point O about a surface parallel to BCBC) are functions of angle . One
could ask: which are the values of angle that make these quantities ( and )
maximum or minimum? These values may be found by differentiating (8.1) with
respect to and setting the derivatives equal to zero.

d x
y
2 sin 2 2 xy cos 2 0
dx 2

d x
y
2 sin 2 xy cos 2 0 . (8.2)

d 2

145
Strength of Materials

We find that
2 xy
tg 2 . (8.3)
x
y


Equation (8.3) gives us two values of (1 and 2 = 1+ ) for which the normal
2
stress has extreme values. This is why, the equation (8.3) is always written as

2 xy
tg 2 1 , 2 . (8.3)
x
y

In other words, the two values of (1 and 2) do give us two perpendicular


directions in the plane about which the normal stress has extreme values (a
maximum value 1 with respect to one direction and a minimum value 2 with respect
to the other one). The two extreme stresses 1 and 2 are thus perpendicular to each
other. They may be obtained by substituting from (8.3) into the expression of
from (8.1). This gives
x
x 2
y 1 2

1 y
4 xy ;
2 2
(8.4)
x

x
y 1 2 2
2 4 .

2 2
y xy

or
x
x 2
y 1 2
1 ,2 y
4 xy . (8.5)
2 2

The two extreme normal stresses are called principal stresses, while the
corresponding directions (given by 1 and 2) are called principal directions.
Returning now to the second equation of (8.1), differentiating it with respect
to and setting the derivative equal to zero, we have

d x y
2 cos 2 2 xy sin 2 0 . (8.6)
d 2

We thus find two perpendicular directions (3 and 4 = 3+ ) for which the shearing
2
stress has extreme values (a maximum value 3 with respect to one direction and a
minimum value 4 with respect to the other one). The two directions are given by
equation:
d y
x
0 tg 2 3 , 4 . (8.7)
d 2 xy

146
Basic elements of the theory of elasticity

Substituting these values of into the second equation of (8.1) we get to

x 2
1 2
3 ,4 y
4 xy , (8.8)
2
or
1
3 ,4 1 2 . (8.9)
2
Since
1
tg 2 1 , 2 , (8.10)
tg 2 3 , 4

the shearing stress reaches its extreme values on planes inclined at 45 with the
planes for which the normal stress reaches its extreme values 1 and 2.
It is to be noted that the extreme condition (8.2) coincides with

x
y
sin 2 xy cos 2 0 .
2

In other words, on planes coinciding with the principal directions, the shearing stress
is zero.
Returning now to the original thin plate of Fig. 8.2, we may facilitate the
visualization of the
stress condition at point
O by considering an
infinitely small cube of
different orientations,
centered at O and the
stresses exerted on each
of the lateral faces of the
cube, (Fig. 8.4). For a
certain orientation of the
cube B1B2B3B4 (given by
the two perpendicular
directions 1 and 2) the
Fig. 8.4 stresses acting on the
lateral faces of the cube
reduce to the principal normal stresses 1 and 2, while the corresponding shearing
stresses become zero (Fig. 8.4).
Rotating this cube by 45o within the plate plane, we get to the cube C1C2C3C4
on whose lateral faces act both the extreme values of the shearing stress
1 2
( max ) and the normal stresses denoted by 3 and 4. In other words, on
2

147
Strength of Materials

planes corresponding to the action of the extreme shearing stresses, the normal
stresses do not become zero.

8.2 SPATIAL STATE OF STRESS

Consider now a body subjected to several external loads P1, P2,,Pi, Pn in


mechanical equilibrium (Fig. 8.5).
To understand the stress
condition created by these loads
at some arbitrary point O
within the body, we isolate an
infinitely small element in form
of a parallelepiped, with its
edges parallel to the coordinate
axes and having the lengths dx,
dy and dz (Fig. 8.6a). We
sketch this element at a larger
scale (Fig. 8.6b).
Generally speaking, the
Fig. 8.5
stress components acting on the
element faces are functions of x, y and z. Thus, the stress components on a pair of
parallel faces are not equal but differ by a differential quantity.

a. b.
Fig. 8.6

For instance, if the average normal stress component on a face is x, then, that on the
parallel face, due to the variation of x, will be

x
x dx .
x

148
Basic elements of the theory of elasticity

Applying the same reasoning like in the


preceding section, one could find a certain
spatial orientation of the element considered, so
that the shearing stresses on the element faces to
be zero and, hence, the corresponding normal
stresses to become the principal stresses 1, 2
and 3 (Fig. 8.7).
One could demonstrate that the principal
stresses 1, 2 and 3 (where 1 2 3 ) may
be computed as the real roods of equation
Fig. 8.7

3 2
I 1 I 2 I3 0 , (8.11)

I1 x
y
z
;



2 2 2

I2 x y
x z
y z
xy xz yz ;


(8.12)


x
xy xz

I 3 yx y
yz .


zx zy z

We get to this result letting dx, dy and dz approach zero (Fig. 8.6b), so that the
element in form of a parallelepiped is contracted to point O.
Using the same analogy we may also reach to the expressions of the extreme
shearing stresses:

1 2 1 3 2 3 (8.13)
1 ,2 ; 1 ,3 ; 2 ,3 .
2 2 2

These shearing stresses correspond to the planes passing through each of the principal
axes of stress and bisecting the right angle between the other two principal axes of
stress. Thus we see that:

the magnitude of the largest shearing stress is equal to half the difference
between the largest and the smallest principal stresses;

the largest shearing stress acts on the plane bisecting the angle between these
two principal stresses and passing through the line of action of the third
principal stress.

149
Strength of Materials

8.3 MONOAXIAL (UNIAXIAL) STATE OF STRESS

With the above presented concepts in mind, we may now return to the
monoaxial state of stress. This is in fact a plane state of stress for which
y
0 and xy yx 0 .
Let us consider again the case of a member under axial loading (Fig. 8.8).

a. b.
Fig. 8.8
From this member we detach an infinitely small element in shape of a prism and, for
convenience, we represent it in plane, separately (Fig. 8.8b). This element is in fact
the one represented in Fig. 8.3, making y = yx = xy= 0. We note therefore that the
conditions of stress in the member may be described as shown in Fig. 8.8b. The only
stresses exerted on the face OC of the prism (which is perpendicular to the x axis) are
the normal stresses x. However, on planes inclined with an angle , normal and
shearing stresses (, ) develop. We thus conclude that the same loading condition
may lead to different interpretations of the stress condition at a given point,
depending upon the orientation of the plane BC (Fig. 8.8b). Using equation (8.3)
2 xy
tg 2 1,2 ,
x
y

with xy y 0 , we find that



1 0 ; 2 1 .
2 2

In other words, the two principal directions are parallel to the axes Ox and Oy. The
corresponding principal stresses 1 and 2 are (Eq. 8.5):

x 4 xy2
x y 1 x 1
1 ,2 y x ,
2 2 2 2

which finally gives:


1 x ;
(8.14)
2 0 .

150
Basic elements of the theory of elasticity

In the same time, using equation (8.9), the maximum shearing stress is given by
1 2 x
max , (8.15)
2 2
acting on a plane inclined at 45o with Oy axis ( = 45o, Fig. 8.8).

8.4 PURE SHEAR STATE

Let us assume that a pure state of stress exists at an arbitrary point O of a body
subjected to several external loads in mechanical equilibrium. In such a case x= y =
0, the pure state of stress being
defined by the stress components xy=
yx as shown in Fig. 8.9. Using
equation (8.3)
2 xy
tg 2 1 , 2 ,
x
y

with x = y= 0, we find that



;
1 4

3 .
2 4 2 4
Fig. 8.9
The principal stresses are given by:


x 2
x y 1 2
1 ,2 y
4 xy xy .
2 2

In other words, this means that a pure shear state is equivalent to a state of
stress consisting in a tensile stress 1 and a compressive stress 2 of the same
magnitude (1=xy; 2 = -xy). The principal stresses 1 and 2 act on planes inclined at
45o and 135o with Ox axis as shown in Fig. 8.9.

8.5 GENERALIZED HOOKES LAW

First of all we have to recall that, for a member in tension or compression


which undergoes small deformations, involving only the straight-line portion of the
corresponding stress-strain diagram, the Hookes law may be written in its simple
form as follows:

E . (8.16)

151
Strength of Materials

As mentioned in the preceding sections, E is called the modulus of elasticity of the


material involved. In other words, the validity of such a law means that the stress is
directly proportional to the strain .
In the same manner, for values of the shearing stress () which do not exceed
the proportional limit in shear (in case of members subjected to shear) the Hookes
law for shearing stress and strain may be written as

G , (8.17)

where G is called the shear modulus of the material involved.


If a rectangular coordinate system Oxyz is being used, the quantities in the
above two relations may become x, y, z, x, y, z, xy, xz, yz, xy, xz, yz and

x E x ;

E ;
y y

E z ;
z
(8.18)
G ;
xy xy

G ;
xz xz

yz G yz
.

Let us now consider a body (made of a homogeneous isotropic material)


subjected to several external loads in mechanical equilibrium P1, P2,, Pk, Pn (Fig.
8.10a).

a. b.
Fig. 8.10

We assume that the most general state of stress occurs within the body, due to
the action of the external loads.

152
Basic elements of the theory of elasticity

This means that, if we isolate


an elementary parallelepiped from
the body, with its edges parallel to
the coordinate axes, normal and
shearing stresses develop on each
face of such a parallelepiped (Fig.
8.10b). For convenience, the
variation of stresses from one face to
the other parallel face has been
neglected.
Let us now consider first only
the effect of the stress component x
Fig. 8.11
(Fig. 8.11).
x
We recall that x causes a normal strain x in the Ox direction and strains equal
E
x
to in each of the Oy and Oz directions. We may write therefore:
E

action of x
x ; y ; z
x x x
x x . (8.19)
E E E


Similarly, the stress component y, if applied separately will cause a strain y
in the
E


y
Oy direction and strains in the other two directions. Finally, the stress
E

component z causes a strain z
in the z direction and strains z
in the Ox and
E E
Oy directions. We write

action of y y y y
x y ; y ; z ; (8.20)
E E E

action of z
x z z
; y
z
; z
z
. (8.21)
E E E

Combining the results obtained, we may write that the components of strain
corresponding to a multiaxial loading of the parallelepiped involved (the state in
which only the normal stresses x, y and z act, being all different from zero is
referred to as a multiaxial loading Fig. 8.12) are

153
Strength of Materials

x y z
x x x x
1
x y z ;
E E E E

x y

y
y
y

y
z

1
y x z ; (8.22)
E E E E
x y

z
z
z

z
z
1
z x y .
E E E E

These relations are referred to as the


generalized Hookes law. They are valid
only as long as the stresses do not exceed
the proportional limit, and as long as the
deformations involved remain small.
It is important to mention again that
a positive value for a normal stress
component signifies tension, while a
negative value signifies compression.
We are now in the position to
complete the generalized Hookes law
corresponding to the state of stress of the
parallelepiped sketched in Fig. 8.10b, by
Fig. 8.12 adding the shearing strains xy, xz, and yz
as follows (Fig. 8.13):
xy yz
xy
; xz

xz
; yz
. (8.23)
G G G

We conclude therefore that relations (8.22) and (8.23) represent the complete
form of the generalized Hookes law for a homogeneous isotropic material.
For a plane state of stress (z = 0, xz= yz = 0) the generalized Hookes law
becomes:
;
1
x
x
y
E

;
1
y
y
x
E
(8.24)
1
z
x
;
E
xy
xy
.
G

A simple examination of the relations (8.22) and (8.23) might lead us to


believe that three distinct constants E, G and must first be determined
experimentally, if we are to predict the deformation caused in a given material by an
arbitrary combination of stresses. Actually, only two of these constants need to be
154
Basic elements of the theory of elasticity

determined experimentally for any given material. But this will be discussed later,
when a relation among E, G and will be found.

Fig. 8.13

8.6 STRAIN ENERGY

Consider a solid deformable body subjected to several external loads P1, P2,,
Pk, Pn in mechanical equilibrium (Fig. 8.14). If the loads applied increase slowly from

155
Strength of Materials

zero up to their nominal values, the loads application points A, B, K, N displace to


different positions A', B', K', N' respectively, an external work L being done in this
way (Fig. 8.14).

Fig. 8.14

The work done by the applied loads P1, P2,,Pk, Pn must result in the increase of
some energy associated with the deformation of the body. This energy is referred to
as the strain energy of the body involved. We denote this energy by U. If the
deformations of the body do not exceed the elastic limit of the material, then, the
storaged strain energy U will be completely released to the surrounding environment
when the external loads P1, P2,, Pk, Pn are removed. In such a case U is called the
elastic strain energy. On the other hand, we may consider that the whole performed
external work is completely transformed into strain energy. We write therefore
L U . (8.25)

8.6.1 ELASTIC STRAIN ENERGY UNDER AXIAL LOADING

Consider a straight rod BC of length , fixed at one end and subjected to an


axial external force P as shown in Fig.
8.15. We assume that the rod material
obeys the Hookes law and the
proportional limit is not exceeded. We
do also assume that the force P is
applied statically (P does slowly
Fig. 8.15
increase from zero up to its nominal
value, as shown in Fig. 8.16).
Since the material does not exceed the proportional limit, during the
application of force P, the axial force N in the rod will increase directly proportional
to the horizontal displacement u of point C as shown in Fig. 8.17a.

156
Basic elements of the theory of elasticity

When N reaches the value P, the


displacement u reaches the final value
of the rod elongation . In the same
manner the normal stress in the rod,
during the application of force P, does
slowly increase and directly
Fig. 8.16 proportional to the normal strain , as
shown in Fig. 8.17b.

a. b.

Fig. 8.17

The work done by force P, as the rod elongates is equal to the hachured area
located under the force deformation diagram (Fig. 8.17a). We may write therefore:

P
L
N du
2
. (8.26)
0

The elastic strain energy under axial loading is


P A A A
U L V , (8.27)
2 2 2 2 2

where V A represents the volume of the rod. Since the Hookes law is valid
( E ), we may also write

2

U V V V . (8.28)
2 2 E 2E
Work and energy are expressed in the same units, obtained by multiplying
units of length by units of force. Thus, if the International System metric units are
used, work and energy are expressed in N m , this unit being called a joule (J).
Returning now to the relations (8.27) and (8.28), we observe that the strain
energy depends upon the dimensions of the rod involved. To eliminate the effect of
size and to direct our attention to the properties of the material we shall define the
concept of elastic strain energy per unit volume. This quantity is referred to as the
elastic strain-energy density and is denoted by U . We write therefore D

2
U
U D . (8.29)
V 2 2E

157
Strength of Materials

In this way the elastic strain energy absorbed by an infinitely small element of
material of volume dV is
2

dU U D dV dV dV . (8.30)
2 2E

Thus, the total strain energy Ut of a body under axial loading is

2

U t
U D dV
2
dV
2E dV . (8.31)
V V V

8.6.2 SHEARING STRESSES ELASTIC STRAIN ENERGY

If a material is subjected to plane shearing stresses (Fig. 8.18) the elastic strain
energy may be computed in a similar manner.

Fig. 8.18

For such a case, the elastic strain energy density may be written therefore as

2

U D . (8.32)
2 2G

Thus, the total elastic strain energy Ut of a body under shearing stresses is

2

U t
2
dV
2G dV . (8.33)
V V

where V is the volume of the body involved.

8.6.3 STRAIN ENERGY FOR A GENERAL STATE OF STRESS

In the preceding sections we have determined the expression of the elastic


strain energy of a body under normal and shearing stresses. Let us now consider a
body subjected to several external loads in mechanical equilibrium (Fig. 8.19a).
158
Basic elements of the theory of elasticity

We do also assume that the most general state of stress develops in the body due to
the action of the external loads. Now we shall isolate an elementary parallelepiped
from the body (Fig. 8.19b). In such a case the general state of stress is characterized

a. b.
Fig. 8.19

by the six stress components x, y, z, xy, xz, and yz (for convenience, in Fig. 8.19b,
the stresses have been represented only on the visible faces of the elementary
parallelepiped considered). If the body behaves linear elastically, the elastic-strain
energy density for a general state of stress may be obtained by adding the expressions
given within the preceding sections. In this way, the elastic strain-energy density may
be expressed as follows:
x y y z xy
xy yz
yz
(8.34)
x z xz xz
U D
.
2 2 2 2 2 2
Recalling the expressions representing the Hookes law for a homogeneous,
elastic and isotropic body:


;
1

x
E
x y z



;
1
y
y
x
z
E

.
1

z z x y
E

xy yz
xy
; xz

xz
; yz

G G G

and substituting for the strain components x, y, z, xy, xz, yz into (8.34), we obtain
the elastic strain-energy density as follows:

159
Strength of Materials

U D

1
2
x

2
y

2
z
x
y
x z
y z

1
2
xy

2
xz

2
yz
. (8.35)
2E E 2G
If the principal axes are used as coordinate axes, the shearing stresses become
zero and (8.35) reduces to
1
U D
2
1

2
2

2
3
1 2
1 3
2 3
. (8.36)
2E E
Thus, the total elastic strain energy (Ut) of a body under the most general stress
condition may be written as
U t

U D dV , (8.37)
V

where V is the volume of the body involved and UD the elastic strain-energy density
mentioned above.

8.6.4 ELASTIC STRAIN-ENERGY DENSITY ASSOCIATED WITH A


CHANGE IN VOLUME. ELASTIC STRAIN-ENERGY DENSITY
ASSOCIATED WITH A DISTORTION (A CHANGE IN SHAPE)

Due to the action of the external loads a solid body changes both its volume and
its shape. In this way we may separate the elastic-strain energy density at a given

a. b. c.
Fig. 8.20

point of the body into two components:


a component Uv associated with a change in volume of the material at that
point;
a component US associated with a change in shape (a distortion) of the material
at the same point.
A given state of stress may be obtained by superposing two states of stress as shown
in Fig. 8.20, where 1, 2, 3 are the principal stresses and ave (or ) is the average
value of the principal stresses:
1
ave

2 3
. (8.38)
3

160
Basic elements of the theory of elasticity

The state of stress described in Fig. 8.20b tends to change the volume of the element,
but not its shape, since all the faces of the element are subjected to the same stress
ave (or ). The state of stress described in Fig. 8.20c tends to change the shape of
the element.
Recalling (8.36), the elastic-strain energy density associated with the state of
stress described in Fig. 8.20b may be written as

UV
1
1
2

2
2

2
3
1 2
1 3
2 3

2E E


1
1
2
2
2
3
2

2E E

3 1 2
2
3 3
2 2 2
3
1 2 1 2 3

2E E 2E 2E 3

1 2
1 .
2
2
3 (8.39)
6E

In the same manner, the elastic strain-energy density associated with the state of
stress described in Fig. 8.20c, may be written as:

U S
U D
UV
1
1
2

2
2

2
3
1 2
1 3
2 3

1 2
1
2

2
2

2
3
,
2E E 6E

which finally gives

1
U S
1
2

2
1
3

2
2
3

2
(8.40)
6E

or

1
U S
2
1

2
2

2
3
1 2
1 3
2 3
. (8.41)
3E

8.6.5 RELATION AMONG E, G AND


Consider a rectangular plate of uniform thickness, made of a homogeneous and
isotropic material (Fig. 8.21). The plate is assumed to be in a plane stress condition,
being subjected to tension about Ox axis ( x 0 ) and to compression about Oy
axis ( y 0 ) as shown in the figure.

161
Strength of Materials

Recalling now the


pure shear state properties
described in section 8.4, the
stress condition mentioned
above is equivalent to a
pure shear state at 45o, as
shown in Fig. 8.21.
Since the elastic-
strain energy densities
associated with the two
states of stress have the
same value, we may write
U I
U II
, (8.42)
Fig. 8.21 D D

where


1
U D I

2
1

2
2

2
3
1 2
1 3
2 3

2E E



1
2
1

2
2
1 2

1
0
2
0

2
0 0
2E E 2E E


2 2 2


0

0

0
1 ,
E E E
and (8.35)

2

2


1 1 xy
x y
2 2 2 2 2 2 0
U D II x y z x y z z xy xz yz
.
2E E 2G 2G 2G

We may write therefore:


2 2
0
1 0
.
E 2G

It thus follows that


E
G . (8.43)
2 1

This relation may be used to determine one of the constants E, G or from the other
two.

162
9. TORSION

Members in torsion are encountered in many engineering applications, when


there is necessary to transmit a twisting couple Mt from one plane to another parallel
plane (Fig. 9.1). The twisting couples are also called torques.

Fig. 9.1

These couples (Fig. 9.1) have a common magnitude and opposite senses. A
case in which torsion is dominant is provided by the transmission shafts, which are
used to transmit power from one point to another, as from a motor to a machine tool,
from an engine to the rear axle of an automobile or from a steam turbine to an
electric generator, etc. (Fig. 9.2).

Fig. 9.2

These shafts may either be solid (Fig. 9.3a) or they may be hollow (Fig. 9.3b).
Strength of Materials

a. b.
Fig. 9.3

If we know the power P which is to be transmitted through the shaft


(expressed in horsepowers - HP) and the frequency of the rotation n (number of
revolutions per minute - rpm ), one could demonstrate that the twisting couple (or
torque) the shaft is subjected to is given by the formula
P
M t 7 , 02 kN m . (9.1)
n

If we express the power in kW (kilowatts), formula (9.1) becomes:


P
M t 9 , 55 kN m . (9.2)
n

9.1 TORSION OF CIRCULAR MEMBERS

9.1.1 STRESSES AND STRAINS


Consider a circular shaft AB of diameter d, subjected to equal and opposite
torques Mt , as shown in Fig. 9.4. We pass a section perpendicular to the axis of the
shaft through some arbitrary point C.

a. b.
Fig. 9.4

The cross section at C will be subjected to elementary shearing forces dF acting


perpendicular to the radius r of the shaft. The elementary shearing force dF
developed at the level of an elementary area dA of the cross-section considered, is
given by the shearing stress at that level multiplied by the area dA (Fig. 9.4b)
164
Torsion

Expressing that the sum of the moments given by the shearing forces dF with
respect to the axis of the shaft is equal in magnitude to the externally applied torque
Mt, we have:
r dF M t, (9.3)
A

where A is the cross-sectional area. Since dF = dA, (9.3) becomes:

r dA M t. (9.4)
A

While the relation (9.4) gives us an important condition which has to be


satisfied by the shearing stresses at any given cross section of the shaft, it does not
tell us how these stresses are distributed over the cross section. This means that the
problem is statically indeterminate, i.e., the distribution of stresses cannot be
determined from statics alone. The integral of (9.4) may be solved only if the
shearing stress distribution law is found.
In deriving the torsion formulas, we make the following assumptions:
- the member in torsion has a constant cross-section;
- the member material is continuous, homogeneous and isotropic;
- Hookes law is available;

a. b.
Fig. 9.5
- the member is loaded by twisting couples in planes that are perpendicular to the
axis of the member;
- stresses do not exceed the proportional limit.
Under the above mentioned assumptions, let us now consider a circular
member subjected to torsion (Fig. 9.5). The deformation in the circular member
subjected to torsion may be easily illustrated if we trace several parallel circles and
straight generatrix lines on the surface of the circular member as shown in Fig. 9.5.
When the circular member represented in Fig. 9.5a is subjected to torsion it will

165
Strength of Materials

deform (Fig. 9.5b). Analysing the mode in which the circular member considered
does deform under the specified loading, we note that:
- when a circular member is subjected to torsion, every cross section remains plane
and undistorted;
- a straight radial line in the section (for example O1A) remains straight after
deformation and rotates with a certain angle ;
- the distance between two cross-sections remains unchanged after deformation.
This means that x = 0, so that, applying the Hooke`s law, we have:

x 0. (9.5)

- due to the action of Mt the rectangle abcd (Fig. 9.5a) changes into a parallelogram
(a`b`c`d` - Fig. 9.5b), the original right angles of abcd modifying with the
quantity (the shearing strain).
It follows therefore that shearing stresses develop on the faces of the
elementary volume a`b`c`d`a``b``c``d`` (Fig. 9.5b). Let us consider now a point C
located on the circumference of a given cross section of the circular member in
torsion (Fig. 9.6).

Fig. 9.6
As mentioned above, at the level of point C, a shearing stress develops, contained
within the cross-sectional plane. We do also assume that has a certain orientation
within the cross-sectional plane (Fig. 9.6). But may be resolved into two
components ( 1 and 2), respectively tangential to the circumference at C and radial.
On the other hand we know that, at a given point, shear cannot take place in one
plane only; an equal shearing stress must be exerted on another plane perpendicular
to the first one. This means that the shearing stress component 2 will be
accompanied by an equal shearing stress 2` acting along Ox direction and located on
the external surface of the member (Fig. 9.6). Since this surface is not loaded, 2` = 0
and, therefore, 2 = 2` = 0.
166
Torsion

In other words, the shearing


stress at point C will be directed along
the tangent to the circumference at C,
perpendicular to the radius OC. We
assume that this property remains
available for any other point of the
cross-section (Fig. 9.7).
We shall now determine the
distribution of the shearing stresses in a
given cross section of a circular shaft
Fig. 9.7 (Fig. 9.8a).
The shaft of diameter d and length , is attached to a fixed support at one end.
When the torque Mt is applied at section O, the shaft will twist with an angle called
the angle of twist.

a. b.
Fig. 9.8

Let us now detach from the shaft an element of length dx, located between two cross
sections of the shaft (I and II Fig. 9.8a). We have sketched this element separately
(Fig. 9.8b).The element considered is also subjected to torsion by the same torque
Mt, the two ends of the element rotating reciprocally with an infinitely small angle of
twist d . For convenience, we shall consider that the end II is fixed while the end I
rotates with angle d . In this way, an arbitrary point C of the end I will displace to
C``, while a point B, located on the same radius, will displace to B``, (Fig. 9.8b).
We may write therefore:
C C `` O 1C d d
tg r , (9.6)
C C` dx dx

where is the corresponding shearing strain (Fig. 9.8b).


167
Strength of Materials

d
Denoting by (the angle of twist per unit length) the quantity , the shearing
dx
strain may be expressed as follows:
r , (9.7)
where r is the distance between the arbitrary investigated point C and the axis x of
the shaft.
Since Hooke`s law applies, we have

G G r , (9.8)
where G is the shear modulus.
Recalling relation (9.4), we may write:
2
M t r dA G r r dA G r dA . (9.9)
A A A

Since G and are constant, we have:


2
M t G r dA G I p (9.10)
A

or
M t
, (9.11)
GI p

4
d
where Ip is the polar moment of inertia of the shaft cross-section I p . With
32

this expression of , relation (9.8) becomes:


M t M t
G r G r r. (9.12)
GI p I p

Equation (9.12) is called the elastic


torsion formula. It shows a linear
distribution of the shearing stresses over an
arbitrary cross-section of a shaft in torsion,
(Fig. 9.9).
Fig. 9.9 shows the shearing stress
distribution in a solid circular shaft for the
cross-sectional points located on Oy axis.
Due to the symmetry of the cross section,
we shall find the same stress distribution
for any other cross-sectional points located
on the same diameter. The shearing stress
at the level of an arbitrary point A of the
Fig. 9.9 shaft cross-section is therefore (Fig. 9.9):

168
Torsion

M t
A r.
I p

Since the shearing stress in the shaft varies linearly with the distance r from the axis
of the shaft, the maximum shearing stress develops on the circumferencial points of
the shaft cross-section. We write therefore:
M t M t M t d
m ax r m ax R . (9.13)
I p I p I p 2

The maximum shearing stress may be also written as


M t M t
m ax , (9.14)
I p W p

rm ax

where Wp is called the polar strength modulus of the involved cross-section. For a
circular cross section of diameter d, the polar strength modulus is
4
d
3
I d
p 32
W p . (9.15)
rm ax d 16
2
To determine whether the shaft may support the applied external torque Mt,
we must compare the maximum value of the shearing stress given by relation (9.14)
with the maximum value of the stress which may be safely applied to the material the
shaft is made of. The strength condition is therefore
M t
m ax a , (9.16)
W p

where ais the maximum allowable shearing stress in the type of the material the
involved shaft is made of. Relation (9.16) applies for circular members in torsion and
it may be used for all specific strength of materials problems.
Even if the formula (9.16) has been derived for a shaft of uniform circular
cross section, it may be also used for a shaft of variable cross section (Fig. 9.10a) or
for a shaft subjected to torques at locations other than its ends (Fig. 9.10b).

a. b.
Fig. 9.10
169
Strength of Materials

Let us now return for a while to the concept of angle of twist, to derive a
relation between this angle and the torque exerted on the shaft (Fig. 9.8a). We recall
that the angle of twist per unit length is given by (9.11):
M t
.
G I p

On the other hand, this quantity has been defined as

d
,
d x

where is the angle of twist. We may write therefore


M t
d d x d x ,
G I p

which gives

M t
d x, (9.17)
G I p
0

where is the length of the shaft involved.


Since the circular shaft shown in Fig. 9.8 has a uniform cross-section
4
d
(I p constant) and is subjected to a constant torque equal to Mt (Mt =
32
constant along the shaft), relation (9.17) becomes:

M t M t
d x (expressed in radians). (9.17')
GI p GI p
0

The relation obtained shows that,


within the elastic range, the angle of twist
is proportional to the torque Mt applied
to the shaft and to the length of the shaft.
This relation (9.17') may be used only if
the material of the shaft is homogeneous
(G = constant), the shaft is loaded only at
its ends and has a uniform cross section. If
the shaft consist of several portions with
various values of the cross sections
diameters and /or is subjected to torque at
locations other than its ends (Fig. 9.11.),
Fig. 9.11 we must divide the shaft into component

170
Torsion

parts which satisfy individually the required conditions of relation (9.17').


For the shaft represented in Fig. 9.11, we may write therefore:

M 0 1 M 0 2M 0 2 M 0 2M 0 3
A D A B B C C D
G I p1 G I p2 G I p3

M 0 1 3M 0 2 3M 0 3
,
GI p1 GI p2 GI p3

where:
4 4 4
d1 d2 d3
I p1 ; I p2 ; I p3 .
32 32 32

In other words, the total angle of twist of the shaft represented in Fig. 9.11, i.e. the
angle through which end A rotates with respect to end B, is obtained by adding
algebraically the angles of twist of each component part.
In case of a shaft with a variable
circular cross section (Fig.
9.12a), relation (9.17') may be
applied only to an element of
infinitely small length dx of the
shaft. Since such an element has
an infinitely small length (dx) we
may neglect the variation of its
cross section from one end to the
other. (Fig. 9.12). This means
that we may apply the formula
(9.17') for this element, as
follows:
M 0 dx
a. b. d , (9.18)
G I p (x)

Fig. 9.12 where:

- d is the infinitely small angle of twist representing the angle through which end
A of the shaft element of Fig. 9.12b rotates with respect to the end B;

- Ip(x) is the polar moment of inertia corresponding to the diameter d(x) of the cross
section of the shaft, located at distance x from the end of the shaft (Fig. 9.12a). In
other words, Ip is a function of x:
4
d x
I p I p x .
32

171
Strength of Materials

Now integrating (9.18) in x, from 0 to , we obtain the total angle of twist of the
shaft:

M 0
dx (9.19)
G I p (x)
0

or, in general:

M t
dx . (9.20)
G I p (x)
0

Formula (9.20) does also apply when Mt has a continuous variation along a circular
member in torsion (Mt = Mt(x), Fig. 9.13a) or when both Mt and Ip vary along the
member involved (Fig. 9.13b).

a. b.
Fig.9.13

Since the shearing stresses over the solid cross sections of circular shafts in torsion
have a linear distribution, in many cases one chooses hollow cross sections, thus
reducing the own weight of the shafts (Fig. 9.14a).
The distribution of the shearing stresses in a hollow cylindrical shaft of inner
diameter d and outer diameter D has been represented in Fig. 9.14b. In such cases,
the maximum shearing stress does also develop on the circumferential points of the
shaft cross section. We may apply therefore the formula (9.14):
M t
m ax ,
W p

where the polar strength modulus Wp is given by:

172
Torsion

a. b.
Fig. 9.14

4 4
4 4 D d
D d 1
I I 32 D
p p 32 32
W p
r m ax D D D
2 2 2

3 4 3
D d D 4
1 1 m , (9.21)
16 D 16

where m expresses the ratio d / D.

Important remarks

Since the shearing stress cannot take place in one plane only (section 4.3), in case
of a circular shaft in torsion, a shearing stress must be also exerted on other plane,
perpendicular to the first one. This is why the shearing stresses over the shaft cross
sections in torsion are always accompanied by equal shearing stresses directed
along the shaft (Fig. 9.15). These longitudinal shearing stresses explain the
specific longitudinal failure of members in torsion, made of different materials,
whose strength is much smaller about longitudinal direction than about normal
direction (normal to the axis of the member). This is for example the case of rods
made of wood, whose failure takes place longitudinally, along the fibres of the
material (Fig. 9.16).

Consider now an element a located on the surface of a circular shaft subjected to


torsion (Fig. 9.17). As presented above, on the faces of such an element, the only
stresses which develop are the shearing stresses max = .

173
Strength of Materials

Fig. 9.15 Fig. 9.16

The element is said to be in pure


shear. Another similar element b,
located on the same surface but
rotated with 45 is subjected to a
tensile stress ( 1 = ) over two of its
faces and to a compressive stress ( 2
= - ) on the other two faces (see the
previous chapter). In general, the
ductile materials fail in shear.
Fig.9.17
Therefore, when subjected to torsion, a member made of a ductile material breaks
along a plane perpendicular to its longitudinal axis (Fig. 9.18a). On the other hand,
brittle materials are weaker in tension than in shear. Thus, when subjected to torsion

a.

Fig. 9.19
a member made of a brittle material
tends to break along surfaces which are
b. perpendicular to the direction about
which tension is maximum ( 1Fig.
9.18b), i.e. along surfaces forming a
45 angle with the axis of the member.
If a circular shaft subjected to
c. torsion presents an abrupt change in
the diameter of its cross section, stress
concentrations occur within the
Fig. 9.18 immediate vicinity of the discontinuity
(Fig. 9.19).
174
Torsion

In such cases, the shearing stresses corresponding to the discontinuity area may be
NOT computed with the above presented formulas. The maximum shearing stress
corresponding to the portion of the shaft with diameter d (and for the immediate
vicinity of the discontinuity) - Fig. 9.19, is given by:

max k n , (9.22)

where n is the nominal value of the maximum shearing stress for the portion
involved (computed with formula (9.14)), and k is a stress concentration factor.
The stress-concentration factor k is usually given in the form of tables or graphs, as
shown in Fig. 9.20.

Fig.9.20

9.1.2 STATICALLY INDETERMINATE SHAFTS

The statically indeterminate problems of torsion represent that kind of problems


in which the internal torques cannot be determined from statics alone (i.e. the
equations of static equilibrium are not sufficient for a solution). In such cases, the
equilibrium equations must be complemented by relations involving the
deformations of the shaft and obtained by considering the geometry of the problem.
The following examples will show how to analyze statically indeterminate shafts.

175
Strength of Materials

Example 1
A shaft AB is attached to fixed supports at both ends and subjected to the torques shown (Fig. 9.21).
Knowing that the entire shaft is made of steel for which G = 8 104 MPa and that M0 = 500 N m, d
= 30 mm and = 500 mm, determine the maximum value of the shearing stress in the shaft ( max)
and the angle of twist of section 1.

Fig.9.21

Due to the action of the two concentrated twisting couples (M0 and 3M0), two reactions develop at
the supports (i.e. the torques MA and MB). The single equilibrium equation which may be written
from statics is:

M 0 M A M 0 3M 0 M B 0 M A M B 2M 0 . (9.23)
t

Since this equation is not sufficient to determine the two unknown torques MA and MB , the problem
is statically indeterminate. However, MA and MB may be determined if we note that the total angle of
twist of shaft AB must be zero, since both of its ends are restrained. We write therefore:

A B 0 A 1 1 2 2 B

M A 2 (M A M 0) (M A M 0 3M 0 )
0, (9.24)
GI p GI p GI p
1 2 2

where:
4 4 4
d (2d ) d
I p and Ip 16 16 I p .
1 32 2 32 32 1

176
Torsion

Thus, (9.24) becomes:

M A 2 (M A M 0) (M A M 0)
0,
G I p G 16 I p G 16 I p
1 1 1
or:

M A M 0 M A M 0
2M A 0
16 16
which finally gives:
1
M A M 0 . (9.25)
34

Substituting this value into the original equilibrium equation, we have:

67
M B M 0. (9.26)
34

With these values, we are now in the position to draw the torques diagram. This has been done in
Fig. 9.21. For the computation of the maximum shearing stress in the shaft, we have to apply
formula (9.14) for each particular portion of the shaft. We write:

1 1 3
M t M 0 500 10
A 1 34 34
A 1 m ax 2 , 77 MPa ; (9.27)
W p 3 3
A 1 d 30
16 16

67 67 3
Mt M0 500 10
2 B 34 34
2 B m ax 23 , 23 MPa . (9.28)
Wp 3 3
2 B (2d ) 60
16 16

Since M t
2 B
M t
1 2
, at the same value of Wp, it follows that 2 B m ax 1 2 m ax , and thus it is
not necessary to compute the value of 1-2 max , when we are looking for the maximum value of the
shearing stress along the shaft.
Comparing (9.27) with (9.28) we conclude that the maximum shearing stress develops at portion
2 - B, and has the value:

m ax 2 B m ax 23 , 23 MPa . (9.29)

The angle of twist of section 1 is:


1 1 3
M 0 2 500 10 2 500
M A 2 34 34 3
1 A 1 2 , 3 10 rad . (9.30)
G I 4 4
p d 4 30
G 8 10
32 32

177
Strength of Materials

Example 2
A shaft AB is attached to fixed supports at both ends and subjected to a uniformly distributed torque
m, as shown in Fig. 9.22. Knowing that the entire shaft is made of steel for which G = 8 104 MPa
and that a = 40 MPa, d = 40 mm and = 500 mm, determine:
a) The maximum allowable value of torque m so that the shaft would not fail;
b) The angle of twist of section 1.

a) From statics we write:


Mt 0 M A M B m . (9.31)

Since the total angle of twist of shaft AB is zero, we have:

A B
0 A 1 1 B


M A mx M A m
dx 0 , (9.32)
GI p GI p
0 1 2

where:
4 4 4
2d d d
Ip 16 16 I ; I .
1 p p
32 32 2 2 32

Fig 9.22
We write therefore:


M A mx M m
A
dx 0
G 16 I P G IP
0 2 2

178
Torsion

16 M A mx d x M A m 0
0

2
1 m 2
M A M A m 0 ,
16 2

which finally gives:


M A 0 , 97 m . (9.33)

Substituting this value into the original equilibrium equation, we obtain:

M B 0 , 03 m . (9.34)

The torques diagram has been represented in Fig. 9.22. The strength condition is:

M t
m ax
A 1 0 , 97 m
m ax 40 ,
A 1 3 a
Wp 2d
A 1
16
which gives:

3 3
2d 40 2 40 40
m 8291 N mm / mm ,
16 0 , 97 16 0 , 97 500
and

Mt
m ax 1 B 0 , 03 m
m ax a 40 ,
1 B Wp 3
1 B d
16
which gives:

3 3
d 1 40 1
m a 40 33510 N mm / mm .
16 0 , 03 16 0 , 03 500

The allowable value of m is given by the smallest of the two quantities computed above.
Therefore, we finally write:

m = 8291 N mm / mm . (9.35)

b) The angle of twist of section 1 is:

M B 0 , 03 m 0 , 03 8291 500 500 3


1 3 , 09 10 rad . (9.36)
GI p 4 4
d 4 40
2 G 8 10
32 32

179
Strength of Materials

9.2 TORSION OF NONCIRCULAR (RECTANGULAR) MEMBERS

The formulas presented in the preceding section, for the distribution of strain
and stress under a torsional loading apply only to circular members. This happens
because the derivation of these formulas has been based upon the assumption that the
cross sections of the member in torsion remain plane and undistorted. But this is not
the case if the cross sections of the member in torsion are noncircular.
For a rectangular member in torsion (for example):
the plane and undistorted cross sections of member before deformation do not
remain plane and undistorted after deformation, Fig. 9.23.
the shearing stresses over the cross
sections of a rectangular member in
torsion may be not assumed to vary
linearly with the distance from the
axis of the member. More than that,
the shearing stress at the corners of
the cross section in such a case is
zero. Consider for example a
rectangular member in torsion
(Fig.9.24). Due to the action of the
torque Mt , shearing stresses develop
over a current cross section of the
Fig. 9.23
member.
We detache a small cubic element of the rectangular member in torsion as shown in
Fig. 9.24.

Fig. 9.24

We assume that the shearing stress developed on the face of the element considered
has an arbitrary orientation within the section. However it may be resolved into two
components ( xy and xz) as shown in Fig. 9.24. On the other hand we know that shear

180
Torsion

cannot take place in one plane only; an equal shearing stress must be exerted on
another plane perpendicular to the first one. This means that the shearing stress
components xy and xz will be accompanied by equal shearing stresses yx and zx
respectively acting along the Ox direction. But, since the external surfaces of the
member are not loaded, we may write:

yx 0 ; zx 0. (9.37)

It follows therefore that :


xy yx 0;
(9.38)
xz zx 0;

and thus:
0. (9.39)

We conclude that there is no shearing stress at the corners of the cross


section of a rectangular member in torsion.
The determination of stresses in rectangular members subjected to torsion
may be done only through the Theory of Elasticity. Since this determination is not
very simple, involving a mathematical reasoning which is beyond the scope of this
text, we shall indicate here only the final results concerning the distribution of the
shearing stresses over the cross sections of rectangular members in torsion.
Consider for example a rectangular member in torsion (Fig. 9.25a).

a. b.
Fig. 9.25

We denote by the length of the member, by b and h, respectively, the


narrower and the wider side of the cross section and by Mt the magnitude of the
torque applied to the member. One could demonstrate that, the shearing stresses over
a current cross section of the member, at points located on the sides and on the axes

181
Strength of Materials

Oz and Oy, have the distribution shown in Fig. 9.25b. The maximum shearing stress
develops at the midpoints of the wider sides of the cross section (points B and B`),
being equal to:
M t M t
m ax 2
, (9.40)
k hb Wt

where Wt is called the torsional strength modulus of the cross section involved.
We note that:
2
Wt k hb , (9.41)

where k is a coefficient which depends only upon the ratio h / b.


The shearing stress developed at the midpoints of the narrower sides of the
cross section is given by (Fig. 9.25b):
'
max k2 max , (9.42)

where k2 is also a coefficient depending only upon the ratio h / b.


The angle of twist per unit length is:
M t M t
3
, (9.43)
G It G k1 h b

where G is the shear modulus and :


3
It k1 h b (9.44)

is the torsional moment of inertia, with k1 a coefficient depending upon the ratio h/b.
The angle of twist may be therefore expressed as:
M t
, (9.45)
G It

or, in general, analogous to the formula (9.16):



M t
d x. (9.45)
G It
0

The coefficients k, k1 and k2 are called Saint-Venants coefficients. They are given in
table 9.1 for a number of values of the ratio h/b.
Table 9.1
Saint-Venants coefficients
h/b 1,00 1,50 1,75 2,00 2,50 3 4 6 8 10
k 0,208 0,231 0,239 0,246 0,258 0,267 0,282 0,299 0,307 0,313 0,333
k1 0,141 0,196 0,214 0,229 0,249 0,263 0,281 0,299 0,307 0,313 0,333
k2 1,000 0,859 0,820 0,795 0,766 0,753 0,745 0,743 0,742 0,742 0,742

182
Torsion

From table 9.1 we note that, for higher values of the ratio h/b, coefficients k and k1
approach 1 . With such values, formula (9.40) becomes:
3

M t M t M t 3M t
m ax 2 2
. (9.46)
Wt k hb 1 2 hb
hb
3

Thus, for a thin-walled member of uniform thickness and arbitrary shape, the
maximum shearing stress is the same as for a rectangular member with a very large
value of h / b and may be determined from (9.46) - Fig. 9.26.

a. b.
Fig. 9.26

Generally speaking, the maximum shearing stress and the angle of twist per unit
length in case of noncircular members in torsion (within the elastic range),
respectively, may be expressed as:

M t
m ax ;
Wt
M t
,
GI t

where Wt and It are given in form of tables, for different kinds of sections.
The strength condition is therefore:

M t
m ax a , (9.47)
Wt

where a is the allowable value of the shearing stress for the material involved.

183
Strength of Materials

Numerical example

A torque M0 is applied, as shown in Fig. 9.27, to a solid steel shaft with a built in end. The shaft
has two distinct portions: a rectangular portion (with sides h and b) of length 1 and a circular
portion (with diameter d) of length 2. Knowing that:
4
b 10 mm , h 15 mm , d 25 mm , G 8 10 MPa , 1 1m, 2 0 ,5 m , a 50 MPa ,
determine:
a) the maximum allowable value of torque M0 which may be safely applied to the shaft;
b) the angle of twist of section 1, 1.

Fig. 9.27
(a) As shown in Fig. 9.27, the torque Mt is constant along the shaft. Since the shaft has two distinct
portions, we must compute the maximum allowable torque which may be safely applied to each
' ''
portion separately ( M t and M t ). Finally, the required value of the torque will be:

' "
M t M 0 min( M t , M t ) (9.48)

The rectangular portion (1-2):


M t M t
m ax 2 a
Wt k hb

' 2 2
M t M t k hb a 0 , 231 15 10 50 17325 N mm 17 , 325 N m.

The circular portion (2-B):

184
Torsion

Mt Mt 16 M t
m ax 3 3 a
Wp d d
16

3 3
" d 25
Mt Mt a 50 153398 N mm 153 , 398 N m.
16 16

Thus, the allowable torque which may be safely applied to the entire shaft is:
' "
M t M min( M t , M t ) 17 , 325 N m.
0

(b) The angle of twist of section 1 is:

M t M t 2 M t M t 2
1 2 1 2 B 1 2 1 2 B
1 1 2 2 B 3 4
G It G Ip G k1 h b d
G
32
3 3
17 , 325 10 1000 17 , 325 10 500
0 , 076 rad .
4 3 4
8 10 0 ,196 15 10 4 25
8 10
32

9.3 THIN-WALLED HOLLOW SHAFTS (TUBES)


As discussed in the preceding section, the torsion of noncircular shafts
requires advanced mathematical methods. However, a simple approximate solution
is possible for the special case of thin-walled hollow shafts.
Let us now consider for example a hollow member of noncircular section
subjected to torsion (Fig. 9.28a). The computation of the shearing stresses
equivalent to the applied torque Mt , requires the following assumptions:

a. b.
Fig. 9.28

185
Strength of Materials

the thickness t of the wall is small compared to the other dimensions of the
member in torsion;

the thickness t of the wall may vary within a cross section of the member, but
remains constant along the axis of the member;

the shape of the member cross section is arbitrary;

since the thickness of the wall is small compared to the other dimensions of the
member in torsion, we may admit that the shearing stresses are uniformly
distributed over the wall thickness, but with a certain variation law along the
contour of the cross section;

since the inner and outer walls of the hollow shaft are free surfaces, the stresses
on them are equal to zero. Thus the shearing stress at any point of a cross section
of the tube is directed along the tangent to the center line Fig. 9.28a. This
reasoning is similarly to that presented in Fig. 9.6.

For the determination of the shearing stress variation law along the center
line of the member cross section, we shall detach from the member an element of the
wall (1212abab - Fig. 9.28b), bounded by two transverse planes at a distance dx
from each other and by two longitudinal planes perpendicular to the wall, at distance
ds from each other (measured along the center line of the wall). The shearing stress
1 across the thickness t1 induces a numerically equal longitudinal stress, as
discussed within section (4.3). In the same manner, across thickness t2, the
corresponding shearing stress 2 is accompanied by a numerically equal longitudinal
stress directed along the x axis, (Fig. 9.28b). Since the element considered is in
mechanical equilibrium, we may write that the summation of all forces acting along x
axis is zero. This means that:

1 t1 dx 2 t2 dx 0 1 t1 2 t2 t constant. (9.49)

Since the element represented in Fig. 9.28b has been chosen arbitrarily, the above
relation tells us that the product t of the shearing stress at a given level (point) of
the cross section and of the wall thickness t at that level is a constant throughout the
member. The term t is called the shear flow. In other words, at points where the
wall thickness has a minimum value, the corresponding shearing stress is maximum
and vice versa.
To relate the shear flow to the applied torque Mt , we consider a small
element of the wall section, of length ds (Fig. 9.29). Since this small element has an
infinitesimal length ds, we may consider that the wall thickness remains
approximately constant along the element (and equal to t).

186
Torsion

Fig. 9.29

The elementary force which develops at the level of this element is expressed
through the product of the shearing stress at this level and the area (tds) of the
element (Fig. 9.29). We write:
dF t ds . (9.50)

The moment of this elementary force about an arbitrary point O within the
cavity of the member is:
dM dF h t h ds (9.51)

Since the externally applied torque Mt represents the sum of all such
elementary moments around the wall section, we may write:

M t dM dF h t h ds (9.52)

The shear flow t being a constant, we have:

M t t h ds t 2 , (9.53)

where is the area bounded by the


center line of the wall cross section (Fig.
9.30). It follows that:

M t
, (9.54)
2 t

Fig. 9.30 which is called the Bredts first formula.

In (9.54) t is the wall thickness at a given point of the member cross section and
the area bounded by the center line (Fig. 9.30). The maximum shearing stress max
develops at points where t has a minimum value. We write therefore:

187
Strength of Materials

M t M t
m ax , (9.55)
2 t m in Wt

where Wt is the torsional strength modulus for the involved member in torsion.
The strength condition becomes:
M t M t
m ax a , (9.56)
2 t m in Wt

where is the allowable value of the shearing stress for the material involved.
a
The angle of twist of a thin-walled hollow shaft may be obtained by using the
method of energy: the work done by the external torque M t will be equal to the
elastic strain energy accumulated by the member in torsion. Consider for example an
element of length dx detached from the hollow member of noncircular section
represented in Fig. 9.28 (Fig. 9.31a).

a. b.
Fig. 9.31

We denote by V the volume of this element of length dx. The elementary work done
by the statically applied torque Mt is:
1
dL M t d , (9.57)
2

where d is the elementary angle of twist (i.e. the angle of twist corresponding to
the elementary shaft of Fig. 9.31a). On the other hand this work is equal to the elastic
strain energy accumulated while the torque Mt is applied. As shown within the
preceding chapter, the elastic strain energy density in case of pure shear is:
2

U D .
2G

188
Torsion

Thus, the elastic strain energy accumulated by the entire element of Fig. 9.31a is:
2

U U D dV dV , (9.58)
2G
V V

where G is the shear modulus and dV is the elementary volume of the element shown
in Fig. 9.31a. This elementary volume has been represented in Fig. 9.31b separately.
Since this elementary volume has an infinitesimal length ds, we may consider that
the wall thickness at its level is approximately constant and equal to t (Fig. 9.31b). It
follows that:
dV t ds dx . (9.59)

We write therefore :
2 2 2
1 1 1
dL M td U dV M td dV M td t dx ds
2 2G 2 2G 2 2G
V V

2 2 2
1 2 1 M t 1 M t 1 M t dx 1
M td t dx ds t dx ds t dx ds d s.
2 2 2
2 2G 2 t 2G 4 t 2G 8G t

We have:
2
1 M t dx 1
M td
2
ds . (9.60)
2 8G t

The angle of twist per unit length is therefore:


d M t 1 M t M t
2
ds
2
, (9.61)
dx 4G t 4G G It
1
ds
t
where:
2
4
It . (9.62)
ds
t
Within the above calculus, Bredts first formula has been applied.
Relation (9.61) is called the Bredts second formula.
If the hollow shaft has a constant thickness t, it follows that:
2 2 2
4 4 4 t
It , (9.63)
ds 1 s
s
t t

s being the length of the center line.


189
Strength of Materials

Numerical example

How many times does the hollow shaft strength of Fig. 9.32,a decrease if a longitudinal split is
done (Fig. 9.32b) and a = 6 t ?

While both shafts are subjected to the same torque Mt , their strength (given by Wt) is completely
different.

a. b.
Fig. 9.32
For the hollow shaft represented in Fig. 9.32a we write :

M t M t M t M t M t
.
m ax 2 2 3
1 Wt 2 t 2a t 2 Gt t 72 t
1

In the second case (Fig. 9.32b) we have in fact a thin-walled member of uniform thickness
whose maximum shearing stress is the same as for a rectangular bar with a very large value of the
ratio h/b. Therefore, we have to use formula (9.46). We write:

M t M t M t 3M t 3M t M t
m ax .
2 1 1 2 2 3
Wt 2 2 4 at 4G t t 8t
2 hb 4a t
3 3

Thus,
Wt 3
1 72 t
9.
3
Wt 8t
2

In other words the strength does decrease 9 times if a longitudinal split is done in the shaft.

190
Torsion

PROBLEMS TO BE ASSIGNED

P.9

P.9.1 A shaft ABC is fixed at one end and subjected to a torque Mt = M0 at the other end as shown
(Fig. P.9.1). Portion AB, having a square cross section, is made of steel while portion BC, with a
circular cross section, is made of aluminum. Knowing that a = =40 mm, d = 70 mm , the allowable
shearing stress of steel all s = 100 MPa, the allowable shearing stress of aluminum all Al = 70
MPa, determine the largest torque M0 which may be applied at A. Neglect the effect of stress
concentration.

Fig. P.9.1

P.9.2 A steel shaft ABC is fixed at one end and subjected to the torques shown. Knowing that =
=0,5 m, all = 90 MPa, M0 = 1,5 kN m and G = 8 104 MPa, draw the torque diagram, determine
the required value of the diameter d and the angle of twist at A.

Fig. P.9.2 Fig. P.9.3


P.9.3 A solid tapered shaft AB, made of steel, is subjected to the torques shown. Knowing that =
0,6 m, all = 100 MPa, d = 30 mm, G = 8 104 MPa:
(a) draw the torque diagram;
(b) determine the largest value of M0 which may be safely applied;
(c) determine the angle of twist at A.
191
Strength of Materials

P.9.4 A shaft ABC having two distinct portions of circular cross section, is fixed at its both ends and
loaded as shown. Determine the maximum value of torque M0 which may be safely applied at B if d
= 30 mm and all = 90 MPa. Do also determine the angle of twist at B if = 0,8 m and G = 8 10 4
MPa.
P.9.5 Two solid shafts made of the
same material, are subjected to the
same torque M0 as shown (Fig.
P.9.5). Determine the ratio of d
and a so that the maximum
shearing stresses in both shafts
would be of the same value.
P.9.6 A brass shaft, having two
distinct portions (of annular and
circular cross sections
respectively), is subjected to a
uniformly distributed twisting
couple m and a torque M0 = 5ml as
shown (Fig. P9.6). Knowing that d
= 40 mm, all = 80 MPa, = 0,6 m,
Fig. P.9.4 G = 3,9 104 MPa:
(a) Determine the torques exerted on the shaft by each of the fixed supports;
(b) Draw the torque diagram;
(c) Determine the largest value of m which may be applied without exceeding the allowable
shearing stress all ;
(d) Determine the angle of twist between 1 and 2.

Fig. P.9.5 Fig. P.9.6

P.9.7 Two tapered shafts AB and BC are bonded together at B and attached to fixed supports at A
and C (Fig. P.9.7). Shaft AB is made of aluminum (with all Al = 75 MPa and GAl = 2,6 104 MPa)
while shaft BC is made of brass (with all B
=90 MPa and GB = 3,9 104 MPa). Knowing that =
0,8m and M0 = 2 kN m:
(a) Determine the torques exerted on the composite shaft by each of the fixed supports;
(b) Draw the torque diagram;
192
Torsion

(c) Determine the required value of d;


(d) Determine the angle of twist at B.

Fig. P.9.7 Fig. P.9.8


P.9.8 A steel shaft of square cross section and an aluminum tube are fixed at one end and connected
to a rigid plate at the other end, as shown (Fig. P.9.8). Determine the maximum value of torque M0
which may be safely applied to the rigid plate and the angle of twist at B if: GAl = 2,6 104 MPa,
GSteel = 8 104 MPa, all Al = 90 MPa, all Steel =100 MPa and = 1,5 m.

Fig. P.9.9 Fig. P.9.10


P.9.9 A steel thin-walled tube of square cross section AB and an aluminum tube BC of annular
cross section are fixed at one end and connected to a rigid plate at B as shown (Fig. P.9.9).
Determine the largest value of torque M0 which may be safely applied at 1 if: d = 80 mm, a = 40
mm, t = 5 mm, all Steel = 100 MPa, all Al = 80 MPa, Gsteel = 8 104 MPa, GAl = 2,6 104 MPa and
= 0,8 m. Do also determine the angle of twist at B.
P.9.10 A solid tapered shaft ABC made of steel is fixed at one end and connected to two steel rods
as shown (Fig. P.9.10). Knowing that d = 30 mm, all Steel =90 MPa, all Steel =180 MPa, d1 = 10
mm, d2 = 16 mm, ESteel = 2 105 MPa, GSteel = 8 104 MPa and = 1 m, determine the largest value
of torque M0 which may be safely applied at B.

193
Strength of Materials

194
10. BENDING

A member is said to be in bending if at its current cross section a bending


moment develops. A member in bending is called beam. The bending may be
classified using different criteria:
a) After the forces position in space, we have:
plane bending: the forces acting on the beam are located within the same
plane, this plane containing the beam axis and one of the principal axes
of inertia of the beam cross sections (Fig. 10.1).
unsymmetric bending: the forces acting on the beam are located within
the same plane, this plane contains the beam axis but contains neither of
the principal axes of inertia of the beam cross sections (Fig. 10.2).

Fig. 10.1 Fig. 10.2


general bending: the forces acting on the beam are not located within the
same plane, but each force does intersect the axis (Ox) of the beam (Fig.
10.3).
b) After the types of internal forces
developing at the current cross section we
have:
pure bending: the internal
forces reduce to a bending
moment (of a constant value,
orientation and sense along
the beam), the shearing force
being zero (Fig. 10.4a and
Fig. 10.4b between points 1
and 2). Fig. 10.3
Strength of Materials

simple bending: the bending moment Mi developed at the current cross


section of the beam is accompanied by the shearing force T (Fig. 10.4b,
between points A-1 and B-2).

a. b.
Fig. 10.4

10.1 PRISMATIC MEMBERS IN PURE BENDING

As discussed above, an example of a member in pure bending is furnished by


the portion 1-2 of the beam shown in Fig. 10.4b. In deriving the pure bending
formulas, we make the following assumptions:
the beam material is continuous, homogeneous, isotropic and elastic;
material modulus of elasticity in tension is equal to that in compression;
the beam cross section is constant along the member involved;
stresses do not exceed the proportional limit (i.e. Hookes law is
available);
the beam is in pure bending;
Bernoullis hypothesis is also valid.
We shall now detach a portion of infinitely small length dx, located between
points 1 and 2, from the beam shown in Fig. 10.4b (Fig. 10.5). This portion is
bounded by the planes (I and II) perpendicular to the axis of the beam. Due to the
action of the external loads the portion considered will bend, but will remain
symmetric with respect to the plane containing the loads applied to the beam.
Moreover, since the bending moment Miz is the same at any cross section, the
portion considered will bend uniformly (Fig. 10.5b). Thus the line CD (for
example) along which the upper face of the member intersects the plane of the
external forces P will have a constant curvature. In other words, the line CD, which
was originally a straight line, will be transformed into a circle of center O1 and so will
the lines CD, CD etc.

196
Bending

On the other hand, we note that, due to the action of the external loads, the
longitudinal fibers of the portion considered will deform: the fibres of the upper part
of the portion considered decrease in length while the fibers of the lower part increase
in length.
It follows that there must exist a surface parallel to the upper and the lower
faces of the beam whose fibers do neither increase nor decrease in length. This
surface is called the neutral surface. The neutral surface intersects the plane of forces
P along an arc of circle CD (Fig. 10.5b) which is called the neutral axis of the beam.

a.
b.
(Scheme of Fig. 10.4b)
Fig. 10.5
Since the longitudinal fibres of the member in pure bending do modify their
length along the longitudinal direction, we conclude that normal stresses x develop
at any cross section of such a member. Using the theory of elasticity or certain
experimental methods, one may verify that the normal component x is the only
nonzero stress component exerted on the current cross-sectional points of the beam.
Let us now return for a while to the bended portion of the beam represented in
Fig. 10.5b. We denote by d the angle made by the two sectional planes I and II
which bound the beam portion considered and by the radius of curvature
corresponding to the neutral axis CD. We shall now consider a fibre CD located at a
distance y bellow the neutral surface. As discussed above, the neutral fibre CD will
not modify its length in bending. The length of fibre CD will therefore remain equal
to dx. We write
dx d , (10.1)
On the other hand, the arbitrary longitudinal fibre CD considered will change its
original length dx with quantity dx . We have
dx dx y d , (10.2)

197
Strength of Materials

It follows that
d (d x ) d yd ,
or
dx yd . (10.3)
The strain of fibre CD considered becomes therefore
x

dx yd y
x . (10.4)
dx d

We conclude that the longitudinal normal strain varies linearly, throughout x

the member, with the distance y from the neutral surface. Moreover, since Hookes
law for uniaxial stress applies, we have:
E y
x E x , (10.5)

where E is the modulus of


elasticity of the material
involved.
Formula (10.5) shows
that, in the elastic range, the
normal stress varies linearly
with the distance from the
neutral surface (Fig. 10.6).
Recalling the relations
between the internal forces and
stresses, applied to the beam
shown in Fig. 10.6, we write: Fig. 10.6
E y E E
N d A 0 d A 0 yd A 0 Sz 0
A A A

Sz 0 . (10.6)
(This means that Oz is a centroidal axis.)
Ey E E
M iy zd A 0 zd A 0 y zd A 0 I zy 0
A A A

I zy 0 . (10.7)
(This means that Oz and Oy are the principal axis of inertia of the beam cross section)

Ey E 2 E
M iz yd A yd A M iz y d A M iz Iz M iz
A A A

198
Bending

1 M iz
(10.8)
EI z

Formula (10.8) (usually called Euler-Bernoulli's formula) gives us the


expression of the curvature of the beam neutral surface as a function of the applied
bending moment Miz, modulus of elasticity E and the second moment of the beam
cross sectional area A with respect to the axis about which the bending moment is
applied (Iz). The curvature is defined as the reciprocal of the radius of curvature . It
measures the deformation of the member caused by the bending moment Miz.
1 M iz
Substituting for into (10.5), we write:
EI z

1 M iz M iz
x E y E y y .
EI z Iz

We have obtained therefore the expression of the normal stress x


at any distance y
from the neutral axis (Oz) as:
M iz
y , (10.9)
Iz

called Naviers formula or elastic flexure formula. We note that the stress is
compressive above the neutral axis (y < 0) and tensile below the neutral axis (y > 0).
This happens only if the applied bending moment is positive (Fig. 10.7a). If the
bending moment at the cross section considered is negative then we shall have
compression bellow and tension above the neutral axis (Fig. 10.7b).
From formula (10.9) it follows that the maximum normal stress develops at
cross sectional points for which the distance y is maximum.

a.

b.
Fig. 10.7

199
Strength of Materials

We write
M iz M iz M iz
m ax y m ax ,
Iz Iz W z (10.10)
y m ax

Iz
where W z is called the elastic cross section modulus with respect to the Oz
y m ax
axis.
The strength condition becomes therefore
M i m ax
m ax a , (10.11)
W z

where is the allowable value of the normal stress for the material involved.
a

The elastic cross section modulus depends upon the shape and dimensions of
the cross - section:

rectangular cross section (Fig. 10.8):

3
bh
2
Iz 12 bh
W z . (10.12)
y m ax h 6
2

Fig. 10.8

circular cross section (Fig.10.9):

4
d
3
Iz 64 d
W z . (10.13)
y m ax d 32
2

Fig. 10.9

200
Bending

annular cross-section (Fig. 10.10):


4 4
D d
Iz 64 64
W z
y m ax D
2
4 4 (10.14)
D d
1
3
64 D D 4
1 c ,
D 32
Fig. 10.10 2

d
where the ratio has been denoted by c.
D

IMPORTANT REMARK
The relatively small number of engineering application where pure bending is
encountered does not justify a complex study in this matter. Fortunately the results
obtained for pure bending may be also applied to simple bending or to other types of
loading as well. An example is presented below.
Numerical example
A steel beam of the cross-section shown, is subjected to several external loads (Fig. 10.11).
Knowing that a 150 MPa :
a) Draw the shear and bending moment diagrams;

Fig. 10.11

201
Strength of Materials

b) Determine the values of Iz and Wz ;


c) Determine the minimum allowable value of dimension b so that the beam would not fail
due to the action of the external loads.
a) For the shear and bending moment diagrams we write

M A 0 10 2 20 30 2 8 4 4 20 YB 8 0 ;

YB 21 kN .

M B 0 10 10 YA 8 20 30 6 8 4 4 20 0 ;

YA 51 kN .

With these values, the shear and bending moment diagrams have been sketched in Fig. 10.11.
b) The second moment of the beam cross - sectional area is:
4
b
3
b 2b 4
Iz 0 , 666 b
4
.
12 64

It follows that
4
Iz 0 , 666 b
Wz 0 , 666 b
3
.
y max b

c) The maximum normal stress is


6
M 89 , 56 10
max
i max
3 all 150 MPa .
Wz 0 , 666 b

Thus
6
89 , 56 10
b 3 97 mm .
0 , 666 150

It is important to note that formula (10.9) has been derived for a member with a
plane of symmetry and a uniform cross section. As presented in chapter 6 or 9, for
beams heaving a variable cross section, the corresponding normal stresses will not
agree with Naviers formula. In such cases a stress concentration occurs. For
example higher stresses develop if the beam cross-section presents a sudden change
(Fig. 10.12). The ratio between the maximum actual normal stress ( ' ) and the max

maximum stress ( max ) computed with formula (10.10) is called the stress
concentration factor:
max
k (10.15)
max

202
Bending

The stress concentration


factor is given in form of
tables or graphs for
different particular cases.
The actual value of the
maximum stress at the
critical cross section may
be then expressed as
m ax k m ax ,
Fig. 10.12
where m ax is computed with Naviers formula.

10.2 SHEARING STRESSES IN BEAMS SUBJECTED TO SIMPLE


BENDING

Let us consider a beam with a vertical plane of symmetry, in simple bending.


This means that, at any cross section, the bending moment (Miz or Miy) is
accompanied by the corresponding shearing force (Ty or Tz respectively) - Fig. 10.13.
Since a shearing force develops at any cross section of the beam, besides the normal
stresses x , shearing stresses do also develop at any elementary area dA of the
cross section. It may be shown that, excepting some particular points of the cross
section, the direction of (with components xy and xz ) cannot be determined
using the strength of materials methods only.

a. b.

Fig. 10.13

203
Strength of Materials

For example, let us detache the element


abcda'b'c'd' from the beam represented in
Fig. 10.13b, Fig. 10.14. We assume that, at
an arbitrary point B located on the
circumference of the beam cross section, the
shearing stress has an arbitrary direction.
It may be resolved however into two
components 1 and 2
as shown in the
figure. But we know that shear cannot take
Fig. 10.14
place in one plane only, an equal shearing
stress occuring on another plane perpendicular to the first one. This means that 1
must be accompanied by 1 ' which is perpendicular to 1 and contained within the
surface aba'b' . Since the face aba'b' of the element considered is a part of the free
surface of the beam, all stresses on this face must be zero. Thus,
'
1 0 . (10.16)

It follows that

1 1 0. (10.17)
Since the stress component 1
equals zero, we conclude that the shearing stress at
point B acts along the tangent at the cross section circumference. This conclusion
remains valid for any particular point of the cross section circumference.
Let us now to return to the current cross section of the beam represented in Fig.
10.13a. As mentioned above, at the level of
this cross section the internal forces are
represented by the bending moment Miz and
the shearing force Ty. Since the effect of Miz
(which gives the normal stresses x
) has
been discussed previously (Navier's formula)
we shall now concentrate our attention on the
shearing stresses. We shall try to investigate
the shearing stresses occuring at points
located on an arbitrary segment mn, parallel
to Oz axis, at distance y from the neutral
surface (i.e. from Oz axis) (Fig. 10.15). As
demonstrated above, at points m and n the
shearing stress is tangent to the cross
section circumference. On the other hand, due
to the symmetry of the cross section with
respect to the plane Oxy, the tangents at
points m and n will intersect the Oy axis at the
Fig. 10.15 same point O1 (Fig. 10.15). Within such a

204
Bending

context, two basic assumptions (usually called Juravski's assumptions) have to be


considered:
for any other point m1 of segment mn, the direction of passes through
point O1;
the shearing stress remains constant for all points located on the
xy

segment mn. Its value depends only upon the distance y from the Oz axis (i.e.
from the neutral surface).
From Fig. 10.15 we have

xz (m1 ) (10.18)
tg 1 xz (m1 ) xy ( m 1 ) tg 1 .
xy (m1 )

At point C, for which 1


0 , we write
xz C xy C tg 1 xy C 0 0 . (10.19)
In other words, the
shearing stress component
xz at any particular point

of segment mn may be
expressed as a function of
xy . The problem is,
therefore, to find out the
mathematical connection
between the shearing
stress component xy and
the shearing force Ty over
the cross section
a. considered. This will be
done bellow.
We shall now
detache from the beam of
Fig. 10.13 a portion of
infinitely small length dx,
bounded by two planes
perpendicular to the axis
of the beam (sections S1
and S2) (Fig. 10.16a). We
do also cut this portion
with a longitudinal plane
passing through an
arbitrary segment mn
(analogous to that of Fig.
b. 10.15) of the beam cross
Fig. 10.16 section. This plane
205
Strength of Materials

(mnm'n') will be, therefore, parallel to the beam neutral surface (Fig. 10.16b).
We shall now retain only that part of the beam located bellow the sectional plane
mnm'n' (Fig. 10.16b). We denote by A1 the area of surface mnq (this area remaining
constant along Ox axis) and by b the width of the beam of distance y from the neutral
axis Oz. As represented in Fig. 10.16b, at the level of the cross section S1, M i M iz
and Ty develop, while, at the level of the cross section S2, M i d M i and T y develop
as cross sectional internal forces. The retained element mnqm'n'q' is subjected to:

the cross sectional shearing stresses xy developed at the level of the cross-
sectional points located on segment mn. As discussed above, these stresses
are accompanied by equal longitudinal shearing stresses yx .

the normal stresses developed at the level of the cross-sectional surfaces


mnq and m'n'q' of area A1.
The axial force N1 developed at the level of the entire mnq surface is therefore
Naviers formula
M i M i M i
N1 d A y1 d A y1 d A Sz, (10.20)
Iz Iz Iz
A1 A1 A1

where S is the first moment of A with respect to the neutral axis Oz while Iz is the
z 1

moment of inertia of the entire cross sectional area of the beam about the neutral axis.
Since an increase of the bending moment ( d M i ) occurs at the level of section S , the 2

axial force developed at surface mnq is therefore


M i dM i Sz
N 2 N1 d N1 . (10.21)
Iz

We shall now write the equilibrium equation of element mnqmnq about Ox axis as
follows:
M i dM i Sz M i
yx bd x Sz 0 .
Iz Iz

We note that
yx bd x is the corresponding axial force given by the average shearing
stress yx over the differential area of width b and length dx.

dM represents the differential change in bending moment


i M i within the
distance dx.
We write:

206
Bending

M i Sz dM i Sz M i
yx bd x Sz 0
Iz Iz Iz

dM i Sz dM i Sz
yx bd x yx
Iz d x b Iz

Ty Sz
yx .
b Iz

In other words, the shearing stress developed at the level of the beam cross
xy

sectional points located on segment mn, at distance y from the neutral axis Oz, may
be expressed through:
Ty Sz
xy yx (Jurawskis formula), (10.22)
b IZ

where
Ty: the shearing force at the beam cross-section considered;
b: the width of the beam cross-section at the level where the shearing
stresses must be derived;
Iz: the moment of inertia of the entire cross-section with respect to Oz
neutral axis;
Sz: the first moment of the area located either above or bellow the level
where the shearing stresses have to be computed.

a. b.
Fig. 10.17

It is important to note that the crosssectional shearing stresses xy (which depend


upon the quantities described in formula (10.22)), are always accompanied by equal
longitudinal shearing stresses which tend to shear the beam about a longitudinal
plane (Fig. 10.17a). In other words, at each such level a reciprocal sliding tendency
occurs (Fig. 10.17b).

207
Strength of Materials

RECTANGULAR SECTIONS
The distribution of shearing
stresses ( xy ) in a rectangular
section may be sketched using
formula (10.22) Fig. 10.18.
We shall compute the
shearing stress value, at a
current level mn located at
distance y1 from the neutral
axis and then we shall
represent the shearing stress
variation along the
rectangular section depth as a Fig. 10.18
function of y1. Thus, at the level mn (Fig. 10.18) we write
Ty Sz T S z
xy
b Iz b Iz

2
T h 1 h T b h 2

3
b y1 y1 3
y1
bh 2 2 2 bh 2 4
b b
12 12
2
6T h 2
= 3
y1 .
bh 4

We write therefore
2
6T h 2

xy 3
y1 xy
( y1 ) . (10.23)
bh 4

This shows that the shearing stresses have a parabolical distribution along the xy

rectangular section depth. The maximum shearing stress develops at the level of the
neutral axis and may be found by substituting y1 with zero in (10.23). We have

2
6T h 3 T 3 T
xy max xy (0 )
3
.
bh 4 2 bh 2 A

This means that, the maximum shearing stress developing at the level of the neutral
axis Oz is
3 T
max
, (10.24)
2 A
where A is the rectangular area ( A b h ). In other words the maximum shearing
stress is 50 % greater than the average shearing stress.

208
Bending

h
We do also note that, for y1 , the shearing stress xy
becomes zero.
2

CIRCULAR SECTIONS

The distribution of shearing stresses in a circular section may be derived in a


similar manner. We shall compute the value of the shearing stress at a current level
mn, at distance y1 from the
neutral axis and then we
shall represent the
shearing stresses variation
along the circular section
depth as a function of y1
(Fig. 10.19).
For convenience we
shall first derive the
current expression of the
first moment of the area A1
located below the current
level mn (hachured area of Fig. 10.19
Fig. 10.19) with respect to the neutral axis Oz as a function of distance y1. We write

S z ( y1 ) mn
ydA . (10.25)
A1

We select the element of area dA in the shape of a thin horizontal strip and thus
reduce the computation of the above double integral to integration in a single variable
(Fig. 10.19). We write
2 2
dA b y dy 2 r y dy ,
where r is the radius of the circular section. We have therefore
r r
2 2 2 2 2 2
S z y1 ydA 2y r y dy r y r y dy
A1 y1 y1
3 3
2 2 2 r 2 2 2
r y 2
y1
r y1 2 .
3 3

With this expression of S z y1 , the shearing stress becomes


3
Ty Sz T Sz T 2 2 2
xy 4
r y1 2
,
bI Z
bI Z 2 2 d 3
2 r y1
64
(d 2r )

209
Strength of Materials

which reduces to
2
4 T y1
xy
1 2 xy
y1 , (10.26)
3 A r

where A is the area of the entire circular section considered.


This shows that the shearing stresses have a parabolic distribution along the
xy

circular section depth. The maximum shearing stress does also develop at the level of
the neutral axis and may be found by substituting y with zero in (10.26). We have
1

4 T
xy max xy
O . (10.27)
3 A

This indicates that the maximum shearing stress in a circular section is 33 % greater
than the average shearing stress. We do also note that, for y 1 r , the shearing stress

xy
becomes zero (Fig. 10.19).

10.3 PREVENTION OF LONGITUDINAL SLIDING IN CASE OF


COMPOSITE SECTIONS

For beams with a long span or subjected to high loads, a high value of the
elastic cross-section modulus is required. In many such cases beams with composite
sections are chosen. If a beam, for example, were composed of two or more layers
placed on each other, bending would produce the effect shown in Fig. 10.20.

Fig. 10.20
The separate layers 1 and 2 would slide past each other and the total strength of
the beam would be the sum of the strengths of the layers. The beam is considerably
weaker than a solid beam of equivalent dimensions.
To increase the strength of the beam shown in Fig. 10.20, the layers are
gripped together by means of several rivets or bolts. These rivets or bolts will prevent
the layers from sliding when bended (Fig.10.21). The built-up beam will work like a
solid beam of equivalent dimensions, and a considerably more effort is required to
make it fail.
On the other hand the rivets or bolts which prevent the layers from sliding
reciprocally will be sheared by longitudinal forces developed at the layers separation
plane. It is important to mention that this horizontal reciprocal sliding tendency does

210
Bending

Fig. 10.21

also occur even if the beam is solid. This may be explained through the longitudinal
shearing stresses yx which develop in simple bending (Fig. 10.17).

Numerical example
The composite beam shown in Fig. 10.22, is made by discontinuous welding. It is subjected
to a uniformly distributed force q 42 kN / m and three concentrated forces P 420 kN .
Knowing that the weld throat depth is a 7 mm , the length of weld c 100 mm and the weld
allowable shearing stress a 80 MPa :

Fig. 10.22
a) Draw the shear and bending-moments diagrams;
b) Compute the values of Iz and Wz for the beam section involved;
c) Determine the maximum values of the normal and shearing stresses ( max and max );
d) Determine the required value of length e at which the discontinuous welds have to be placed so
that the composite section would not fail in bending;
e) Draw the normal and shearing stresses distribution over the cross section 2left ;
f) Compute the values of the principal stresses 1 and 2 at point K of section 2right.

211
Strength of Materials

Solution
a. The shear and bending moments diagrams have been sketched in Fig. 10.22.
b. Within the computation of the cross-sectional Iz and Wz we shall neglect the weld area. We write
therefore
3 3
240 20 2 10 800 8 4
Iz 410 20 240 2 20 , 40 10 mm ;
12 12

8
Iz 20 , 40 10 5 3
W z 48 , 57 10 mm .
y m ax 420

c. The maximum normal stress is


6
M imax 735 10
max 5
151 , 32 MPa .
W z 48 , 57 10
The maximum shearing stress develops at the level of the neutral surface of the beam (Oz
axis) at the cross section where the shear is maximum. From Juravskis formula we have:
3
T max Sz 462 10 20 240 410 400 10 200
max 8
62 , 68 MPa .
b Iz 10 20 , 40 10

d. The welds are subjected to longitudinal shearing forces, which are a direct consequence of the
shearing stresses yx which develop on longitudinal planes at levels mn or m n Fig. 10.22.
As we already know, the longitudinal shearing stresses yx xy
may be computed with
Juravskis formula:
T Sz
yx xy .
b Iz

For a covering computation we shall chose:


3
T T max 462 10 N ;
b b min 10 mm ;
8 4
Iz 20 , 40 10 mm ;
3
S z 20 240 410 mm .
Two welds (the upper right and left welds for example) have to cover the longitudinal
shearing force developed on a rectangular area of dimensions b e . This force is obtained from the
shearing stress yx multiplied by this area:

T S T Sz e
F yx b e
z
b e . (10.28)
b Iz Iz

On the other hand, this force has to be supported by the two welds of length c. We write

T S e
z
2a c 2a a ,
Iz

which finally gives:

212
Bending

e 216 mm .
e. The normal and shearing stresses distribution at section 2left may be obtained using Naviers and
Juravsky's formulas. The results are given bellow.

Fig. 10.23
At the level of point K (Fig. 10.22) two kinds of stresses develop:
A normal stress x , perpendicular to the cross section at K;

A shearing stress xy , acting within the cross-sectional plane, and of a constant value for
all cross-sectional points located at level mn Fig. 10.22.
The principal stresses at K for the cross section 2right are:
x y 1 2 2
1, 2 x y 4 xy ,
2 2

where
y 0 .
We write
6
M i 735 10
x (k ) y
8
410 147 , 72 MPa ;
Iz 20 , 40 10
3
T S z 210 10 10 240 415
xy (k )
8
0 , 42 MPa .
b Iz 240 20 , 40 10

It follows that
147 , 72 1 2 2
1, 2 147 , 72 4 0 , 42 ,
2 2

which reduces to

1 0 , 00119 MPa ;

2 147 , 721 MPa .

10.4 BEAMS OF CONSTANT STRENGTH

The above presented matters concerning bending have been limited so far to
prismatic beams of constant cross sections. If the applied bending moment is constant

213
Strength of Materials

along the beam involved, the stresses developed at any cross section will have the
same value and distribution (Fig. 10.24).
Since the maximum normal stress max usually controls the design of the
beam, the strength condition
max a

(where a is the allowable value of the


normal stress for the material involved)
for the beam shown in Fig. 10.24 is met in
the same manner at each particular cross
section. If the applied bending moment
varies along the beam, the strength
condition must insure that the stresses in
the critical section(s) are at most

Fig. 10.24

equal to the allowable values of the normal and shearing stresses (Fig. 10.25). It
follows that, in all other sections, the stresses will be smaller (or much smaller)
than the allowable stresses.

Fig. 10.25

A prismatic beam therefore, is almost


always overdesigned, and considerable
savings of material may be realized by using

214
Bending

nonprismatic beams of variable cross sections.


The problem is, therefore, to design such beams for which the elastic cross -
sectional modulus W z (or W y ) to vary along the beam ( W z = W z (x) or W z = W z (x)), so
that at any cross section, to have
M i
x
a
. (10.29)
W z
x

It follows that
M i
x
W z (x) . (10.30)
a

If we know the variation law of the bending moment as a function of x ( M i M i ( x ) )


and the stress allowable value of the material involved - a , we may thus design the
geometry of the beam which to lead to an optimum use of the material. A beam
designed in this manner is referred to as a beam of constant strength.
Let us now return to the beam represented in Fig. 10.26 in a simplified manner.
The bending moment diagram shows a linear variation of M i with respect to x:
M i
P x .

This beam may be designed as a beam


of constant strength if we vary the cross
section in a continuous way along the
length of the member. Usually, there
Fig. 10.26 are two ways to do this in practical
applications:
to vary the depth of the beam at constant width
or
to vary the width of the beam at constant depth.
Let us now consider the first case (Fig. 10.26). The strength condition at any cross
section of the beam is:
M i (x)
max a .
W z (x)

215
Strength of Materials

It follows therefore that


2
P x b h (x)
W z (x) ,
a
6
which gives
6P x
h(x) . (10.31)
b a

Formula (10.31) shows us a parabolic variation of the beam cross section depth as a
function of x. The geometry of such a beam has been represented in Fig. 10.26. On
the other hand, it is important to note that, for cross sections located in the vicinity of
the external force P application point, the cross sectional area decreases too much and
does not meet the shearing stresses strength condition requirement
3T 3P
( max a
, A: the cross-sectional area; a
: the allowable value of the
2A 2A
shearing stresses).
This is why we have to adopt a constant cross section which to insure the
shearing stresses strength condition for a certain length x of the beam (Fig. 10.26). 0

We denote by y the depth of the beam throughout this portion. We write therefore
0

3 T 3 P
max a
,
2 A 2 b y0

which gives
3P 6P x0
y0 h x0 .
2b a
b a

This means that


2 2
9P 6P x0 9P b a 3 P a
2 2
x0
2 2 2
.
4b a
b a 4b a
6P 8 b a

In conclusion, the beam of constant strength (with a variable depth and a


3 P a
constant width) must have a constant cross section for a length x0 2
,
8 b a

following than a parabolic


variation of the depth, as shown in
Fig. 10.26. The beam of constant
strength designed in this manner
provides savings of material of
aprox. 30 %, with respect to a
prismatic beam (Fig. 10.27).
The volume of the beam of
constant strength is (Fig. 10.27)
Fig. 10.27
216
Bending

2
V abc .
3
In the second case we may keep the depth constant and vary the width of the

beam (Fig. 10.28). We shall follow the


same reasoning. The only difference
compared with the preceding case
consists in the variation of the width
instead of that of the depth.
The strength condition at any cross
Fig. 10.28 section of the beam is:
M i
x
max a .
W z
x

It follows therefore that


2
P x b x h
W z (x) ,
a
6

which gives
6P x
b x 2
. (10.32)
h a

Formula (10.32) shows us a linear variation of the beam cross section width as a
function of x. The geometry of such a beam has been represented in Fig. 10.28. On
the other hand, as specified in the preceding case, the shearing stresses strength
condition requirement must also be met. We write therefore (Fig. 10.28):
3 T 3 P
max a
,
2 A 2 h b1

which gives
3 P 6P x1
b1 2
.
2 h a h a

217
Strength of Materials

This means that


2
3P h a
h a
x1 .
2h a
6P 4 a

In conclusion, the beam of constant strength (with a variable width and a


h a
constant depth) must have a constant cross section for a length x1 ,
4 a

following then a linear variation of the width, as shown in Fig. 10.28. The beam of
constant strength designed in this
manner provides savings of
material of aprox. 50 % with
respect to a prismatic beam (Fig.
10.29).
The volume of the beam of
constant strength is (Fig. 10.29).
1
Fig. 10.29
V abc .
2

10.5 UNSYMMETRIC BENDING

A beam is said to be under unsymmetric bending if at any cross section the


bending moment vector is not directed along a principal centroidal axis of the cross
section (Fig. 10.30).

Fig. 10.30

GENERAL CASE OF UNSYMMETRIC BENDING STATE OF STRESS


Consider a straight beam of constant cross section subjected to unsymmetric
bending. A centroidal rectangular coordinate system zOy is attached to the beam
cross sections (Fig. 10.31). Due to the action of the applied bending moment
M i (with components M iz and M iy ), normal stresses develop at the level of each

particular elementary area d A of the beam cross section. To derive the unsymmetric
bending formulas, we make the following assumptions:

218
Bending

stresses do not exceed the proportional limit (i.e. Hookes law is


available;
the beam is in unsymmetric bending;
Bernoullis hypothesis is also valid.
The stresses developed
may be obtained by
superposing the stresses
corresponding to the bending
moment components M iz and
M iy , as long as the
conditions of applicability of
the principle of superposition
are satisfied. Within the
context of the above
presented assumptions, we
may consider that the normal
strain of the beam
Fig. 10.31 longitudinal fibres does
linearly depends upon the
coordinates z and y. We may write therefore:
x
C1 z C2 y C3 , (10.33)
where C1, C2 and C3 are constants.
Since the Hooke s law conditions are satisfied we have
x
E x
E C1 z C2 y C3
(10.34)
E C1 z E C2 y E C3 C1 z C2 y C3 .

where C1, C2 and C3 are also constants.


The relationships among internal forces and stresses within the involved beam
cross section may be written as follows

N d A 0; zd A M ; yd A M .
iy iz (10.35)
A A A

We may now substitute C1z C2y C3 into (10.44) and write

C1z C2y C3 d A 0;
A

C1z C2y C3 zd A M iy
;
A

C1z C2y C3 y d A M iz
.
A

219
Strength of Materials

C1 zd A C2 yd A C3 d A 0;
A A A

2
C1 z d A C2 yz d A C3 zd A M iy
; (10.36)
A A A

2
C1 zy d A C2 y d A C3 yd A M iz
.
A A A

It follows that

C 1 I zy C2Iz C 3S z M iz
;

C1I y
C 2 I zy C3S y
M iy
; (10.37)
C1S y
C2S z C3A 0.

Since Oz and Oy are centroidal axes we have

Sz yd A 0;
A

S y
zd A 0.
A

In these conditions, the last relation of (10.37) gives

C3 A 0 C3 0. (10.38)

We may now retain the first two relations of (10.37), with C3 = 0, and (10.34) in a
single system of three equations:

C 1 I zy C2Iz M iz
;

C1I y
C 2 I zy M iy
; (10.39)
C1z C2y .

This may be considered a system of three equations with two unknowns (C1 and C2).
The condition of compatibility in accordance with Rouches theorem is therefore

I zy Iz M iz

I y
I zy M iy
0 , (10.40)
z y

220
Bending

where
I zy Iz
0 .
Iy I zy

It follows that
Iy I zy I zy Iz I zy Iz
M iz
M iy
0 ,
z y z y Iy I zy

which finally reduces to:


yI y
z I zy y I zy z Iz
2
M iz 2
M iy , (10.41)
IzI y
I zy IzI y
I zy

The condition 0 , leads to the equation of the cross section neutral axis (N.A):

I y
y I zy z M iz
I zy y Iz z M iy
0 . (10.42)

10.6 GENERAL BENDING

A beam is said to be under general bending if the forces acting on the beam are not
located within the same plane, but the support of each force does intersect the Ox axis
of the beam (Fig. 10.3). In such cases the beam deforms after a certain curve in space.
The neutral axes of different cross sections will not be located within the same plane
and there will be NO neutral plane.
To resolve such cases the following steps have to be covered:

The external forces have to be resolved about two principal planes (Oz and Oy);

The bending moment diagrams must be then drawn for each of the two planes
(Oz and Oy);

The critical sections (where the resultant bending moment has maximum
values) must be established;

For each critical section the maximum normal stress has to be computed and
then compared with the normal stress allowable value of the material involved.

221
Strength of Materials

PROBLEMS TO BE ASSIGNED
P.10
P.10.1 For the beams shown (Fig. P.10.1):
(a) Draw the shearing force and bending moment diagrams;
(b) Determine the cross-sections centroidal points, Iz and Wz ;
(c) Determine the required dimensions of the cross-sections (t=?) if a = 180 MPa.

a.

b.

c.

d.

Fig. P.10.1

222
Bending

P.10.2 For the beam shown in Fig. P.10.2, determine the largest value of the uniformly distributed
force q which may be applied without exceeding either of the following allowable stresses: a = -
90 MPa and a = + 30 MPa.

Fig. P.10.2

P.10.3 For the beam shown in Fig. P.10.3:


(a) Draw the shear and bending moment diagram;
(b) Determine the centroid, Iz and Wz of the cross-section;
(c) Determine the allowable uniformly distributed load q if a = + 40 MPa in tension and a =-
120 MPa in compression.

Fig. P.10.3

P.10.4 A steel I - shaped beam must support the loading shown (Fig. P.10.4). Knowing that all =
180 MPa, select the lightest I- shaped profile required.

Fig. P.10.4

P.10.5. For the beam and loading shown (Fig. P.10.5):


(a) Draw the shear and bending moment diagrams;
(b) Determine the required value of t if a = 200 MPa ;
(c) Draw the normal and shearing stresses diagrams at sections Aleft and Bleft.

223
Strength of Materials

Fig. P.10.5

P.10.6 A cantilever beam AB has a constant depth h = 20 mm and a variable width b as shown
(Fig. P.10.6). Locate the cross section where the normal stress has a maximum value and determine
this value.

Fig. P.10.6

P.10.7 Three beams of the same cross section are pin connected at B and C and loaded as shown
(Fig. P.10.7). Determine the required value of the uniformly distributed load q knowing that a =
160 MPa.

Fig. P.10.7

224
Appendix I - TYPICAL MOMENTS OF INERTIA
Cross-sectional shape Computation Cross-sectional shape Computation formulas
formulas

3
bh (a b)h h 2a b
Iz ; A ;e ;
12 2 3 a b
3 3 2 2
hb h a 4 ab b
Iy ; Iz ;
12 36 a b
h 2 2
Iy (a b )( a b ).
48

2 3 2
A a ; h (B 2 ab )
4 Iz ;
a 36 B
Iz Iy ;
2 2
12 h B (B ab )
Iy
2
36 B d (B d ) (B 2 ab )
h 2
I zy (B 2 d )( B 2 ab ) .
12 B

3
bh bh
A ; Iz ;
2 12
3 2 2
h bh bh ( b c )
e ;Iz Iy ;
2 36 12
3 2
hb h bc
Iy I zy .
48 12

bh Hexagon
A ;
2
3 5 5 4
bh Iz Iy a ;
Iz ; 16
36
3
hb
Iy
36

Polygon with n equal sides


bh nR
4

A ; I sin (2 cos );
2 24
3
bh na
4
sin (2 cos )
Iz ; I
12 96 (1 cos )
2

4
n = 3, I a 3 / 96

247
Appendix II - It AND Wt FORMULAS FOR DIFFERENT TYPES OF
CROSS-SECTIONAL SHAPES OF MEMBERS IN TORSION
No Cross-sectional shape It Wt Remarks

max
1 R
4
d
4
R
3
d
3
occurs on circumference

2 32 2 16

(D d ) (D d )
4 4 4 4

2 max
32 16 D
occurs on circumference

a b
3 3

3 ab
2
nb
2
We denote by
a b
2 2
a
n b
3 3 2 2 n 1
b
n 1
2

n ( b o b1 )
3 4 4
n ( b o b1 )
4 4
We denote by
ao a1
n 1
2

4 2 bo n 1
bo b1

4
2
2 t min - the area bounded by
the center line of the wall
ds
5 cross section;
t(s) max occurs where t is
minimum (t = tmin)

We admit that

6 2 r t
3
2 r t
2
t r

max = const.

2b h
2 2
max occurs where t is
7 2 bht minimum (t = tmin)
b h min

t1 t2
tmin = min (t1, t2)

249
Appendix II (continued)

No Cross-sectional It Wt Remarks
shape

0,296R4 0,348R3 0 - at corners

max - at the diameter midpoint


8

max - at sides midpoint


4
0,0216a a
3

9 20

max - at sides midpoint


10 0,141a4 0,208a3 0 at corners

1,04a4= 0,977a3=0,188b3 max - at sides midpoint


11 = 0,108b4

max - at sides midpoint


Khb3 K1hb2
0 at corners
12
h/ 1,0 1,5 2,0 3,0 4,0 8,0 10,0
b
k 0,141 0,196 0,229 0,263 0,281 0,298 0,307 0,31 0,33
k1 0,208 0,231 0,246 0,267 0,282 0,299 0,307 0,31 0,33
k2 1,0 0,859 0,796 0,753 0,745 0,743 0,743 0,74 0,33

t
1 1 0
ht
3
ht
2
h
13 3 3 t h

1
h
3 It t max h
14 [ t ( s )] ds
3 0 t max

249
Appendix II (continued)

No Cross-sectional It Wt Remarks
shape

15
2 2 t h
rt rt
3 2

3 3

16 hit i
3
It
i
t max

Profile L [ T I
0,99 1,12 1,12 1,31 1,17

R D R D
4 4 3 3

c1 c1 c2 c2
17 2 32 2 16

r/D 0 0,05 0,1 0,2 0,3 0,4 0,5


c1 1 0,98 0,93 0,78 0,59 0,40 0,24
c2 0,5 0,52 0,52 0,49 0,42 0,33 0,24

18 d
4


32

0 , 01 D 0 ,17 dD
4 3

19

249
APPENDIX III - PROPERTIES OF ROLLED-STEEL SHAPES

ROLLED-STEEL SHAPES
I SHAPED PROFILE
STAS 564 80
R Flange inner radius
(dimensions and geometrical characteristics) r Flange outer radius
h Depth Iz, Iy Axial moments of inertia
b Width Wz, Wy Strength moduli
d Web thickness iz, iy Radii of gyration
t Thickness Sz - First moment of the half-cross-sectional area

Designation Dimensions [mm] Cross- Designation


sectional Axis z - z Axis y -y Sz[cm3]
H b t d=R r area Iz[cm4] Wz[cm3] iz[cm] Iy[cm4] Wy[cm3] iy[cm]
A[cm2]
I8 80 42 5,9 3,9 2,3 7,58 77,8 19,5 3,2 6,29 3,0 0,91 11,4 I8
I 10 100 50 6,8 4,5 2,7 10,6 171 34,2 4,01 12,2 4,88 1,07 19,9 I 10
I 12 120 58 7,7 5,1 3,1 14,2 328 54,6 4,81 21,5 7,41 1,23 31,8 I 12
I 14 140 66 8,6 5,7 3,4 18,3 573 81,9 5,61 36,2 10,71 1,40 47,7 I 14
I 16 160 74 9,5 6,3 3,8 22,8 935 117 6,4 54,7 14,8 1,55 68,0 I 16
I 18 180 82 10,4 6,9 4,1 27,9 1450 161 7,2 81,3 19,8 1,71 93,4 I 18
I 20 200 90 11,3 7,5 4,5 33,5 2140 214 8,0 117 26,0 1,87 125 I 20
I 22 220 98 11,92 8,1 4,9 39,6 3060 278 8,80 162 33,1 2,02 162 I 22
I 24 240 106 12,80 8,7 5,2 46,1 4250 354 9,59 221 41,7 2,20 206 I 24
I 26 260 113 13,77 9,4 5,6 53,4 5740 442 10,4 288 51,0 2,32 257 I 26
I 28 280 119 14,85 10,1 6,1 61,1 7590 542 11,1 364 61,2 2,45 316 I 28
I 30 300 125 15,82 10,8 6,5 69,1 9800 653 11,9 451 72,2 2,56 381 I 30
I 32 320 131 16,92 11,5 6,9 77,8 12510 782 12,7 555 84,7 2,67 457 I 32
I 36 360 143 19,05 13,0 7,8 97,1 19610 1090 14,2 818 114 2,90 638 I 36
I 40 400 155 21,10 14,4 8,6 118 29210 1460 15,7 1160 149 3,13 857 I 40

249
249
APPENDIX III - PROPERTIES OF ROLLED-STEEL SHAPES

ROLLED-STEEL SHAPES
r Legs inner radius
ANGLES - Equal Legs
r1 Legs outer radius
STAS 424 80 Iz, Iy Axial moments of inertia
Wz, Wy Strength moduli
(dimensions and static quantities for bending) Iz, Iy Radii of gyration
a Leg Width I or I1, I or I2 Principal moments of inertia,
g Leg Thickness maximum and minimum
C Centroid Sz - First moment of the half-cross-sectional
e Distance between the centroid and the outer area
side
z, y Centroidal Axes
, or 1,2 - principal centroidal axes

Cross-sectional Cross- Distance between axes, Inertial Geometrical Quantities


Size and dimensions [mm] sectional [mm] zz y-y 11 2-2
Thickness [mm] a g r r1 area e u v1 v2 Iz=Iy Wz=Wy iz=iz I=I1 i I=I2 W i Izy
[cm2] [cm4] [cm3] [cm] [cm4] [cm] [cm4] [cm3] [cm] [cm4]
20 x 20 x 3 20 3 3,5 2 1,12 0,60 1,41 0,84 0,70 0,39 0,28 0,59 0,61 0,74 0,16 0,19 0,38 0,25
20 x 20 x 4 20 4 3,5 2 1,45 0,64 1,41 0,90 0,71 0,41 0,36 0,58 0,77 0,73 0,21 0,23 0,38 0,28
25 x 25 x 3 25 3 3,5 2 1,42 0,72 1,77 1,02 0,87 0,80 0,45 0,75 1,26 0,94 0,33 0,32 0,48 0,465
25 x 25 x 4 25 4 3,5 2 1,85 0,76 1,77 1,08 0,89 1,01 0,58 0,74 1,60 0,93 0,43 0,40 0,48 0,585
25 x 25 x 5 25 5 3,5 2 2,26 0,80 1,77 1,13 0,91 1,20 0,71 0,73 1,89 0,91 0,52 0,46 0,48 0,685
30 x 30 x 4 30 4 5 2,5 2,27 0,88 2,12 1,24 1,05 1,80 0,85 0,89 2,85 1,12 0,75 1,61 0,58 1,05
30 x 30 x 5 30 5 5 2,5 2,78 0,92 2,12 1,30 1,07 2,16 1,04 0,88 3,41 1,41 0,92 0.70 0,57 1,245
35 x 35 x 4 35 4 5 2,5 2,67 1,00 2,47 1,42 1,24 2,95 1,18 1,05 4,68 1,33 1,23 0,86 0,68 1,725
35 x 35 x 5 35 5 5 2,5 3,28 1,04 2,47 1,48 1,25 3,56 1,45 1,04 5,64 1,31 1,49 1,01 0,67 2,075
40 x 40 x 4 40 4 6 3 3,08 1,12 2,83 1,64 1,42 5,43 1,91 1,20 8,60 1,51 2,26 1,37 0,78 2,62
40 x 40 x 5 40 5 6 3 3,79 1,16 2,83 1,64 1,42 5,43 1,91 1,20 8,60 1,51 2,26 1,37 0,77 3,17
45 x 45 x 5 45 5 7 3,5 4,30 1,28 3,18 1,81 1,58 7,84 2,43 1,35 12,4 1,70 3,25 1,80 0,87 4,575
45 x 45 x 6 45 6 7 3,5 5,08 1,32 3,18 1,87 1,59 9,16 2,88 1,34 14,5 1,69 3,82 2,04 0,87 5,34

251
APPENDIX III (continued)
Cross-sectional Cross- Distance between axes, Inertial Geometrical Quantities
Size and dimensions [mm] sectional [mm] zz y-y 11 2-2
Thickness [mm] a g r r1 area e u v1 v2 Iz=Iy Wz=Wy iz=iz I=I1 i I=I2 W i Izy
[cm2] [cm4] [cm3] [cm] [cm4] [cm] [cm4] [cm3] [cm] [cm4]
50 x 50 x 5 50 5 7 3,5 4,80 1,40 3,54 1,98 1,76 12,80 3,61 1,50 17,4 1,90 4,54 2,59 0,97 6,43
50 x 50 x 6 50 6 7 3,5 5,69 1,45 3,54 2,01 1,77 12,80 3,61 1,50 20,4 1,89 5,33 2,61 0,97 7,535
50 x 50 x 7 50 7 7 3,5 6,56 1,49 3,54 3,10 1,78 14,60 4,16 1,49 23,1 1,88 6,10 2,91 0,96 8,5
60 x 60 x 5 60 5 8 4 5,82 1,64 4,24 2,32 2,11 19,4 4,45 1,82 30,7 2,30 8,02 3,45 1,17 11,34
60 x 60 x 6 60 6 8 4 6,91 1,69 4,24 2,39 2,11 22,8 5,29 1,82 36,2 2,29 9,43 3,95 1,17 13,385
60 x 60 x 8 60 8 8 4 9,63 1,77 4,24 2,50 2,14 29,2 6,89 1,80 46,2 2,26 12,1 4,86 1,16 17,05
60 x 60 x 10 60 10 8 4 11,1 1,85 4,24 2,61 2,17 34,9 8,41 1,78 55,1 2,23 14,8 5,67 1,16 20,15
70 x 70 x 6 70 6 9 4,5 8,13 1,93 4,95 2,73 2,46 36,9 7,27 2,13 53,5 2,68 15,2 5,69 1,37 19,15
70 x 70 x 7 70 7 9 4,5 9,40 1,97 4,95 2,79 2,47 42,4 8,41 2,12 67,1 2,67 17,5 6,27 1,36 24,8
70 x 70 x 8 70 8 9 4,5 10,60 2,01 4,95 2,85 2,49 47,5 9,52 2,11 75,3 2,66 19,7 6,91 1,36 27,8
70 x 70 x 10 70 10 9 4,5 13,10 2,09 4,95 2,96 2,52 57,2 11,70 2,09 90,5 2,63 23,9 8,09 1,35 33,3
80 x 80 x 6 80 6 10 5 9,35 2,17 5,66 3,07 2,82 55,8 9,57 2,44 88,5 3,08 23,1 7,55 1,56 32,7
80 x 80 x 8 80 8 10 5 12,30 2,26 5,66 3,19 2,82 72,2 12,6 2,43 115 3,06 29,8 9,36 1,55 42,6
80 x 80 x 10 80 10 10 5 15,1 2,34 5,66 3,30 2,85 87,5 15,4 2,41 139 3,03 36,3 11,0 1,55 51,35
90 x 90 x 8 90 8 11 5,5 13,9 2,50 6,36 3,53 3,17 104 16,1 2,74 166 3,45 43,1 12,2 1,76 61,45
90 x 90 x 9 90 9 11 5,5 15,5 2,54 6,36 3,59 3,18 116 18,0 27,4 184 3,45 47,8 13,3 1,76 68,1
90 x 90 x 11 90 11 11 5,5 18,7 2,62 6,36 3,70 3,21 138 21,6 2,72 218 3,41 57,1 15,4 1,75 80,45
100 x 100 x 8 100 8 12 6 15,5 2,74 7,07 3,87 3,52 145 19,9 3,06 230 3,85 59,8 15,4 1,96 85,1
100 x 100 x 10 100 10 12 6 19,2 2,82 7,07 3,99 3,54 177 24,6 3,04 280 3,83 72,9 18,3 1,95 103,55
100 x 100 x 12 100 12 12 6 22,7 2,90 7,07 4,11 3,57 207 29,1 3,02 328 3,80 85,7 20,9 1,94 121,15
120 x 120 x 10 120 10 13 6,5 23,2 3,31 8,49 4,69 4,23 313 36,0 3,67 497 4,63 129 27,5 2,36 184
120 x 120 x 12 120 12 13 6,5 27,5 3,40 8,49 4,80 4,26 368 42,7 3,65 584 4,60 151 31,5 2,35 216,5
130 x 130 x 12 130 12 14 7 30,3 3,64 9,19 5,15 4,60 472 50,4 3,97 750 5,00 194 37,7 2,54 278
130 x 130 x 14 130 14 14 7 34,7 3,72 9,19 5,26 4,63 540 58,2 3,94 857 4,97 223 42,4 2,53 317
140 x 140 x 12 140 12 15 7,5 32,5 3,90 9,90 5,50 5,04 602 59,7 4,31 957 5,43 248 44,9 2,76 354,5
140 x 140 x 14 140 14 15 7,5 37,6 3,98 9,90 5,61 5,07 689 68,8 4,30 1094 5,42 284 50,5 2,74 405
150 x 150 x 14 150 14 16 8 40,3 4,21 10,6 5,95 5,31 845 78,2 4,58 1340 5,77 347 58,3 2,94 496,5
150 x 150 x 16 150 16 16 8 45,7 4,29 10,6 6,07 5,34 949 88,7 4,56 1510 5,74 391 64,4 2,93 559,5
160 x 160 x 12 160 12 17 8,5 37,4 4,39 11,3 6,19 5,74 913 78,6 4,94 1450 6,23 376 60,5 3,17 537
160 x 160 x 14 160 14 17 8,5 43,3 4,47 11,3 6,30 5,77 1046 90,8 4,92 1662 6,20 431 68,1 3,16 615,5

251
APPENDIX III - PROPERTIES OF ROLLED-STEEL SHAPES
ROLLED-STEEL SHAPES
ANGLES - Unequal Legs
STAS 425 - 80 r Legs inner radius
(dimensions and static quantities for bending) r1 Legs outer radius
a Big leg size Iz, Iy Axial mements of inertia
b Small leg size I or I1, I or I2 principal moments of inertia,
g Legs thickness maximum i minimum
C Centroid Wz, Wy Strength moduli
ez, ey Distance between the centroid and the iz, iy Radii of gyration
outer leg side Sz - First moment of the half-cross-sectional area
z, y Centroidal axes
, or 1,2 - Principal centroidal axes

Size and Cross-sectional Cross- Inertial Geometrical Quantities


Distance between axes, [mm]
Thickness dimensions [mm] sectional
tg
area A y-y z-z - -
[mm] [cm2] Iy Wy iy Iz Wz iz I i I i Izy
a b g r r1 ez ez u1 u2 v1 v2 v3
[cm4] [cm3] [cm] [cm4] [cm3] [cm] [cm4] [cm] [cm4] [cm] [cm4]
30x20x3 3 1,43 0,99 0,50 2,05 1,51 0,86 1,04 0,56 0,427 1,25 0,62 0,93 0,44 0,29 0,55 1,43 1,00 0,26 0,42 0,43
30 20 3,5 2
30x20x4 4 1,86 1,03 0,54 2,02 1,52 0,91 1,04 0,58 0,427 1,59 0,81 1,92 0,55 0,38 0,55 1,81 0,99 0,33 0,42 0,54
40x20x3 3 1,73 1,42 0,44 2,61 1,77 0,79 1,19 0,46 0,257 2,80 1,09 0,27 0,47 0,30 0,52 2,96 1,31 0,31 0,42 0,64
40 20 3,5 2
40x20x4 4 2,26 1,47 0,48 2,58 1,80 0,83 1,17 0,50 0,252 3,59 1,42 1,26 0,60 0,39 0,51 3,80 1,30 0,39 0,42 0,81
45x30x4 4 2,86 1,48 0,74 3,06 2,23 1,21 1,58 0,80 0,434 5,77 1,91 1,42 2,05 0,91 0,85 6,63 1,52 1,19 0,65 1,98
45 30 4,5 2
45x30x5 5 3,52 1,52 0,78 3,04 2,26 1.27 1.58 0,83 0,429 6,98 2,35 1,41 2,47 1,11 0,84 8,00 1,51 1,45 0,64 2,37
60x30x5 5 4,29 2,15 0,68 3,89 2,67 1,20 1,77 0,72 0,256 15,6 4,04 1,90 2,60 1,12 0,78 16,5 1,96 1,69 0,63 3,56
60 30 6 3
60x30x6 6 5,08 2,20 0,72 3,86 2,69 1,25 1,75 0,74 0,252 18,2 4,78 1,89 3,02 1,32 0,.77 19,2 1,95 1,99 0,63 4,06
60x40x5 5 4,79 1,96 0,97 4,10 3,01 1,68 2,1 1,10 0,434 17,2 4,25 1,89 6,11 2,02 1,13 19,8 2,03 3,54 0,86 5,99
60x40x6 60 40 6 6 3 5,68 2,00 1,01 4,08 3,02 1,72 2,10 1,12 0,431 20,1 5,03 1,88 7,12 2,38 1,12 23,1 2,02 4,15 0,86 6,91
60x40x7 7 6,55 2,04 1,05 4,06 3,03 1,77 2,09 1,14 0,427 22,9 5,79 1,87 8,07 2,74 1,11 26,3 2,00 4,75 0,85 7,86
65x50x6 6 6,58 2,04 1,29 4,52 3,60 2,15 2,39 1,50 0,575 27,2 6,1 2,03 14,0 3,77 1,46 33,8 2,27 7,43 1,06 11,37
65x50x7 65 50 7 6,5 3,5 7,60 2,08 1,33 4,50 3,62 2,19 2,39 1,52 0,572 31,1 7,03 2,02 15,9 4,34 1,45 38,5 2,25 8,51 1,06 13,9
65x50x8 8 8,60 2,11 1,37 4,49 3,64 2,23 2,39 1,54 0,569 34,8 7,93 2,01 17,7 4,89 1,44 43,0 2,23 9,57 1,50 14,3

253
APPENDIX III (continued)

Size and Cross-sectional Cross- Inertial Geometrical Quantities


Distance between axes, [mm]
Thickness dimensions [mm] sectional
tg
area A y-y z-z - -
[mm] [cm2] Iy Wy iy Iz Wz iz I i I i Izy
a b g r r1 ez ez u1 u2 v1 v2 v3
[cm4] [cm3] [cm] [cm4] [cm3] [cm] [cm4] [cm] [cm4] [cm] [cm4]
65x50x9 9 6,5 3,5 9,58 2,15 1,49 4,48 3,63 2,28 2,36 1,57 0,567 38,2 8,77 2,00 19,4 5,39 1,42 47,0 2,25 10,5 1,05 15,7
75x50x7 75 50 7 6,5 3,5 8,31 2,48 1,25 5,10 3,77 2,13 2,63 1,38 0,433 46,4 9,24 2,36 16,5 4,39 1,41 53,3 2,53 9,57 1,07 15,09
80x60x7 80 60 7 8 4 9,38 2,51 1,52 5,55 3,79 2,17 2,92 1,40 0,546 59,0 10,7 2,51 28,4 6,34 1,74 72,00 2,77 15,4 1,28 23,7
80x65x6 6 8,41 2,39 1,65 5,61 4,63 2,69 2,94 2,01 0,649 52,9 9,41 2,51 31,2 6,44 1,93 68,5 2,85 15,6 1,35 24,2
80x65x8 8 11,0 2,47 1,73 5,59 4,65 2,79 9,94 2,05 0,645 68,1 12,3 2,49 40,1 8,41 1,91 88,0 2,82 20,3 1,36 30,9
80x65x10 10 13,6 2,55 1,81 5,56 1,68 2,90 2,95 2,11 0,640 82,2 15,1 2,46 48,3 10,3 1,89 10,6 2,79 24,8 1,35 36,8
90x60x6 60 6 8 8,69 2,89 1,41 6,14 4,50 2,46 3,16 1,60 0,442 71,7 11,7 2,87 25,8 5,61 1,72 82,8 3,09 14,6 1,30 25,2
90 4
90x60x8 8 11,4 2,97 1,49 6,11 4,54 2,56 3,15 1,69 0,437 92,5 15,4 2,85 33,0 7,31 1,70 107 3,06 19,0 1,29 32,1
100x50x8 8 11,4 3,59 1,12 6,49 4,44 2,00 2,96 1,18 0,257 116 18,1 3,18 19,5 5,04 1,31 123 3,28 12,7 1,05 26,7
100 50 9 4,5
100x50x10 10 14,1 3,67 1,2 6,43 4,40 2,08 2,93 1,22 0,253 141 22,2 3,16 23,4 6,17 1,29 149 3,25 15,4 1,05 31,6
100x75x7 7 11,9 3,06 1,83 6,96 5,42 3,10 3,61 2,18 0,553 118 17,0 3,15 56,9 10,0 2,19 145 3,49 30,1 1,59 48,7
100x75x9 100 75 9 10 5 15,1 3,15 1,91 6,91 5,45 3,22 3,63 2,22 0,549 148 21,5 3,13 71,0 12,7 2,17 181 3,47 37,8 1,59 60,5
100x75x11 11 18,2 3,23 1,99 6,87 5,49 3,32 3,65 2,27 0,545 176 25,9 3,11 84,0 15,3 2,15 214 3,44 45,1 1,58 71,3
120x80x8 8 15,5 3,83 1,87 8,23 5,99 3,27 4,23 2,16 0,437 226 27,6 3,82 80,8 13,2 2,28 260 4,10 46,6 1,73 79,4
120x80x10 120 80 10 11 5,5 19,1 3,92 1,95 8,18 6,03 3,37 4,21 2,19 0,435 276 34,1 3,80 98,1 16,2 2,26 317 4,07 56,8 1,72 96,5
120x80x12 12 22,7 4,00 2,03 8,14 6,06 3,46 4,20 2,25 0,432 323 40,4 3,77 114 19,1 2,24 371 4,04 66,6 1,71 111
150x90x10 10 23,2 5,00 2,04 10,1 7,05 3,60 5,03 2,24 0,360 533 53,3 4,80 146 21,0 2,51 591 5,05 88,3 1,95 161
150 90 12,5 6,5
150x90x12 12 27,5 5,08 2,12 10,1 7,10 3,70 5,00 2,30 0,358 627 63,3 4,77 171 24,8 2,49 694 5,02 104 1,94 189
150x100x10 10 24,5 4,80 2,34 10,3 7,50 4,10 5,25 2,68 0,442 552 54,1 4,78 198 25,8 2,86 637 5,13 112 2,15 191
150x100x12 150 100 12 13 6,5 28,7 4,89 2,42 10,2 7,53 4,19 5,24 2,73 0,439 650 64,2 4,76 232 30,6 2,84 719 5,10 132 2,15 227
150x100x14 14 33,2 4,97 2,50 10,2 7,56 4,28 5,23 2,77 0,435 744 74,1 4,73 264 35,2 2,82 856 5,07 152 2,14 257

253
APPENDIX III - PROPERTIES OF ROLLED-STEEL SHAPES

ROLLED-STEEL SHAPES
U SHAPED PROFILE R Flange inner radius
STAS 564 80 r Flange outer radius
Iz, Iy Axial moments of inertia
(dimensions and geometrical characteristics) Wz, Wy Strength moduli
h Depth iz, iy Radii of gyration
b Width Sz - First moment of the half-cross-sectional area
t Thickness

Cross- Static quantities for bending


Dimensions [mm] sectional
Designation z-z y -y
h b d t R r area Iz[cm4] Wz[cm ] iz[cm] Iy[cm ] Wy[cm3]
3 4
iy[cm] Sz[cm3] ey[cm]
2
A[cm ]
U5 50 38 5 7 3,5 7,12 26,4 10,6 1,92 9,12 3,75 1,13 - 1,37
U6,5 65 42 5,5 7,28 7,5 4 9,03 57,5 17,7 2,52 14,1 5,07 1,25 - 1,42
U8 80 45 6 7,76 4 11,0 106 26,5 3,10 19,4 6,36 1,33 15,9 1,45
U 10 100 50 6 8,26 4,5 13,5 205 41,2 3,91 29,3 8,49 1,47 24,5 1,55
U 12 120 55 7 8,72 4,5 17,0 361 60,7 4,62 43,2 11,1 1,59 36,3 1,60
U 14 140 60 7 9,72 5 20,4 605 86,4 5,45 62,7 14,8 1,75 51,4 1,75
U 16 160 65 7,5 10,20 5,5 24,0 925 116 6,21 85,3 18,3 1,89 68,8 1,84
U 18 180 70 8 10,68 5,5 28,0 1350 150 6,95 114 22,4 2,02 89,6 1,92
U 20 200 75 8,5 11,16 6 32,2 1910 191 7,70 148 27,0 2,14 114 2,01
U 22 220 80 9 12,14 6,5 37,4 2690 245 8,48 197 33,6 2,30 146 2,14
U 24 240 85 9,5 12,62 6,5 42,3 3600 300 9,22 248 39,6 2,42 179 2,23
U 26 260 90 10 13,60 7 48,3 4820 371 9,99 317 47,7 2,56 221 2,36
U 30 300 100 10 15,60 8 58,8 8030 535 11,7 495 67,8 2,90 316 2,70

254
STRENGTH OF MATERIALS
TYPICAL KEY WORDS AND PHRASES

ENGLISH - ROMANIAN ROMANIAN - ENGLISH


A A
ABSCISSA abscis ABSCIS - abscissa
ALGEBRAIC SUM - nsumare algebric ALAM - brass
ALLOWABLE LOAD - sarcin capabil ALIAJ - alloy
ALLOWABLE STRESS tensiune ALIERE - alloying
admisibil ALUMINIU - aluminum
ALLOWABLE - STRESS METHOD - ALUNECARE - slip
metoda tensiunii admisibile ALUNGIRE - elongation
ALLOY - aliaj AMPLITUDINE - amplitude
ALLOYING - aliere ANALIZA - analysis
ALUMINUM - aluminiu ANALOGIE CU MEMBRANA
ALUMINUM PIPE eav de aluminiu (TORSIUNE) - membrane analogy
ALUMINUM SHELL - manta (cma) ANEXA - appendix
de aluminiu ANSAMBLU MOMENT FOR
AMPLITUDE amplitudine (TORSOR) - force-couple system
ANALYSIS analiz APLICAREA FORMULEI LUI EULER
ANALYTICAL FUNCTION funcie PENTRU ALTE CAZURI DE
analitic REZEMARE ALE BARELOR -
ANGLE - unghi extension of Eulers formula to columns
ANGLE OF TWIST - unghi de torsiune with other end conditions
ANGULAR VELOCITY - viteza APLICAT DIRECT LA VALOAREA
unghiular NOMINAL (DESPRE O FOR,
ANNEALED STEEL oel recopt ETC) - fully applied
ANNULAR CROSS SECTION seciune ARBORE COTIT crankshaft
inelar ARBORE DE TRANSMISIE -
APPENDIX anex transmission shaft
ARC OF CIRCLE - arc de cerc ARBORE DE TRANSMISIE
AVERAGE STRESS - tensiune medie CILINDRIC MASIV (PLIN) solid
AVERAGE VALUE - valoare medie cylindrical transmission shaft
AXIAL LOAD - for axial ARBORI STATIC NEDETERMINAI-
AXIAL LOADING - solicitare axial statically indeterminate shafts
AXLE - ax ARC CU SPIR STRNS - close
coiled spring
B ARC DE CERC - arc of circle
BALL SOCKET - cuzinet, loca sferic ARC DE TORSIUNE torsional spring
BAR bar ARC LAMELAR - leaf spring
Strength of Materials

ENGLISH - ROMANIAN ROMANIAN ENGLISH


BAUSCHINGER EFFECT - efectul ARCE DE CERC CONCENTRICE-
Bauschinger concentric arcs of circle
BEAM grind ARIA SECIUNII TRANSVERSALE-
BEAM OF CONSTANT STRENGTH - cross-sectional area
grind de egal rezisten ARTICULAIE - pin-connection
BEAM UNIT WIDTH - grind de lime ARTICULAIE PLASTIC - plastic
unitate hinge
BEARING rulment AX-axle
BEARING FRAME - cadru portant AXA DE COORDONATE - coordinate
BEARING STRESS - tensiune (presiune) axis
de contact AXA NEUTR - neutral axis
BEARING SURFACE suprafa de AXE CENTRALE - centroidal axes
contact AXE CENTRALE PRINCIPALE DE
BENDING ncovoiere INERIE ALE SECIUNII
BENDING MOMENT - moment de TRANSVERSALE - principal centroidal
ncovoiere axes of the cross section
BENDING-MOMENT DIAGRAM - AXE DE COORDONATE coordinate
diagram de moment ncovoietor axes
BIAXIAL STRESS CONDITION - stare AXE PRINCIPALE - principal axes
de tensiune biaxial
BINDING fretare
BOLT bol B
BOUNDARY CONDITIONS condiii
BAND, CUREA - strap
de contur (de rezemare) BAR - bar
BRASS alam BAR ZVELT - slender member (bar)
BRASS LAYER - strat de alamBAR (ZVELT CE POATE
BREAKING STRENGTH rezistena la
PREZENTA PERICOLUL PIERDERII
rupere STABILITII), COLOAN, STLP
BRITTLE MATERIALS materiale
column
fragile BAR (ZVELT) NCASTRAT LA
BRONZE bronz CAPETE - column with two fixed ends
BUCKLE (TO-) - flamba (a-) BAR CONIC - tapered bar
BUCKLING flambaj BAR CURB - curved member
BUCKLING COEFFICIENT coeficient
BAR DE LUNGIME MIC - stubby
de flambaj column
BAR DREAPT - straight bar
C BAR PRISMATIC - prismatic bar
CABLE cablu BAR ZVELT DUBLU
CANTILEVER BEAM - grind n ARTICULAT LA CAPETE -pin-ended
consol column
CAP - capac, calot BARE (ZVELTE) CE FLAMBEAZ N
CAST IRON font DOMENIUL ELASTIC - long columns

226
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


CASTIGLIANOS THEOREM - teorema BARE (ZVELTE) CE FLAMBEAZ N
lui Castigliano DOMENIUL ELASTO-PLASTIC -
CENTRE OF CURVATURE - centru de intermediate columns
curbur BARE (ZVELTE) DE LUNGIME
CENTRIC AXIAL LOAD - for axial MIC, LA CARE NU APARE
aplicat n centrul de greutate PERICOLUL DE PIERDERE A
CENTRIC LOAD - sarcin (for) STABILITII - short columns
aplicat n centrul de greutate BARE CU PEREI SUBIRI - thin-
CENTROID - centru de greutate walled members
CENTROIDAL AXES - axe centrale BETON concrete
CHECK (TO-) - verifica (a-) BOL - bolt
CIRCULAR CROSS SECTION BRAD - fir
seciune circular BRONZ - bronze
CIRCULAR HOLE - gaur circular
CIRCULAR WIRE srm (fir) de
seciune circular
CLEARANCE interstiiu C
CLOCKWISE - n sensul acelor de CABLU - cable
ceasornic CABLU DE OEL - steel cable
CLOSE-COILED SPRING - arc cu spir CADRAN dial gage
strns CADRU frame
COEFFICIENT OF FRICTION CADRU PORTANT - bearing frame
coeficient de frecare CAL PUTERE - horsepower
COEFFICIENT OF THERMAL CAPAC, CALOT - cap
EXPANSION - coeficient de dilatare CAPACITATEA DE A ACUMULA
liniar ENERGIE energy-absorbing capacity
COLD-WORKED - prelucrat la rece CAPT LIBER - free end
COLLAR manon, colier CAUZE NATURALE IMPREVIZIBILE
COLUMN - bar (zvelt ce poate unpreventable natural causes
prezenta pericolul pierderii stabilitii), CENTRU DE CURBUR - centre of
coloan, stlp curvature
COLUMN WITH TWO FIXED ENDS - CENTRU DE GREUTATE - centroid
bar (zvelt) ncastrat la capete CENTRU DE FORFECARE - shear
COMPRESSED-AIR TANK - recipient center
cu aer comprimat CERCUL LUI MOHR - Mohrs circle
COMPRESSIVE STRESS - tensiune de CERCUL LUI MOHR PENTRU
compresiune STAREA PLAN DE DEFORMAIE-
CONCENTRATED LOAD for Mohrs circle for plane strain
concentrat CERCUL LUI MOHR PENTRU
CONCENTRIC ARCS OF CIRCLE - STAREA PLAN DE TENSIUNE-
arce de cerc concentrice Mohrs circle for plane stress
CONCRETE - beton CHERESTEA timber
CONDITIONS OF EQUILIBRIUM - CILINDRU TUBULAR -hollow cylinder

227
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


condiii de echilibru COEFICIENT (FACTOR) DE
CONE - con CONCENTRARE A TENSIUNILOR -
CONNECTION legatur stress-concentration factor
CONSTANTS OF INTEGRATION - COEFICIENT DE DILATARE
constante de integrare LINIAR coefficient of thermal
COORDINATE AXES - axe de expansion
coordonate COEFICIENT DE FLAMBAJ - buckling
CORNERS OF THE SECTION - coefficient
colurile seciunii COEFICIENT DE FRECARE-coefficient
CORROSION - coroziune of friction
CORROSION RESISTANCE
COEFICIENT DE SIGURAN - factor
rezistena la coroziune of safety
COUNTERCLOCKWISE n sens invers COEFICIENT DE ZVELTEE -
acelor de ceasornic (sens trigonometric)
slenderness ratio
COUPLE - cuplu, moment COEFICIENT EFECTIV DE
CRACK fisur ZVELTEE - effective slenderness ratio
CRACK GROWTH creterea fisurii COEFICIENI DE INFLUEN -
CRADLE -jgheab, cuzinet de reazem influence coefficients
CRANE - macara COEFICIENTUL LUI POISSON-
CRANKSHAFT - arbore cotit Poissons ratio
CREEP - fluaj COLURILE SECIUNII - corners of
CRITICAL LOAD - sarcin critic the section
CRITICAL STRESS - tensiune critic COMPONENTELE TENSIUNII - stress
CROSS-SECTIONAL AREA - aria
components
seciunii transversale CON - cone
CURVATURE curbur CONCENTRATORI DE TENSIUNE -
CURVE curb stress concentrators
CURVED MEMBER - bar curb CONDIII DE CONTUR (DE
CUT (TO-) seciona (a-) REZEMARE) - boundary conditions
CYCLIC LOADING - sarcin ciclic CONDIII DE ECHILIBRU - conditions
CYLINDRICAL BRASS VESSEL - vas of equilibrium
cilindric din alam CONDUCTA DE EVACUARE-penstock
CYLINDRICAL PRESSURE VESSELS CONSTANTE DE INTEGRARE -
-vase de presiune cilindrice constants of integration
CONTRACIE TRANSVERSAL -
D transverse contraction
DASHED LINE - linie punctat COORDONATE RECTANGULARE-
DECAY putrezire rectangular coordinates
DEFLECTION - sgeat CORNIER DIN OEL - steel angle
DEFLECTION AND SLOPE AT POINT COROZIUNE corrosion
A - sgeata i rotirea n punctul A CORP RIGID rigid body
DEFLECTION OF BEAMS deformaia CORP SOLID solid body
grinzilor CRETEREA FISURII - crack growth

228
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


DEFLECTION OF BEAMS deformaia CRIC HIDRAULIC - hydraulic jack
grinzilor (supuse la ncovoiere) CRITERIUL LUI MOHR - Mohrs
DEFORMABLE STRUCTURE - criterion
structur deformabil CUI nail
DEFORMATION deformaie CUPLU, MOMENT - couple
DEGREE OF ACCURACY - grad de CURB - curve
precizie CURBUR - curvature
DEGREES CELSIUS (FAHRENHEIT) - CUZINET, LOCA SFERIC -ball socket
grade Celsius (Fahrenheit) CURGERE (A MATERIALULUI) -
DENSITY densitate yielding
DESIGN proiectare
DESIGN OF TRANSMISSION SHAFTS D
- proiectarea arborilor de transmisie
DECHIDERE (A UNEI GRINZI) - span
DIAGRAM diagram DEFECT DE STRUCTUR - flaw
DIAL GAGE cadran DEFORMAIA GRINZILOR (SUPUSE
DIAMETER diametru LA NCOVOIERE) - deflection of beams
DIEDRAL ANGLE - unghi diedru DEFORMAIA PERMANENT -
DIMENSIONLESS QUANTITY -
permanent deformation
mrime adimensional DEFORMAIE - deformation
DIRECTLY PROPORTIONAL - directDEFORMAIE PLASTIC - plastic
proporional deformation
DEFORMAIE SPECIFIC - strain
DISCONTINUITY - discontinuitate
DISPLACEMENT - deplasare DEFORMAIE SPECIFIC CAUZAT
DISTANCE - distan DE VARIAII DE TEMPERATUR-
DISTRIBUTED FORCES fore thermal strain
distribuite DEFORMAIE SPECIFIC
DISTRIBUTED LOADS sarcini LONGITUDINAL - normal strain
distribuite DEFORMAIE SPECIFIC
DOUBLE INTEGRAL - integral dubl
TRANSVERSAL - lateral strain
DOUBLE SHEAR - forfecare dublDEFORMAIE SPECIFIC
DRAWING, SKETCH desen TRANSVERSAL (LUNECARE
DUCTILE MATERIALS materiale SPECIFIC) - shearing strain
ductile DEFORMAIE TOTAL - total
DYNAMIC LOAD - sarcin dinamic
deformation
DENSITATE density
E DEPLASARE displacement
ECCENTRIC AXIAL LOAD for DEPLASARE RELATIV - relative
axial aplicat excentric displacement
ECCENTRICAL - excentric DERIVAT PARIAL - partial
ECCENTRICITY - excentricitate derivative
EDGE OF THE BEAM - marginea DESCRCARE - unloading
(muchia ) grinzii DESEN - drawing, sketch

229
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


EFFECTIVE LENGTH lungime de DIAGRAM - diagram
flambaj DIAGRAM DE FORE TIETOARE-
EFFECTIVE SLENDERNESS RATIO - shear diagram
coeficient efectiv de zveltee DIAGRAM DE MOMENT
ELASTIC CORE - miez elastic NCOVOIETOR - bending-moment
ELASTIC CURVE fibr medie diagram
deformat DIAGRAMA TENSIUNE-
ELASTIC FLEXURE FORMULA - DEFORMAIE SPECIFIC-stress-strain
relaia lui Navier (ncovoiere) diagram
ELASTIC LIMIT limit de elasticitate DIAMETRU diameter
ELASTIC RANGE domeniu elastic DIAMETRU EXTERIOR-outer diameter
ELASTIC SECTION MODULUS - DIAMETRU INTERIOR- inner diameter
modulul de rezisten la ncovoiere al DIRECT PROPORIONAL - directly
seciunii (n domeniul elastic) proportional
ELASTOPLASTIC MATERIAL - DISCONTINUITATE - discontinuity
material elastoplastic DISPUNEREA PE LUNGIME A
ELECTRIC GENERATOR - generator SUDURILOR SE FACE DIN...N mm
electric - longitudinal spacing of the welds is
ELEMENTARY FORCES fore ofmm
elementare DISTAN - distance
ELEMENTARY INTERNAL FORCES - DISTRIBUIE LINEAR - linear
eforturi elementare distribution
ELEMENTARY WORK - lucru mecanic DISTRIBUIE PARABOLIC -
elementar parabolical distribution
ELLIPSE elips DISTRUGERE failure
ELLIPTIC CROSS SECTION seciune DOMENIU ELASTIC - elastic range
transversal eliptic
ELONGATION alungire
EMPIRICAL FORMULAS - formule E
empirice ECHILIBRU equilibrium
ENDURANCE LIMIT rezistena la ECHILIBRU INIIAL - original
oboseal equilibrium
ENERGY - energie ECHIVALEN - equivalence
ENERGY-ABSORBING CAPACITY ECONOMII DE MATERIAL -savings of
capacitatea de a acumula energie material
ENGINEERING PRACTICE - practica ECRUISARE strain-hardening
inginereasc ECUAIE - equation
EQUATION ecuaie ECUAIE DIFERENIAL DE
EQUILIBRIUM echilibru ORDINUL DOI - second-order linear
EQUILIBRIUM POSITION poziie de differential equation
echilibru ECUAIE DIFERENIAL NEOMO-
EQUIVALENCE echivalen GEN LINIAR DE ORDINUL DOI,
CU COEFICIENI CONSTANI-

230
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


EQUIVALENT STATIC LOAD sarcin - linear nonhomogeneous differential
static echivalent equation of the second order with
EULERS FORMULA - formula lui constant coefficients
Euler ECUAIE DIFERENIAL OMOGE-
EXPERIMENTAL EQUIPMENT - N LINIAR - linear homogeneous
instalaie experimental differential equation
EXTENSION OF EULERS FORMULA EFECTUL BAUSCHINGER -
TO COLUMNS WITH OTHER END Bauschinger effect
CONDITIONS aplicarea formulei lui EFORTURI ELEMENTARE -
Euler pentru alte cazuri de rezemare ale elementary internal forces
barelor ELEMENT (COMPONENT AL UNEI
EXTERNAL FORCES - fore exterioare STRUCTURI) - member
ELEMENTE DIN LEMN - wooden
members
F ELICE, SPIRAL (N SPAIU) - helix
FACTOR OF SAFETY coeficient de ELIPS-ellipse
siguran ENERGIA DE DEFORMAIE LA
FAILURE ruptur, rupere, avarie, NCOVOIERE - strain energy in bending
distrugere, cedare, defectare, ntrerupere,ENERGIA DE DEFORMAIE LA
etc. TORSIUNE - strain energy in torsion
FATIGUE oboseal ENERGIE SPECIFIC DE DEFORMA-
FATIGUE CRACK - fisura de oboseal IE - strain energy density
FICTITIOUS OR DUMMY LOAD - ENERGIE - energy
sarcin fictiv ENERGIE CINETIC - kinetic energy
FIGURE figur ENERGIE DE DEFORMAIE - strain
FIR brad energy
FIRST MOMENT (OF AN AREA) - ENERGIE POTENIAL - potential
moment static energy
FIBER fibr EPRUVET - specimen
FIXED SUPPORT ncastrare EPRUVET CU SUPRAFA
FLANGED JOINT - mbinare cu flane LUSTRUIT - polished specimen
FLAT END - terminaie aplatizat EROARE PROCENTUAL - percentage
FLAW - defect de structur error
FLEXURAL RIGIDITY - produsul EI EVIDENIEREA GRAFIC A
(ncovoiere) TUTUROR SARCINILOR CE
FLUID - fluid SOLICIT UN ANUMIT CORP N
FLYWHELL - volant ECHILIBRU MECANIC, INCLUSIV
FORCE for REACIUNILE - free-body diagram
FORCE-COUPLE SYSTEM - ansamblu EXCENTRIC - eccentrical
moment-for (torsor) EXCENTRICITATE - eccentricity
FORGED COMPONENT - pies forjat
FORMULA - formul F
FRACTURE rupere FALC (DE MENGHIN, ETC) - jaw

231
Strength of Materials

ENGLISH - ROMANIAN ROMANIAN ENGLISH


FRACTURE CRITERIA - teorii de FIABILITATE reliability
rupere FIBR - fiber
FRACTURE MECHANICS - mecanica FIBRA MEDIE DEFORMAT - elastic
ruperii curve
FRAME - cadru FIER PUR - pure-iron
FREE END - capt liber FIGUR - figure
FREE SURFACE - suprafa liber FIR wire
FREE-BODY DIAGRAM evidenierea FIR DE NYLON - nylon thread
grafic a tuturor sarcinilor ce solicit un FISUR - crack
anumit corp n echilibru mecanic, FISUR DE OBOSEAL - fatigue crack
inclusiv reaciunile FISURI MICROSCOPICE - microscopic
FREQUENCY frecven cracks
FREQUENCY OF ROTATION - viteza FLAMBA (A-) -buckle (to-)
de rotaie FLAMBAJ buckling
FREQUENTLY ENCOUNTERED FLUAJ - creep
frecvent ntlnit FLUID fluid
FRUSTRUM OF A CIRCULAR CONE FLUXUL TENSIUNILOR TANGENI-
- trunchi de con ALE - shear flow
FULLY APPLIED aplicat direct la FONT - cast iron
valoarea nominal (despre o for ,..,etc) FORFECARE - shear
FUNCTION OF X funcie de x FORFECARE DUBL - double shear
FORFECARE SIMPL - single shear
G FORMUL - formula
GAP - interstiiu, imperfeciune de FORMULA LUI EULER - Eulers
montaj formula
GEAR - roat dinat FORMULE EMPIRICE - empirical
GENERAL STATE OF STRESS - stare formulas
complex de tensiune FOR - force
GENERALIZED HOOKES LAW - FOR AXIAL - axial load
legea lui Hooke generalizat FOR AXIAL APLICAT N
GLASS sticl CENTRUL DE GREUTATE centric
GLUED JOINT mbinare prin lipire axial load
GRAIN - grunte FOR AXIAL APLICAT
GRAPH grafic EXCENTRIC - eccentric axial loading
GROUND sol FOR CAPABIL - largest allowable
force
H FOR CONCENTRAT-concentrated
HALF-CYLINDER jumtate de cilindru load
HEAT TREATMENT tratament termic FOR CRESCTOARE - increasing
HEIGHT nlime load
HELIX - elice, spiral (n spaiu) FOR ORIZONTAL -horizontal load
HEXAGON hexagon FOR TIETOARE - shearing force
HIGH-CARBON STEEL oel cu FORE DISTRIBUITE - distributed

232
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


coninut ridicat de carbon forces
HIGH-STRENGTH STEEL oel de FORE ELEMENTARE - elementary
rezisten ridicat forces
HINGED BEAM - grind cu articulaie FORE EXTERIOARE - external forces
HINT - indicaie (n rezolvarea unei FORE TRANSVERSALE - transverse
probleme) forces; transverse load
HOLE gaur FRECVENT NTLNIT - frequently
HOLLOW CYLINDER - cilindru tubular encountered
HOMOGENEOUS MATERIAL - FRECVEN - frequency
material omogen FRETARE binding
HOOKES LAW - legea lui Hooke FUNCIE ANALITIC - analytical
HOOP STRESS - tensiune circumfe- function
renial (vase cu perei subiri) FUNCIE DE GRADUL 2 - second
HORIZONTAL DEFLECTION sgeata degree function
pe direcie orizontal FUNCIE DE GRADUL 3 - third degree
HORIZONTAL LOAD - for orizontal function
HORIZONTAL PLANE - plan orizontal FUNCIE DE x - function of x
HORSEPOWER - cal putere FUNCIE TREAPT - step function
HYDRAULIC JACK - cric hidraulic FUNCII SINGULARE - singularity
HYDROSTATIC PRESSURE presiune functions
hidrostatic
HYPOTENUSE ipotenuz G
GAUR - hole
I GAUR CIRCULAR - circular hole
IMPACT FACTOR multiplicator de GAUR CONIC - taper hole
impact GENERATOR ELECTRIC - electric
IMPACT LOADING sarcin aplicat generator
cu oc GRIND EXTRUDAT CU PEREI
IMPULSE - impuls SUBIRI- thin-walled extruded beam
INCREASING LOAD - for cresctoare GTUIRE - necking
INERTIA - inerie GRAD DE PRECIZIE - degree of
INFLUENCE COEFFICIENTS - accuracy
coeficieni de influen GRADE CELSIUS (FAHRENHEIT)
INNER DIAMETER - diametru interior degrees celsius (fahrenheit)
INTEGRATION integrare GRAFIC - graph
INTERMEDIATE COLUMNS - bare GRUNTE - grain
(zvelte) ce flambeaz n domeniul elasto- GREUTATE PROPRIE - own weight
plastic GREUTATE SPECIFIC - specific
INTERSECT (TO-) intersecta (a-) weight
INVERSELY PROPORTIONAL - invers GRIND - beam
proporional GRIND CU ARTICULAIE - hinged
ISOTROPIC MATERIAL - material beam
izotrop GRIND CU INIM PLIN -web beam

233
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


J GRIND CU ZBRELE - truss
JAW - falc (de menghin, etc) GRIND DE EGAL REZISTEN-
JUST TO THE RIGHT OF POINT B n beam of constant strength
seciunea B dreapta GRIND DE LIME UNITATE-beam
unit width
K GRIND DE LEMN - wooden beam
KERN smbure central GRIND DE SECIUNE DREPTUN-
KINETIC ENERGY - energie cinetic GHIULAR CU LIME MIC-
narrow rectangular beam
GRIND DIN BETON ARMAT -
L reinforced concrete beam
LABORATORY laborator GRIND DIN PROFIL CU TALP
LARGEST ALLOWABLE FORCE - LAT- wide-flanged beam
for capabil GRIND EXTRUDAT CU PEREI
LATERAL STRAIN deformaie SUBIRI - thin-walled extruded beam
specific transversal GRIND N CONSOL - cantilever
LAW lege beam
LEAF SPRING - arc lamelar GRIND SIMPLU REZEMAT -simply
LENGTH lungime supported beam
LEVER prghie GRIND SIMPLU STATIC NEDETER-
LINEAR DISTRIBUTION distribuie MINAT - statically indeterminate beam
liniar to the first degree
LINEAR HOMOGENEOUS GRIND SUPRANCRCAT -
DIFFERENTIAL EQUATION - ecuaie overstressed beam
diferenial omogen liniar GRINZI STATIC NEDETERMINATE -
LINEAR NONHOMOGENEOUS statically indeterminate beams
DIFFERENTIAL EQUATION OF THE GRINZI ZVELTE - slender beams
SECOND ORDER WITH CONSTANT GROSIME thickness
COEFFICIENTS - ecuaie diferenial
neomogen liniar de ordinul doi, cu H
coeficieni constani HEXAGON hexagon
LOAD POINT OF APPLICATION -
punct de aplicaie a sarcinii I
LOADING ncrcare IMPULS - impulse
LOCATION - loc poziie INDICAIE (N REZOLVAREA UNEI
LOGARITHMIC SCALE scar PROBLEME) - hint
logaritmic INEL - ring
LONG COLUMNS - bare (zvelte) ce INERIE inertia
flambeaz n domeniul elastic INSTABIL - unstable
LONG CRACK macrofisur INSTALAIE EXPERIMENTAL -
LONGITUDINAL SHEARING experimental equipment
STRESSES - tensiuni tangeniale INTEGRAL DUBL - double integral
longitudinale INTEGRARE integration
INTEGRARE N RAPORT CU O
234
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


LONGITUDINAL SPACING OF THE SINGUR VARIABIL - simple
WELDS IS OFmm - dispunerea pe integration
lungime a sudurilor se face din...n...mm INTERSECTA (A-) - intersect (to-)
LOW - CARBON STEEL - oel cu INTERSTIIU - clearance
coninut sczut de carbon INTERSTIIU, IMPERFECIUNE DE
LOW - STRENGTH STEEL - oel de MONTAJ gap
rezisten sczut INVERS PROPORIONAL - inversely
L-SHAPED MACHINE ELEMENT - proportional
pies n form de L IPOTENUZA hypotenuse
IPOTEZA SIMPLIFICATOARE -
M simplifying assumption
MACHINE COMPONENTS - organe de
maini
MACHINE TOOL - main unealt MBINARE CAP LA CAP - splice
MACROSCOPIC CRACKS - fisuri MBINARE CU FLANE - flanged joint
macroscopice MBINARE PRIN LIPIRE - glued joint
MASS - masa MPINGE (A-) - push (to-)
MATERIAL - material N SECIUNEA B DREAPTA - just to
MAXIMUM ALLOWABLE SPEED - the right of point B
vitez maxim admisibil N SENS INVERS ACELOR DE
MAXIMUM DEFLECTION - sgeata CEASORNIC (SENS TRIGONOM.)
maxim - counterclockwise
MAXIMUM-DISTORTION-ENERGY N SENSUL ACELOR DE CEASOR-
CRITERION - teoria energiei maxime NIC - clock-wise
modificatoare de form NLIME - height
MAXIMUM-NORMAL-STRESS NCRCARE - loading
CRITERION -teoria tensiunii normale NCASTRARE - fixed support
maxime NCERCARE - test
MAXIMUM-SHEARING-STRENGTH NCERCARE LA TRACIUNE - tensile
CRITERION - teoria tensiunii tangeniale test
maxime NCOVOIERE - bending
MAXWELLS RECIPROCAL NCOVOIERE OBLIC - unsymmetric
THEOREM - teorema reciprocitii bending
deplasrilor (Maxwell) NCOVOIERE PUR - pure bending
MECHANICAL WORKING - prelucrare NFURARE - wrap
mecanic NSUMARE ALGEBRIC - algebraic
MEMBER - element (component al unei sum
structuri) NTINDE (A-), TRAGE (A-) - stretch
MEMBRANE - membran NVELI - shell
MEMBRANE ANALOGY - analogie cu
membrana (torsiune) J
METHODS OF ANALYSIS - metode de JGHEAB, CUZINET DE REAZEM
analiz cradle

235
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


MICROSCOPIC CRACKS - fisuri JUMTATE DE CILINDRU - half-
microscopice cylinder
MILD STEEL - oel moale (cu coninut
sczut de carbon)
MODULUS OF ELASTICITY - modul
L
de elasticitate LABORATOR - laboratory
MODULUS OF RIGIDITY - modul de LIME - width
rigiditate LEGTUR - connection
MOHRS CIRCLE - cercul lui Mohr LEGE - law
MOHRS CIRCLE FOR PLANE LEGEA LUI HOOKE GENERALIZA-
STRESS - cercul lui Mohr pentru starea T - generalized Hookes law
plan de tensiune LEGEA LUI HOOKE - Hookes law
MOHRS CRITERION - criteriul lui LIMITA DE ELASTICITATE - elastic
Mohr limit
MOHRS CIRCLE FOR PLANE LIMITA DE PROPORIONALITATE -
STRAIN - cercul lui Mohr pentru starea proportional limit
plan de deformaie LINIE DREAPT - straight line
MOMENT OF INERTIA OF THE LINIE PUNCTAT - dashed line
SECTION ABOUT THE PRINCIPAL LOC POZIIE - location
CENTROIDAL AXIS - moment de LOVI (A-) - strike (to-)
inerie al seciunii n raport cu axa LUCRU MECANIC - work
central principal a seciunii LUCRU MECANIC ELEMENTAR -
elementary work
LUNGIME - length
N LUNGIME DE FLAMBAJ - effective
NAIL - cui length
NARROW RECTANGULAR BEAM -
grinda de seciune dreptunghiular cu
lime mic M
NECKING - gtuire MACARA - crane
NEGATIVE (POSITIVE) SIGN - semn MACROFISUR - long crack
negativ (pozitiv) MANON, COLIER - collar
NEGLECTING THE TERM MANTA (CMA) DE AL-aluminum
CONTAINING - neglijnd termenul ce-l shell
conine pe MARC TENSOMETRIC -strain gage
NEUTRAL AXIS - ax neutr MARGINEA (MUCHIA ) GRINZII-edge
NEUTRAL SURFACE - suprafa neutr of the beam
NORMAL STRAIN - deformaie MRIME ADIMENSIONAL -
specific longitudinal dimensionless quantity
NORMAL STRESS - tensiune normal MRIME SCALAR - scalar quantity
NUMERICAL VALUE - valoare MRIME VECTORIAL - vectorial
numeric quantity
NUT - piuli MASA - mass
NYLON THREAD - fir de nylon

236
Strength of Materials typical key words and phrases

ENGLISH - ROMANIAN ROMANIAN ENGLISH


O MAIN DE NCERCRI LA
OBLIQUE PLANE - plan nclinat TORSIUNE - torsion testing machine
OBLIQUE SECTION - seciune oblic MAINA UNEALT - machine tool
ORDINATE - ordonat MATERIAL - material
ORIGINAL EQUILIBRIUM - echilibru MATERIAL ELASTOPLASTIC -
iniial elastoplastic material
OUTER DIAMETER - diametru exterior MATERIAL IZOTROPIC - isotropic
OVERSTRESSED BEAM - grind material
suprancrcat MATERIAL OMOGEN - homogeneous
OWN WEIGHT - greutate proprie material
MATERIALE DUCTILE - ductile
P materials
PARABOLA - parabol MATERIALE FRAGILE - brittle
PARABOLICAL DISTRIBUTION - materials
distribuie parabolic MECANICA RUPERII - fracture
PARABOLOID OF REVOLUTION - mechanics
paraboloid de revoluie MEMBRANA - membrane
PARALLEL-AXIS THEOREM -teorema MENGHIN - vice (vise)
axei paralele (relaiile lui Steiner) METODA TENSIUNII ADMISIBILE -
PARALLELOGRAM - paralelogram allowable - stress method
PARENTHESE -parantez METODE DE ANALIZ - methods of
PARTIAL DERIVATIVE - derivat analysis
parial MICROFISUR - short crack
PEG - tift, cep MIEZ DE OEL - steel core
PENSTOCK - conducta de evacuare MIEZ ELASTIC - elastic core
PERCENTAGE ERROR - eroarea MODUL DE ELASTICITATE -modulus
procentual of elasticity
PERMANENT DEFORMATION - MODUL DE ELASTICITATE TRANS-
deformaie permanent VERSAL (G) -shear modulus
PERPENDICULAR - perpendicular MODUL DE REZISTEN LA
PIN-CONNECTION - articulaie NCOVOIERE AL SECIUNII (N
PIN-ENDED COLUMN - bar zvelt DOMENIUL ELASTIC) - elastic section
dublu articulat la capete modulus
PIPE - eav MODUL DE RIGIDITATE - modulus of
PLANE OF ARBITRARY rigidity
ORIENTATION - plan cu o orientare MOMENT DE NCOVOIERE-bending
oarecare moment
PLANE OF SYMMETRY - plan de MOMENT DE INERIE AL
simetrie SECIUNII N RAPORT CU AXA
PLANE STATE OF STRAIN - stare CENTRAL PRINCIPAL A
plan de deformaie SECIUNII - moment of inertia of the
PLANE STATE OF STRESS -stare plan section about the principal centroidal axis
de tensiune
237
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


PLANE STRAIN STATE - starea plan MOMENT DE INERIE POLAR - polar
de deformaie moment of inertia
PLANE STRESS STATE - starea plan MOMENT DE TORSIUNE - torque
de tensiune (twisting couple)
PLANK - scndur MOMENT STATIC - first moment (of an
PLASTIC DEFORMATION - deformaie area)
plastic MULTIPLICATOR DE IMPACT-impact
PLASTIC HINGE - articulaie plastic factor
PLASTIC ZONE - zona plastic
PLATE - plac N
PLOT (TO-) - trasa (a-) NEGLIJND TERMENUL CE-L
PLYWOOD - placaj CONINE PE- neglecting the term
POINT OF INFLECTION - punct de containing
inflexiune NIT - rivet
POISSONS RATIO - coeficientul lui NIVELUL TENSIUNII - stress level
Poisson
POLAR MOMENT OF INERTIA - O
moment de inerie polar OBOSEAL - fatigue
POLISHED SPECIMEN - epruvet cu ORDONAT - ordinate
suprafa lustruit ORGANE DE MAINI - machine
POLYSTYRENE - polistiren components
POOR MAINTENANCE - proasta OEL - steel
ntreinere (a utilajului, structurii, etc. ) OEL ALIAT, CLIT I REVENIT -
PORTION - poriune (de bara, etc) quenched and tempered alloy steel
POST - reazem, suport OEL CLIT - quenched-steel
POTENTIAL ENERGY - energie OEL CU CONINUT RIDICAT DE
potenial CARBON - high-carbon steel
POWER - putere OEL DE REZISTEN RIDICAT-
PRINCIPAL AXES - axe principale high-strength steel
PRINCIPAL CENTROIDAL AXES OF OEL DE REZISTEN SCZUT -
THE CROSS SECTION - axe centrale low-strength steel
principale de inerie ale seciunii OEL LAMINAT - rolled steel
transverasale OEL MOALE (CU CONINUT
PRINCIPAL PLANES AND SCZUT DE CARBON) - mild steel
PRINCIPAL STRESSES - plane OEL RECOPT - annealed steel
principale i tensiuni principale OEL REVENIT - tempered steel
PRINCIPAL STRESSES - tensiuni
principale P
PRISMATIC BAR - bar prismatic PARABOL - parabola
PROBLEMS TO BE ASSIGNED - PARABOLOID DE REVOLUIE -
probleme propuse spre rezolvare paraboloid of revolution

238
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


PROJECTION - proiecie PARALELOGRAM - parallelogram
PROPAGATE (TO-) - propaga (a se -) PARANTEZ - parenthese
PROPORTIONAL LIMIT - limit de PTRAT CU LATURA a =- square of
proporionalitate side a =
PULL (TO-) -(a) trage PERETE DE GROSIME CONSTANT-
PULLEY - roat de transmisie uniform wall thickness
PURE BENDING - ncovoiere pur PERPENDICULAR - perpendicular
PURE-IRON - fier pur PIATR - stone
PUSH (TO-) - a mpinge PIES FORJAT - forged component
PIES N FORM DE L - L shaped
Q machine element
QUENCHED, TEMPERED ALLOY PIES LAMINAT - rolled component
STEEL - oel aliat, clit i revenit PRGHIE - lever
QUENCHED-STEEL - oel clit PIULI - nut
R PLAC - plate
RADIAN - radian PLACAJ - plywood
RADIUS - raz PLAN CU O ORIENTARE OARECARE
RADIUS OF CURVATURE - raz de - plane of arbitrary orientation
curbur PLAN DE SIMETRIE - plane of
RADIUS OF GYRATION - raz de symmetry
inerie PLAN NCLINAT - oblique plane
RATIO - raport PLAN ORIZONTAL - horizontal plane
REACTION - reaciune PLAN VERTICAL - vertical plane
REASONING - raionament PLANE PRINCIPALE I TENSIUNI
RECTANGULAR COORDINATES - PRINCIPALE - principal planes and
coordonate rectangulare principal stresses
RECTANGULAR CROSS SECTION - POLISTIREN - polystyrene
seciune dreptunghiular PORIUNE (DE BAR, ETC) - portion
REDUNDANT REACTION - reaciune POZIIE DE ECHILIBRU - equilibrium
suplimentar (static nedeterminat) position
REINFORCED CONCRETE BEAM - PRACTICA INGINEREASC -
grind din beton armat engineering practice
RELATIONS AMONG LOAD, SHEAR PRELUCRARE MECANIC -
AND BENDING MOMENT - relaii mechanical working
ntre sarcin, for tietoare i moment PRELUCRAT LA RECE - cold-worked
ncovoietor PRESIUNE HIDROSTATIC -
RELATIVE DISPLACEMENT - hydrostatic pressure
deplasare relativ PRINCIPIUL LUI SAINT-VENANT -
RELIABILITY - fiabilitate Saint-Venants principle
REPEATED LOADING - sarcin ciclic PRINCIPIUL SUPRAPUNERII DE
RESIDUAL STRESSES - tensiuni EFECTE - superposition method
remanente

239
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


RESILIENCE - rezilien PROASTA NTREINERE (A
RESULTANT - rezultant UTILAJULUI, STRUCTURII, ETC. ) -
RHOMBUS - romb poor maintenance
RIGHT-HAND SUPPORT - suportul din PROBLEM REZOLVAT - sample
dreapta problem
RIGID BODY - corp rigid PROBLEME PROPUSE SPRE
RING - inel REZOLVARE - problems to be assigned
RIVET - nit PRODUSUL EI (NCOVOIERE) -
ROD - tij flexural rigidity
ROD UNDER ITS OWN WEIGHT - tij PROFIL CU TALP LAT - wide-
sub greutate proprie flanged
ROLLED COMPONENT -pies laminat PROFILE DIN OEL LAMINAT -
ROLLED STEEL - oel laminat rolled-steel shapes
ROLLED-STEEL SHAPES - profile din PROIECTARE - design
oel laminat PROIECTAREA ARBORILOR DE
ROLLING - rostogolire TRANSMISIE - design of transmission
RULE - regul shafts
RUPTURE - rupere PROIECIE - projection
RUSTING - ruginire PROPAGA (A SE -) - propagate (to-)
PUNCT DE INFLEXIUNE - point of
S inflection
SAINT-VENANTS PRINCIPLE - PUNCTUL DE APLICAIE AL SAR-
principiul lui Saint-Venant CINII - load point of application
SAMPLE PROBLEM problem PUTERE - power
rezolvat PUTREZIRE - decay
SAVINGS OF MATERIAL - economii
de material R
SCALAR QUANTITY - mrime scalar RADIAN - radian
SECOND DEGREE FUNCTION - RAPORT - ratio
funcie de gradul 2 RAIONAMENT - reasoning
SECOND - ORDER LINEAR RAZ - radius
DIFFERENTIAL EQUATION - ecuaie RAZ DE CURBUR - radius of
diferenial de ordinul doi curvature
SEMIMAJOR AXIS - semiaxa mare (a RAZ DE INERIE - radius of gyration
unei elipse) REACIUNE - reaction
SEMIMINOR AXIS - semiaxa mic (a REACIUNE SUPLIMENTAR
unei elipse) (STATIC NEDETERMINAT) -
SHAFT-DISK-BELT ARRANGEMENT redundant reaction
-sistem arbore - roat de curea - curea REAZEM, SUPORT - post
SHEAR - forfecare RECIPIENT CU AER COMPRIMAT-
SHEAR CENTER - centrul de forfecare compressed-air tank
SHEAR DIAGRAM - diagram de fore REGUL - rule
tietoare

240
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


SHEAR FLOW - fluxul tensiunilor RELAIA LUI NAVIER
tangeniale (NCOVOIERE) - elastic flexure formula
SHEAR MODULUS - modul de RELAII NTRE SARCIN, FOR
elasticitate transversal (G) TIETOARE I MOMENT NCOVO-
SHEARING FORCE - for tietoare IETOR - relations among load, shear and
SHEARING STRAIN - deformaie bending moment
specific transversal (lunecare specific) REZILIEN - resilience
SHEARING STRESS - tensiune REZISTENA LA COROZIUNE -
tangenial corrosion resistance
SHEARING STRESSES ON REZISTENA LA CURGERE - yield
TRANSVERSE SECTION - tensiuni strength
tangeniale pe seciuni transversale REZISTENA LA OBOSEAL -
SHELL - nveli endurance limit
SHORT COLUMNS - bare (zvelte) de REZISTENA LA RUPERE - breaking
lungime mic, la care nu apare pericolul strength
de pierdere a stabilitii REZISTENA MATERIALELOR -
SHORT CRACK - microfisur strength of materials
SIMILAR TRIANGLES - triunghiuri REZULTANT - resultant
asemenea RIGIDITATE - stiffness
SIMPLE INTEGRATION - integrare n ROAT - wheel
raport cu o singur variabil ROAT DE TRANSMISIE - pulley
SIMPLIFYING ASSUMPTION - ipotez ROAT DINAT - gear
simplificatoare ROMB - rhombus
SIMPLY SUPPORTED BEAM - grind ROSTOGOLIRE - rolling
simplu rezemat ROTIRE - slope
SINGLE SHEAR - forfecare simpl ROZET TENSOMETRIC - strain
SINGULARITY FUNCTIONS - funcii rosette
singulare RUGINIRE - rusting
SLENDER BEAMS - grinzi zvelte RULMENT - bearing
SLENDER MEMBER - bar (element) RUPERE - fracture; rupture
zvelt RUPTURA, RUPERE, AVARIE, DIS-
SLENDERNESS RATIO - coeficient de TRUGERE, CEDARE, DEFECTARE,
zveltee NTRERUPERE, ETC. - failure
SLIP - alunecare
SLOPE - pant
SLOWLY INCREASING LOAD - S
sarcin uor cresctoare SGEAT - deflection
SMOOTH HORIZONTAL SURFACE - SGEAT MAXIM - maximum
suprafa orizontal neted deflection
SOLID BODY - corp solid SGEAT PE DIRECIE ORIZON-
SOLID CYLINDRICAL SHAFT - arbore TAL - horizontal deflection
de transmisie cilindric masiv (plin) SGEAT PE DIRECIE VERTICAL
SOLUTION - soluie (la o problem) -vertical deflection

241
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


SPAN - deschidere (a unei grinzi) SGEATA I ROTIREA N PUNCTUL
SPECIFIC WEIGHT - greutate specific A - deflection and slope at point A
SPECIMEN - epruvet AIB - washer
SPHERICAL PRESSURE VESSELS - SARCIN (FOR) APLICAT N
vase de presiune sferice CENTRUL DE GREUTATE-centric load
SPLICE - mbinare cap la cap SARCIN APLICAT CU OC -impact
SQUARE CROSS SECTION - seciune loading
transversal ptratic SARCIN CAPABIL - allowable load
SQUARE OF SIDE a=- ptrat cu SARCIN CICLIC - cyclic load;
latura a= repeated loading
STABILITY - stabilitate SARCIN CRITIC - critical load
STABILITY OF STRUCTURES - SARCIN DINAMIC - dynamic load
stabilitatea structurilor SARCIN FICTIV - fictitious or
STABLE - stabil () dummy load
STRAIN GAGE - marc tensometric SARCIN NECUNOSCUT - unknown
STATICALLY INDETERMINATE - load
static nedeterminat () SARCIN STATIC ECHIVALENT -
STATICALLY INDETERMINATE equivalent static load
BEAMS - grinzi static nedeterminate SARCIN UOR CRESCTOARE -
STATICALLY INDETERMINATE slowly increasing load
BEAM TO THE FIRST DEGREE - SARCINI DISTRIBUITE - distributed
grinda simplu static nedeterminate loads
STATICALLY INDETERMINATE SCAR LOGARITMIC - logarithmic
SHAFTS - arbori static nedeterminai scale
STATICS - statica SCNDUR - plank
STEAM TURBINE - turbin cu abur SECIONA (A-) - cut (to-)
STEEL - oel SECIUNE CIRCULAR -circular cross
STEEL ANGLE - cornier din oel section
STEEL CABLE - cablu de oel SECIUNE DREPTUNGHIULAR -
STEEL CORE - miez de oel rectangular cross section
STEP FUNCTION - funcie treapt SECIUNE INELAR - annular cross
STIFFNESS - rigiditate section
STONE - piatr SECIUNE OBLIC - oblique section
STRAIGHT BAR - bar dreapt SECIUNE TRANSVERSAL CONS-
STRAIGHT LINE - linie dreapt TANT - uniform cross-section
STRAIN - deformaie specific SECIUNE TRANSVERSAL DE
STRAIN ENERGY - energie de FORM TRAPEZOIDAL - trapezoidal
deformaie cross section
STRAIN ENERGY DENSITY - energie SECIUNE TRANSVERSAL DE
specific de deformaie FORM TRIUNGHIULAR - triangular
STRAIN ENERGY IN BENDING - cross section
energie de deformaie la ncovoiere

242
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


STRAIN ENERGY IN TORSION - SECIUNE TRANSVERSAL ELIP-
energie de deformaie la torsiune TIC - elliptic cross section
STRAIN ROSETTE -rozet tensometric SECIUNE TRANSVERSAL P-
STRAIN-HARDENING - ecruisare TRATIC - square cross section
STRAP - band, curea SECIUNE TRANSVERSAL VA-
STRENGTH OF MATERIALS - RIABIL - variable cross section
rezistena materialelor SEMIAXA MARE (A UNEI ELIPSE) -
STRESS - tensiune semimajor axis
STRESS AT A POINT - tensiunea dintr- SEMIAXA MIC (A UNEI ELIPSE) -
un punct semiminor axis
STRESS COMPONENTS -componentele SEMN NEGATIV (POZITIV) - negative
tensiunii (positive) sign
STRESS CONCENTRATORS SMBURE CENTRAL - kern
-
concentratori de tensiune SIMETRIE - symmetry
STRESS LEVEL - nivelul tensiunii SRM (FIR) DE SECIUNE CIRCU-
STRESS-CONCENTRATION FACTOR LAR - circular wire
- coeficient (factor) de concentrare aSISTEM ARBORE - ROAT DE
tensiunilor CUREA-CUREA - shaft-disk-belt
STRESSES UNDER COMBINED arrangement
LOADINGS - tensiuni n cazul SOL - ground
solicitrilor compuse SOLICITARE AXIAL - axial loading
STRESS-STRAIN DIAGRAM SOLUIE (LA O PROBLEM)-solution
-
diagrama tensiune-deformaie specificSTABIL - stable
STRETCH (TO-) - ntinde, trage (a-) STABILITATE - stability
STRIKE (TO-) -lovi (a-) STABILITATEA STRUCTURILOR -
STRUCTURE - structur stability of structures
STUBBY COLUMN - bar de lungime STARE COMPLEX DE TENSIUNE -
mic general state of stress
SUPERPOSITION METHOD -principiul STARE DE TENSIUNE BIAXIAL -
suprapunerii de efecte biaxial stress condition
SYMMETRY - simetrie STARE PLAN DE DEFORMAIE -
plane state of strain
T STARE PLAN DE TENSIUNE - plane
TABLE - tabel state of stress
TAPER HOLE - gaur conic STARE SPAIAL DE TENSIUNE -
TAPERED BAR - bar conic three-dimensional state of stress
TEMPERATURE CHANGES - variaii STARE PLAN DE DEFORMAIE -
de temperatur plane state of strain
TEMPERED STEEL - oel revenit STATIC NEDETERMINAT () -
TENSILE STRESS -tensiune de traciune statically indeterminate
TENSILE TEST - ncercare la traciune STATICA statics
TEST - ncercare STICL - glass

243
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


THEORY OF ELASTICITY - teoria STRAT DE ARAM - brass layer
elasticitii STRUCTUR - structure
THERMAL STRAIN - deformaie STRUCTUR DEFORMABIL -
specific cauzat de variaii de deformable structure
temperatur STRUCTURI BIDIMENSIONALE-two-
THICKNESS - grosime dimensional structures
THIN-WALLED EXTRUDED BEAM - SUDUR - weld
grind extrudat cu perei subiri SUPORTUL DIN DREAPTA -right-hand
THIN-WALLED HOLLOW SHAFTS - support
tuburi cu perei subiri SUPRAFA DE CONTACT - bearing
THIN-WALLED MEMBERS - bare cu surface
perei subiri SUPRAFA LIBER - free surface
THIRD DEGREE FUNCTION - funcie SUPRAFA NEUTR -neutral surface
de gradul 3 SUPRAFA ORIZONTAL NETE-
THREE-DIMENSIONAL STATE OF D - smooth horizontal surface
STRESS - stare spaial de tensiune
TIMBER - cherestea
TORQUE (TWISTING COUPLE) - TIFT, CEP peg
moment de torsiune T
TORSION - torsiune TABEL - table
TORSION TESTING MACHINE - TENSIUNE - stress
main de ncercri la torsiune TENSIUNE (PRESIUNE) DE CON-
TORSIONAL SPRING - arc de torsiune TACT - bearing stress
TOTAL DEFORMATION - deformaie TENSIUNE ADMISIBIL - allowable
total stress
TRANSMISSION SHAFT - arbore de TENSIUNE CIRCUMFERENIAL
transmisie (VASE CU PEREI SUBIRI) - hoop
TRANSVERSE FORCES - fore stress
transversale TENSIUNE CRITIC - critical stress
TRANSVERSE CONTRACTION - TENSIUNE DE COMPRESIUNE -
contracie transversal compressive stress
TRANSVERSE LOAD - sarcin TENSIUNE DE TRACIUNE - tensile
transversal stress
TRAPEZOIDAL CROSS SECTION - TENSIUNE MEDIE - average stress
seciune transversal de form TENSIUNE NORMAL - normal stress
trapezoidal TENSIUNE TANGENIAL - shearing
TRIANGULAR CROSS SECTION - stress
seciune transversal de form TENSIUNEA DINTR-UN PUNCT-stress
triunghiular at a point
TRUSS grind cu zbrele TENSIUNI N CAZUL SOLICITRI -
LOR COMPUSE - stresses under
combined loading

244
Strength of Materials typical key words and phrases

ENGLISH ROMANIAN ROMANIAN ENGLISH


TWO-DIMENSIONAL STRUCTURES - TENSIUNI PRINCIPALE - principal
structuri bidimensionale stresses
TENSIUNI REMANENTE - residual
U stresses
UNIFORM CROSS-SECTION - seciune TENSIUNI TANGENIALE LONGI -
transversal constant TUDINALE - longitudinal shearing
UNIFORM WALL THICKNESS - perete stresses
de grosime constant TENSIUNI TANGENIALE PE SEC -
UNIT - unitate TIUNI TRANSVERSALE - shearing
UNKNOWN LOAD - sarcin stresses on transverse sections
necunoscut TEOREMA AXEI PARALELE
UNLOADING - descrcare (RELAIILE LUI STEINER) - parallel-
UNPREVENTABLE NATURAL axis theorem
CAUSES - cauze naturale imprevizibile TEOREMA LUI CASTIGLIANO -
UNSTABLE - instabil () Castiglianos theorem
UNSYMMETRIC BENDING - TEOREMA RECIPROCITII DEPLA-
ncovoiere oblic SRILOR (MAXWELL)- Maxwells
reciprocal theorem
V TEORIA ELASTICITII - theory of
VARIABLE CROSS SECTION -seciune elasticity
transversal variabil TEORIA ENERGIEI MAXIME MODI-
VECTOR - vector FICATOARE DE FORM - maximum-
VECTORIAL QUANTITY - mrime distortion-energy criterion
vectorial TEORIA TENSIUNII NORMALE
VELOCITY - vitez MAXIME - maximum-normal-stress
VERTICAL DEFLECTION - sgeata pe criterion
direcie vertical TEORIA TENSIUNII TANGENIALE
VERTICAL PLANE - plan vertical MAXIME - maximum-shearing-strength
VIBRATION - vibraie criterion
VICE (VISE) - menghin TEORII DE CURGERE - yield criteria
VOLUME - volum TEORII DE RUPERE - fracture criteria
TERMINAIE APLATIZAT -flat end
W TIJ-rod
WASHER - aib TIJ SUB GREUTATE PROPRIE - rod
WEB BEAM - grind cu inim plin under its own weight
WELD - sudur TORSIUNE - torsion
WHEEL - roat TRAGE (A-) -pull (to-)
WIDE-FLANGED - profil cu talp lat TRASA (A-) -plot (to-)
WIDE-FLANGED BEAM - grind din TRATAMENT TERMIC - heat treatment
profil cu talp lat TRIUNGHIURI ASEMENEA - similar
WIDTH - lime triangles
WIRE - fir TRUNCHI DE CON - frustum of a
WOODEN BEAM grind de lemn circular cone

245
Strength of Materials

ENGLISH ROMANIAN ROMANIAN ENGLISH


WOODEN MEMBERS - elemente din TUBURI CU PEREI SUBIRI - thin-
lemn walled hollow shafts
WORK - lucru mecanic TURBIN CU ABUR - steam turbine
WRAP - nfurare

Y EAV - pipe
YIELD CRITERIA - teorii de curgere EAV DE ALUMINIU-aluminum pipe
YIELD STRENGTH - rezistena la
curgere
U
YIELDING - curgere (a materialului) UNGHI - angle
UNGHI DE TORSIUNE - angle of twist
UNGHI DIEDRU - diedral angle
UNITATE - unit

V
VALOARE MEDIE - average value
VALOARE NUMERIC - numerical
value
VARIAII DE TEMPERATUR -
temperature changes
VAS CILINDRIC DIN ALAM -
cylindrical brass vessel
VASE DE PRESIUNE CILINDRICE -
cylindrical pressure vessels
VASE DE PRESIUNE SFERICE -
spherical pressure vessels
VECTOR - vector
VERIFICA (A-) - check (to-)
VIBRAIE - vibration
VITEZ - velocity
VITEZ DE ROTAIE - frequency of
rotation
VITEZ MAXIM ADMISIBIL -
maximum allowable speed
VITEZ UNGHIULAR - angular
velocity
VOLANT - flywhell
VOLUM - volume

Z
ZONA PLASTIC - plastic zone

246
Strength of Materials typical key words and phrases

247
BIBLIOGRAFIE

1. BEER, P., F., JHONSON, R., E. Mechanics of materials, Mc Graw Hill Inc., SUA, 1992.
2. BIA, C., ILIE, V. Rezistena Materialelor i Teoria elasticitii, Editura didactic i
Pedagogic, Bucureti, 1983.
3. BI, C., RADU, N., Gh., CIOFOAIA, V. Elemente de mecanica ruperii, Editura Macarie,
Trgovite, 1997.
4. BI, C. Puncte de vedere asupra oboselii mecanice, Editura Universitii Transilvania,
Braov, 2001.
5. BOLFA, T., ROCA, C., DUMITRIU, N., BI, C. Rezistena Materialelor, Braov, 1996.
6. BROEK, D. Elementary EngineeringFracture Mechanics, Martinus Nijhoff Publishers,
London, 1982.
7. BUZDUGAN, Gh. Rezistena materialelor, Editura Academiei, Bucureti, 1986.
8. CIOFAIA, V. Rezistena Materialelor i elemente de construcii industriale, Reprografia
Universitii din Braov, 1987.
9. CIOFOAIA, V., BOTI, M., DOGARU, F., CURTU, I. Metoda elementelor finite, Editura
Infomarket, Braov, 2001.
10. CIOFOAIA, V., CURTU, I. Teoria elasticitii corpurilor izotrope i anizotrope,
Universitatea Transilvania, Braov, 2000.
11. CIOFOAIA, V., TALPOI, A., BI, C. Teoria elasticitii i plasticitii, Braov, 1995.
12. CIOFOAIA, V., ULEA, M. - Teoria elasticitii i rezistena materialelor, Braov, 1992.
13. CURTU, I., CRIAN, R. - Rezistena materialelor i teoria elasticitii, curs i aplicaii,
partea I, Reprografia Universitii Transilvania, Braov, 1997.
14. CURTU, I., ROCA, C. 2288 probleme de rezistena materialelor, Reprografia Universitii
Transilvania, Braov, 1991.
15. CURTU, I., CRIAN, L. R., BI, C. - Rezistena materialelor i teoria elasticitii, curs i
aplicaii, partea a II - a, Universitatea Transilvania, Braov, 1998.
16. CURTU, I., BI, C. - Rezistena materialelor i teoria elasticitii, curs i aplicaii, partea
a III - a, Universitatea Transilvania, Braov, 2000.
17. CURTU, I., BI, C. - Rezistena materialelor i teoria elasticitii, curs i aplicaii, partea
a IV - a, Universitatea Transilvania, Braov, 2001.
18. DEUTSCH, I., GOIA, I., CURTU, I., NEAMU, T., SPERCHEZ, Fl. Probleme de rezistena
Materialelor, Ediia I, E.D.P., Bucureti,1979.
19. DEUTSCH, I., GOIA, I., CURTU, I., NEAMU, T., SPERCHEZ, Fl. Probleme de rezistena
Materialelor, Ediia a II - a, E.D.P., Bucureti,1983.
20. DEUTSCH, I., GOIA, I., NEAMU, T., SPERCHEZ, Fl. Probleme de rezistena
Materialelor, E.D.P., Bucureti,1980.
21. DIETER, G., E. Metalurgie mecanic, Editura Tehnic, 1970.
22. GOIA, I., A. Rezistena materialelor, Vol. I, Editura Transilvania, Braov, 2000.
23. MUNTEANU, M., Gh., RADU, N., Gh., POPA, Al., V. - Rezistena materialelor I, Reprografia
Universitii din Braov, 1989.
24. MUNTEANU, M., Gh., RADU, N., Gh., POPA, Al., V. - Rezistena materialelor II,
Reprografia Universitii din Braov, 1989.
25. NSTSESCU, V., BRSAN, Gh. - Rezistena materialelor Probleme vol. 1 i 2, Editura
Academiei Tehnice Militare, Bucureti, 1997.
26. RADU, Gh., N., MUNTEANU, M., Gh., BI, C. - Rezistena materialelor i elemente de teoria
elasticitii, Vol. I, Editura Macarie, Trgovite, 1995.
27. RADU, Gh., N., MUNTEANU, M., Gh., BI, C. - Rezistena materialelor i elemente de teoria
elasticitii, Vol. II, Editura Macarie, Trgovite, 1995.
28. TIMOSHENKO, S. History of Strength of Materials, Mc. Graw Hill, Book Company Inc.,
SUA, 1953.
29. ZHILUN, XU. Applied Elasticity, John Wiley & Soons, SUA, 1992.

Das könnte Ihnen auch gefallen