SlideShare une entreprise Scribd logo
1  sur  150
Introduction
External loads
Surface forces
Body Forces
Equilibrium
Internal loading / Internal Forces
Sign Conventions
Axial forces, SF, BM
General State of stress
General condition of loading
Components of stress
Analysis and design of any structure or machine involve two major
questions: (a) Is the structure strong enough to with­stand the loads
applied to it and (b) is it stiff enough to avoid excessive deformations
and deflections? In Statics, the members of a structure were treated as
rigid bodies; but actually all materials are deformable and this property
will henceforth be taken into account. Thus Strength of Materials may
be regarded as the statics of deformable or elastic bodies.
BRANCHES OF ENGINEERING MECHANICS
Both the strength and stiffness of a structural member are functions of
its size and shape and also of certain physical properties of the material
from which it is made. These physical properties of materials are
largely determined from experimental studies of their behavior in a
testing machine. The study of Strength of Materials is aimed at
predicting just how these geometric and physical properties of a
structure will influence its behavior under service conditions. The
applications of the subject are broad in scope and will be found in all
branches of engineering.
The primary objective of the Strength of material / Mechanic of
Material is the development of relationships between the loads applied
to a non­rigid body and the internal forces and deformations induced in
the body.
       Mechanics of materials is a branch of mechanics that develops
relationships between the external loads applied to a deformable body
and the intensity of internal forces acting within the body. This subject
is also concerned with computing the deformations of the body, and it
provides a study of the body's stability when the body is subjected to
external forces.
The forces that act on a structure include the applied
loads and the resulting reaction forces.
      The applied loads are the known loads that act on a
structure. They can be the result of the structure’s own
weight, occupancy loads, environmental loads, and so on.
The reactions are the forces that the supports exert on a
structure. They are considered to be part of the external
forces applied.
Forces on a structure can arise from many sources, such as the
structure's own weight, any objects placed on it, wind pressure and so
forth. Force is a vector quantity, that is, it has both magnitude and
direction. The SI unit of force is the Newton (N), which is defined as
the force required to impart an acceleration of one metre per second per
second to a mass of one kilogram (that is, 1 N = 1 kg m/s2). An object
placed on a structure will thus impart a vertical force equal to its mass
multiplied by the acceleration due to gravity (g = 9.81 m/s2).
Surface Forces. As the name implies, surface forces are caused by the
dir contact of one body with the surface of another. In all cases these
forces distributed over the area of contact between the bodies, Fig. 1­
1a. In particular, if this area is small in comparison with the total
surface area of the body then the surface force may be idealized as a
single concentrated force, which is applied to a point on the body, Fig.
1­1a. For example, this might be done to represent the effect of the
ground on the wheels of a bicycle when studying the loading on the
bicycle. If the surface loading is applied along a narrow area, the
loading may be idealized as a linear distributed load, w(s). Here the
loading is measured as having an intensity of force/length along the
area and is represented graphically by a series of arrows along the line s
Concentrated force
   idealization




Linear distributed
load idealization




                     Fig. 1-1 (a)
Fig. 1-1 (b)
s, Fig. 1­1a. The loading along the length of a beam is a typical
example of where this idealization is often applied, Fig. 1­1b. The
resultant force FR of w(s) is equivalent to the area under the distributed
loading curve, and this resultant acts through the centroid C or
geometric center of this area.
Body Force. A body force occurs when one body exerts a force on
another body without direct physical contact between the bodies.
Examples include the effects caused by the earth's gravitation or its
electromagnetic field. Although body forces affect each of the particles
composing the body, these forces are normally represented by a single
concentrated force acting on the body. In the case of gravitation, this
force is called the weight of the body and acts through the body's center
of gravity.
The forces on a body can also give rise to moments, which tend to cause
the body to rotate about an axis. The moment of a force about an axis is
simply equal to the magnitude of the force multiplied by the
perpendicular distance from the axis to the line of action of the force.
Consider, for example, the lever AB shown in Figure 2.1 (a). The effect
of the force P acting at B is to impart both a direct force P and a
moment M = Pa on the hinge at A, as shown in Figure 2.1 (b).
       For a general case of forces and moments in space each force in
the resultant of three forces Fx, Fy and Fz and similarly each moment is
the resultant of three couples Mx, My and Mz.
SUMMARY
Type of force system                            possible resultants
Collinear ……………………………………………                     Force
concurrent , coplanar …………………………….              Force
parallel , coplanar………………………………….               Force or a couple
 Nonconcurrent , nonparallel ,coplanar …………     Force or a couple
concurrent , noncoplanar…………………………              Force
parallel , noncoplanar …………………………….             Force or a couple
Nonconcurrent , nonparallel , noncoplanar……..   Force or a couple , or a force and a couple
The surface forces that develop at the supports or points of
support between bodies are called reactions. For two­dimensional
problems, i.e., bodies subjected to coplanar force systems, the supports
most commonly encountered are shown in Table 1­1. Note the symbol
used to represent each support and the type of reactions it exerts on its
contacting member. One possible way to determine a type of support
reaction is to imagine the attached member as being translated or
rotated in a particular direction. If the support prevents translation in a
given direction, then a force must be developed on the member in that
direction. Likewise, if rotation is prevented, a couple moment must be
exerted on the member. For example, a roller support only prevents
translation in the contact direction, perpendicular or normal to the
surface. Hence, the roller exerts a normal force F on the member at the
point of contact. Since the member is free to rotate about the roller, a
couple moment cannot be developed by the roller on the member at the
point of contact.
       Remember that the concentrated forces and couple moment
shown in Table 1­1 actually represent the resultants of distributed
surface forces that exist between each support and its contacting
member. Although it is these resultants that are determined in practice,
it is generally not important to determine the actual surface load
distribution, since the area over which it acts is considerably smaller
than the total surface area of the contacting member.



2.8    SUPPORTS AND EXTERNAL REACTIONS

In structural analysis terminology the consequence of subjecting a
structure to an action is termed the "response" of the structure. A
structure subjected to an action (whether static or dynamic) will move
(translate and/or rotate) indefinitely unless the action is resisted in some
way. "Supports" are provided as the means of preventing free
movement of the structure. A statically loaded structure will deform
into a new shape but remain attached to its supports and at rest in its
new position. A dynamically loaded structure will remain attached to
supports but its deformed shape will change as a function of time.
"Damping" will eventually bring it to rest but a finite length of time is
subjected to simultaneously applied loads and come to rest
instantaneously. Supports prevent free motion of the structure as a
whole by developing forces to counteract the load actions. The
counteracting forces are commonly termed "external reactions" or
simply "reactions."
Equilibrium of a body requires both a balance of forces, to prevent the
body from translating or moving along a straight or curved path, and a
balance of moments, to prevent the body from rotating. These
conditions can be expressed mathematically by the two vector
equations
                          ΣF =0
                                                               (1–1)
                      Σ MO = 0
Here, Σ F represents the sum of all the forces acting on the body, and Σ
MO is the sum of the moments of all the forces about any point O either
on or off the body. If an x, y, z coordinate system is established with the
origin at point O, the force and moment vectors can be resolved into
components along the coordinate axes and the above two equations can
be written in scalar form as six equations, namely,

  Σ Fx = 0            Σ Fy = 0            Σ Fz = 0
                                                                  (1–2)
 ΣMx = 0             ΣMy = 0             ΣMz = 0
Often in engineering practice the loading on a body can be
represented as a system of coplanar forces. If this is the case, and the
forces lie in the x–y plane, then the conditions for equilibrium of the
body can be specified by only three scalar equilibrium equations; that
is,


                          Σ Fx = 0

                          Σ Fy = 0                              (1–3)

                        Σ MO = 0
In this case, if point O is the origin of coordinates, the moments will be
directed along the z axis, which is perpendicular to the plane that
contains the forces.
       Successful application of the equations of equilibrium requires
complete specification of all the known and unknown forces that act on
the body. The best way to account for these forces is to draw the body's
free-body diagram. Obviously, if the free­body diagram is drawn
correctly, the effects of all the applied forces and couple moments can
be accounted for when the equations of equilibrium are written.
Consider a body of arbitrary shape acted upon by the forces shown in
Fig. 1­2. In statics, we would start by determining the resultant of the
applied forces 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, may apply inertia
forces to bring about dynamic equilibrium. Such cases are discussed
later under dynamic loading. For the present, we consider only cases
involving static equilibrium.
       In strength of materials, we make an additional investigation of
the internal distribution of the forces. This is done by passing an
Figure 1­2 Exploratory section a-a through loaded member.
exploratory section a­a through the body and exposing the internal
forces acting on the exploratory section that are necessary to maintain
the equilibrium of either segment. In general, the internal forces reduce
to a force and a couple that, for convenience, are resolved into
components that are normal and tangent to the section, as shown in Fig.
1­3.
       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 yet
ready to show why this is so, we shall prove it as we progress; in
particular, we shall prove it for normal forces in the next article. If the x
axis is normal to the section, the section is known as the x surface or,
more briefly, the x face.
Figure 1-3
Components of internal effects on exploratory section a-a.
The notation used in Fig. 1­3 identifies both the exploratory section and
the direction of the force or moment component. The first subscript
denotes the face on which the component acts; the second subscript
indicates the direction of the particular component. Thus Pxy is the force
on the x face acting in the y direction.

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

Pxx   Axial force. This component measures the pulling (or pushing)
      action perpendicular to the section. A pull rep­resents a tensile
      force that tends to elongate the member, whereas a push is a
      compressive force that tends to shorten it. It is often denoted by P.
Pxy, Pxz   Shear forces. These are components of the total resistance
           to sliding the portion to one side of the exploratory section
           past the other. The resultant shear force is usually
           designated by V, and its components by Vy and Vz to
           identify their directions.
Mxx        Torque. This component measures the resistance· to
           twisting the member and is commonly given the symbol T.
Mxy, Mxz   Bending moments.        These   components measure the
           resistance to bending the member about the y or z axes and
           are often denoted merely by My or Mz.
2.3 Internal forces in structures
So far, we have concentrated on the external forces applied to structures
the applied loads and the support reactions. In order for the structure to
transmit the external loads to the ground, internal forces must be
developed within the individual members. The aim of the design
process is to produce a structure that is capable of carrying all these
internal forces, which may take the form of axial forces, shear forces,
bending moments or torques.
        Consider first a two­dimensional beam where the applied forces
and reactions all lie in a single plane (Figure 2.14(a)). The internal
forces at a point in the structure can be found by splitting it at that point
and drawing free body diagrams for the two sides (Figure 2.14(b)). The
requirements of equilibrium state that not only must the resultant force
on the entire structure be zero, but the resultant on any segment of it
must also be zero. It is therefore clear that there must be forces acting at
the cut point, as shown. These are drawn on the free body diagrams of
the segmented structure as though they were external loads, but they are
in fact the internal forces in the beam. The forces can be thought of as
the external forces that would have to be applied to the cut beam in
order to produce the same deformations as in the original uncut beam.
The forces shown are an axial force T, a transverse force S, known as a
shear force, and a bending moment M.
For equilibrium at the cut point, the forces acting on the faces
either side of the cut must be equal and opposite; this means that, when
the two segments are put together to form the complete structure, there
is no resultant external load at that point.
        For a member in three­dimensional space, a total of six internal
forces must be considered, as shown in Figure 2.15. Here, there is again
an axial force T in the x direction, and the resultant shear force has been
resolved into components Sy and Sz parallel to the y and z axes

respectively. There are also moments about each of the three axes: My
tends to cause the structure to bend in the horizontal (x ­ z) plane; Mz
causes bending in the vertical (x ­ y) plane; Mx causes the member to
y




        x
z




            Figure 2.15
being, however, we will restrict ourselves to two­dimensional systems.

2.3.1 Sign convention for internal forces

Before defining the internal forces more fully, we need to extend the
sign convention introduced earlier. For a two­dimensional system,
positive forces act in the positive x and y directions, and a positive
moment about the z-axis acts from the positive x towards the positive y-
axis, that is anti­clockwise. We can also define a positive face of a
member as one whose outward normal is in a positive axis direction.
Thus, for the beam segment in Figure 2.16(a), the right­hand face is a
positive x-face, the top surface is a positive y-face and the other two
faces are negative.
We can now define a positive internal force as one which acts
either in a positive direction on a positive face or in a negative
direction on a negative face. Conversely, a negative force either acts in
a negative direction on a positive face or vice versa.
EQUILIBRIUM AND STRESS
We will not attempt a thorough review of equilibrium methods of statics, since
presumably the reader is familiar with analytical concepts that lead to computation of
equivalent­force systems and reactions for load­carrying members. For now the
discussion will be conceptual. Later, certain topics of statics will be reviewed as
considered pertinent.
         The beam­type structure of Fig. 1.1 is considered to be in equilib­rium with
six actions occurring at each end of the beam. There are six and only six possible
actions that can occur: three forces and three couples, which are evaluated using the
equilibrium equations of statics. A section passed through the beam anywhere along
its axis can be viewed as shown in Fig. 1.2. Internal forces and couples act on the cut
section as illustrated and have magnitudes necessary to produce equi­librium for the
free body. Specifically, we classify the forces and couples as in Fig. 1.3. The force
z




                                                          x

                   FIGURE 1.1
Beam­type structure showing the six actions of statics.
y
vector acting along the beam axis is shown in Fig. 1.3a and causes either tension or
compression on the section, depending upon its direction. The remaining two force
vectors (Fig. 1.3b) produce shear loading on the cut section that is characterized by
the forces acting tangent to the cut section as opposed to acting normal to the cut
section as in the case of the axial force. The three couples of Fig. 1.2 are illustrated in
Figs. 1.3c­e as vectors. The axial­couple vector of Fig. 1.3c represents a twisting
couple whose direction is determined using the right­hand­screw rule. The twisting
couple, referred to as torque, causes a shear action to occur on the cut cross section.
The couples of Figs. 1.3d and e are referred to as bending moments, and the vector
representation is interpreted as illustrated. It turns out that these couples cause a
combination of tension and com­pression on the cut section.
         The early chapters of this text are devoted to analytical methods for
computing the magnitude and direction of these six actions and then the computation
of the corresponding stresses. In particular, Chapter 2 deals with the action of Fig.
1.3a. Chapter 4 is devoted to the twisting couple of Fig. 1.3c applied to members of
circular cross section. Analytical methods for computing shear forces and bending
moments of Figs. 1.3b, d, and e are discussed in Chapter 5. Chapter 6 deals with the
methods for computing stress caused by the bending moments of Figs. 1.3d and e, and
Chapter 7 is devoted to the derivation of computational methods for the shear stress
produced by the shear forces of Fig. 1.3b. Load­carrying members subject to various
combina­tions of the forces and couples are analyzed in Chapter 8. This analysis is the
culmination of the study of equivalent force systems and their associated stresses.
INTERNAL STRESSES AND STRESS
   RESULTANTS (INTERNAL FORCES)


       Framed members subjected to load must develop internal
stresses to resist the loads and prevent a material failure. The integrated
effects of stresses are force quantities of a particular magnitude and
direction. A frequent terminology used to refer to these force quantities
is stress resultants. “Internal forces” is an alternative terminology.
Internal force, that can be created in framed members are categorized
into four types.
1.Axial force
        2.shear force
        3.flexural moment
        4.torsional moment

The determination of these force quantities is one of the basic aims of
structural analysis. The particular force quantities created in a given
structure depend upon its behavior. For a planar frame member, the
significant internal forces are the hear and axial direct forces and the
flexural moment. Figure 2.11 depicts the manner in which these
generalized forces are developed by the internal stresses. The resultants
of the shear stress distribution τ, and the axial stress distribution σ1,
also a major cause of deflections, and this subject is also discussed as is
deflection due to shear. Internal forces developed in each type of
framed structure are presented in Table 2.1.
Table 2.1. Internal Forces (Stress Resultants) for the
               Basic Structural Types

                  Schematic of Assumed
Structural Type                                        Description
                     Member Forces



Plane truss and                               Only axial     Force,   F 1,   is
  space truss                                 significant.
                                         F1



                                              Only in­plane shear force F1 ,
                                              flexural moment F2 , and axial
Beam and plane
    frame                                     force F3, are significant. For
                                              beams, axial force generally
                                     F3       does not exist.
                        F2
Schematic of Assumed
Structural Type                                              Description
                     Member Forces

                                    F1             Only shear force F1 and
                                                   flexural moment F2 (both
     Grid                                          transverse to the plane of the
                                                   grid) and the torsional moment
                      F2                      F3   F3 are significant.

                                F2 F5
                                                   All stress resultants are signif­
                                                   icant: axial force F1, shear
 Space Frame                                       forces F2 and F3 torsional
                      F3                           moments F4, and flexural
                                              F1
                                                   moments F5 and F6·
                           F6            F4
PROBLEM: ANALYZE THE TRUSS BY STIFFNESS
METHOD
ANALYSIS OF GRIDS

Definition:

A grid is a structure that has loads applied perpendicular to its plane.
The members are assumed to be rigidly connected at the joints.
A very basic example of a grid structure is floor system as shown
y-axis

                                                                                     Wzzj,δ zzj
                                                Wzzi,δ zzi

                                                                               Wzj,δ zj
                                     Wzi,δ zi

                                                                                                             x-axis
Wxxi,δ xxi   Wxi,δ xi                                                              Wxxj,δ xxj     Wxj,δ xj


                        Wyi,δ yi
                  is




                                                                   Wyj,δ yj
                  x
              z-a




                        Wyyi,δ yyi                                Wyyj,δ yyj

                                     FORCES AND DISPLACEMENT IN
                                        LOCAL COORDINATES
when integrated over the member cross section, are the transverse shear
force V and the longitudinal (or axial) force P, respectively. Flexural
stresses are the source of two internal forces. Tensile and compressive
region of this stress distribution,σ2 produce the compressive force C and
tensile force T, respectively. Unlike the resultant P. which acts at the
centroid. the resultants T and C are separated by a distance, and neither
of their lines of action coincides with the centroid. Consequently, the
net effect of T and C is the creation of an internal resisting moment, M.
       Shear and moment resultants that develop in an individual
framed member exhibit a rela­tionship to each other and to the
transverse loads applied to the member. These fundamental
interrelationships will be formulated in Chapter 3. Flexural moment is
It is common to refer to loads as “forces that are applied to a structure.”
"Applied" is taken to mean “having an identifiable location or point of
application.” Each of the six types of framed structures were illustrated
in Fig. 1.4. Inspection of that figure indicates the applied load types that
are included in each of the six idealized models.

1. Planar truss. Concentrated loads applied in two orthogonal
     directions at the joints.
2.    Space truss. Concentrated loads applied in three ol1hogonal
     directions at the joints.
3. Beams, a. Transverse loads and flexural moments that are applied
     along the member and in the plane of the beam. These can be
     concentrated or distributed in nature, b. concentrated moments
applied at the joints and acting in the plane of the beam.
• Grid. a. Concentrated or distributed loads (transverse to the plane of
  the grid) and flexural moments applied along the member length. b,
  Concentrated or distributed torsional moments acting along the
  member length. c, Concentrated loads (transverse to the plane of the
  frame) and out­of­plane moments applied at the joints.
• Planar frame. a, Transverse loads and flexural moments that are
  applied along the member and in the plane of the frame. These can
  be concentrated or distributed in nature. b, Concentrated loads in two
  orthogonal directions and in­plane mo­ments applied at the joints
  and in the plane of the frame.
• Space frame. a. Transverse loads and flexural moments that are
applied along the member length. These may be concentrated or
    distributed in nature and act in any plane passing through the entire
    member. b, Concentrated or distributed torsional moments acting
    along the member length. c, Concentrated loads and moments
    applied at the joints and in any of the three orthogonal planes.

       Various load categories were described in Section 1.6. The
weight of objects (either dead or live load) and the hydrostatic pressure
of water are two examples of applied loads. In each case there is
contact between the load source and the loaded structure.
       In a strict technical sense. loads arc not al­ways applied to the
structure. Frequently. structures are subjected to phenomena not com­
monly referred to as loads. Temperature change and shrinkage of
material are two examples. Each of these phenomena cause a structure
to experience strain and stress and consequently, to deform. These are
the same kinds of effects as caused by applied loads. When such effects
are included, it is conventional to refer to the general category of loads
as "actions." In this text the terms "load" and "action" are treated as
synonomous, both meaning any effect that causes stress and/or strain in
a structure.
        any action must satisfy Newton’s first law, i.e., it must cause a
reaction. The fundamental concepts of how a structure an action are
developed in the balance of this chapter.
General State of stress
      As we have stated, the six actions can combine to
produce a combination of stresses. The question of exactly
what is stress should be answered in a very simplistic manner
that is also quite exact in a theoretical sense. We idealize a
load­carrying body as shown in Fig. 1.4 and say that the body
is a continuum. Essentially, a continuum is a
y




z                         x




           FIGURE 1.4
        Idealized body.
collection of material particles, and its exact size is not important for this discussion.
Materials are made up of clusters of molecules, and every material has a definite
molecular structure. On a microscopic scale a material is composed of a space with
atoms at specific locations. The continuum model is a collection of many molecules
and is large enough that the individual molecular interactions for the material can be
ignored and the total of all molecular interactions can be averaged and the continuum
can be assigned some overall gross property to describe its behavior. The continuum
is considered to be quite large compared to an atomistic model; however, it can still
be imagined to be small enough to be of differential size. In other words, we can
effect the mathematical concept of the limit of some quantity with respect to a length
dimension. We assume that as we find the limiting value of a quantity as a length
parameter approaches zero that we do not violate the material assumption of a
continuum; the continuum still exists even though it may be of differential size.
Consider the continuum of Fig. 1.4 to be in statical equilibrium and imagine
a slice taken through the continuum, as shown in Fig. 1.5. The continuum is still in
equilibrium since internal forces and couples at the slice can be viewed as external
balancing forces and couples. We are concerned with forces acting on the smooth, cut
surface that are illustrated as acting on individual, small surface areas. On a small
scale these forces are considered to be acting in a direction either normal or tangent to
their respective areas. The forces originate from the molecu­lar interactions at the
microscopic level; however, as we have stated, we assume that microscopic forces are
averaged and their resultants act on the individual areas. We do not consider small
couples acting on each element even though they may exist. It has been demonstrated,
both experimentally and analytically, that small­body couples can be ignored for the
general theory of mechanics of materials.
         Every force depicted in Fig. 1.5 can be resolved into one normal component
and two shear components. The definition of normal and shear arises naturally;
normal
y




FIGURE 1.5
forces act normally to their area and shear forces act tangentially to their area.
Visualize one small area ΔA as shown in Fig. 1.5b. The force ΔF is shown in
components ΔFx, ΔFy, and ΔFz. Normal stress is denoted by σ (sigma) and is defined
as the normal force per unit area; hence, the terminology normal stress. Normal stress
is given as

                        ∆Fx
                   σx =                                                          (1.1)
                        ∆A
The Greek letter τ (tau) is usually used for shear stress and is the shear force divided
by the area, or


                            ∆Fy
                    τy =                                                         (1.2)
                             ∆A
and

                          ∆Fz
                     τz =                                                         (1.3)
                          ∆A
  A more complete description is given in Chapter 9 concerning shear and normal
stresses that act at a point such as the small area of Fig. 1.5b.
           Previously, it was noted that each action illustrated in Fig. 1.3 produced
either a normal stress or shear stress. In every case we begin with a definition of
stress, either normal or shear, as given by Eqs. (1.1), (1.2), and (1.3), and establish a
theory that describes how a particular action causes a stress. The distribution of stress
that occurs on the cross section of the member because of the action of forces and
couples is dependent upon the way the load­carrying member deforms. In the case of
the axial force, Fig. 1.3a, the member either elongates or shortens in the axial
direction. If we assume a uniform deformation, meaning that every point on the cross
section deforms an equal amount parallel to the beam axis, then we must investigate
the limitation of the theory subject to that assumption.
         It turns out that each of the six actions produces some corresponding
deformation of the member, and prior to developing a theory for stress distribution we
must establish the deformation characteristics of the member when subjected to a
particular load. Chapters 2­9 are devoted to the study of concepts that have been
briefly introduced in this section.
Fig. 1-1   (a)
Fig. 1-2 (b)
Internal Loadings. One of the most important applications of statics in
the analysis of problems involving mechanics of materials is to be able
to determine the resultant force and moment acting within a body,
which are necessary to hold the body together when the body is
subjected to external loads. For example, consider the body shown in
Fig. 1-2a, which is held in equilibrium by the four external forces.* In
order to obtain the internal Loadings acting on a specific region within
the body, it is necessary to use the method of sections. This requires
that an imaginary section or "cut" be made through the region where
the internal loadings are to be determined. The two parts of the body
are
         * The body's weight is not shown, since it is assumed to be quite small, and therefore
negligible compared with the other loads.
then separated, and a free­body diagram of one of the parts is drawn. If
we consider the section shown in Fig. 1-2a, then the resulting free­body
diagram of the bottom part of the body is shown in Fig. 1­2b. Here it
can be seen that there is actually a distribution of internal force acting
on the "exposed" area of the section. These forces represent the effects
of the material of the top part of the body acting on the adjacent
material of the bottom part. Although this exact distribution may be
unknown, we can use statics to determine the resultant internal force
and moment, FR and MRo this distribution exerts at a specific point O on
the sectioned area, Fig. 1-2c.
      Since the entire body is in equilibrium, then each part of the body
is also in equilibrium. Consequently, FR and MRo can be determined by
applying Eqs. 1­1 to anyone of the two parts of the sectioned body.
When doing so, note that FR acts through point 0, although its computed
value will not depend on the location of this point. On the other hand,
MRo does depend on this location, since the moment arms must extend
from O to the line of action of each force on the free­body diagram. It
will be shown in later portions of the text that point O is most often
chosen at the centroid of the sectioned area, and so we will always
choose this location for O, unless otherwise stated. Also, if a member is
long and slender, as in the case of a rod or beam, the section to be
considered is generally taken perpendicular to the longitudinal axis of
the member. This section is referred to as the cross section.
Fig. 1-2 (c)
Fig. 1-2 (d)
Later in this text we will show how to relate the resultant
internal force and moment to the distribution of force on the sectioned
area, Fig. 1­2b, and thereby develop equations that can be used for
analysis and design of the body. To do this, however, the components
of FR and MRo acting both normal or perpendicular to the sectioned area
and within the plane of the area, must be considered. If we establish x,
y, z axes with origin at point O, as shown in Fig; 1­2d, then FR and MRo
can each be resolved into three components. Four different types of
loadings can then be defined as follows:
Nz is called the normal force, since it acts perpendicular to the area.
This force is developed when the external loads tend to push or pull on
V is called the shear force, and it can be determined from its two
components using vector addition, V = Vx + Vy. The shear force lies in
the plane of the area and is developed when the external loads tend to
cause the two segments of the body to slide over one another.
Tz is called the torsional moment or torque. It is developed when the
external loads tend to twist one segment of the body with respect to the
other.
M is called the bending moment. It is determined from the vector
addition of its two components, M = Mx + My. The bending moment is
caused by the external loads that tend to bend the body about an axis
lying within the plane of the area.
In this text, note that representation of a moment or torque is
shown in three dimensions as a vector with an associated curl, Fig. I­2d.
By the right-hand rule, the thumb gives the arrowhead sense of the
vector and the fingers or curl indicate the tendency for rotation (twist or
bending).
Fig. 1-3
F3




Fig. 1-3
Each of the six unknown x, y, z components of force and
moment shows in Fig. 1­2d can be determined directly from the six
equations of equilibrium, that is, Eqs. 1­2, applied to either segment of
the body. If, however, the body is subjected to a coplanar system of
forces, then only normal­force, shear, and bending­moment components
will exist at the section. To show this, consider the body in Fig. 1­3a. If
it is in equilibrium, then the internal resultant components, acting at the
indicated section, can be determined by first "cutting" the body into two
parts, as shown in Fig. 1­3b, and then applying the equations of
equilibrium to one of the parts. Here the internal resultants consist of
the normal force N, shear force V, and bending moment Mo. These
loadings must be equal in magnitude and opposite in direction on each
of the sectioned parts (Newton's third law). Furthermore, the magnitude
of each unknown is determined by applying the three equations of
equilibrium to either one of these parts, Eqs. 1­3. If we use the x, y, z
coordinate axes, with origin established at point O, as shown on the left
segment, then a direct solution for N can be obtained by applying ΣFx =

0, and V can be obtained directly from ΣFy = 0. Finally, the bending

moment Mo can be determined directly by summing moments about

point O (the z axis), Σ Mo = 0, in order to eliminate the moments caused
by the unknowns N and V.
The method of sections is used to determine the internal loadings at a
point located on the section of a body. These resultants are statically
equivalent to the forces that are distributed over the material on the
sectioned area. If the body is static, that is, at rest or moving with
constant velocity, the resultants must be in equilibrium with the
external loadings acting on either one of the sectioned segments of the
body.
        We will now present a procedure that can be used for applying
the method of sections to determine the internal resultant normal force,
shear force, bending moment, and torsional moment at a specific
Support Reactions. First decide which segment of the body is to be
considered. Then before the body is sectioned, it will be necessary to
determine the support reactions or the reactions at the connections only
on the chosen segment of the body. This is done by drawing the free­
body diagram for the entire body, establishing a coordinate system, and
then applying the equations of equilibrium to the body.


Free-Body Diagram. Keep all external distributed loadings, couple
moments, torques, and forces acting on the body in their exact
locations, then pass an imaginary section through the body at the point
where the internal resultant loadings are to be determined. If the body
represents a member of a structure or mechanical device, this section is
often taken perpendicular to the longitudinal axis of the member. Draw
a free­body diagram of one of the' 'cut" segments and indicate the
unknown resultants N, V, M, and T at the section. In most cases, these
resultants are placed at the point representing the geometric center or
centroid of the sectioned area. In particular, if the member is subjected
to a coplanar system of forces, only N, V, and M act at the centroid.
       Establish the x, y, z coordinate axes at the centroid and show the
resultant components acting along the axes.


Equations of Equilibrium. Apply the equations of equilibrium to
obtain the unknown resultants. Moments should be summed at the
section, about the axes where the resultants act. Doing this eliminates
unknown forces N and V and allows a direct solution for M (and T). If
the solution of the equilibrium equations yields a negative value for a
resultant, the assumed directional sense of the resultant is opposite to
that shown on the free­body diagram.
The following examples illustrate this procedure numerically and also
provide a review of some of the important principles of statics. Since
statics plays such an important role in the analysis of problems in
mechanics of materials, it is highly recommended that one solves as
many problems as possible of those that follow these examples.
In Sec. 1.2 we showed how to determine the internal resultant force and
moment acting at a specified point on the sectioned area of a body, Fig.
1­9a. It was stated that these two loadings represent the resultant effects
of the actual distribution of force acting over the sectioned area, Fig. 1­
9b. Obtaining this distribution of internal loading, however, is one of
the major problems in mechanics of materials.
Later it will be shown that to solve this problem it will be
necessary to study how the body deforms under load, since each
internal force distribution will deform the material in a unique way.
Before this can be done, however, it is first necessary to develop a
means for describing the internal force distribution at each point on the
sectioned area. To do this we will establish the concept of stress.
       Consider the sectioned area to be subdivided into small areas,
such as the one ΔA shown shaded in Fig. 1­9b. As we reduce ΔA to a
smaller and smaller size, we must make two assumptions regarding the
properties of the material. We will consider the material to be
continuous, that is, to consist of a continuum or uniform distribution of
matter having no voids, rather than being composed of a finite number
of distinct atoms or molecules. Further­more, the material must be
cohesive, meaning that all portions of it are connected together, rather
than having breaks, cracks, or separations. Now, as the subdivided area
ΔA of this continuous­cohesive material is reduced to one of
infinitesimal size, the distribution of force acting over the entire
sectioned area will consist of an infinite number of forces, each acting
on an element ΔA located at a specific point on the sectioned area. A
typical finite yet very small force ΔF, acting on its associated area ΔA,
is shown in Fig. 1­9c. This force, like all the others, will have a unique
direction, but for further discussion we will replace it by two of its
components, namely, ΔFn and ΔFt, which are taken normal and tangent
to the area, respectively. As the area Δ A approaches zero, so do the
force ΔF and its components; however, the quotient of the force and
area will, in general, approach a finite limit. This quotient is called
stress, and as noted, it describes the intensity of the internal force on a
specific plane (area) passing through a point.
Fig. 1.9 (a)
(b)              (c)


      Fig. 1.9
Normal Stress. The intensity of force, or force per unit area, acting
normal to ΔA is defined as the normal stress, σ(sigma). Mathematically
it can be expressed as

                                  ∆Fn
                         σ = lim                                     (1–4)
                             ∆A→0 ∆A

If the normal force or stress "pulls" on the area element ΔA as shown in
Fig. I­9c, it is referred to as tensile stress, whereas if it "pushes" on ΔA
it is called compressive stress.
Shear Stress. Likewise, the intensity of force, or force per unit area,
acting tangent to ΔA is called the shear stress, τ(tau). This component
is expressed mathematically as
Fig. 1.9
∆Ft
                          τ = lim                                  (1–5)
                              ∆A→0 ∆A


In Fig. I­9c, note that the orientation of the area ΔA completely
specifies the direction of ΔFn, which is always perpendicular to the area.

On the other hand, each shear force ΔFt can act in an infinite number of
directions within the plane of the area. Provided, however, the direction
of ΔF is known, then the direction of ΔFt can be established by
resolution of ΔF as shown in the figure.
Cartesian Stress Components. To specify further the direction of the
shear stress, we will resolve it into rectangular components, and to do
this we will make reference to x, y, z coordinate axes, oriented as shown
in Fig. 1–10a. Here the element of area ΔA = Δx Δy and the three
Cartesian components of ΔF are shown in Fig. 1­10b. We can now
express the normal­stress component as

                             ∆Fz
                  σ z = lim
                        ∆A→0 ∆A
Fig. 1.10 (a)
(b)
                  (c)



      Fig. 1.10
z




F1

         (b)


               Fig. 1.11
z




          (d)



(c)         Fig. 1.11
and the two shear­stress components as

                              ∆Fx
                  τ zx = lim
                         ∆A→0 ∆A


                                  ∆Fy
                  τ zy = lim
                           ∆A→0    ∆A

The subscript notation z in σz is used to reference the direction of the
outward normal line, which specifies the orientation of the area ~A.
Two subscripts are used for the shear­stress components, τzx and τzy.
The z specifies the orientation of the area, and x and y refer to the
direction lines for the shear stresses.
To summarize these concepts, the intensity of the internal force
at a point in a body must be described on an area having a specified
orientation. This intensity can then be measured using three
components of stress acting on the area. The normal component acts
normal or perpendicular to the area, and the shear components act
within the plane of the area. These three stress components are shown
graphically in Fig. 1­l0c.
       Now consider passing another imaginary section through the
body parallel to the x–z plane and intersecting the front side of the
element shown in Fig.1­10a. The resulting free­body diagram is shown
in Fig. 1­11a. Resolving the force acting on the area ΔA = Δx Δz into its
rectangular components, and then determining the intensity of these
force components, leads to the normal stress and shear­stress
components shown in Fig. 1­11b. Using the same notation as before, the
subscript y in σy, τyx, and τyz refers to the direction of the normal line

associated with the orientation of the area, and x and z in τyx and τyz
refer to the corresponding direction lines for the shear stress. Lastly,
one more section of the body parallel to the y–z plane, as shown in Fig.
1­11c, gives rise to normal stress σx and shear stresses τxy and τxz, Fig. 1­
11d. If we continue in this manner, using corresponding parallel x
planes, we can "cut out" a cubic volume element of material that
represents the state of stress acting around the chosen point in the body,
Fig. 1­12.
Equilibrium Requirements. Although each of the six faces of the
element in Fig. 1­12 will have three components of stress acting on it, if
the stress around the point is constant, some of these stress components
can be related by satisfying both force and moment equilibrium for the
element. To show the relationships between the components we will
consider a free­body diagram of the element, Fig. 1­13a. This element
has a volume of ΔV = Δx Δy Δz, and in accordance with Eqs. 1­4 and 1­
5, the forces acting on each face are determined from the product of the
average stress times the area of the face. For simplicity, we have not
labeled the "dashed" forces acting on the "hidden" sides of the element.
Instead, to view, and thereby label, some of these forces, the element is
shown from a front view in Fig. 1­13b. Here it should be noted that the
z




x
    Fig. 1.13 (a) Element free-body diagram
z




Fig. 1.13 (b) Element free-body diagram
force components on the "hidden" sides of the element are designated
with stresses having primes, and these forces are shown in the opposite
direction to their counterparts acting on the opposite faces of the
element.
       If we now consider force equilibrium in the y direction, we have
CONCEPT OF STRESS AT A GENERAL
     POINT IN AN ARBITRARILY LOADED
                 MEMBER


In Arts. 1­3 and 1­4 the concept of stress was introduced by considering
the internal force distribution required to satisfy equilibrium in a
portion of a bar under centric load. The nature of the force distribution
led to uniformly distributed normal and shearing stresses on transverse
planes through the bar. In more complicated structural members or
machine components the stress distributions will not be uniform on
arbitrary internal planes; there­fore. a more general concept of the state
of stress at a point is needed.
Consider a body of arbitrary shape that is in equilibrium under
the action of a system of applied forces. The nature of the internal force
distribu­tion at an arbitrary interior point O can be studied by exposing
an interior plane through O as shown in Fig. 1­13a. The force
distribution required on such an interior plane to maintain equilibrium
of the isolated part of the body, in general, will not be uniform;
however, any distributed force acting on the small area ΔA surrounding
the point of interest O can be replaced by a statically equivalent
resultant force ΔFn through O and a couple ΔMn. The subscript n
indicates that the resultant force and couple are associated with a
particular plane through O­namely, the one having an outward normal
in the n direction at O. For any other plane through O the values of ΔF
and ΔM could be different. Note that the line of action of ΔFn or ΔMn

may not coincide with the direction of n. If the resultant force ΔFn is
divided by the area ΔA, an average force per unit area (average
resultant stress) is obtained. As the area ΔA is made smaller and
smaller, the couple ΔMn vanishes as the force distribution becomes
more and more uniform. In the limit a quantity known as the stress
vector4 or resultant stress is obtained. Thus,


       4
           The component of a tensor on a plane is a vector; therefore, on
a particular plane, the stresses can be treated as vectors.
FIG. 1—13
FIG. 1—13
FIG. 1—13
∆Fn
                         S n = lim
                               ∆A→0 ∆A


In Art. 1­3 it was pointed out that materials respond to components of
the stress vector rather than the stress vector itself. In particular, the
com­ponents normal and tangent to the internal plane were important.
As shown in Fig. 1-13b the resultant force ΔFn can be resolved into the

components ΔFnn and ΔFnt. A normal stress σn and a shearing stress τn
are then defined as
                                   ∆Fnn
                        σ n = lim
                              ∆A→0 ∆A
∆Fnt
                        τ n = lim
                               ∆A→0 ∆A


For purposes of analysis it is convenient to reference stresses to some
coordinate system. For example, in a Cartesian coordinate system the
stresses on planes having outward normals in the x, y. and z directions
are usually chosen. Consider the plane having an outward normal in the
x direc­tion. In this case the normal and shear stresses on the plane will
be σx and τx, respectively. Since τx, in general, will not coincide with

the y or z axes, it must be resolved into the components τxy and τxz, as
shown in Fig. 1­13c. Unfortunately the state of stress at a point in a
stress vector since the stress vector itself depends on the orientation of
the plane with which it is associated. An infinite number of planes can
be passed through the point, resulting in an infinite number of stress
vectors being associated with the point. Fortunately it can be shown
(see Art. 1­9) that the specification of stresses on three mutually
perpendicular planes is sufficient to describe com­pletely the state of
stress at the point. The rectangular components of stress vectors on
planes having outward normals in the coordinate directions are shown
in Fig. 1­14. The six faces of the small element are denoted by the
directions of their outward normals so that the positive x face is the one
whose outward normal is in the direction of the positive x axis. The
coordi­nate axes x, y, and z are arranged as a right­hand system. The
FIG. 1—14
sign convention for stresses is as follows. Normal stresses (indicated by
the symbol σ and a single subscript to indicate the plane on which the
stress acts) are positive if they point in the direction of the outward
normal. Thus normal stresses are positive if tensile. Shearing stresses
are denoted by the symbol τ followed by two subscripts; the first
subscript designates the plane on which the shearing stress acts and the
second the coordinate axis to which it is parallel. Thus, τxy is the
shearing stress on an x plane parallel to the z axis. A positive shearing
stress points in the positive direction of the coordinate axis of t he
second subscript if it acts on a surface with an outward normal in the
positive direction. Conversely, if the outward normal of the surface is in
the negative direction, then the positive shearing stress points in the
negative direction of the coordinate axis of the second subscript. The
stresses shown on the element in Fig. 1­14 are all positive.
STRESS UNDER GENERAL LOADING
   CONDITIONS; COMPONENTS OF STRESS


         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.
         Consider a body subjected to several loads P 1, P2, etc. (Fig. 1.32). To
understand the stress condition created by these loads at some point Q within the
body, we shall first pass a section through Q, using a plane parallel to the yz plane.
The portion of the body to the left of the section is subjected to some of the original
loads, and to normal and shearing forces distributed over the section. We shall denote
by ΔFx and ΔVx, respectively, the normal and the shearing forces acting on a small
Fig. 1.32
(a)               (b)


      Fig. 1.33
area ΔA surrounding point Q (Fig. 1.33a). Note that the superscript x is used to
indicate that the forces ΔFx and ΔVx act on a surface per­pendicular to the x axis.
While the normal force ΔFx has a well­defined direction, the shearing force ΔVx may
have any direction in the plane of the section. We therefore resolve ΔVx into two
component forces, ΔVxy and ΔVxz in directions parallel to the y and z axes,
respectively (Fig. 1.33 b). Dividing now the magnitude of each force by the area ΔA,
and letting ΔA approach zero, we define the three stress components shown in Fig.
1.34:
                                 x
                    ∆F
          σ x = lim
               ∆A→0 ∆A

                             ∆V yx                ∆Vzx
          τ xy = lim                   τ xz = lim                          (1.18)
                    ∆A→0 ∆A                  ∆A→0 ∆A
y       τxy




                    x
z
        Fig. 1.34
We note that the first subscript in σx, τxy and τxz is used to indicate that the stresses
under consideration are exerted on a surface perpendicular to the x axis. The second
subscript in τxy and τxz identifies the direction of the component. The normal stress σx
is positive if the corresponding arrow points in the positive x direction, i.e., if the
body is in tension, and negative otherwise. Similarly, the shearing stress components
τxy and τxz are positive if the corresponding arrows point, respectively, in the positive y
and z directions.
         The above analysis may also be carried out by considering the portion of
body located to the right of the vertical plane through Q (Fig. 1.35). The same
magnitudes, but opposite directions, are obtained for the normal and shearing forces
ΔFx, ΔVxy and ΔVxz Therefore, the same values are also obtained for the
corresponding stress components, but since the section in Fig. 1.35 now faces the
negative x axis, a positive sign for σx will indicate that the corresponding arrow points
corresponding arrows point, respectively, in the negative y and z directions, as shown
in Fig. 1.35.
                                   y




                                       σx




          Fig. 1.35
                                                  z
Passing a section through Q parallel to the zx plane, we define in the same
manner the stress components, σy, τyz, and τyx Finally, a section through Q parallel to

the xy plane yields the components σz, τzx and τzy.
         To facilitate the visualization of the stress condition at point Q, we shall
consider a small cube of side a centered at Q and the stresses ex­erted on each of the
six faces of the cube (Fig. 1.36). The stress com­ponents shown in the figure are σx, σy

and σz which represent the nor­mal stress on faces respectively perpendicular to the x,

y, and z axes, and the six shearing stress components τxy τxz etc. We recall that, ac­

cording to the definition of the shearing stress components, τxy represents the y
component of the shearing stress exerted on the face perpendicular to the x axis, while
τyx represents the x 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. 1.36,
and that equal and opposite stress components act on the hidden faces. While the
y




z                   x
        Fig. 1.36
Fig. 1.37
involved is small and vanishes as side a of the cube approaches zero.
         Important relations among the shearing stress components will now be
derived. Let us consider the free­body diagram of the small cube centered at point Q
(Fig. 1.37). The normal and shearing forces acting on the various faces of the cube are
obtained by multiplying the corresponding stress components by the area ΔA of each
face. We first write the following three equilibrium equations:


            ∑F    x    =0      ∑F    y       =0   ∑F    z   =0                   (1.19)

Since forces equal and opposite to the forces actually shown in Fig. 1.37 are acting on
the hidden faces of the cube, it is clear that Eqs. (1.19) are satisfied. Considering now
the moments of the forces about axes Qx', Qy', and Qz' drawn from Q in directions
respectively parallel to the x, y, and z axes, we write the three additional equations

            ∑M     x   =0      ∑M        y   =0   ∑M        z   =0               (1.20)
Fig. 1.38
Using a projection on the x'y' plane (Fig. 1.38), 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 (positive) moment (τxy ΔA)a, the other of

clockwise (negative) moment –(τxy ΔA)a. The last of the three Eqs. (1.20) yields,
therefore,
             + ↑ ∑Mz = 0:                (τ   xy   ∆A) a − (τ yx ∆A) a = 0   (1.21)
from which we conclude that
                                   τ xy = τ yx

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 shearing stress
exerted on a face perpendicular to the y axis. From the remaining two equations
(1.20), we derive in a similar man­ner the relations

               similarly, τ yz = τ zy                and τ yz = τ zy         (1.22)
We conclude from Eqs. (1.21) and (1.22) that only six stress com­ponents are
required to define the condition of stress at a given point Q, instead of nine as
originally assumed. These six components are ux, uy, uz, T XY' Tyz, and T ZX. 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. For
example, considering again the bolt of Fig. 1.29 and a small cube at the center Q of
the bolt (Fig. 1.39a), we find that shearing stresses of equal mag­nitude must be
exerted on the two horizontal faces of the cube and on the two faces that are
perpendicular to the forces P and P' (Fig. 1.39b). Before concluding our discussion of
stress components, let us con­sider again the case of a member under axial loading. If
we consider a small cube with faces respectively parallel to the faces of the member
and recall the results obtained in Sec. 1.11, we find that the conditions of stress in the
member may be described as shown in Fig. lAOa; the only stresses are normal
stresses U x exerted on the faces of the cube which are perpendicular to the x axis.
(a)               (b)


      Fig. 1.39
Fig. 1.40
is ro­tated by 45° about the z axis so that its new orientation matches the orientation
of the sections considered in Fig. 1.31c and d, we conclude that normal and shearing
stresses of equal magnitude are exerted on four faces of the cube (Fig. 1.40b). We
thus observe that the same loading condition may lead to different interpretations of
the stress situation at a given point, depending upon the orientation of the element
considered. More will be said about this in Chap 7.
Stress, Strain, and Their Relationships

2.1 Introduction

 The concepts of stress and strain are two of the most important concepts within the
subject of mechanics of materials or mechanics of deformable bodies. They are
discussed in detail in this chapter, particularly as they relate to two­dimensional
situations.
         In the case of stress as well as in the case of strain, emphasis is placed on the
use of the semigraphical procedure known as the Mohr’s circle solution. The
underlying mathematical concepts leading to Mohr's circle are developed and
discussed. Examples are solved to illustrate the use of this powerful semigraphical
method of solution, which will be used throughout this book whenever problems are
encountered dealing with stress and strain analysis.
The treatment given to the concepts of stress and strain in this book differs
from that in other books in several important respects, the most significant of which is
the fact that the sign convention adopted for strain is compatible with the sign
convention for stress as they relate to the construction of the corresponding Mohr's
circles. This approach is advantageous in that it makes the construction of the stress
and strain Mohr's circles identical.
         The discussion relating stress to strain in this chapter is limited to the range
of material behavior within which the strain varies linearly with stress. This procedure
frees the students from information which, although very important, is extraneous for
the time being. A more complete discussion of material behavior is provided in
Chapters 3 and 4.
2.2 Concept of Stress at a Point

         If a body is subjected to external forces, a system of internal forces is
developed. These internal forces tend to separate or bring closer together the material
particles that make up the body. Consider, for example, the body shown in Figure
2.1(a), which is subjected to the external forces Fl, F2, ... , Fi. Consider an imaginary
plane that cuts the body into two parts, as shown. Internal forces are transmitted from
one part of the body to the other through this imaginary plane. Let the free­body
diagram of the lower part of the body be constructed as shown in Figure 2,1 (b). The
forces F1, F2, and F3 are held in equilibrium by the action of an internal system of
forces distributed in some manner through the surface area of the imaginary plane.
This system of internal forces may be represented by a single resultant force R and/or
by a couple. For the sake of simplicity in introducing the concept of stress, only the
force R is assumed to exist. In general, the force R may be decomposed into a
Figure 2.1
Figure 2.1
component Fn perpendicular to the plane and known as the normal force, and a

component Ft, parallel to the plane and known as the shear force.

         If the area of the imaginary plane is to be A, then Fn / A and Ft / A represent,
respectively, average values of normal and shear forces per unit area called stresses.
These stresses, however, are not, in general, uniformly distributed throughout the area
under consideration, and it is therefore desirable to be able to determine the
magnitude of both the normal and shear stresses at any point within the area. If the
normal and shear forces acting over a differential element of area ΔA in the
neighborhood of point O are ΔFn and ΔFt respectively, as shown in Figure 2.1 (b),
then the normal stress σ and the shearing stress, are given by the following
expressions:                            ∆Fn
                          σ = lim
                                ∆A→0    ∆A
                                                                                 (2.1)
                                        ∆Ft
                          τ = lim
                               ∆A → 0   ∆A
In the special case where the components Fn and Fr are uniformly distributed over
the entire area A, then σ = Fn/A and τ = Ft/A.
         Note that a normal stress acts in a direction perpendicular to the plane on
which it acts and it can be either tensile or compressive. A tensile normal stress is one
that tends to pull the material particles away from each other, while a compressive
normal stress is one that tends to push them closer together. A shear stress, on the
other hand, acts parallel to the plane on which it acts and tends to slide (shear)
adjacent planes with respect to each other. Also note that the units of stress (σ or τ)
consist of units of force divided by units of area. Thus, in the British gravitational
system of measure, such units as pounds per square inch (psi) and kilopounds per
square inch (ksi) are common. In the metric (SI) system of measure, the unit that has
been proposed for stress is the Newton per square meter (N/m 2), which is called the
pascal and denoted by the symbol Pa. Because the Pascal is a very small quantity,
another SI unit that is widely used is the mega Pascal (106 pascals) and is denoted by
the symbol MPa. This unit may also be written as MN/m2.
Components of Stress

         In the most general case, normal and shear stresses at a point in a body may
be considered to act on three mutually perpendicular planes. This most general state
of stress is usually referred to as triaxial. It is convenient to select planes that are
normal to the three coordinates axes x, y, and z and designate them as the X, Y, and Z
planes, respectively. Consider these planes as enclosing a differential volume of
material in the neighborhood of a given point in a stressed body. Such a volume of
material is depicted in Figure 2.2 and is referred to as a three-dimensional stress
element. On each of the three mutually perpendicular planes of the stress element,
there acts a normal stress, and a shear stress which is represented by its two
perpendicular components.
         The notation for stresses used in this text consists of affixing one subscript to
a normal stress, indicating the plane on which it is acting, and two subscripts to a
shear stress, the first of which designates the plane on which it is acting and the
second its direction. For example, σx is a normal stress acting on the X plane, τxy is a

shear stress acting on the X plane and pointed in the positive y direction, and τxz is a
shear stress acting in the X plane and pointed in the positive z direction.
         It is observed from Figure 2.2 that three stress components exist on each of
the three mutually perpendicular planes that define the stress element. Thus there
exists a total of nine stress components that must be specified in order to define
completely the state of stress at any point in the body. By considerations of the
equilibrium of the stress element, it can easily be shown that, τxy = τyx, τxz = τzx, and τyz

= τzy so that the number of stress components required to completely define the state
of stress at a point is reduced to six.
By convention, a normal stress is positive if it points in the direction of the outward
normal to the plane. Thus a positive normal stress produces tension and a negative
normal stress produces compression. A component of shear stress is positive if it is
pointed along the positive direction of the coordinate axis and if the outward normal
to its plane is also in the positive direction of the corresponding axis. If, however, the
outward normal is in the negative direction of the coordinate axis, a positive shear
stress will also be in the negative direction of the corresponding axis. The stress
components shown in Figure 2.2 are all positive. It should be noted, however, that
such a sign convention for shear stress is rather cumbersome. It is only used in the
analysis of triaxial stress problems that are usually dealt with in advanced courses
such as the theory of elasticity.
A complete study of the triaxial or three­dimensional state of stress is beyond the
scope of this chapter, and the analysis that follows is limited to the special case in
which the stress components in one direction are all zero. For example, if all the stress
components in the z direction are zero (i.e., τxz = τyz = τz = 0), the stress condition
reduces to a biaxial or two­dimensional state of stress in the xy plane. This state of
stress is referred to as plane stress. Fortunately, many of the problems encountered in
practice are such that they can be considered plane stress problems.


2.4 Analysis of Plane Stress
         As mentioned previously the state of stress known as plane stress is one in
which all the stress components in one direction vanish. Thus, if it is assumed that all
the components in the z direction shown in Figure 2.2 are zero (i.e., τxz = τyz = τz = 0),
the stress element shown in Figure 2.3(a) is obtained and it is the most general plane
stress condition that can exist. It should be observed that a stress element is in reality
a schematic representation of two sets of perpendicular planes passing through a point
and that the element degenerates into a point in the limit when both dx and dy
approach zero.
From considerations of the equilibrium of forces on the stress element in Figure
2.3(a), it can be shown that τxy = τyx. Thus assume that the depth of the stress element
into the paper is a constant equal to h. Since, by definition, force is the product of
stress and the area over which it acts, then a summation of moments of all forces
about a z axis through point O leads to the following equation:


              τ xy = (h dy )dx − τ yx = (h dx)dy = 0
from which


                                   τ xy = τ yx
STRESS AT A POINT
Stress at a point is terminology that means exactly what it says. Refer to Fig.
8.2 of Example 8.1. Stress at a point C would imply that the stresses at point C
are to be computed. The point must be drawn large enough for it to be
visualized. Therefore the concept of a stress block or material element is
necessary. The stress block is the point enlarged for the practical purpose of
drawing it and is referred to as a stress element since its actual size is
elemental. Point C of Fig. 8.2 is shown enlarged in Fig. 8.la. Physically, the
element can be visualized as a small square near the outside surface of the
beam located at point C.
        The stresses are identified as acting on the edge of the elemental
square and is the standard concept of stress at a point in two dimensions.
(a)

FIGURE 8.1
(a)Element viewed along the z
   axis;
(b) positive stresses acting on an         x
   element;                          (b)
(c)


FIGURE 8.1
(c) element viewed along the x axis;
(d) element viewed along the y axis.   (d)
Actually, the elemental square is an elemental cube located at point C.
The cube is shown in Fig. 8.1b. The subscript notation identifies the direction
of the stress at the face, or area, of the cube on which it acts. For instance, the
stress σxx is the normal stress acting on the face perpendicular to the x axis.
The first subscript identifies the face, or area, of the cube. The second
subscript identifies the direction of the stress. A shear stress on the face of the
cube normal to the y axis and acting in the x direction would be τyx as shown
in Fig. 8.1b. The complete three­dimensional description of stress at a point is
illustrated in the figure.
         The discussion in the previous chapters concerning shear stresses on
mutually perpendicular planes would imply that τxy = τyx τyz = τzy and τxz = τzx.
The subscript notation lends itself toward identifying this equivalence. The
nine components shown in Fig. 8.lb are conveniently presented using the
following format.
(8.5)




        The use of the boldface σ will represent the stress at a point. Equation
(8.5) represents the stress, σ, as a nine­component quantity and is referred to
as a stress tensor, which has certain mathematical properties that are useful in
advanced studies.
Stress at a point will be viewed two dimensionally in this chapter.
        Figure 8.la is actually the cube as it is viewed along the z axis;
therefore, even though the stresses appear to be applied along the edge of the
element, they are actually applied to a surface that is perpendicular to the
plane of the page. The stress components of Fig. 8.la will be written as

Contenu connexe

Tendances

Solution Manual for Structural Analysis 6th SI by Aslam Kassimali
Solution Manual for Structural Analysis 6th SI by Aslam KassimaliSolution Manual for Structural Analysis 6th SI by Aslam Kassimali
Solution Manual for Structural Analysis 6th SI by Aslam Kassimaliphysicsbook
 
Module 3 SFD &BMD PART-1
Module 3 SFD &BMD PART-1Module 3 SFD &BMD PART-1
Module 3 SFD &BMD PART-1Akash Bharti
 
STRENGTH OF MATERIALS for beginners
STRENGTH OF MATERIALS for  beginnersSTRENGTH OF MATERIALS for  beginners
STRENGTH OF MATERIALS for beginnersmusadoto
 
Mechanics of materials
Mechanics of materialsMechanics of materials
Mechanics of materialsSelf-employed
 
Static and Kinematic Indeterminacy of Structure.
Static and Kinematic Indeterminacy of Structure.Static and Kinematic Indeterminacy of Structure.
Static and Kinematic Indeterminacy of Structure.Pritesh Parmar
 
Shear force and bending moment
Shear force and bending moment Shear force and bending moment
Shear force and bending moment temoor abbasa
 
Slope Deflection Method
Slope Deflection MethodSlope Deflection Method
Slope Deflection MethodMahdi Damghani
 
Lecture 7 stress distribution in soil
Lecture 7 stress distribution in soilLecture 7 stress distribution in soil
Lecture 7 stress distribution in soilDr.Abdulmannan Orabi
 
centroid & moment of inertia
centroid & moment of inertiacentroid & moment of inertia
centroid & moment of inertiasachin chaurasia
 
Determinate structures
 Determinate structures Determinate structures
Determinate structuresTarun Gehlot
 
Fundamentals of structural analysis
Fundamentals of structural analysisFundamentals of structural analysis
Fundamentals of structural analysisKathan Sindhvad
 
Reinforced concrete deck girder bridge
Reinforced concrete deck girder bridgeReinforced concrete deck girder bridge
Reinforced concrete deck girder bridgeMarian Joyce Macadine
 
Examples on soil classification
Examples on soil classification Examples on soil classification
Examples on soil classification Malika khalil
 
Iii design-of-steel-structures-unit-2
Iii design-of-steel-structures-unit-2Iii design-of-steel-structures-unit-2
Iii design-of-steel-structures-unit-2saibabu48
 

Tendances (20)

Solution Manual for Structural Analysis 6th SI by Aslam Kassimali
Solution Manual for Structural Analysis 6th SI by Aslam KassimaliSolution Manual for Structural Analysis 6th SI by Aslam Kassimali
Solution Manual for Structural Analysis 6th SI by Aslam Kassimali
 
Module 3 SFD &BMD PART-1
Module 3 SFD &BMD PART-1Module 3 SFD &BMD PART-1
Module 3 SFD &BMD PART-1
 
STRENGTH OF MATERIALS for beginners
STRENGTH OF MATERIALS for  beginnersSTRENGTH OF MATERIALS for  beginners
STRENGTH OF MATERIALS for beginners
 
Mechanics of materials
Mechanics of materialsMechanics of materials
Mechanics of materials
 
Static and Kinematic Indeterminacy of Structure.
Static and Kinematic Indeterminacy of Structure.Static and Kinematic Indeterminacy of Structure.
Static and Kinematic Indeterminacy of Structure.
 
Shear force and bending moment
Shear force and bending moment Shear force and bending moment
Shear force and bending moment
 
Slope Deflection Method
Slope Deflection MethodSlope Deflection Method
Slope Deflection Method
 
Lecture notes in influence lines
Lecture notes in influence linesLecture notes in influence lines
Lecture notes in influence lines
 
consistent deformation
consistent deformationconsistent deformation
consistent deformation
 
Method of joints
Method of jointsMethod of joints
Method of joints
 
Lecture 7 stress distribution in soil
Lecture 7 stress distribution in soilLecture 7 stress distribution in soil
Lecture 7 stress distribution in soil
 
1
11
1
 
centroid & moment of inertia
centroid & moment of inertiacentroid & moment of inertia
centroid & moment of inertia
 
Determinate structures
 Determinate structures Determinate structures
Determinate structures
 
Lecture 2 bearing capacity
Lecture 2 bearing capacityLecture 2 bearing capacity
Lecture 2 bearing capacity
 
Fundamentals of structural analysis
Fundamentals of structural analysisFundamentals of structural analysis
Fundamentals of structural analysis
 
Slope deflection method
Slope deflection methodSlope deflection method
Slope deflection method
 
Reinforced concrete deck girder bridge
Reinforced concrete deck girder bridgeReinforced concrete deck girder bridge
Reinforced concrete deck girder bridge
 
Examples on soil classification
Examples on soil classification Examples on soil classification
Examples on soil classification
 
Iii design-of-steel-structures-unit-2
Iii design-of-steel-structures-unit-2Iii design-of-steel-structures-unit-2
Iii design-of-steel-structures-unit-2
 

En vedette

Video lectures for mba
Video lectures for mbaVideo lectures for mba
Video lectures for mbaEdhole.com
 
Free Ebooks Download ! Edhole.com
Free Ebooks Download ! Edhole.comFree Ebooks Download ! Edhole.com
Free Ebooks Download ! Edhole.comEdhole.com
 
Structures And Forces Science Power Point
Structures And Forces Science Power PointStructures And Forces Science Power Point
Structures And Forces Science Power Pointdallas9261
 
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUS
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUSCatálogo Corporativo Grupo IMPULS - Revista Impuls PLUS
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUSDavid Keshabyan
 
Philadelphia Industrial Development Corporation - John Grady, President @Open...
Philadelphia Industrial Development Corporation - John Grady, President @Open...Philadelphia Industrial Development Corporation - John Grady, President @Open...
Philadelphia Industrial Development Corporation - John Grady, President @Open...OpenAccessPHL
 
My Organization’S Profile1
My Organization’S Profile1My Organization’S Profile1
My Organization’S Profile1lynnfcampbell
 
William Web Portfolio
William Web PortfolioWilliam Web Portfolio
William Web Portfoliochung3271
 
Capitulo I, II y III Circuitos electricos
Capitulo I, II y III Circuitos electricos Capitulo I, II y III Circuitos electricos
Capitulo I, II y III Circuitos electricos MariRizcala
 
Budget for Children in West Bengal 2004-05 to 2008-09
Budget for Children in West Bengal 2004-05 to 2008-09Budget for Children in West Bengal 2004-05 to 2008-09
Budget for Children in West Bengal 2004-05 to 2008-09HAQ: Centre for Child Rights
 
Списак највећих дужника предузетника на дан 30.12.2011.г.
Списак највећих дужника предузетника на дан 30.12.2011.г. Списак највећих дужника предузетника на дан 30.12.2011.г.
Списак највећих дужника предузетника на дан 30.12.2011.г. Knjazevac
 
Periodismo y social media ¿oportunidad o amenaza? (parte I)
Periodismo y social media ¿oportunidad o amenaza? (parte I)Periodismo y social media ¿oportunidad o amenaza? (parte I)
Periodismo y social media ¿oportunidad o amenaza? (parte I)Comunica2 Campus Gandia
 
Store Media Cencosud
Store Media CencosudStore Media Cencosud
Store Media CencosudJavier Russo
 
Vence a phobos
Vence a phobosVence a phobos
Vence a phoboslopezd
 
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbH
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbHTrends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbH
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbHNicole Szigeti
 
Extreme Networks SDN Innovation Challenge
Extreme Networks SDN Innovation ChallengeExtreme Networks SDN Innovation Challenge
Extreme Networks SDN Innovation ChallengeUS-Ignite
 

En vedette (20)

Sapienza naito-25-06-13
Sapienza naito-25-06-13Sapienza naito-25-06-13
Sapienza naito-25-06-13
 
Video lectures for mba
Video lectures for mbaVideo lectures for mba
Video lectures for mba
 
Free Ebooks Download ! Edhole.com
Free Ebooks Download ! Edhole.comFree Ebooks Download ! Edhole.com
Free Ebooks Download ! Edhole.com
 
Structures And Forces Science Power Point
Structures And Forces Science Power PointStructures And Forces Science Power Point
Structures And Forces Science Power Point
 
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUS
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUSCatálogo Corporativo Grupo IMPULS - Revista Impuls PLUS
Catálogo Corporativo Grupo IMPULS - Revista Impuls PLUS
 
SCPA_Newsletter_3Q2015
SCPA_Newsletter_3Q2015SCPA_Newsletter_3Q2015
SCPA_Newsletter_3Q2015
 
Los Gadgets
Los GadgetsLos Gadgets
Los Gadgets
 
Philadelphia Industrial Development Corporation - John Grady, President @Open...
Philadelphia Industrial Development Corporation - John Grady, President @Open...Philadelphia Industrial Development Corporation - John Grady, President @Open...
Philadelphia Industrial Development Corporation - John Grady, President @Open...
 
My Organization’S Profile1
My Organization’S Profile1My Organization’S Profile1
My Organization’S Profile1
 
William Web Portfolio
William Web PortfolioWilliam Web Portfolio
William Web Portfolio
 
Srpski jezik - Vežba 1
Srpski jezik - Vežba 1Srpski jezik - Vežba 1
Srpski jezik - Vežba 1
 
Capitulo I, II y III Circuitos electricos
Capitulo I, II y III Circuitos electricos Capitulo I, II y III Circuitos electricos
Capitulo I, II y III Circuitos electricos
 
Budget for Children in West Bengal 2004-05 to 2008-09
Budget for Children in West Bengal 2004-05 to 2008-09Budget for Children in West Bengal 2004-05 to 2008-09
Budget for Children in West Bengal 2004-05 to 2008-09
 
Списак највећих дужника предузетника на дан 30.12.2011.г.
Списак највећих дужника предузетника на дан 30.12.2011.г. Списак највећих дужника предузетника на дан 30.12.2011.г.
Списак највећих дужника предузетника на дан 30.12.2011.г.
 
Periodismo y social media ¿oportunidad o amenaza? (parte I)
Periodismo y social media ¿oportunidad o amenaza? (parte I)Periodismo y social media ¿oportunidad o amenaza? (parte I)
Periodismo y social media ¿oportunidad o amenaza? (parte I)
 
Store Media Cencosud
Store Media CencosudStore Media Cencosud
Store Media Cencosud
 
Ponal UNE
Ponal UNEPonal UNE
Ponal UNE
 
Vence a phobos
Vence a phobosVence a phobos
Vence a phobos
 
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbH
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbHTrends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbH
Trends auf dem ECM-Markt - Bernhard Zöller, Zöller & Partner GmbH
 
Extreme Networks SDN Innovation Challenge
Extreme Networks SDN Innovation ChallengeExtreme Networks SDN Innovation Challenge
Extreme Networks SDN Innovation Challenge
 

Similaire à Lecture no.1

Lecture 4 - Resultant of Forces - Part 1.pptx
Lecture 4 - Resultant of Forces - Part 1.pptxLecture 4 - Resultant of Forces - Part 1.pptx
Lecture 4 - Resultant of Forces - Part 1.pptxMarjorieJeanAnog
 
engineering mechanics - statics and dynamics
engineering mechanics - statics and dynamicsengineering mechanics - statics and dynamics
engineering mechanics - statics and dynamicsVelmuruganV15
 
A preparation for interview engineering mechanics
A preparation for interview  engineering mechanicsA preparation for interview  engineering mechanics
A preparation for interview engineering mechanicsDr. Ramesh B
 
moments couples and force couple systems by ahmad khan
moments couples and force couple systems by ahmad khanmoments couples and force couple systems by ahmad khan
moments couples and force couple systems by ahmad khanSelf-employed
 
Engineering mechanics
Engineering mechanics Engineering mechanics
Engineering mechanics Ashish Mishra
 
enggmechanicsbya-170923041240-converted.pptx
enggmechanicsbya-170923041240-converted.pptxenggmechanicsbya-170923041240-converted.pptx
enggmechanicsbya-170923041240-converted.pptxswathirani7
 
coplanar forces res comp of forces - for merge
coplanar forces res   comp of forces - for mergecoplanar forces res   comp of forces - for merge
coplanar forces res comp of forces - for mergeEkeeda
 
Engineering Mechanics Pdf
Engineering Mechanics PdfEngineering Mechanics Pdf
Engineering Mechanics PdfEkeeda
 
Comp of forces
Comp of forcesComp of forces
Comp of forcesEkeeda
 
Coplanar forces res & comp of forces - for merge
Coplanar forces res & comp of forces - for mergeCoplanar forces res & comp of forces - for merge
Coplanar forces res & comp of forces - for mergeEkeeda
 
Momento en estructuras
Momento en estructurasMomento en estructuras
Momento en estructurasRol D
 
module 3 (Mechanics)
module 3 (Mechanics)module 3 (Mechanics)
module 3 (Mechanics)Nexus
 
Engineering Physics
Engineering PhysicsEngineering Physics
Engineering PhysicsDileepCS
 

Similaire à Lecture no.1 (20)

force.docx
force.docxforce.docx
force.docx
 
Lecture 4 - Resultant of Forces - Part 1.pptx
Lecture 4 - Resultant of Forces - Part 1.pptxLecture 4 - Resultant of Forces - Part 1.pptx
Lecture 4 - Resultant of Forces - Part 1.pptx
 
Ctm 154[1]
Ctm 154[1]Ctm 154[1]
Ctm 154[1]
 
engineering mechanics - statics and dynamics
engineering mechanics - statics and dynamicsengineering mechanics - statics and dynamics
engineering mechanics - statics and dynamics
 
A preparation for interview engineering mechanics
A preparation for interview  engineering mechanicsA preparation for interview  engineering mechanics
A preparation for interview engineering mechanics
 
12475602.ppt
12475602.ppt12475602.ppt
12475602.ppt
 
12475602.ppt
12475602.ppt12475602.ppt
12475602.ppt
 
moments couples and force couple systems by ahmad khan
moments couples and force couple systems by ahmad khanmoments couples and force couple systems by ahmad khan
moments couples and force couple systems by ahmad khan
 
Engineering mechanics
Engineering mechanics Engineering mechanics
Engineering mechanics
 
enggmechanicsbya-170923041240-converted.pptx
enggmechanicsbya-170923041240-converted.pptxenggmechanicsbya-170923041240-converted.pptx
enggmechanicsbya-170923041240-converted.pptx
 
coplanar forces res comp of forces - for merge
coplanar forces res   comp of forces - for mergecoplanar forces res   comp of forces - for merge
coplanar forces res comp of forces - for merge
 
Engineering Mechanics Pdf
Engineering Mechanics PdfEngineering Mechanics Pdf
Engineering Mechanics Pdf
 
Comp of forces
Comp of forcesComp of forces
Comp of forces
 
Coplanar forces res & comp of forces - for merge
Coplanar forces res & comp of forces - for mergeCoplanar forces res & comp of forces - for merge
Coplanar forces res & comp of forces - for merge
 
Momento en estructuras
Momento en estructurasMomento en estructuras
Momento en estructuras
 
module 3 (Mechanics)
module 3 (Mechanics)module 3 (Mechanics)
module 3 (Mechanics)
 
Strength of materials
Strength of materialsStrength of materials
Strength of materials
 
DYNAMICS OF MACHINES UNIT-1 BY Mr.P.RAMACHANDRAN/AP/MECH/KIT/CBE
DYNAMICS OF MACHINES UNIT-1 BY Mr.P.RAMACHANDRAN/AP/MECH/KIT/CBEDYNAMICS OF MACHINES UNIT-1 BY Mr.P.RAMACHANDRAN/AP/MECH/KIT/CBE
DYNAMICS OF MACHINES UNIT-1 BY Mr.P.RAMACHANDRAN/AP/MECH/KIT/CBE
 
Basic Principles of Statics
Basic Principles of StaticsBasic Principles of Statics
Basic Principles of Statics
 
Engineering Physics
Engineering PhysicsEngineering Physics
Engineering Physics
 

Lecture no.1

  • 1. Introduction External loads Surface forces Body Forces Equilibrium Internal loading / Internal Forces Sign Conventions Axial forces, SF, BM General State of stress General condition of loading Components of stress
  • 2. Analysis and design of any structure or machine involve two major questions: (a) Is the structure strong enough to with­stand the loads applied to it and (b) is it stiff enough to avoid excessive deformations and deflections? In Statics, the members of a structure were treated as rigid bodies; but actually all materials are deformable and this property will henceforth be taken into account. Thus Strength of Materials may be regarded as the statics of deformable or elastic bodies.
  • 4. Both the strength and stiffness of a structural member are functions of its size and shape and also of certain physical properties of the material from which it is made. These physical properties of materials are largely determined from experimental studies of their behavior in a testing machine. The study of Strength of Materials is aimed at predicting just how these geometric and physical properties of a structure will influence its behavior under service conditions. The applications of the subject are broad in scope and will be found in all branches of engineering.
  • 5. The primary objective of the Strength of material / Mechanic of Material is the development of relationships between the loads applied to a non­rigid body and the internal forces and deformations induced in the body. Mechanics of materials is a branch of mechanics that develops relationships between the external loads applied to a deformable body and the intensity of internal forces acting within the body. This subject is also concerned with computing the deformations of the body, and it provides a study of the body's stability when the body is subjected to external forces.
  • 6. The forces that act on a structure include the applied loads and the resulting reaction forces. The applied loads are the known loads that act on a structure. They can be the result of the structure’s own weight, occupancy loads, environmental loads, and so on. The reactions are the forces that the supports exert on a structure. They are considered to be part of the external forces applied.
  • 7. Forces on a structure can arise from many sources, such as the structure's own weight, any objects placed on it, wind pressure and so forth. Force is a vector quantity, that is, it has both magnitude and direction. The SI unit of force is the Newton (N), which is defined as the force required to impart an acceleration of one metre per second per second to a mass of one kilogram (that is, 1 N = 1 kg m/s2). An object placed on a structure will thus impart a vertical force equal to its mass multiplied by the acceleration due to gravity (g = 9.81 m/s2).
  • 8. Surface Forces. As the name implies, surface forces are caused by the dir contact of one body with the surface of another. In all cases these forces distributed over the area of contact between the bodies, Fig. 1­ 1a. In particular, if this area is small in comparison with the total surface area of the body then the surface force may be idealized as a single concentrated force, which is applied to a point on the body, Fig. 1­1a. For example, this might be done to represent the effect of the ground on the wheels of a bicycle when studying the loading on the bicycle. If the surface loading is applied along a narrow area, the loading may be idealized as a linear distributed load, w(s). Here the loading is measured as having an intensity of force/length along the area and is represented graphically by a series of arrows along the line s
  • 9. Concentrated force idealization Linear distributed load idealization Fig. 1-1 (a)
  • 11. s, Fig. 1­1a. The loading along the length of a beam is a typical example of where this idealization is often applied, Fig. 1­1b. The resultant force FR of w(s) is equivalent to the area under the distributed loading curve, and this resultant acts through the centroid C or geometric center of this area. Body Force. A body force occurs when one body exerts a force on another body without direct physical contact between the bodies. Examples include the effects caused by the earth's gravitation or its electromagnetic field. Although body forces affect each of the particles composing the body, these forces are normally represented by a single concentrated force acting on the body. In the case of gravitation, this force is called the weight of the body and acts through the body's center of gravity.
  • 12. The forces on a body can also give rise to moments, which tend to cause the body to rotate about an axis. The moment of a force about an axis is simply equal to the magnitude of the force multiplied by the perpendicular distance from the axis to the line of action of the force. Consider, for example, the lever AB shown in Figure 2.1 (a). The effect of the force P acting at B is to impart both a direct force P and a moment M = Pa on the hinge at A, as shown in Figure 2.1 (b). For a general case of forces and moments in space each force in the resultant of three forces Fx, Fy and Fz and similarly each moment is the resultant of three couples Mx, My and Mz.
  • 13. SUMMARY Type of force system possible resultants Collinear …………………………………………… Force concurrent , coplanar ……………………………. Force parallel , coplanar…………………………………. Force or a couple Nonconcurrent , nonparallel ,coplanar ………… Force or a couple concurrent , noncoplanar………………………… Force parallel , noncoplanar ……………………………. Force or a couple Nonconcurrent , nonparallel , noncoplanar…….. Force or a couple , or a force and a couple
  • 14.
  • 15.
  • 16. The surface forces that develop at the supports or points of support between bodies are called reactions. For two­dimensional problems, i.e., bodies subjected to coplanar force systems, the supports most commonly encountered are shown in Table 1­1. Note the symbol used to represent each support and the type of reactions it exerts on its contacting member. One possible way to determine a type of support reaction is to imagine the attached member as being translated or rotated in a particular direction. If the support prevents translation in a given direction, then a force must be developed on the member in that direction. Likewise, if rotation is prevented, a couple moment must be
  • 17. exerted on the member. For example, a roller support only prevents translation in the contact direction, perpendicular or normal to the surface. Hence, the roller exerts a normal force F on the member at the point of contact. Since the member is free to rotate about the roller, a couple moment cannot be developed by the roller on the member at the point of contact. Remember that the concentrated forces and couple moment shown in Table 1­1 actually represent the resultants of distributed surface forces that exist between each support and its contacting member. Although it is these resultants that are determined in practice, it is generally not important to determine the actual surface load distribution, since the area over which it acts is considerably smaller
  • 18. than the total surface area of the contacting member. 2.8 SUPPORTS AND EXTERNAL REACTIONS In structural analysis terminology the consequence of subjecting a structure to an action is termed the "response" of the structure. A structure subjected to an action (whether static or dynamic) will move (translate and/or rotate) indefinitely unless the action is resisted in some way. "Supports" are provided as the means of preventing free movement of the structure. A statically loaded structure will deform into a new shape but remain attached to its supports and at rest in its new position. A dynamically loaded structure will remain attached to
  • 19. supports but its deformed shape will change as a function of time. "Damping" will eventually bring it to rest but a finite length of time is subjected to simultaneously applied loads and come to rest instantaneously. Supports prevent free motion of the structure as a whole by developing forces to counteract the load actions. The counteracting forces are commonly termed "external reactions" or simply "reactions."
  • 20. Equilibrium of a body requires both a balance of forces, to prevent the body from translating or moving along a straight or curved path, and a balance of moments, to prevent the body from rotating. These conditions can be expressed mathematically by the two vector equations ΣF =0 (1–1) Σ MO = 0
  • 21. Here, Σ F represents the sum of all the forces acting on the body, and Σ MO is the sum of the moments of all the forces about any point O either on or off the body. If an x, y, z coordinate system is established with the origin at point O, the force and moment vectors can be resolved into components along the coordinate axes and the above two equations can be written in scalar form as six equations, namely, Σ Fx = 0 Σ Fy = 0 Σ Fz = 0 (1–2) ΣMx = 0 ΣMy = 0 ΣMz = 0
  • 22. Often in engineering practice the loading on a body can be represented as a system of coplanar forces. If this is the case, and the forces lie in the x–y plane, then the conditions for equilibrium of the body can be specified by only three scalar equilibrium equations; that is, Σ Fx = 0 Σ Fy = 0 (1–3) Σ MO = 0
  • 23. In this case, if point O is the origin of coordinates, the moments will be directed along the z axis, which is perpendicular to the plane that contains the forces. Successful application of the equations of equilibrium requires complete specification of all the known and unknown forces that act on the body. The best way to account for these forces is to draw the body's free-body diagram. Obviously, if the free­body diagram is drawn correctly, the effects of all the applied forces and couple moments can be accounted for when the equations of equilibrium are written.
  • 24. Consider a body of arbitrary shape acted upon by the forces shown in Fig. 1­2. In statics, we would start by determining the resultant of the applied forces 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, may apply inertia forces to bring about dynamic equilibrium. Such cases are discussed later under dynamic loading. For the present, we consider only cases involving static equilibrium. In strength of materials, we make an additional investigation of the internal distribution of the forces. This is done by passing an
  • 25. Figure 1­2 Exploratory section a-a through loaded member.
  • 26. exploratory section a­a through the body and exposing the internal forces acting on the exploratory section that are necessary to maintain the equilibrium of either segment. In general, the internal forces reduce to a force and a couple that, for convenience, are resolved into components that are normal and tangent to the section, as shown in Fig. 1­3. 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 yet ready to show why this is so, we shall prove it as we progress; in particular, we shall prove it for normal forces in the next article. If the x axis is normal to the section, the section is known as the x surface or, more briefly, the x face.
  • 27. Figure 1-3 Components of internal effects on exploratory section a-a.
  • 28. The notation used in Fig. 1­3 identifies both the exploratory section and the direction of the force or moment component. The first subscript denotes the face on which the component acts; the second subscript indicates the direction of the particular component. Thus Pxy is the force on the x face acting in the y direction. Each component reflects a different effect of the applied loads on the member and is given a special name, as follows: Pxx Axial force. This component measures the pulling (or pushing) action perpendicular to the section. A pull rep­resents a tensile force that tends to elongate the member, whereas a push is a compressive force that tends to shorten it. It is often denoted by P.
  • 29. Pxy, Pxz Shear forces. These are components of the total resistance to sliding the portion to one side of the exploratory section past the other. The resultant shear force is usually designated by V, and its components by Vy and Vz to identify their directions. Mxx Torque. This component measures the resistance· to twisting the member and is commonly given the symbol T. Mxy, Mxz Bending moments. These components measure the resistance to bending the member about the y or z axes and are often denoted merely by My or Mz.
  • 30. 2.3 Internal forces in structures So far, we have concentrated on the external forces applied to structures the applied loads and the support reactions. In order for the structure to transmit the external loads to the ground, internal forces must be developed within the individual members. The aim of the design process is to produce a structure that is capable of carrying all these internal forces, which may take the form of axial forces, shear forces, bending moments or torques. Consider first a two­dimensional beam where the applied forces and reactions all lie in a single plane (Figure 2.14(a)). The internal forces at a point in the structure can be found by splitting it at that point
  • 31. and drawing free body diagrams for the two sides (Figure 2.14(b)). The requirements of equilibrium state that not only must the resultant force on the entire structure be zero, but the resultant on any segment of it must also be zero. It is therefore clear that there must be forces acting at the cut point, as shown. These are drawn on the free body diagrams of the segmented structure as though they were external loads, but they are in fact the internal forces in the beam. The forces can be thought of as the external forces that would have to be applied to the cut beam in order to produce the same deformations as in the original uncut beam. The forces shown are an axial force T, a transverse force S, known as a shear force, and a bending moment M.
  • 32. For equilibrium at the cut point, the forces acting on the faces either side of the cut must be equal and opposite; this means that, when the two segments are put together to form the complete structure, there is no resultant external load at that point. For a member in three­dimensional space, a total of six internal forces must be considered, as shown in Figure 2.15. Here, there is again an axial force T in the x direction, and the resultant shear force has been resolved into components Sy and Sz parallel to the y and z axes respectively. There are also moments about each of the three axes: My tends to cause the structure to bend in the horizontal (x ­ z) plane; Mz causes bending in the vertical (x ­ y) plane; Mx causes the member to
  • 33. y x z Figure 2.15
  • 34. being, however, we will restrict ourselves to two­dimensional systems. 2.3.1 Sign convention for internal forces Before defining the internal forces more fully, we need to extend the sign convention introduced earlier. For a two­dimensional system, positive forces act in the positive x and y directions, and a positive moment about the z-axis acts from the positive x towards the positive y- axis, that is anti­clockwise. We can also define a positive face of a member as one whose outward normal is in a positive axis direction. Thus, for the beam segment in Figure 2.16(a), the right­hand face is a positive x-face, the top surface is a positive y-face and the other two faces are negative.
  • 35. We can now define a positive internal force as one which acts either in a positive direction on a positive face or in a negative direction on a negative face. Conversely, a negative force either acts in a negative direction on a positive face or vice versa.
  • 36. EQUILIBRIUM AND STRESS We will not attempt a thorough review of equilibrium methods of statics, since presumably the reader is familiar with analytical concepts that lead to computation of equivalent­force systems and reactions for load­carrying members. For now the discussion will be conceptual. Later, certain topics of statics will be reviewed as considered pertinent. The beam­type structure of Fig. 1.1 is considered to be in equilib­rium with six actions occurring at each end of the beam. There are six and only six possible actions that can occur: three forces and three couples, which are evaluated using the equilibrium equations of statics. A section passed through the beam anywhere along its axis can be viewed as shown in Fig. 1.2. Internal forces and couples act on the cut section as illustrated and have magnitudes necessary to produce equi­librium for the free body. Specifically, we classify the forces and couples as in Fig. 1.3. The force
  • 37. z x FIGURE 1.1 Beam­type structure showing the six actions of statics.
  • 38. y
  • 39.
  • 40. vector acting along the beam axis is shown in Fig. 1.3a and causes either tension or compression on the section, depending upon its direction. The remaining two force vectors (Fig. 1.3b) produce shear loading on the cut section that is characterized by the forces acting tangent to the cut section as opposed to acting normal to the cut section as in the case of the axial force. The three couples of Fig. 1.2 are illustrated in Figs. 1.3c­e as vectors. The axial­couple vector of Fig. 1.3c represents a twisting couple whose direction is determined using the right­hand­screw rule. The twisting couple, referred to as torque, causes a shear action to occur on the cut cross section. The couples of Figs. 1.3d and e are referred to as bending moments, and the vector representation is interpreted as illustrated. It turns out that these couples cause a combination of tension and com­pression on the cut section. The early chapters of this text are devoted to analytical methods for computing the magnitude and direction of these six actions and then the computation of the corresponding stresses. In particular, Chapter 2 deals with the action of Fig.
  • 41. 1.3a. Chapter 4 is devoted to the twisting couple of Fig. 1.3c applied to members of circular cross section. Analytical methods for computing shear forces and bending moments of Figs. 1.3b, d, and e are discussed in Chapter 5. Chapter 6 deals with the methods for computing stress caused by the bending moments of Figs. 1.3d and e, and Chapter 7 is devoted to the derivation of computational methods for the shear stress produced by the shear forces of Fig. 1.3b. Load­carrying members subject to various combina­tions of the forces and couples are analyzed in Chapter 8. This analysis is the culmination of the study of equivalent force systems and their associated stresses.
  • 42. INTERNAL STRESSES AND STRESS RESULTANTS (INTERNAL FORCES) Framed members subjected to load must develop internal stresses to resist the loads and prevent a material failure. The integrated effects of stresses are force quantities of a particular magnitude and direction. A frequent terminology used to refer to these force quantities is stress resultants. “Internal forces” is an alternative terminology. Internal force, that can be created in framed members are categorized into four types.
  • 43. 1.Axial force 2.shear force 3.flexural moment 4.torsional moment The determination of these force quantities is one of the basic aims of structural analysis. The particular force quantities created in a given structure depend upon its behavior. For a planar frame member, the significant internal forces are the hear and axial direct forces and the flexural moment. Figure 2.11 depicts the manner in which these generalized forces are developed by the internal stresses. The resultants of the shear stress distribution τ, and the axial stress distribution σ1,
  • 44. also a major cause of deflections, and this subject is also discussed as is deflection due to shear. Internal forces developed in each type of framed structure are presented in Table 2.1.
  • 45. Table 2.1. Internal Forces (Stress Resultants) for the Basic Structural Types Schematic of Assumed Structural Type Description Member Forces Plane truss and Only axial Force, F 1, is space truss significant. F1 Only in­plane shear force F1 , flexural moment F2 , and axial Beam and plane frame force F3, are significant. For beams, axial force generally F3 does not exist. F2
  • 46. Schematic of Assumed Structural Type Description Member Forces F1 Only shear force F1 and flexural moment F2 (both Grid transverse to the plane of the grid) and the torsional moment F2 F3 F3 are significant. F2 F5 All stress resultants are signif­ icant: axial force F1, shear Space Frame forces F2 and F3 torsional F3 moments F4, and flexural F1 moments F5 and F6· F6 F4
  • 47. PROBLEM: ANALYZE THE TRUSS BY STIFFNESS METHOD
  • 48. ANALYSIS OF GRIDS Definition: A grid is a structure that has loads applied perpendicular to its plane. The members are assumed to be rigidly connected at the joints. A very basic example of a grid structure is floor system as shown
  • 49. y-axis Wzzj,δ zzj Wzzi,δ zzi Wzj,δ zj Wzi,δ zi x-axis Wxxi,δ xxi Wxi,δ xi Wxxj,δ xxj Wxj,δ xj Wyi,δ yi is Wyj,δ yj x z-a Wyyi,δ yyi Wyyj,δ yyj FORCES AND DISPLACEMENT IN LOCAL COORDINATES
  • 50. when integrated over the member cross section, are the transverse shear force V and the longitudinal (or axial) force P, respectively. Flexural stresses are the source of two internal forces. Tensile and compressive region of this stress distribution,σ2 produce the compressive force C and tensile force T, respectively. Unlike the resultant P. which acts at the centroid. the resultants T and C are separated by a distance, and neither of their lines of action coincides with the centroid. Consequently, the net effect of T and C is the creation of an internal resisting moment, M. Shear and moment resultants that develop in an individual framed member exhibit a rela­tionship to each other and to the transverse loads applied to the member. These fundamental interrelationships will be formulated in Chapter 3. Flexural moment is
  • 51. It is common to refer to loads as “forces that are applied to a structure.” "Applied" is taken to mean “having an identifiable location or point of application.” Each of the six types of framed structures were illustrated in Fig. 1.4. Inspection of that figure indicates the applied load types that are included in each of the six idealized models. 1. Planar truss. Concentrated loads applied in two orthogonal directions at the joints. 2. Space truss. Concentrated loads applied in three ol1hogonal directions at the joints. 3. Beams, a. Transverse loads and flexural moments that are applied along the member and in the plane of the beam. These can be concentrated or distributed in nature, b. concentrated moments
  • 52. applied at the joints and acting in the plane of the beam. • Grid. a. Concentrated or distributed loads (transverse to the plane of the grid) and flexural moments applied along the member length. b, Concentrated or distributed torsional moments acting along the member length. c, Concentrated loads (transverse to the plane of the frame) and out­of­plane moments applied at the joints. • Planar frame. a, Transverse loads and flexural moments that are applied along the member and in the plane of the frame. These can be concentrated or distributed in nature. b, Concentrated loads in two orthogonal directions and in­plane mo­ments applied at the joints and in the plane of the frame. • Space frame. a. Transverse loads and flexural moments that are
  • 53. applied along the member length. These may be concentrated or distributed in nature and act in any plane passing through the entire member. b, Concentrated or distributed torsional moments acting along the member length. c, Concentrated loads and moments applied at the joints and in any of the three orthogonal planes. Various load categories were described in Section 1.6. The weight of objects (either dead or live load) and the hydrostatic pressure of water are two examples of applied loads. In each case there is contact between the load source and the loaded structure. In a strict technical sense. loads arc not al­ways applied to the structure. Frequently. structures are subjected to phenomena not com­ monly referred to as loads. Temperature change and shrinkage of
  • 54. material are two examples. Each of these phenomena cause a structure to experience strain and stress and consequently, to deform. These are the same kinds of effects as caused by applied loads. When such effects are included, it is conventional to refer to the general category of loads as "actions." In this text the terms "load" and "action" are treated as synonomous, both meaning any effect that causes stress and/or strain in a structure. any action must satisfy Newton’s first law, i.e., it must cause a reaction. The fundamental concepts of how a structure an action are developed in the balance of this chapter.
  • 55. General State of stress As we have stated, the six actions can combine to produce a combination of stresses. The question of exactly what is stress should be answered in a very simplistic manner that is also quite exact in a theoretical sense. We idealize a load­carrying body as shown in Fig. 1.4 and say that the body is a continuum. Essentially, a continuum is a
  • 56. y z x FIGURE 1.4 Idealized body.
  • 57. collection of material particles, and its exact size is not important for this discussion. Materials are made up of clusters of molecules, and every material has a definite molecular structure. On a microscopic scale a material is composed of a space with atoms at specific locations. The continuum model is a collection of many molecules and is large enough that the individual molecular interactions for the material can be ignored and the total of all molecular interactions can be averaged and the continuum can be assigned some overall gross property to describe its behavior. The continuum is considered to be quite large compared to an atomistic model; however, it can still be imagined to be small enough to be of differential size. In other words, we can effect the mathematical concept of the limit of some quantity with respect to a length dimension. We assume that as we find the limiting value of a quantity as a length parameter approaches zero that we do not violate the material assumption of a continuum; the continuum still exists even though it may be of differential size.
  • 58. Consider the continuum of Fig. 1.4 to be in statical equilibrium and imagine a slice taken through the continuum, as shown in Fig. 1.5. The continuum is still in equilibrium since internal forces and couples at the slice can be viewed as external balancing forces and couples. We are concerned with forces acting on the smooth, cut surface that are illustrated as acting on individual, small surface areas. On a small scale these forces are considered to be acting in a direction either normal or tangent to their respective areas. The forces originate from the molecu­lar interactions at the microscopic level; however, as we have stated, we assume that microscopic forces are averaged and their resultants act on the individual areas. We do not consider small couples acting on each element even though they may exist. It has been demonstrated, both experimentally and analytically, that small­body couples can be ignored for the general theory of mechanics of materials. Every force depicted in Fig. 1.5 can be resolved into one normal component and two shear components. The definition of normal and shear arises naturally; normal
  • 60. forces act normally to their area and shear forces act tangentially to their area. Visualize one small area ΔA as shown in Fig. 1.5b. The force ΔF is shown in components ΔFx, ΔFy, and ΔFz. Normal stress is denoted by σ (sigma) and is defined as the normal force per unit area; hence, the terminology normal stress. Normal stress is given as ∆Fx σx = (1.1) ∆A The Greek letter τ (tau) is usually used for shear stress and is the shear force divided by the area, or ∆Fy τy = (1.2) ∆A
  • 61. and ∆Fz τz = (1.3) ∆A A more complete description is given in Chapter 9 concerning shear and normal stresses that act at a point such as the small area of Fig. 1.5b. Previously, it was noted that each action illustrated in Fig. 1.3 produced either a normal stress or shear stress. In every case we begin with a definition of stress, either normal or shear, as given by Eqs. (1.1), (1.2), and (1.3), and establish a theory that describes how a particular action causes a stress. The distribution of stress that occurs on the cross section of the member because of the action of forces and couples is dependent upon the way the load­carrying member deforms. In the case of the axial force, Fig. 1.3a, the member either elongates or shortens in the axial direction. If we assume a uniform deformation, meaning that every point on the cross
  • 62. section deforms an equal amount parallel to the beam axis, then we must investigate the limitation of the theory subject to that assumption. It turns out that each of the six actions produces some corresponding deformation of the member, and prior to developing a theory for stress distribution we must establish the deformation characteristics of the member when subjected to a particular load. Chapters 2­9 are devoted to the study of concepts that have been briefly introduced in this section.
  • 63. Fig. 1-1 (a)
  • 65. Internal Loadings. One of the most important applications of statics in the analysis of problems involving mechanics of materials is to be able to determine the resultant force and moment acting within a body, which are necessary to hold the body together when the body is subjected to external loads. For example, consider the body shown in Fig. 1-2a, which is held in equilibrium by the four external forces.* In order to obtain the internal Loadings acting on a specific region within the body, it is necessary to use the method of sections. This requires that an imaginary section or "cut" be made through the region where the internal loadings are to be determined. The two parts of the body are * The body's weight is not shown, since it is assumed to be quite small, and therefore negligible compared with the other loads.
  • 66. then separated, and a free­body diagram of one of the parts is drawn. If we consider the section shown in Fig. 1-2a, then the resulting free­body diagram of the bottom part of the body is shown in Fig. 1­2b. Here it can be seen that there is actually a distribution of internal force acting on the "exposed" area of the section. These forces represent the effects of the material of the top part of the body acting on the adjacent material of the bottom part. Although this exact distribution may be unknown, we can use statics to determine the resultant internal force and moment, FR and MRo this distribution exerts at a specific point O on the sectioned area, Fig. 1-2c. Since the entire body is in equilibrium, then each part of the body is also in equilibrium. Consequently, FR and MRo can be determined by
  • 67. applying Eqs. 1­1 to anyone of the two parts of the sectioned body. When doing so, note that FR acts through point 0, although its computed value will not depend on the location of this point. On the other hand, MRo does depend on this location, since the moment arms must extend from O to the line of action of each force on the free­body diagram. It will be shown in later portions of the text that point O is most often chosen at the centroid of the sectioned area, and so we will always choose this location for O, unless otherwise stated. Also, if a member is long and slender, as in the case of a rod or beam, the section to be considered is generally taken perpendicular to the longitudinal axis of the member. This section is referred to as the cross section.
  • 70. Later in this text we will show how to relate the resultant internal force and moment to the distribution of force on the sectioned area, Fig. 1­2b, and thereby develop equations that can be used for analysis and design of the body. To do this, however, the components of FR and MRo acting both normal or perpendicular to the sectioned area and within the plane of the area, must be considered. If we establish x, y, z axes with origin at point O, as shown in Fig; 1­2d, then FR and MRo can each be resolved into three components. Four different types of loadings can then be defined as follows: Nz is called the normal force, since it acts perpendicular to the area. This force is developed when the external loads tend to push or pull on
  • 71. V is called the shear force, and it can be determined from its two components using vector addition, V = Vx + Vy. The shear force lies in the plane of the area and is developed when the external loads tend to cause the two segments of the body to slide over one another. Tz is called the torsional moment or torque. It is developed when the external loads tend to twist one segment of the body with respect to the other. M is called the bending moment. It is determined from the vector addition of its two components, M = Mx + My. The bending moment is caused by the external loads that tend to bend the body about an axis lying within the plane of the area.
  • 72. In this text, note that representation of a moment or torque is shown in three dimensions as a vector with an associated curl, Fig. I­2d. By the right-hand rule, the thumb gives the arrowhead sense of the vector and the fingers or curl indicate the tendency for rotation (twist or bending).
  • 75. Each of the six unknown x, y, z components of force and moment shows in Fig. 1­2d can be determined directly from the six equations of equilibrium, that is, Eqs. 1­2, applied to either segment of the body. If, however, the body is subjected to a coplanar system of forces, then only normal­force, shear, and bending­moment components will exist at the section. To show this, consider the body in Fig. 1­3a. If it is in equilibrium, then the internal resultant components, acting at the indicated section, can be determined by first "cutting" the body into two parts, as shown in Fig. 1­3b, and then applying the equations of equilibrium to one of the parts. Here the internal resultants consist of the normal force N, shear force V, and bending moment Mo. These loadings must be equal in magnitude and opposite in direction on each
  • 76. of the sectioned parts (Newton's third law). Furthermore, the magnitude of each unknown is determined by applying the three equations of equilibrium to either one of these parts, Eqs. 1­3. If we use the x, y, z coordinate axes, with origin established at point O, as shown on the left segment, then a direct solution for N can be obtained by applying ΣFx = 0, and V can be obtained directly from ΣFy = 0. Finally, the bending moment Mo can be determined directly by summing moments about point O (the z axis), Σ Mo = 0, in order to eliminate the moments caused by the unknowns N and V.
  • 77. The method of sections is used to determine the internal loadings at a point located on the section of a body. These resultants are statically equivalent to the forces that are distributed over the material on the sectioned area. If the body is static, that is, at rest or moving with constant velocity, the resultants must be in equilibrium with the external loadings acting on either one of the sectioned segments of the body. We will now present a procedure that can be used for applying the method of sections to determine the internal resultant normal force, shear force, bending moment, and torsional moment at a specific
  • 78. Support Reactions. First decide which segment of the body is to be considered. Then before the body is sectioned, it will be necessary to determine the support reactions or the reactions at the connections only on the chosen segment of the body. This is done by drawing the free­ body diagram for the entire body, establishing a coordinate system, and then applying the equations of equilibrium to the body. Free-Body Diagram. Keep all external distributed loadings, couple moments, torques, and forces acting on the body in their exact locations, then pass an imaginary section through the body at the point where the internal resultant loadings are to be determined. If the body represents a member of a structure or mechanical device, this section is
  • 79. often taken perpendicular to the longitudinal axis of the member. Draw a free­body diagram of one of the' 'cut" segments and indicate the unknown resultants N, V, M, and T at the section. In most cases, these resultants are placed at the point representing the geometric center or centroid of the sectioned area. In particular, if the member is subjected to a coplanar system of forces, only N, V, and M act at the centroid. Establish the x, y, z coordinate axes at the centroid and show the resultant components acting along the axes. Equations of Equilibrium. Apply the equations of equilibrium to obtain the unknown resultants. Moments should be summed at the section, about the axes where the resultants act. Doing this eliminates
  • 80. unknown forces N and V and allows a direct solution for M (and T). If the solution of the equilibrium equations yields a negative value for a resultant, the assumed directional sense of the resultant is opposite to that shown on the free­body diagram. The following examples illustrate this procedure numerically and also provide a review of some of the important principles of statics. Since statics plays such an important role in the analysis of problems in mechanics of materials, it is highly recommended that one solves as many problems as possible of those that follow these examples.
  • 81. In Sec. 1.2 we showed how to determine the internal resultant force and moment acting at a specified point on the sectioned area of a body, Fig. 1­9a. It was stated that these two loadings represent the resultant effects of the actual distribution of force acting over the sectioned area, Fig. 1­ 9b. Obtaining this distribution of internal loading, however, is one of the major problems in mechanics of materials.
  • 82. Later it will be shown that to solve this problem it will be necessary to study how the body deforms under load, since each internal force distribution will deform the material in a unique way. Before this can be done, however, it is first necessary to develop a means for describing the internal force distribution at each point on the sectioned area. To do this we will establish the concept of stress. Consider the sectioned area to be subdivided into small areas, such as the one ΔA shown shaded in Fig. 1­9b. As we reduce ΔA to a smaller and smaller size, we must make two assumptions regarding the properties of the material. We will consider the material to be continuous, that is, to consist of a continuum or uniform distribution of matter having no voids, rather than being composed of a finite number
  • 83. of distinct atoms or molecules. Further­more, the material must be cohesive, meaning that all portions of it are connected together, rather than having breaks, cracks, or separations. Now, as the subdivided area ΔA of this continuous­cohesive material is reduced to one of infinitesimal size, the distribution of force acting over the entire sectioned area will consist of an infinite number of forces, each acting on an element ΔA located at a specific point on the sectioned area. A typical finite yet very small force ΔF, acting on its associated area ΔA, is shown in Fig. 1­9c. This force, like all the others, will have a unique direction, but for further discussion we will replace it by two of its components, namely, ΔFn and ΔFt, which are taken normal and tangent to the area, respectively. As the area Δ A approaches zero, so do the
  • 84. force ΔF and its components; however, the quotient of the force and area will, in general, approach a finite limit. This quotient is called stress, and as noted, it describes the intensity of the internal force on a specific plane (area) passing through a point.
  • 86. (b) (c) Fig. 1.9
  • 87. Normal Stress. The intensity of force, or force per unit area, acting normal to ΔA is defined as the normal stress, σ(sigma). Mathematically it can be expressed as ∆Fn σ = lim (1–4) ∆A→0 ∆A If the normal force or stress "pulls" on the area element ΔA as shown in Fig. I­9c, it is referred to as tensile stress, whereas if it "pushes" on ΔA it is called compressive stress. Shear Stress. Likewise, the intensity of force, or force per unit area, acting tangent to ΔA is called the shear stress, τ(tau). This component is expressed mathematically as
  • 89. ∆Ft τ = lim (1–5) ∆A→0 ∆A In Fig. I­9c, note that the orientation of the area ΔA completely specifies the direction of ΔFn, which is always perpendicular to the area. On the other hand, each shear force ΔFt can act in an infinite number of directions within the plane of the area. Provided, however, the direction of ΔF is known, then the direction of ΔFt can be established by resolution of ΔF as shown in the figure. Cartesian Stress Components. To specify further the direction of the shear stress, we will resolve it into rectangular components, and to do this we will make reference to x, y, z coordinate axes, oriented as shown
  • 90. in Fig. 1–10a. Here the element of area ΔA = Δx Δy and the three Cartesian components of ΔF are shown in Fig. 1­10b. We can now express the normal­stress component as ∆Fz σ z = lim ∆A→0 ∆A
  • 92. (b) (c) Fig. 1.10
  • 93. z F1 (b) Fig. 1.11
  • 94. z (d) (c) Fig. 1.11
  • 95. and the two shear­stress components as ∆Fx τ zx = lim ∆A→0 ∆A ∆Fy τ zy = lim ∆A→0 ∆A The subscript notation z in σz is used to reference the direction of the outward normal line, which specifies the orientation of the area ~A. Two subscripts are used for the shear­stress components, τzx and τzy. The z specifies the orientation of the area, and x and y refer to the direction lines for the shear stresses.
  • 96. To summarize these concepts, the intensity of the internal force at a point in a body must be described on an area having a specified orientation. This intensity can then be measured using three components of stress acting on the area. The normal component acts normal or perpendicular to the area, and the shear components act within the plane of the area. These three stress components are shown graphically in Fig. 1­l0c. Now consider passing another imaginary section through the body parallel to the x–z plane and intersecting the front side of the element shown in Fig.1­10a. The resulting free­body diagram is shown in Fig. 1­11a. Resolving the force acting on the area ΔA = Δx Δz into its rectangular components, and then determining the intensity of these
  • 97. force components, leads to the normal stress and shear­stress components shown in Fig. 1­11b. Using the same notation as before, the subscript y in σy, τyx, and τyz refers to the direction of the normal line associated with the orientation of the area, and x and z in τyx and τyz refer to the corresponding direction lines for the shear stress. Lastly, one more section of the body parallel to the y–z plane, as shown in Fig. 1­11c, gives rise to normal stress σx and shear stresses τxy and τxz, Fig. 1­ 11d. If we continue in this manner, using corresponding parallel x planes, we can "cut out" a cubic volume element of material that represents the state of stress acting around the chosen point in the body, Fig. 1­12.
  • 98.
  • 99. Equilibrium Requirements. Although each of the six faces of the element in Fig. 1­12 will have three components of stress acting on it, if the stress around the point is constant, some of these stress components can be related by satisfying both force and moment equilibrium for the element. To show the relationships between the components we will consider a free­body diagram of the element, Fig. 1­13a. This element has a volume of ΔV = Δx Δy Δz, and in accordance with Eqs. 1­4 and 1­ 5, the forces acting on each face are determined from the product of the average stress times the area of the face. For simplicity, we have not labeled the "dashed" forces acting on the "hidden" sides of the element. Instead, to view, and thereby label, some of these forces, the element is shown from a front view in Fig. 1­13b. Here it should be noted that the
  • 100. z x Fig. 1.13 (a) Element free-body diagram
  • 101. z Fig. 1.13 (b) Element free-body diagram
  • 102. force components on the "hidden" sides of the element are designated with stresses having primes, and these forces are shown in the opposite direction to their counterparts acting on the opposite faces of the element. If we now consider force equilibrium in the y direction, we have
  • 103. CONCEPT OF STRESS AT A GENERAL POINT IN AN ARBITRARILY LOADED MEMBER In Arts. 1­3 and 1­4 the concept of stress was introduced by considering the internal force distribution required to satisfy equilibrium in a portion of a bar under centric load. The nature of the force distribution led to uniformly distributed normal and shearing stresses on transverse planes through the bar. In more complicated structural members or machine components the stress distributions will not be uniform on arbitrary internal planes; there­fore. a more general concept of the state of stress at a point is needed.
  • 104. Consider a body of arbitrary shape that is in equilibrium under the action of a system of applied forces. The nature of the internal force distribu­tion at an arbitrary interior point O can be studied by exposing an interior plane through O as shown in Fig. 1­13a. The force distribution required on such an interior plane to maintain equilibrium of the isolated part of the body, in general, will not be uniform; however, any distributed force acting on the small area ΔA surrounding the point of interest O can be replaced by a statically equivalent resultant force ΔFn through O and a couple ΔMn. The subscript n indicates that the resultant force and couple are associated with a particular plane through O­namely, the one having an outward normal in the n direction at O. For any other plane through O the values of ΔF
  • 105. and ΔM could be different. Note that the line of action of ΔFn or ΔMn may not coincide with the direction of n. If the resultant force ΔFn is divided by the area ΔA, an average force per unit area (average resultant stress) is obtained. As the area ΔA is made smaller and smaller, the couple ΔMn vanishes as the force distribution becomes more and more uniform. In the limit a quantity known as the stress vector4 or resultant stress is obtained. Thus, 4 The component of a tensor on a plane is a vector; therefore, on a particular plane, the stresses can be treated as vectors.
  • 109. ∆Fn S n = lim ∆A→0 ∆A In Art. 1­3 it was pointed out that materials respond to components of the stress vector rather than the stress vector itself. In particular, the com­ponents normal and tangent to the internal plane were important. As shown in Fig. 1-13b the resultant force ΔFn can be resolved into the components ΔFnn and ΔFnt. A normal stress σn and a shearing stress τn are then defined as ∆Fnn σ n = lim ∆A→0 ∆A
  • 110. ∆Fnt τ n = lim ∆A→0 ∆A For purposes of analysis it is convenient to reference stresses to some coordinate system. For example, in a Cartesian coordinate system the stresses on planes having outward normals in the x, y. and z directions are usually chosen. Consider the plane having an outward normal in the x direc­tion. In this case the normal and shear stresses on the plane will be σx and τx, respectively. Since τx, in general, will not coincide with the y or z axes, it must be resolved into the components τxy and τxz, as shown in Fig. 1­13c. Unfortunately the state of stress at a point in a
  • 111. stress vector since the stress vector itself depends on the orientation of the plane with which it is associated. An infinite number of planes can be passed through the point, resulting in an infinite number of stress vectors being associated with the point. Fortunately it can be shown (see Art. 1­9) that the specification of stresses on three mutually perpendicular planes is sufficient to describe com­pletely the state of stress at the point. The rectangular components of stress vectors on planes having outward normals in the coordinate directions are shown in Fig. 1­14. The six faces of the small element are denoted by the directions of their outward normals so that the positive x face is the one whose outward normal is in the direction of the positive x axis. The coordi­nate axes x, y, and z are arranged as a right­hand system. The
  • 113. sign convention for stresses is as follows. Normal stresses (indicated by the symbol σ and a single subscript to indicate the plane on which the stress acts) are positive if they point in the direction of the outward normal. Thus normal stresses are positive if tensile. Shearing stresses are denoted by the symbol τ followed by two subscripts; the first subscript designates the plane on which the shearing stress acts and the second the coordinate axis to which it is parallel. Thus, τxy is the shearing stress on an x plane parallel to the z axis. A positive shearing stress points in the positive direction of the coordinate axis of t he second subscript if it acts on a surface with an outward normal in the positive direction. Conversely, if the outward normal of the surface is in the negative direction, then the positive shearing stress points in the
  • 114. negative direction of the coordinate axis of the second subscript. The stresses shown on the element in Fig. 1­14 are all positive.
  • 115. STRESS UNDER GENERAL LOADING CONDITIONS; COMPONENTS OF STRESS 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. Consider a body subjected to several loads P 1, P2, etc. (Fig. 1.32). To understand the stress condition created by these loads at some point Q within the body, we shall first pass a section through Q, using a plane parallel to the yz plane. The portion of the body to the left of the section is subjected to some of the original loads, and to normal and shearing forces distributed over the section. We shall denote by ΔFx and ΔVx, respectively, the normal and the shearing forces acting on a small
  • 117. (a) (b) Fig. 1.33
  • 118. area ΔA surrounding point Q (Fig. 1.33a). Note that the superscript x is used to indicate that the forces ΔFx and ΔVx act on a surface per­pendicular to the x axis. While the normal force ΔFx has a well­defined direction, the shearing force ΔVx may have any direction in the plane of the section. We therefore resolve ΔVx into two component forces, ΔVxy and ΔVxz in directions parallel to the y and z axes, respectively (Fig. 1.33 b). Dividing now the magnitude of each force by the area ΔA, and letting ΔA approach zero, we define the three stress components shown in Fig. 1.34: x ∆F σ x = lim ∆A→0 ∆A ∆V yx ∆Vzx τ xy = lim τ xz = lim (1.18) ∆A→0 ∆A ∆A→0 ∆A
  • 119. y τxy x z Fig. 1.34
  • 120. We note that the first subscript in σx, τxy and τxz is used to indicate that the stresses under consideration are exerted on a surface perpendicular to the x axis. The second subscript in τxy and τxz identifies the direction of the component. The normal stress σx is positive if the corresponding arrow points in the positive x direction, i.e., if the body is in tension, and negative otherwise. Similarly, the shearing stress components τxy and τxz are positive if the corresponding arrows point, respectively, in the positive y and z directions. The above analysis may also be carried out by considering the portion of body located to the right of the vertical plane through Q (Fig. 1.35). The same magnitudes, but opposite directions, are obtained for the normal and shearing forces ΔFx, ΔVxy and ΔVxz Therefore, the same values are also obtained for the corresponding stress components, but since the section in Fig. 1.35 now faces the negative x axis, a positive sign for σx will indicate that the corresponding arrow points
  • 121. corresponding arrows point, respectively, in the negative y and z directions, as shown in Fig. 1.35. y σx Fig. 1.35 z
  • 122. Passing a section through Q parallel to the zx plane, we define in the same manner the stress components, σy, τyz, and τyx Finally, a section through Q parallel to the xy plane yields the components σz, τzx and τzy. To facilitate the visualization of the stress condition at point Q, we shall consider a small cube of side a centered at Q and the stresses ex­erted on each of the six faces of the cube (Fig. 1.36). The stress com­ponents shown in the figure are σx, σy and σz which represent the nor­mal stress on faces respectively perpendicular to the x, y, and z axes, and the six shearing stress components τxy τxz etc. We recall that, ac­ cording to the definition of the shearing stress components, τxy represents the y component of the shearing stress exerted on the face perpendicular to the x axis, while τyx represents the x 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. 1.36, and that equal and opposite stress components act on the hidden faces. While the
  • 123. y z x Fig. 1.36
  • 125. involved is small and vanishes as side a of the cube approaches zero. Important relations among the shearing stress components will now be derived. Let us consider the free­body diagram of the small cube centered at point Q (Fig. 1.37). The normal and shearing forces acting on the various faces of the cube are obtained by multiplying the corresponding stress components by the area ΔA of each face. We first write the following three equilibrium equations: ∑F x =0 ∑F y =0 ∑F z =0 (1.19) Since forces equal and opposite to the forces actually shown in Fig. 1.37 are acting on the hidden faces of the cube, it is clear that Eqs. (1.19) are satisfied. Considering now the moments of the forces about axes Qx', Qy', and Qz' drawn from Q in directions respectively parallel to the x, y, and z axes, we write the three additional equations ∑M x =0 ∑M y =0 ∑M z =0 (1.20)
  • 127. Using a projection on the x'y' plane (Fig. 1.38), 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 (positive) moment (τxy ΔA)a, the other of clockwise (negative) moment –(τxy ΔA)a. The last of the three Eqs. (1.20) yields, therefore, + ↑ ∑Mz = 0: (τ xy ∆A) a − (τ yx ∆A) a = 0 (1.21) from which we conclude that τ xy = τ yx 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 shearing stress exerted on a face perpendicular to the y axis. From the remaining two equations (1.20), we derive in a similar man­ner the relations similarly, τ yz = τ zy and τ yz = τ zy (1.22)
  • 128. We conclude from Eqs. (1.21) and (1.22) that only six stress com­ponents are required to define the condition of stress at a given point Q, instead of nine as originally assumed. These six components are ux, uy, uz, T XY' Tyz, and T ZX. 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. For example, considering again the bolt of Fig. 1.29 and a small cube at the center Q of the bolt (Fig. 1.39a), we find that shearing stresses of equal mag­nitude must be exerted on the two horizontal faces of the cube and on the two faces that are perpendicular to the forces P and P' (Fig. 1.39b). Before concluding our discussion of stress components, let us con­sider again the case of a member under axial loading. If we consider a small cube with faces respectively parallel to the faces of the member and recall the results obtained in Sec. 1.11, we find that the conditions of stress in the member may be described as shown in Fig. lAOa; the only stresses are normal stresses U x exerted on the faces of the cube which are perpendicular to the x axis.
  • 129. (a) (b) Fig. 1.39
  • 131. is ro­tated by 45° about the z axis so that its new orientation matches the orientation of the sections considered in Fig. 1.31c and d, we conclude that normal and shearing stresses of equal magnitude are exerted on four faces of the cube (Fig. 1.40b). We thus observe that the same loading condition may lead to different interpretations of the stress situation at a given point, depending upon the orientation of the element considered. More will be said about this in Chap 7.
  • 132. Stress, Strain, and Their Relationships 2.1 Introduction The concepts of stress and strain are two of the most important concepts within the subject of mechanics of materials or mechanics of deformable bodies. They are discussed in detail in this chapter, particularly as they relate to two­dimensional situations. In the case of stress as well as in the case of strain, emphasis is placed on the use of the semigraphical procedure known as the Mohr’s circle solution. The underlying mathematical concepts leading to Mohr's circle are developed and discussed. Examples are solved to illustrate the use of this powerful semigraphical method of solution, which will be used throughout this book whenever problems are encountered dealing with stress and strain analysis.
  • 133. The treatment given to the concepts of stress and strain in this book differs from that in other books in several important respects, the most significant of which is the fact that the sign convention adopted for strain is compatible with the sign convention for stress as they relate to the construction of the corresponding Mohr's circles. This approach is advantageous in that it makes the construction of the stress and strain Mohr's circles identical. The discussion relating stress to strain in this chapter is limited to the range of material behavior within which the strain varies linearly with stress. This procedure frees the students from information which, although very important, is extraneous for the time being. A more complete discussion of material behavior is provided in Chapters 3 and 4.
  • 134. 2.2 Concept of Stress at a Point If a body is subjected to external forces, a system of internal forces is developed. These internal forces tend to separate or bring closer together the material particles that make up the body. Consider, for example, the body shown in Figure 2.1(a), which is subjected to the external forces Fl, F2, ... , Fi. Consider an imaginary plane that cuts the body into two parts, as shown. Internal forces are transmitted from one part of the body to the other through this imaginary plane. Let the free­body diagram of the lower part of the body be constructed as shown in Figure 2,1 (b). The forces F1, F2, and F3 are held in equilibrium by the action of an internal system of forces distributed in some manner through the surface area of the imaginary plane. This system of internal forces may be represented by a single resultant force R and/or by a couple. For the sake of simplicity in introducing the concept of stress, only the force R is assumed to exist. In general, the force R may be decomposed into a
  • 137. component Fn perpendicular to the plane and known as the normal force, and a component Ft, parallel to the plane and known as the shear force. If the area of the imaginary plane is to be A, then Fn / A and Ft / A represent, respectively, average values of normal and shear forces per unit area called stresses. These stresses, however, are not, in general, uniformly distributed throughout the area under consideration, and it is therefore desirable to be able to determine the magnitude of both the normal and shear stresses at any point within the area. If the normal and shear forces acting over a differential element of area ΔA in the neighborhood of point O are ΔFn and ΔFt respectively, as shown in Figure 2.1 (b), then the normal stress σ and the shearing stress, are given by the following expressions: ∆Fn σ = lim ∆A→0 ∆A (2.1) ∆Ft τ = lim ∆A → 0 ∆A
  • 138. In the special case where the components Fn and Fr are uniformly distributed over the entire area A, then σ = Fn/A and τ = Ft/A. Note that a normal stress acts in a direction perpendicular to the plane on which it acts and it can be either tensile or compressive. A tensile normal stress is one that tends to pull the material particles away from each other, while a compressive normal stress is one that tends to push them closer together. A shear stress, on the other hand, acts parallel to the plane on which it acts and tends to slide (shear) adjacent planes with respect to each other. Also note that the units of stress (σ or τ) consist of units of force divided by units of area. Thus, in the British gravitational system of measure, such units as pounds per square inch (psi) and kilopounds per square inch (ksi) are common. In the metric (SI) system of measure, the unit that has been proposed for stress is the Newton per square meter (N/m 2), which is called the pascal and denoted by the symbol Pa. Because the Pascal is a very small quantity, another SI unit that is widely used is the mega Pascal (106 pascals) and is denoted by the symbol MPa. This unit may also be written as MN/m2.
  • 139. Components of Stress In the most general case, normal and shear stresses at a point in a body may be considered to act on three mutually perpendicular planes. This most general state of stress is usually referred to as triaxial. It is convenient to select planes that are normal to the three coordinates axes x, y, and z and designate them as the X, Y, and Z planes, respectively. Consider these planes as enclosing a differential volume of material in the neighborhood of a given point in a stressed body. Such a volume of material is depicted in Figure 2.2 and is referred to as a three-dimensional stress element. On each of the three mutually perpendicular planes of the stress element, there acts a normal stress, and a shear stress which is represented by its two perpendicular components. The notation for stresses used in this text consists of affixing one subscript to a normal stress, indicating the plane on which it is acting, and two subscripts to a
  • 140. shear stress, the first of which designates the plane on which it is acting and the second its direction. For example, σx is a normal stress acting on the X plane, τxy is a shear stress acting on the X plane and pointed in the positive y direction, and τxz is a shear stress acting in the X plane and pointed in the positive z direction. It is observed from Figure 2.2 that three stress components exist on each of the three mutually perpendicular planes that define the stress element. Thus there exists a total of nine stress components that must be specified in order to define completely the state of stress at any point in the body. By considerations of the equilibrium of the stress element, it can easily be shown that, τxy = τyx, τxz = τzx, and τyz = τzy so that the number of stress components required to completely define the state of stress at a point is reduced to six.
  • 141.
  • 142. By convention, a normal stress is positive if it points in the direction of the outward normal to the plane. Thus a positive normal stress produces tension and a negative normal stress produces compression. A component of shear stress is positive if it is pointed along the positive direction of the coordinate axis and if the outward normal to its plane is also in the positive direction of the corresponding axis. If, however, the outward normal is in the negative direction of the coordinate axis, a positive shear stress will also be in the negative direction of the corresponding axis. The stress components shown in Figure 2.2 are all positive. It should be noted, however, that such a sign convention for shear stress is rather cumbersome. It is only used in the analysis of triaxial stress problems that are usually dealt with in advanced courses such as the theory of elasticity. A complete study of the triaxial or three­dimensional state of stress is beyond the scope of this chapter, and the analysis that follows is limited to the special case in which the stress components in one direction are all zero. For example, if all the stress
  • 143. components in the z direction are zero (i.e., τxz = τyz = τz = 0), the stress condition reduces to a biaxial or two­dimensional state of stress in the xy plane. This state of stress is referred to as plane stress. Fortunately, many of the problems encountered in practice are such that they can be considered plane stress problems. 2.4 Analysis of Plane Stress As mentioned previously the state of stress known as plane stress is one in which all the stress components in one direction vanish. Thus, if it is assumed that all the components in the z direction shown in Figure 2.2 are zero (i.e., τxz = τyz = τz = 0), the stress element shown in Figure 2.3(a) is obtained and it is the most general plane stress condition that can exist. It should be observed that a stress element is in reality a schematic representation of two sets of perpendicular planes passing through a point and that the element degenerates into a point in the limit when both dx and dy approach zero.
  • 144. From considerations of the equilibrium of forces on the stress element in Figure 2.3(a), it can be shown that τxy = τyx. Thus assume that the depth of the stress element into the paper is a constant equal to h. Since, by definition, force is the product of stress and the area over which it acts, then a summation of moments of all forces about a z axis through point O leads to the following equation: τ xy = (h dy )dx − τ yx = (h dx)dy = 0 from which τ xy = τ yx
  • 145.
  • 146. STRESS AT A POINT Stress at a point is terminology that means exactly what it says. Refer to Fig. 8.2 of Example 8.1. Stress at a point C would imply that the stresses at point C are to be computed. The point must be drawn large enough for it to be visualized. Therefore the concept of a stress block or material element is necessary. The stress block is the point enlarged for the practical purpose of drawing it and is referred to as a stress element since its actual size is elemental. Point C of Fig. 8.2 is shown enlarged in Fig. 8.la. Physically, the element can be visualized as a small square near the outside surface of the beam located at point C. The stresses are identified as acting on the edge of the elemental square and is the standard concept of stress at a point in two dimensions.
  • 147. (a) FIGURE 8.1 (a)Element viewed along the z axis; (b) positive stresses acting on an x element; (b)
  • 148. (c) FIGURE 8.1 (c) element viewed along the x axis; (d) element viewed along the y axis. (d)
  • 149. Actually, the elemental square is an elemental cube located at point C. The cube is shown in Fig. 8.1b. The subscript notation identifies the direction of the stress at the face, or area, of the cube on which it acts. For instance, the stress σxx is the normal stress acting on the face perpendicular to the x axis. The first subscript identifies the face, or area, of the cube. The second subscript identifies the direction of the stress. A shear stress on the face of the cube normal to the y axis and acting in the x direction would be τyx as shown in Fig. 8.1b. The complete three­dimensional description of stress at a point is illustrated in the figure. The discussion in the previous chapters concerning shear stresses on mutually perpendicular planes would imply that τxy = τyx τyz = τzy and τxz = τzx. The subscript notation lends itself toward identifying this equivalence. The nine components shown in Fig. 8.lb are conveniently presented using the following format.
  • 150. (8.5) The use of the boldface σ will represent the stress at a point. Equation (8.5) represents the stress, σ, as a nine­component quantity and is referred to as a stress tensor, which has certain mathematical properties that are useful in advanced studies. Stress at a point will be viewed two dimensionally in this chapter. Figure 8.la is actually the cube as it is viewed along the z axis; therefore, even though the stresses appear to be applied along the edge of the element, they are actually applied to a surface that is perpendicular to the plane of the page. The stress components of Fig. 8.la will be written as