Vous êtes sur la page 1sur 34

You can support Wikipedia by making a tax-deductible donation.

Stress (mechanics)
From Wikipedia, the free encyclopedia

Jump to: navigation, search


Continuum mechanics

[show]Laws
[show]Solid mechanics
[show]Fluid mechanics
[show]Scientists
This box: view talk edit

Figure 1.1. Stress in a loaded deformable material body assumed as a continuum.

Figure 1.2. Axial stress in a prismatic bar axially loaded


In continuum mechanics, the concept of stress, introduced by Cauchy around 1822, is a measure
of the average amount of force exerted per unit area of a surface within a deformable body on
which internal forces act (Figure 1.1). In other words, it is a measure of the intensity or internal
distribution of the total internal forces acting within a deformable body across imaginary
surfaces. These internal forces are produced between the particles in the body as a reaction to
external forces applied on the body. External forces are either surface forces or body forces.
Because the loaded deformable body is assumed as a continuum, these internal forces are
distributed continuously within the volume of the material body, i.e. the stress distribution in the
body is expressed as a piecewise continuous function of space coordinates and time.
The SI unit for stress is the pascal (symbol Pa), which is equivalent to one newton (force) per
square meter (unit area). The unit for stress is the same as that of pressure, which is also a
measure of force per unit area. Engineering quantities are usually measured in megapascals
(MPa) or gigapascals (GPa). In imperial units, stress is expressed in pounds-force per square inch
(psi) or kilopounds-force per square inch (ksi).
For the simple case of a body axially loaded, e.g., a prismatic bar subjected to tension or
compression by a force passing through its centroid (Figure 1.2), the stress , or intensity of the
distribution of internal forces, can be obtained by dividing the total tensile or compressive force
by the cross-sectional area where it is acting upon. In this case the stress is represented
by a scalar called engineering stress or nominal stress that represents an average stress (
)
over the area, meaning that the stress in the cross section is uniformly distributed. Thus, we have

In general, however, the stress is not uniformly distributed over a cross section of a material
body, and consequently the stress at a point on a given area is different than the average stress
over the entire area. Therefore, it is necessary to define the stress not at a given area but at a
specific point in the body (Figure 1.1). According to Cauchy, the stress at any point in an object,

assumed to be a continuum, is completely defined by the nine components


tensor known as the Cauchy stress tensor, :

of a second order

The Cauchy stress tensor obeys the tensor transformation law under a change in the system of
coordinates. A graphical representation of this transformation law is the Mohr's circle for stress.
The Cauchy stress tensor is used for stress analysis of material bodies experiencing small
deformations. For large deformations, also called finite deformations, other measures of stress
are required, such as the first and second Piola-Kirchhoff stress tensors, the Biot stress tensor,
and the Kirchhoff stress tensor.
According to the principle of conservation of linear momentum, if the continuum body is in
static equilibrium it can be demonstrated that the components of the Cauchy stress tensor in
every material point in the body satisfy the equilibrium equations (Cauchy's equations of motion
for zero acceleration). At the same time, according to the principle of conservation of angular
momentum, equilibrium requires that the summation of moments with respect to an arbitrary
point is zero, which leads to the conclusion that the stress tensor is symmetric, thus having only
six independent stress components, instead of the original nine.
There are certain invariants associated with the stress tensor, whose values do not depend upon
the coordinate system chosen, or the area element upon which the stress tensor operates. These
are the three eigenvalues of the stress tensor, which are called the principal stresses.
The determination of the internal distribution of stresses, i.e., stress analysis, is required in
engineering, e.g., civil engineering and mechanical engineering, for the study and design of
structures, e.g., tunnels, dams, mechanical parts, and structural frames among others, under
prescribed or expected loads. To determine the distribution of stress in the structure it is
necessary to solve a boundary-value problem by specifying the boundary conditions, i.e.
displacements and/or forces on the boundary. Constitutive equations, such as Hooke's Law for
linear elastic materials, are used to describe the stress:strain relationship in these calculations. A
boundary-value problem based on the theory of elasticity is applied to structures expected to
deform elastically, i.e. infinitesimal strains, under design loads. When the loads applied to the
structure induce plastic deformations, the theory of plasticity is implemented.
Approximate solutions for boundary-value problems can be obtained through the use numerical
methods such as the Finite Element Method, the Finite Difference Method, and the Boundary
Element Method, which are implemented in computer programs. Analytical or close-form
solutions can be obtained for simple geometries, constitutive relations, and boundary conditions.

The stress analysis can be simplified in cases where the physical dimensions and the distribution
of loads allows the structure to be assumed as one-dimensional or two-dimensional. For a twodimensional analysis a plane stress or a plane strain condition can be assumed.
Alternatively, experimental determination of stresses can be carried out using the photoelastic
method.
In design of structures, calculated stresses are restricted to be less than an specified allowable
stress, also known as working or designed stress, that is chosen as some fraction of the yield
strength or of the ultimate strength of the material which the structure is made of. The ratio of the
ultimate stress to the allowable stress is defines as the factor of safety. Laboratory tests are
usually performed on material samples in order to determine the yield strength and the ultimate
strength that the material can withstand before failure.
Solids, liquids, and gases have stress fields. Static fluids support normal stress but will flow
under shear stress. Moving viscous fluids can support shear stress (dynamic pressure). Solids can
support both shear and normal stress, with ductile materials failing under shear and brittle
materials failing under normal stress. All materials have temperature dependent variations in
stress related properties, and non-Newtonian materials have rate-dependent variations.

Contents
[hide]
x
x

x
x
x
x
x
x

x
x

1 Definition of stress
o 1.1 Stress in a prismatic bar
2 Cauchy's stress principle
o 2.1 Relationship stress vector - stress tensor
o 2.2 Transformation rule of the stress tensor
o 2.3 Normal and shear stresses
3 Equilibrium equations and symmetry of the stress tensor
4 Principal stresses and stress invariants
5 Maximum and minimum shear stress
6 Stress deviator tensor
o 6.1 Invariants of the stress deviator tensor
7 Octahedral stresses
8 Analysis of stress
o 8.1 Uniaxial stress
o 8.2 Plane stress
o 8.3 Plane strain
9 Stress transformation in plane stress and plane strain
10 Mohr's circle for stress
o 10.1 Mohr's circle for plane stress or plane strain
o 10.2 Mohr's circle for a general three-dimensional state of stresses
11 Alternative measures of stress
o 11.1 Piola-Kirchhoff stress tensor

x
x

12 See also
13 References

11.1.1 1st Piola-Kirchhoff stress tensor


11.1.2 2nd Piola-Kirchhoff stress tensor

[edit] Definition of stress


In the mechanics of a continuum, a material body can be acted upon by external forces that
produce motion and which are of two kind: surface forces and body forces.
Surface forces or contact forces, designated by (force per unit area), act on the bounding
surface as a result of mechanical contact between bodies, or they may also represent the force
which an imaginary surface within the body exerts on the adjacent surface. The intensity of
surface forces is related, i.e. is inversely proportional, to the area of contact, as will be seen in
this article.
Body forces, such as gravitational forces, electromagnetic forces, and inertial forces, are forces
distributed over the entire volume of a body, i.e. acting on every point in the body. In the case of
gravitational and inertial forces, the intensity of the force depends on or is proportional to the
mass density of the material[clarification needed]. These two forces are specified in terms of force per
unit mass ( ) or per unit volume ( ). These two specifications are related through the material
density by the equation
. Similarly, the intensity of electromagnetic forces depends
upon the strength (electric charge) of the electromagnetic field.
These acting external forces (surface and body forces) are then transmitted from point to point
within the material body, leading to the generation of internal forces. The transmission of such
forces is governed by the conservation laws of linear and angular momenta (Newton's Second
Law of motion). For bodies in static equilibrium, these laws are related to the principles of
equilibrium of forces and moments, respectively.
The measure of the intensity of this internal forces acting within the material body across
imaginary surfaces is called stress. In other words, stress is a measure of the average quantity of
force exerted per unit area of the surface on which these internal forces act. For example, if we
compare a force applied to a small area and a distributed load of the same resulting magnitude
applied to a larger area, we find that the effects or intensities of these two forces are locally
different because the stresses are not the same.
Stress is related to deformations in the body. This relationship is expressed through constitutive
equations.

[edit] Stress in a prismatic bar

Figure 1.3. Normal stress in a prismatic bar. The stress or force distribution in the cross section
of the bar is not necessarily uniform. However, an average normal stress
can be used

Figure 1.4. Shear stress in a prismatic bar. The stress or force distribution in the cross section of
the bar is not necessarily uniform. However, an average shear stress
is not a good
approximation.
Let's first examine the simple case of a prismatic bar subjected to an axial force
, which can
be producing either tension or compression (Figures 1.2 and 1.3). Considering a cross sectional
area perpendicular to the axis of the bar, we can find from equilibrium of forces that the resultant
normal force is equal to
. The intensity of internal forces, or stress , in the cross sectional
area can then be obtained by dividing the total normal force
, e.g. tensile force if acting
outward to the plane or compressive force if acting inward to the plane, by the cross-sectional
area where it is acting upon. In this case the stress is a scalar quantity called engineering or
nominal stress that represents an average stress (
) over the area, i.e. the stress in the cross
section is uniformly distributed. Thus, we have

A different type of stress is obtained when transverse forces


are applied to the prismatic bar
as show in Figure 1.4. Considering the same cross section as before, from static equilibrium, the
internal force has a magnitude equal to
and in opposite direction parallel to the cross section.
is called the shear force. Dividing the shear force
by the area of the cross section we

obtain the shear stress. In this case the shear stress is a scalar quantity representing an average
shear stress (
) in the section, i.e. the stress in the cross section is uniformly distributed.

In general, however, the stress is not uniformly distributed over the cross section of a material
body, and consequently the stress at a point on a given area is different from the average stress
over the entire area. In Figure 1.3, the normal stress is observed in two planes
and
of the axially loaded prismatic bar. The stress on plane
, which is closer to the
point of application of the load , varies more across the cross section than that of plane
. However, if the cross sectional area of the bar is very small, e.g. a slender bar, the
variation of stress across the area is small and the normal stress can be approximated by
.
On the other hand, the variation of shear stress across the section of a prismatic bar cannot be
assumed uniform.
Therefore, it is necessary to define the stress at a specific point in the surface.

[edit] Cauchy's stress principle

Figure 2.1 Internal forces in a body


Let's consider a material body in equilibrium subjected to surface forces and body forces per
unit of volume, with an imaginary plane dividing the body into two segments (Figure 2.1). A
small area
in one of the segments, passing through a point , and with a normal unit vector
is acted upon by a force
resulting from the action of the material on one side of the area
(left segment) onto the other side (right segment).
The distribution of force on the area
moment
at due to the force
states that as

is, however, not always uniform, as there may be a


, as shown in Figure 2.1. Cauchy's stress principle

becomes very small and tends to zero the ratio

becomes

and the moment

vanishes. The resultant vector

traction vector

at point

is defined as the stress vector or

associated with a plane with a normal vector

This equation means that the stress vector depends on the location in the body and the orientation
of the plane on which it is acting.
By Newton's third law of motion, the stress vectors acting on opposite sides of the same surface
are equal in magnitude and opposite in direction. Thus,

The state of stress at a point in the body is then defined by all the stress vectors
associated with all planes (infinite number of planes) that pass through that point. However,
according to Cauchy's fundamental theorem, by just knowing the stress vectors on three mutually
perpendicular planes, the stress vector on any other plane passing through that point can be found
through coordinate transformation equations.
Depending on the orientation of the plane under consideration, the stress vector may not
necessarily be perpendicular to that plane, and can be resolved into two components:
x

where
x

one normal to the plane, called normal stress

is the normal component of the force

to the differential area

and the other parallel to this plane, called the shearing stress

where
is the tangential component of the force
to the differential surface area
shear stress can be further decomposed into two mutually perpendicular vectors.

. The

Figure 2.2 Components of stress in three dimensions


Assuming a material element (Figure 2.2) with planes perpendicular to the coordinate axes of a
Cartesian coordinate system, the stress vectors associated with each of the element planes, i.e.
,
, and
can be decomposed into a normal component and two shear
components, i.e. components in the direction of the three coordinate axes. For the particular case
of a surface with normal unit vector oriented in the direction of the
-axis, the normal stress is
denoted by
, and the two shear stresses are denoted as
and
:

In index notation this is

The nine components


of the stress vectors are the components of a second-order Cartesian
tensor called the Cauchy stress tensor, which completely defines the state of stresses at a point
and it is given by

where
,

, and

are normal stresses, and

, and

are shear stresses.

The first index indicates that the stress acts on a plane normal to the axis, and the second
index denotes the direction in which the stress acts. A stress component is positive if it acts in
the positive direction of the coordinate axes, and if the plane where it acts has an outward normal
vector pointing in the positive coordinate direction.
The Voigt notation representation of the Cauchy stress tensor takes advantage of the symmetry
of the stress tensor to express the stress as a 6-dimensional vector of the form

The Voigt notation is used extensively in representing stress-strain relations in solid mechanics
and for computational efficiency in numerical structural mechanics software.

[edit] Relationship stress vector - stress tensor


The stress vector
at any point associated with a plane of normal vector can be expressed
as a function of the stress vectors on the planes perpendicular to the coordinate axes, i.e. in terms
of the components of the stress tensor
. In tensor form this is:

To prove the expression, we consider a tetrahedron with three faces oriented in the coordinate
planes, and with an infinitesimal area
oriented in an arbitrary direction specified by a normal
vector (Figure 2.3). The stress vector on this plane is denoted by
. The stress vectors
acting on the faces of the tetrahedron are denoted as
,
, and
, and are by
definition the components of the stress tensor
. This tetrahedron is sometimes called the
Cauchy tetrahedron. From equilibrium of forces, i.e. Newton's second law, we have

Figure 2.3. Stress vector acting on a plane with normal vector n.


A note on the sign convention: The tetrahedron is formed by slicing a parallelepiped along an
arbitrary plane n. So, the force acting on the plane n is the reaction exerted by the other half of
the parallelepiped and has an opposite sign.
where the right hand side of the equation represent the product of the mass enclosed by the
tetrahedron and its acceleration: is the density, is the acceleration, and is the height of the
tetrahedron, considering the plane as the base. The area of the faces of the tetrahedron
perpendicular to the axes can be found by projecting
into each face (dot product):

and

Here
, is proportional to the square of the linear dimension of the tetrahedron and
to
the third power. Thus, in the limit when the tetrahedron shrinks to a point, the RHS of the above
equation approaches zero and,

or, equivalently,

Or, in matrix form we have

This equation expresses the components of the stress vector acting on an arbitrary plane with
normal vector at a given point in terms of the components of the stress tensor,
, at that
point.

[edit] Transformation rule of the stress tensor


It can be shown that the stress tensor is a second order tensor, which is a statement of how it
transforms under a change of the coordinate system. From an

system to an

components
in the initial system are transformed into the components
according to the tensor transformation rule (Figure 2.4):

where

is the rotation matrix with components

. In matrix form this is

system, the
in the new system

Figure 2.4 Transformation of the stress tensor


Expanding the matrix operation, and simplifying some terms by taking advantage of the
symmetry of the stress tensor, we have:

A graphical representation of this transformation of stresses, for a two-dimensional (plane stress


and plane strain) and a general three-dimensional state of stresses, is the Mohr's circle for
stresses

[edit] Normal and shear stresses


The magnitude of the normal stress component,
, of any stress vector
acting on an
arbitrary plane with normal vector at a given point in terms of the component of the stress
tensor
is the dot product of the stress vector and the normal vector, thus

The magnitude of the shear stress component, , acting in the plane formed by the two vectors
and , can then be found using the Pythagorean theorem, thus

where

[edit] Equilibrium equations and symmetry of the stress


tensor

Figure 4. Continuum body in equilibrium


When a body is in equilibrium the components of the stress tensor in every point of the body
satisfy the equilibrium equations,

[show]Derivation of equilibrium equations

At the same time, equilibrium requires that the summation of moments with respect to an
arbitrary point is zero, which leads to the conclusion that the stress tensor is symmetric, i.e.

[show]Derivation of symmetry of the stress tensor

However, in the presence of couple-stresses, i.e. moments per unit volume, the stress tensor is
non-symmetric. This also is the case when the Knudsen number is close to one,
, or

the continuum is a Non-Newtonian fluid, which can lead to rotationally non-invariant fluids,
such as polymers.

[edit] Principal stresses and stress invariants


The components
of the stress tensor depend on the orientation of the coordinate system at the
point under consideration. However, the stress tensor itself is a physical quantity and as such, it
is independent of the coordinate system chosen to represent it. There are certain invariants
associated with every tensor which are also independent of the coordinate system. For example, a
vector is a simple tensor of rank one. In three dimensions, it has three components. The value of
these components will depend on the coordinate system chosen to represent the vector, but the
length of the vector is a physical quantity (a scalar) and is independent of the coordinate system
chosen to represent the vector. Similarly, every second rank tensor (such as the stress and the
strain tensors) has three independent invariant quantities associated with it. One set of such
invariants are the principal stresses of the stress tensor, which are just the eigenvalues of the
stress tensor. Their direction vectors are the principal directions or eigenvectors. When the
coordinate system is chosen to coincide with the eigenvectors of the stress tensor, the stress
tensor is represented by a diagonal matrix:

where
,
, and
, are the principal stresses. These principal stresses may be combined to
form three other commonly used invariants, , , and , which are the first, second and third
stress invariants, respectively. The first and third invariant are the trace and determinant
respectively, of the stress tensor. Thus, we have

Because of its simplicity, working and thinking in the principal coordinate system is often very
useful when considering the state of the elastic medium at a particular point.
[show]Derivation of principal stresses and stress invariants

[edit] Maximum and minimum shear stress


...

[edit] Stress deviator tensor

The stress tensor

can be expressed as the sum of two other stress tensors:

1. a mean hydrostatic stress tensor or volumetric stress tensor or mean normal stress
tensor,
, which tends to change the volume of the stressed body; and
2. a deviatoric component called the stress deviator tensor,
, which tends to distort it.

where

is the mean stress given by

The deviatoric stress tensor can be obtained by subtracting the hydrostatic stress tensor from the
stress tensor:

[edit] Invariants of the stress deviator tensor


As it is a second order tensor, the stress deviator tensor also has a set of invariants, which can be
obtained using the same procedure used to calculate the invariants of the stress tensor. It can be
shown that the principal directions of the stress deviator tensor
are the same as the principal
directions of the stress tensor
. Thus, the characteristic equation is

where ,
and
are the first, second, and third deviatoric stress invariants, respectively.
Their values are the same (invariant) regardless of the orientation of the coordinate system
chosen. These deviatoric stress invariants can be expressed as a function of the components of
or its principal values , , and , or alternatively, as a function of
or its principal
values
,
, and
. Thus,

Because

, the stress deviator tensor is in a state of pure shear.

A quantity called the equivalent stress or von Mises stress is commonly used in solid mechanics.
The equivalent stress is defined as

[edit] Octahedral stresses

Figure 6. Octahedral stress planes

Considering the principal directions as the coordinate axes, a plane whose normal vector makes
equal angles with each of the principal axes (i.e. having direction cosines equal to
) is
called an octahedral plane. There are a total of eight octahedral planes (Figure 6). The normal
and shear components of the stress tensor on these planes are called octahedral normal stress
and octahedral shear stress
, respectively.
Knowing that the stress tensor of point O (Figure 6) in the principal axes is

the stress vector on an octahedral plane is then given by:

The normal component of the stress vector at point O associated with the octahedral plane is

which is the mean normal stress or hydrostatic stress. This value is the same in all eight
octahedral planes. The shear stress on the octahedral plane is then

[edit] Analysis of stress


The analysis of stress within a body implies the determination at each point of the body of the
magnitudes of the nine stress components. In other words, it is the determination of the internal
distribution of stresses. A stress analysis is required in engineering, e.g., civil engineering and

mechanical engineering, for the study and design of structures, e.g., tunnels, dams, mechanical
parts, and structural frames among others, under prescribed or expected loads.
To determine the distribution of stress in the structure it is necessary to solve a boundary-value
problem by specifying the boundary conditions, i.e. displacements and/or forces on the
boundary. Constitutive equations, such as e.g. Hooke's Law for linear elastic materials, are used
to describe the stress:strain relationship in these calculations. A boundary-value problem based
on the theory of elasticity is applied to structures expected to deform elastically, i.e. infinitesimal
strains, under design loads. When the loads applied to the structure induce plastic deformations,
the theory of plasticity is implemented.
Approximate solutions for boundary-value problems can be obtained through the use numerical
methods such as the Finite Element Method, the Finite Difference Method, and the Boundary
Element Method, which are implemented in computer programs. Analytical or close-form
solutions can be obtained for simple geometries, constitutive relations, and boundary conditions.
Alternatively, experimental determination of stresses can be carried out using the photoelastic
method.
In design of structures, calculated stresses are restricted to be less than an specified allowable
stress, also known as working or designed stress, that is chosen as some fraction of the yield
strength or of the ultimate strength of the material which the structure is made of. The ratio of the
ultimate stress to the allowable stress is defined as the factor of safety. Laboratory test are
usually performed on material samples in order to determine the yield strength and the ultimate
strength that the material can withstand before failure.
All real objects occupy a three-dimensional space. The stress analysis can be simplified in cases
where the physical dimensions and the loading conditions allows the structure to be assumed as
one-dimensional or two-dimensional. For a two-dimensional analysis a plane stress or a plane
strain condition can be assumed.

[edit] Uniaxial stress


If two of the dimensions of the object are very large or very small compared to the others, the
object may be modelled as one-dimensional. In this case the stress tensor has only one
component and is indistinguishable from a scalar. One-dimensional objects include a piece of
wire loaded at the ends and a metal sheet loaded on the face and viewed up close and through the
cross section.
When a structural element is elongated or compressed, its cross-sectional area changes by an
amount that depends on the Poisson's ratio of the material. In engineering applications, structural
members experience small deformations and the reduction in cross-sectional area is very small
and can be neglected, i.e., the cross-sectional area is assumed constant during deformation. For
this case, the stress is called engineering stress or nominal stress. In some other cases, e.g.,
elastomers and plastic materials, the change in cross-sectional area is significant, and the stress

must be calculated assuming the current cross-sectional area instead of the initial cross-sectional
area. This is termed true stress and is expressed as
,
where
is the nominal (engineering) strain, and
is nominal (engineering) stress.
The relationship between true strain and engineering strain is given by
.
In uniaxial tension, true stress is then greater than nominal stress. The converse holds in
compression.

[edit] Plane stress

Figure 7.1 Plane stress state in a continuum.


A state of plane stress exist when one of the three principal
, stresses is zero. This
usually occurs in structural elements where one dimension is very small compared to the other

two, i.e. the element is flat or thin. In this case, the stresses are negligible with respect to the
smaller dimension as they are not able to develop within the material and are small compared to
the in-plane stresses. Therefore, the face of the element is not acted by loads and the structural
element can be analyzed as two-dimensional, e.g. thin-walled structures such as plates subject to
in-plane loading or thin cylinders subject to pressure loading. The other three non-zero
components remain constant over the thickness of the plate. The stress tensor can then be
approximated by:

.
The corresponding strain tensor is:

in which the non-zero


term arises from the Poisson's effect. This strain term can be
temporarily removed from the stress analysis to leave only the in-plane terms, effectively
reducing the analysis to two dimensions.

[edit] Plane strain


If one dimension is very large compared to the others, the principal strain in the direction of the
longest dimension is constrained and can be assumed as zero, yielding a plane strain condition.
In this case, though all principal stresses are non-zero, the principal stress in the direction of the
longest dimension can be disregarded for calculations. Thus, allowing a two dimensional
analysis of stresses, e.g. a dam analyzed at a cross section loaded by the reservoir.

[edit] Stress transformation in plane stress and plane strain


Consider a point

in a continuum under a state of plane stress, or plane strain, with stress

components
and all other stress components equal to zero (Figure 7.1, Figure 8.1).
From static equilibrium of an infinitesimal material element at (Figure 8.2), the normal stress
and the shear stress on any plane perpendicular to the - plane passing through with a
unit vector making an angle of with the horizontal, i.e.
is the direction cosine in the
direction, is given by:

These equations indicate that in a plane stress or plane strain condition, one can determine the
stress components at a point on all directions, i.e. as a function of , if one knows the stress
components
on any two perpendicular directions at that point. It is important to
remember that we are considering a unit area of the infinitesimal element in the direction parallel
to the - plane.

Figure 8.1 - Stress transformation at a point in a continuum under plane stress conditions.

Figure 8.2 - Stress components at a plane passing through a point in a continuum under plane
stress conditions.
The principal directions (Figure 8.3), i.e. orientation of the planes where the shear stress
components are zero, can be obtained by making the previous equation for the shear stress
equal to zero. Thus we have:

and we obtain

This equation defines two values


obtained by finding the angle

which are

apart (Figure 8.3). The same result can be

which makes the normal stress

a maximum, i.e.

The principal stresses


and
, or minimum and maximum normal stresses
and
,
respectively, can then be obtained by replacing both values of
into the previous equation for
. This can be achieved by rearranging the equations for
and , first transposing the first
term in the first equation and squaring both sides of each of the equations then adding them.
Thus we have

where

which is the equation of a circle of radius centered at a point with coordinates


called Mohr's circle. But knowing that for the principal stresses the shear stress
obtain from this equation:

,
, then we

Figure 8.3 - Transformation of stresses in two dimensions, showing the planes of action of
principal stresses, and maximum and minimum shear stresses.
When
the infinitesimal element is oriented in the direction of the principal planes, thus
the stresses acting on the rectangular element are principal stresses:
and
.
Then the normal stress
and shear stress acting on a plane making an angle of with the
principal directions can be obtained by making
. Thus we have

Then the maximum shear stress

Then the minimum shear stress

occurs when

occurs when

, i.e.

, i.e.

(Figure 8.3):

(Figure 8.3):

[edit] Mohr's circle for stress


The Mohr's circle, named after Christian Otto Mohr, is a two-dimensional graphical
representation of the state of stress at a point. The abscissa,
, and ordinate, , of each point

on the circle are the normal stress and shear stress components, respectively, acting on a
particular cut plane with a unit vector with components
. In other words, the
circumference of the circle is the locus of points that represent state of stress on individual planes
at all their orientations.
Karl Culmann was the first to conceive a graphical representation for stresses while considering
longitudinal and vertical stresses in horizontal beams during bending. Mohr's contribution
extended the use of this representation for both two- and three-dimensional stresses and
developed a failure criterion based on the stress circle.

[edit] Mohr's circle for plane stress or plane strain

Figure 9.1. Mohr's circle for plane stress and plane strain conditions (double angle approach).

Figure 9.2. Mohr's circle for plane stress and plane strain conditions (Pole approach). Any
straight line drawn from the pole will intersect the Mohr circle at a point that represents the state
of stress on a plane inclined at the same orientation (parallel) in space as that line.
If we know the stress components
.
, and
at a point for any two perpendicular
planes in a continuum body under plane stress, or plane strain (Figures 8.1 and 8.2) we can
construct the Mohr circle of stress. Once the Mohr circle is drawn one can use it to find the stress
state on any other plane passing through that point in the body.
According to the sign convention for engineering mechanics, in disciplines such as mechanical
engineering and structural engineering, which is the one used in this article, for the construction
of the Mohr circle the normal stresses are positive if they are outward to the plane of action
(tension), and shear stresses are positive if they rotate clockwise about the point in consideration.
In geomechanics, i.e. soil mechanics and rock mechanics, however, normal stresses are
considered positive when they are inward to the plane of action (compression), and shear stresses
are positive if they rotate counterclockwise about the point in consideration.
To construct the Mohr circle of stress for a state of plane stress, or plane strain, first we plot two
points in the
space corresponding to the known stress components on both perpendicular
planes, i.e.
and
(Figure 9.1 and 9.2). We then connect points and
by a straight line and find the midpoint which corresponds to the intersection of this line
with the
axis. Finally, we draw a circle with diameter
and centre at .
As demonstrated in the previous section, the radius

of the circle is

, and the coordinates of its centre are

The principal stresses are then the abscissa of the points of intersection of the circle with the
axis (note that the shear stresses are zero for the principal stresses).
Using the Mohr circle one can find the stress components
on any other plane with a
different orientation that passes through point . For this, two approaches can be used:
The first approach relies on the fact that the angle

between two planes passing through

half the angle between the lines joining their corresponding stress points

is

on the Mohr

circle and the centre of the circle (Figure 9.1). In other words, the stresses
acting on a
plane at an angle counterclockwise to the plane on which
acts is determined by traveling
counterclockwise around the circle from the known stress point
and angle
at the centre of the circle (Figure 9.1).

a distance subtending

The second approach involves the determination of a point on the Mohr circle called the pole or
the origin of planes. Any straight line drawn from the pole will intersect the Mohr circle at a
point that represents the state of stress on a plane inclined at the same orientation (parallel) in
space as that line. Therefore, knowing the stress components and on any particular plane, one
can draw a line parallel to that plane through the particular coordinates
and on the Mohr
circle and find the pole as the intersection of such line with the Mohr circle. As an example, let's
assume we have a state of stress with stress components
,
, and
, as shown on Figure
9.2. First, we can draw a line from point parallel to the plane of action of
, or, if we choose
otherwise, a line from point parallel to the plane of action of
. The intersection of any of
these two lines with the Mohr circle is the pole. Once the pole has been determined, to find the
state of stress on a plane making an angle with the vertical, or in other words a plane having its
normal vector forming an angle with the horizontal plane, then we can draw a line from the
pole parallel to that plane (See Figure 9.2). The normal and shear stresses on that plane are then
the coordinates of the point of intersection between the line and the Mohr circle.

[edit] Mohr's circle for a general three-dimensional state of stresses

Figure 7. Mohr's circle for a three-dimensional state of stress


To construct the Mohr's circle for a general three-dimensional case of stresses at a point, the
values of the principal stresses
first evaluated, as explained previously.

and their principal directions

must be

Considering the principal axes as the coordinate system, instead of the general
,
,
coordinate system, and assuming that
, then the normal and shear components
of the stress vector
, for a given plane with unit vector , satisfy the following equations

Knowing that

, we can solve for

, which yields

Since
satisfy

, and

is non-negative, the numerators from the these equations

as the denominator

and

as the denominator

and

as the denominator

and

These expressions can be rewritten as

which are the equations of the three Mohr's circles for stress
,
coordinates

, and
,

, and

, with radii

, and their centres with


,

, respectively.

These equations for the Mohr's circles show that all admissible stress points

lie on these

circles or within the shaded area enclosed by them (see Figure 7). Stress points
satisfying the equation for circle

lie on, or outside circle

. Stress points

the equation for circle


lie on, or inside circle
. And finally, stress points
satisfying the equation for circle
lie on, or outside circle
.

[edit] Alternative measures of stress


Main article: Stress measures

satisfying

The Cauchy stress tensor is not the only measure of stress that is used in practice. Other
measures of stress include the first and second Piola-Kirchhoff stress tensors, the Biot stress
tensor, and the Kirchhoff stress tensor.

[edit] Piola-Kirchhoff stress tensor


In the case of finite deformations, the Piola-Kirchhoff stress tensors are used to express the
stress relative to the reference configuration. This is in contrast to the Cauchy stress tensor which
expresses the stress relative to the present configuration. For infinitesimal deformations or
rotations, the Cauchy and Piola-Kirchhoff tensors are identical. These tensors take their names
from Gabrio Piola and Gustav Kirchhoff.
[edit] 1st Piola-Kirchhoff stress tensor
Whereas the Cauchy stress tensor, , relates forces in the present configuration to areas in the
present configuration, the 1st Piola-Kirchhoff stress tensor,
relates forces in the present
configuration with areas in the reference ("material") configuration.

where

is the deformation gradient and

is the Jacobian determinant.

In terms of components with respect to an orthonormal basis, the first Piola-Kirchhoff stress is
given by

Because it relates different coordinate systems, the 1st Piola-Kirchhoff stress is a two-point
tensor. In general, it is not symmetric. The 1st Piola-Kirchhoff stress is the 3D generalization of
the 1D concept of engineering stress.
If the material rotates without a change in stress state (rigid rotation), the components of the 1st
Piola-Kirchhoff stress tensor will vary with material orientation.
The 1st Piola-Kirchhoff stress is energy conjugate to the deformation gradient.
[edit] 2nd Piola-Kirchhoff stress tensor
Whereas the 1st Piola-Kirchhoff stress relates forces in the current configuration to areas in the
reference configuration, the 2nd Piola-Kirchhoff stress tensor relates forces in the reference
configuration to areas in the reference configuration. The force in the reference configuration is
obtained via a mapping that preserves the relative relationship between the force direction and
the area normal in the current configuration.

In index notation with respect to an orthonormal basis,

This tensor is symmetric.


If the material rotates without a change in stress state (rigid rotation), the components of the 2nd
Piola-Kirchhoff stress tensor will remain constant, irrespective of material orientation.
The 2nd Piola-Kirchhoff stress tensor is energy conjugate to the Green-Lagrange finite strain
tensor.

[edit] See also


x
x
x
x
x
x
x
x
x
x
x
x
x
x

Bending
Linear elasticity
Kelvin probe force microscope
Residual stress
Shot peening
Strain
Strain tensor
Stress-energy tensor
Stress-strain curve
Stress concentration
Virial stress
Von Mises stress
Yield stress
Yield surface

[edit] References
x
x
x
x
x
x

Dieter, G. E. (3 ed.). (1989). Mechanical Metallurgy. New York: McGraw-Hill. ISBN 007-100406-8.
Love, A. E. H. (4 ed.). (1944). Treatise on the Mathematical Theory of Elasticity. New
York: Dover Publications. ISBN 0-486-60174-9.
Marsden, J. E., & Hughes, T. J. R. (1994). Mathematical Foundations of Elasticity. New
York: Dover Publications. ISBN 0-486-67865-2.
L.D.Landau and E.M.Lifshitz. (1959). Theory of Elasticity.
Beer, Ferdinand Pierre; Elwood Russell Johnston, John T. DeWolf (1992). Mechanics of
Materials. McGraw-Hill Professional. ISBN 0071129391.
Mase, George E. (1970). Continuum Mechanics. McGraw-Hill. ISBN 0070406634.
http://books.google.ca/books?id=bAdg6yxC0xUC&rview=1.

x
x

Mase, G. Thomas; George E. Mase (1999). Continuum Mechanics for Engineers (Second
ed.). CRC Press. ISBN 0-8493-1855-6.
http://books.google.ca/books?id=uI1ll0A8B_UC&rview=1.
Rees, David (2006). Basic Engineering Plasticity - An Introduction with Engineering and
Manufacturing Applications. Butterworth-Heinemann. ISBN 0750680253.
http://books.google.ca/books?id=4KWbmn_1hcYC.
Brady, B.H.G.; E.T. Brown (1993). Rock Mechanics For Underground Mining (Third
ed.). Kluwer Academic Publisher. ISBN 0412475502.
Timoshenko, Stephen P.; James Norman Goodier (1970). Theory of Elasticity (Third ed.).
McGraw-Hill International Editions. ISBN 0-07-085805-5.

Retrieved from "http://en.wikipedia.org/wiki/Stress_(mechanics)"


Categories: Continuum mechanics | Classical mechanics | Tensors | Materials science | Elasticity
(physics) | Plasticity | Solid mechanics | Mechanics
Hidden categories: All pages needing cleanup | Wikipedia articles needing clarification from July
2009
Views
x
x
x
x

Article
Discussion
Edit this page
History

Personal tools
x
x

Try Beta

Log in / create account

Navigation
x
x
x
x
x

Main page
Contents
Featured content
Current events
Random article

Search
Go

Sear c h

Interaction
x
x
x

About Wikipedia
Community portal
Recent changes

x
x
x

Contact Wikipedia
Donate to Wikipedia
Help

Toolbox
x
x
x
x
x
x
x

What links here


Related changes
Upload file
Special pages
Printable version
Permanent link
Cite this page

Languages
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x


Azrbaycan

Dansk
Deutsch
Eesti
Espaol
Espaol

Franais
Galego

Italiano

Latvieu
Magyar
Nederlands
Norsk (bokml)

Polski

Simple English
Slovenina
Slovenina
/ Srpski
Svenska
Suomi
Ting Vit

x
x

This page was last modified on 8 September 2009 at 14:05.


Text is available under the Creative Commons Attribution-ShareAlike License;
additional terms may apply. See Terms of Use for details.
Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a non-profit
organization.
Privacy policy
About Wikipedia
Disclaimers

x
x
x

Vous aimerez peut-être aussi