Jump to content

Stress (mechanics): Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Mohr's circle: section moved one level up
deleted sections of One dimensional and Two dimensional stress. (please see discussion)
Line 265: Line 265:
</math>
</math>
Because <math>s_{kk}=0 \,</math>, the stress deviator tensor is in a state of pure shear.
Because <math>s_{kk}=0 \,</math>, the stress deviator tensor is in a state of pure shear.

==Stress in one-dimensional bodies==
All real objects occupy three-dimensional space. However, if two dimensions are very large or very small compared to the others, the object may be modelled as [[List of structural elements|one-dimensional]]. This simplifies the mathematical modelling of the object. One-dimensional objects include a piece of wire loaded at the ends and viewed from the side, and a metal sheet loaded on the face and viewed up close and through the cross section.

For one-dimensional objects, the stress tensor has only one component and is indistinguishable from a scalar. The simplest definition of stress, σ = ''F''/''A'', where A is the ''initial'' cross-sectional area prior to the application of the load (or [[force]]) ''F'', is called '''engineering stress''' or '''nominal stress'''. However, when any material is stretched, its cross-sectional area may change by an amount that depends on the [[Poisson's ratio]] of the material. Engineering stress neglects this change in area. The stress axis on a [[stress-strain curve|stress-strain graph]] is often engineering stress, even though the sample may undergo a substantial change in cross-sectional area during testing.

'''True stress''' is an alternative definition in which the initial area is replaced by the current area. In engineering applications, the initial area is always known, and so calculations using nominal stress are generally easier. For small deformation, such as in practical material usage, the reduction in cross-sectional area is small and the distinction between nominal and true stress is insignificant; so the change of cross-sectional area could be assumed to be a constant value. This is not so for the large deformations typical of [[elastomer]]s and [[plasticity (physics)|plastic]] materials when the change in cross-sectional areas can be significant.

In one dimension, conversion between true stress and nominal (engineering) stress is given by

:<math>\sigma_{true} = (1 + \epsilon_e)(\sigma_e) \,</math>,

where
:ε<sub>e</sub> is nominal (engineering) [[strain (materials science)|strain]], and
:σ<sub>e</sub> is nominal (engineering) stress.

The relationship between true strain and engineering strain is given by

:<math>\epsilon_{true} = ln(1 + \epsilon_e) \,</math>.

In uniaxial tension, true stress is then greater than nominal stress. The converse holds in compression.

Example:&nbsp; A [[steel]] [[Screw#Bolt|bolt]] of [[diameter]] 5 [[millimeter|mm]] has a cross-sectional [[area]] of 19.6 mm<sup>2</sup>. A load of 50 [[newton|N]] induces a stress (force distributed over the cross section) of σ = 50/19.6 = 2.55 [[megapascal|MPa]] (N/mm<sup>2</sup>). This can be thought of as each square millimeter of the bolt supporting 2.55 N of the total load. In another bolt with half the diameter, and hence a quarter the cross-sectional area, carrying the same 50 N load, the stress will be quadrupled (10.2 MPa).

The ultimate [[tensile strength]] is a property of a material and is usually determined experimentally from a uniaxial [[tensile test]]. It allows the calculation of the load that would cause [[fracture]]. The [[compressive strength]] is a similar property for compressive loads. The [[yield (engineering)|yield strength]] is the value of stress causing [[plasticity (physics)|plastic deformation]].

==Stress in two-dimensional bodies==
[[Image:Cracks at Sunrise-on-Sea, Eastern Cape.jpg|thumb|200px|[[Fracture|Crack]]s in rock resulting from stress.]]
All real objects occupy 3-dimensional space. However, if one dimension is very large or very small compared to the others, the object may be modelled as two-dimensional. This simplifies the mathematical modelling of the object. Two-dimensional objects include a piece of wire '''loaded on the sides and viewed up close and through the cross-section''' and a metal sheet '''loaded in-plane and viewed face-on'''.

Notice that the same physical, three-dimensional object can be modelled as one-dimensional, two-dimensional or even three-dimensional, depending on the loading and viewpoint of the observer.

===Plane stress===
[[Plane stress]] is a two-dimensional state of stress (Figure 2). This 2-D state models well the state of stresses in a flat, thin plate loaded in the plane of the plate. Figure 2 shows the stresses on the ''x''- and ''y''-faces of a differential element. Not shown in the figure are the stresses in the opposite faces and the external forces acting on the material. Since moment [[Mechanical equilibrium|equilibrium]] of the differential element shows that the shear stresses on the perpendicular faces are equal, the 2-D state of stresses is characterized by three independent stress components (σ<sub>x</sub>, σ<sub>y</sub>, τ<sub>xy</sub>). Note that forces perpendicular to the plane can be abbreviated. For example, σ<sub>x</sub> is an abbreviation for σ<sub>xx</sub>. This notation is described further below.
[[Image:stress 3.PNG|thumbnail|right|Figure 2&nbsp; Stresses normal and tangent to faces]]

See also [[plane strain]].

===Principal stresses in 2-D===
[[Augustin Louis Cauchy]] was the first to demonstrate that at a given point, it is always possible to locate two [[orthogonal]] planes in which the shear stress vanishes. These planes in which the normal forces are acting are called the ''principal planes'', while the normal stresses on these planes are the ''principal stresses''. They are the [[eigenvalue, eigenvector and eigenspace|eigenvalue]]s of the stress tensor and are orthogonal because the stress tensor is [[symmetric tensor|symmetric]] (as per the [[spectral theorem]]). Eigenvalues are invariants with respect to choice of basis and are the roots of the [[Cayley–Hamilton theorem]] (although the term 'the' invariants usually means (I1,I2,I3)). [[Mohr's circle]] is a graphical method of extracting the principal stresses in a 2-dimensional stress state. The maximum and minimum principal stresses are the maximum and minimum possible values of the normal stresses. The maximum principal stress controls brittle fracture.

The two dimensional Cauchy stress tensor is defined as:

:<math>\sigma_{ij}=
\left[{\begin{matrix}
{\sigma _x } & {\tau _{xy}} \\
{\tau _{xy}} & {\sigma _y } \\
\end{matrix}}\right].
</math>

Then principal stresses <math> \sigma_1, \sigma_2</math> are equal to:

:<math>\sigma _1 = \frac {\sigma _x + \sigma _ y}{2} + \sqrt{ \left( \frac {\sigma _x - \sigma _ y}{2} \right)^2 + {\tau _{xy}}^2 }</math>

:<math>\sigma _2 = \frac {\sigma _x + \sigma _ y}{2} - \sqrt{ \left( \frac {\sigma _x - \sigma _ y}{2} \right)^2 + {\tau _{xy}}^2 }.</math>

Those formulas have geometrical interpretation in the form of Mohr Circle presented in section below.


==Mohr's circle for stresses==<!-- This section is linked from [[Concrete]] -->
==Mohr's circle for stresses==<!-- This section is linked from [[Concrete]] -->

Revision as of 23:31, 9 February 2008

In Continuum mechanics, stress is a measure of the average amount of force exerted per unit area. It is a measure of the intensity of the total internal forces acting within a body across imaginary internal surfaces, as a reaction to external applied forces and body forces. It was introduced into the theory of elasticity by Cauchy around 1822. Stress is a concept that is based on the concept of continuum. In general, stress is expressed as

where

is the average stress, also called engineering or nominal stress, and
is the force acting over the area .

The SI unit for stress is the pascal (symbol Pa), which is a shorthand name for one Newton (Force) per square metre (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).

As with force, stress cannot be measured directly but is usually inferred from measurements of strain and knowledge of elastic properties of the material. Devices capable of measuring stress indirectly in this way are strain gauges and piezoresistors.

Stress types

There are two basic stresses: normal stress, , acting normal to the surface under consideration, and shearing stress, , acting parallel to the stressed surface. All other stresses produced on a body by different load conditions are similar or a combination of them. An axial stress is a normal stress produced when a force acts parallel to the major axis of a body, e.g. column. If the forces pull the body producing an elongation it is termed tensile stress; or compressive stresses if the forces push the body compressing it. Bending stresses, e.g. produced on a bent beam, are a combination of tensile and compressive stresses. Torsional stresses, e.g. produced on twisted shafts, are shearing stresses. The term normal stress in rheology is called extensional stress, and in acoustics is called longitudinal stress.

Solids, liquids and gases have stress fields. Static fluids support normal stress (hydrostatic pressure) 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.

The von Mises stress is derived from the distortion energy theory and is a simple way to combine stresses in three dimensions to calculate failure criteria of ductile materials. In this way, the strength of material in a 3-D state of stress can be compared to a test sample that was loaded in one dimension.

Cauchy's stress principle

Figure 1. Internal forces in a body
Figure 2. Components of stress in three dimensions
Figure 3. Stress vector acting on a plane with normal vector n

Cauchy's stress principle asserts that when a continuum body is acted on by forces, i.e. surface forces and body forces, there are internal reactions (forces) throughout the body acting between the material points.

Considering a body subjected to surface forces and body forces per unit of volume, with an imaginary plane dividing the body into two segments (Figure 1). A small area in one of the segments, passing through a point , and with a normal vector is acted upon by a force resulting from the action of the material in one side of the area (right segment) onto the other side (left segment). The distribution of force on is, however, not always uniform, as there may be a moment at due to the force , as shown in the Figure. As becomes very small and tends to zero the ratio becomes , and the moment vanishes. The vector is defined as the stress vector at point associated with a plane with a normal vector :

By Newton's third law, the stress vectors acting upon opposit sides of the same surface are equal in magnitude and of opposite direction. Thus,

The stress vector, not necessarily being perpendicular to the plane on which it acts, can be resolved into two components: one normal to the plane, called normal stress, and the other parallel to this plane, called the shearing stress. The latter, can be further decomposed into two mutually perpendicular vectors.

The state of stress at a point would be defined by all the stress vectors associated with all planes (infinite number of planes) that pass through that point. However, by just knowing the stress vectors on three mutually perpendicular planes, the stress vector on any plane passing through that point can be found through coordinate transformation equations. Assuming a material element (Figure 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 decompose into components in the direction of the three coordinate axes:

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 stresess.

The first index indicates 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 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. stress tensor . For this, 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 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 . From equilibrium of forces, i.e. Newton's second law, we have

where the right hand side of the equation represent the body forces acting on the tetrahedron: 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 projecting into each face (dot product):

Thus, taking the limit when and replacing the previous equations, we have

or as a scalar equation

Equilibrium equation and symmetry of the stress tensor

The state of stress as defined by the stress tensor is an equilibrium state if the following conditions are satisfied:



are the components of the tensor, and f 1 , f 2 , and f 3 are the body forces (force per unit volume).

These equations can be compactly written using Einstein notation as:

The equilibrium conditions may be derived from the condition that the net force on an infinitesimal volume element must be zero. Consider an infinitesimal cube aligned with the , , and axes, with one corner at and the opposite corner at and having each face of area . Consider just the faces of the cube which are perpendicular to the axis. The area vector for the near face is and for the far face it is . The net stress force on these two opposite faces is

A similar calculation can be carried out for the other pairs of faces. The sum of all the stress forces on the infinitesimal cube will then be

Since the net force on the cube must be zero, it follows that this stress force must be balanced by the force per unit volume on the cube (e.g., due to gravitation, electromagnetic forces, etc.) which yields the equilibrium conditions written above.

Equilibrium also requires that the resultant moment on the cube of material must be zero. Taking the moment of the forces above about any suitable point, it follows that, for equilibrium in the absence of body moments

.

The stress tensor is then symmetric and the subscripts can be written in either order.

Another approach to find the symmetry of the stress tensor

The fact that the stress is a symmetric tensor follows from some simple considerations. The force on a small volume element will be the sum of all the stress forces over the surface area of that element. Suppose we have a volume element in the form of a long bar with a triangular cross section, where the triangle is a right triangle. We can neglect the forces on the ends of the bar, because they are small compared to the faces of the bar. Let be the vector area of one face of the bar, be the area of the other, and be the area of the "hypotenuse face" of the bar. It can be seen that

Let's say is the force on area and likewise for the other faces. Since the stress is by definition the force per unit area, it is clear that

The total force on the volume element will be:

Suppose that the volume element contains mass, at a constant density. The important point is that if we make the volume smaller, say by halving all lengths, the area will decrease by a factor of four, while the volume will decrease by a factor of eight. As the size of the volume element goes to zero, the ratio of area to volume will become infinite. The total stress force on the element is proportional to its area, and so as the volume of the element goes to zero, the force/mass (i.e. acceleration) will also become infinite, unless the total force is zero. In other words:

This, along with the second equation above, proves that the function is a linear vector operator (i.e. a tensor). By an entirely analogous argument, we can show that the total torque on the volume element (due to stress forces) must be zero, and that it follows from this restriction that the stress tensor must be symmetric.

However, there are two fundamental ways in which this mode of thinking can be misleading. First, when applying this argument in tandem with the underlying assumption from continuum mechanics that the Knudsen number is strictly less than one, then in the limit , the symmetry assumptions in the stress tensor may break down. This is the case of Non-Newtonian fluid, and can lead to rotationally non-invariant fluids, such as polymers. The other case is when the system is operating on a purely finite scale, such as is the case in mechanics where Finite deformation tensors are used.

Principal stresses and stress invariants

The components of the stress tensor depend on the orientation of the plane that passes through the point under consideration. This would lead to the incorrect conclusion that the state of stress at a point in a body depends on the viewpoint of the observer, i.e. orientation of the coordinate system. However, every tensor, including stress, has invariants that do not depend on the choice of viewpoint. This means that the stress components seen by one observer are related, via the tensor transformation relations, to those seen by any other observer. The length of a first-order tensor, i.e a vector, is a simple example.

At every point in a stressed body there are at least three planes, called principal planes, with normal vectors , called principal directions, where the corresponding stress vector is parallel or in the same direction as the normal vector and where there are no normal shear stresses . Thus,

where is a constant of proportionality, and in this particular case corresponds to the magnitudes of the normal stress vectors or principal stresses.

Knowing that and , we have

This is a homogenous system, i.e. equal to zero, of three linear equations where are the unknowns. To obtain a nontrivial (non-zero) solution for , the determinant matrix of the coefficients must be equal to zero, i.e. the system is singular. Thus,

Expanding the determinant leads to the characteristic equation

where

, and are the first, second, and third stress invariants, respectively. Their values are the same (invariant) regardless of the orientation of the coordinate system chosen.

The characteristic equation has three real roots , i.e. not imaginary due to the symmetry of the stress tensor. The three roots , , and are the eigenvalues or principal stresses, and they are the roots of the Cayley–Hamilton theorem. For each eigenvalue, there is a non-trivial solution for in the equation . These solutions are the principal directions or eigenvectors defining the plane where the principal stresses act. The principal stresses and principal directions characterize the stress at a point and are independent on the orientation of the coordinate system.

If we choose a coordinate system with axes oriented to the principal directions, then the normal stresses will be the principal stresses. Thus, we have

Stress deviator tensor

The stress tensor can be expressed as the sum of two stress tensors:

  1. a mean hydrostatic 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:

As it is a second order tensor, the stress deviator tensor also has a set of invariants, which can be obtain using the same procedure used to calculate the stress invariants. 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. This deviatoric stress invariantes can be expressed as a function of the components of or its principal values , , and , or alternatevily, as a function of or its principal values , , and . Thus,

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

Mohr's circle for stresses

Mohr's circle is a graphical representation of any 2-D stress state. It was named for Christian Otto Mohr. Mohr's circle may also be applied to three-dimensional stress. In this case, the diagram has three circles, two within a third.

Mohr's circle is used to find the principal stresses, maximum shear stresses, and principal planes. For example, if the material is brittle, the engineer might use Mohr's circle to find the maximum component of normal stress (tension or compression); and for ductile materials, the engineer might look for the maximum shear stress.

See also

Books

  • Dieter, G. E. (3 ed.). (1989). Mechanical Metallurgy. New York: McGraw-Hill. ISBN 0-07-100406-8.
  • 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.

External links