Acoustic resonance

From formulasearchengine
Revision as of 20:44, 27 January 2014 by en>InverseHypercube (→‎Resonance of a sphere of air (Vented))
Jump to navigation Jump to search
Wood is an example of an orthotropic material. Material properties in three perpendicular directions (axial, radial, and circumferential) are different.

An orthotropic material has two or three mutually orthogonal twofold axes of rotational symmetry so that its mechanical properties are, in general, different along each axis. Orthotropic materials are thus anisotropic; their properties depend on the direction in which they are measured. An isotropic material, in contrast, has the same properties in every direction.

One common example of an orthotropic material with two axis of symmetry would be a polymer reinforced by parallel glass or graphite fibers. The strength and stiffness of such a composite material will usually be greater in a direction parallel to the fibers than in the transverse direction. Another example would be a biological membrane, in which the properties in the plane of the membrane will be different from those in the perpendicular direction. Such materials are sometimes called transverse isotropic.

A familiar example of an orthotropic material with three mutually perpendicular axis is wood, in which the properties (such as strength and stiffness) along its grain and in each of the two perpendicular directions are different. Hankinson's equation provides a means to quantify the difference in strength in different directions. Another example is a metal which has been rolled to form a sheet; the properties in the rolling direction and each of the two transverse directions will be different due to the anisotropic structure that develops during rolling.

It is important to keep in mind that a material which is anisotropic on one length scale may be isotropic on another (usually larger) length scale. For instance, most metals are polycrystalline with very small grains. Each of the individual grains may be anisotropic, but if the material as a whole comprises many randomly oriented grains, then its measured mechanical properties will be an average of the properties over all possible orientations of the individual grains.

Orthotropy in physics

Anisotropic material relations

Material behavior is represented in physical theories by constitutive relations. A large class of physical behaviors can be represented by linear material models that take the form of a second-order tensor. The material tensor provides a relation between two vectors and can be written as

where are two vectors representing physical quantities and is the second-order material tensor. If we express the above equation in terms of components with respect to an orthonormal coordinate system, we can write

Summation over repeated indices has been assumed in the above relation. In matrix form we have

Examples of physical problems that fit the above template are listed in the table below.[1]

Problem
Electrical conduction Electrical current
Electric field
Electrical conductivity
Dielectrics Electrical displacement
Electric field
Electric permittivity
Magnetism Magnetic induction
Magnetic field
Magnetic permeability
Thermal conduction Heat flux
Temperature gradient
Thermal conductivity
Diffusion Particle flux
Concentration gradient
Diffusivity
Flow in porous media Weighted fluid velocity
Pressure gradient
Fluid permeability

Condition for material symmetry

The material matrix has a symmetry with respect to a given orthogonal transformation () if it does not change when subjected to that transformation. For invariance of the material properties under such a transformation we require

Hence the condition for material symmetry is (using the definition of an orthogonal transformation)

Orthogonal transformations can be represented in Cartesian coordinates by a matrix given by

Therefore the symmetry condition can be written in matrix form as

Orthotropic material properties

An orthotropic material has three orthogonal symmetry planes. If we choose an orthonormal coordinate system such that the axes coincide with the normals to the three symmetry planes, the transformation matrices are

It can be shown that if the matrix for a material is invariant under reflection about two orthogonal planes then it is also invariant under reflection about the third orthogonal plane.

Consider the reflection about the plane. Then we have

The above relation implies that . Next consider a reflection about the plane. We then have

That implies that . Therefore the material properties of an orthotropic material are described by the matrix

Orthotropy in linear elasticity

Anisotropic elasticity

In linear elasticity, the relation between stress and strain depend on the type of material under consideration. This relation is known as Hooke's law. For anisotropic materials Hooke's law can be written as[2]

where is the stress tensor, is the strain tensor, and is the elastic stiffness tensor. If the tensors in the above expression are described in terms of components with respect to an orthonormal coordinate system we can write

where summation has been assumed over repeated indices. Since the stress and strain tensors are symmetric, and since the stress-strain relation in linear elasticity can be derived from a strain energy density function, the following symmetries hold for linear elastic materials

Because of the above symmetries, the stress-strain relation for linear elastic materials can be expressed in matrix form as

An alternative representation in Voigt notation is

or

The stiffness matrix in the above relation satisfies point symmetry.[3]

Condition for material symmetry

The stiffness matrix satisfies a given symmetry condition if it does not change when subjected to the corresponding orthogonal transformation. The orthogonal transformation may represent symmetry with respect to a point, an axis, or a plane. Orthogonal transformations in linear elasticity include rotations and reflections, but not shape changing transformations and can be represented, in orthonormal coordinates, by a matrix given by

In Voigt notation, the transformation matrix for the stress tensor can be expressed as a matrix given by[3]

The transformation for the strain tensor has a slightly different form because of the choice of notation. This transformation matrix is

It can be shown that .

The elastic properties of a continuum are invariant under an orthogonal transformation if and only if[3]

Stiffness and compliance matrices in orthotropic elasticity

An orthotropic elastic material has three orthogonal symmetry planes. If we choose an orthonormal coordinate system such that the axes coincide with the normals to the three symmetry planes, the transformation matrices are

We can show that if the matrix for a linear elastic material is invariant under reflection about two orthogonal planes then it is also invariant under reflection about the third orthogonal plane.

If we consider the reflection about the plane, then we have

Then the requirement implies that[3]

The above requirement can be satisfied only if

Let us next consider the reflection about the plane. In that case

Using the invariance condition again, we get the additional requirement that

No further information can be obtained because the reflection about third symmetry plane is not independent of reflections about the planes that we have already considered. Therefore, the stiffness matrix of an orthotropic linear elastic material can be written as

The inverse of this matrix is commonly written as[4]

where is the Young's modulus along axis , is the shear modulus in direction on the plane whose normal is in direction , and is the Poisson's ratio that corresponds to a contraction in direction when an extension is applied in direction .

Bounds on the moduli of orthotropic elastic materials

The strain-stress relation for orthotropic linear elastic materials can be written in Voigt notation as

where the compliance matrix is given by

The compliance matrix is symmetric and must be positive definite for the strain energy density to be positive. This implies from Sylvester's criterion that all the principal minors of the matrix are positive,[5] i.e.,

where is the principal submatrix of .

Then,

We can show that this set of conditions implies that[6]

or

However, no similar lower bounds can be placed on the values of the Poisson's ratios .[5]

See also

References

43 year old Petroleum Engineer Harry from Deep River, usually spends time with hobbies and interests like renting movies, property developers in singapore new condominium and vehicle racing. Constantly enjoys going to destinations like Camino Real de Tierra Adentro.

Further reading

  1. Milton, G. W., 2002, The Theory of Composites, Cambridge University Press.
  2. Lekhnitskii, S. G., 1963, Theory of Elasticity of an Anisotropic Elastic Body, Holden-Day Inc.
  3. 3.0 3.1 3.2 3.3 Slawinski, M. A., 2010, Waves and Rays in Elastic Continua: 2nd Ed., World Scientific. http://samizdat.mines.edu/wavesandrays/WavesAndRays.pdf
  4. Boresi, A. P, Schmidt, R. J. and Sidebottom, O. M., 1993, Advanced Mechanics of Materials, Wiley.
  5. 5.0 5.1 Ting, T. C. T. and Chen, T., 2005, Poisson's ratio for anisotropic elastic materials can have no bounds,, Q. J. Mech. Appl. Math., 58(1), pp. 73-82.
  6. Ting, T. C. T., 1996, Positive definiteness of anisotropic elastic constants,, Mathematics and Mechanics of Solids, 1, pp. 301-314.