Compatibility (mechanics)

Last updated

In continuum mechanics, a compatible deformation (or strain) tensor field in a body is that unique tensor field that is obtained when the body is subjected to a continuous, single-valued, displacement field. Compatibility is the study of the conditions under which such a displacement field can be guaranteed. Compatibility conditions are particular cases of integrability conditions and were first derived for linear elasticity by Barré de Saint-Venant in 1864 and proved rigorously by Beltrami in 1886. [1]

Contents

In the continuum description of a solid body we imagine the body to be composed of a set of infinitesimal volumes or material points. Each volume is assumed to be connected to its neighbors without any gaps or overlaps. Certain mathematical conditions have to be satisfied to ensure that gaps/overlaps do not develop when a continuum body is deformed. A body that deforms without developing any gaps/overlaps is called a compatible body. Compatibility conditions are mathematical conditions that determine whether a particular deformation will leave a body in a compatible state. [2]

In the context of infinitesimal strain theory, these conditions are equivalent to stating that the displacements in a body can be obtained by integrating the strains. Such an integration is possible if the Saint-Venant's tensor (or incompatibility tensor) vanishes in a simply-connected body [3] where is the infinitesimal strain tensor and

For finite deformations the compatibility conditions take the form

where is the deformation gradient.

Compatibility conditions for infinitesimal strains

The compatibility conditions in linear elasticity are obtained by observing that there are six strain-displacement relations that are functions of only three unknown displacements. This suggests that the three displacements may be removed from the system of equations without loss of information. The resulting expressions in terms of only the strains provide constraints on the possible forms of a strain field.

2-dimensions

For two-dimensional, plane strain problems the strain-displacement relations are

Repeated differentiation of these relations, in order to remove the displacements and , gives us the two-dimensional compatibility condition for strains

The only displacement field that is allowed by a compatible plane strain field is a plane displacement field, i.e., .

3-dimensions

In three dimensions, in addition to two more equations of the form seen for two dimensions, there are three more equations of the form

Therefore, there are 34=81 partial differential equations, however due to symmetry conditions, this number reduces to six different compatibility conditions. We can write these conditions in index notation as [4]

where is the permutation symbol. In direct tensor notation

where the curl operator can be expressed in an orthonormal coordinate system as .

The second-order tensor

is known as the incompatibility tensor, and is equivalent to the Saint-Venant compatibility tensor

Compatibility conditions for finite strains

For solids in which the deformations are not required to be small, the compatibility conditions take the form

where is the deformation gradient. In terms of components with respect to a Cartesian coordinate system we can write these compatibility relations as

This condition is necessary if the deformation is to be continuous and derived from the mapping (see Finite strain theory). The same condition is also sufficient to ensure compatibility in a simply connected body.

Compatibility condition for the right Cauchy-Green deformation tensor

The compatibility condition for the right Cauchy-Green deformation tensor can be expressed as

where is the Christoffel symbol of the second kind. The quantity represents the mixed components of the Riemann-Christoffel curvature tensor.

The general compatibility problem

The problem of compatibility in continuum mechanics involves the determination of allowable single-valued continuous fields on simply connected bodies. More precisely, the problem may be stated in the following manner. [5]

Figure 1. Motion of a continuum body. Displacement of a continuum.svg
Figure 1. Motion of a continuum body.

Consider the deformation of a body shown in Figure 1. If we express all vectors in terms of the reference coordinate system , the displacement of a point in the body is given by

Also

What conditions on a given second-order tensor field on a body are necessary and sufficient so that there exists a unique vector field that satisfies

Necessary conditions

For the necessary conditions we assume that the field exists and satisfies . Then

Since changing the order of differentiation does not affect the result we have

Hence

From the well known identity for the curl of a tensor we get the necessary condition

Sufficient conditions

Figure 2. Integration paths used in proving the sufficiency conditions for compatibility. Compatibility mechanics.png
Figure 2. Integration paths used in proving the sufficiency conditions for compatibility.

To prove that this condition is sufficient to guarantee existence of a compatible second-order tensor field, we start with the assumption that a field exists such that . We will integrate this field to find the vector field along a line between points and (see Figure 2), i.e.,

If the vector field is to be single-valued then the value of the integral should be independent of the path taken to go from to .

From Stokes' theorem, the integral of a second order tensor along a closed path is given by

Using the assumption that the curl of is zero, we get

Hence the integral is path independent and the compatibility condition is sufficient to ensure a unique field, provided that the body is simply connected.

Compatibility of the deformation gradient

The compatibility condition for the deformation gradient is obtained directly from the above proof by observing that

Then the necessary and sufficient conditions for the existence of a compatible field over a simply connected body are

Compatibility of infinitesimal strains

The compatibility problem for small strains can be stated as follows.

Given a symmetric second order tensor field when is it possible to construct a vector field such that

Necessary conditions

Suppose that there exists such that the expression for holds. Now

where

Therefore, in index notation,

If is continuously differentiable we have . Hence,

In direct tensor notation

The above are necessary conditions. If is the infinitesimal rotation vector then . Hence the necessary condition may also be written as .

Sufficient conditions

Let us now assume that the condition is satisfied in a portion of a body. Is this condition sufficient to guarantee the existence of a continuous, single-valued displacement field ?

The first step in the process is to show that this condition implies that the infinitesimal rotation tensor is uniquely defined. To do that we integrate along the path to , i.e.,

Note that we need to know a reference to fix the rigid body rotation. The field is uniquely determined only if the contour integral along a closed contour between and is zero, i.e.,

But from Stokes' theorem for a simply-connected body and the necessary condition for compatibility

Therefore, the field is uniquely defined which implies that the infinitesimal rotation tensor is also uniquely defined, provided the body is simply connected.

In the next step of the process we will consider the uniqueness of the displacement field . As before we integrate the displacement gradient

From Stokes' theorem and using the relations we have

Hence the displacement field is also determined uniquely. Hence the compatibility conditions are sufficient to guarantee the existence of a unique displacement field in a simply-connected body.

Compatibility for Right Cauchy-Green Deformation field

The compatibility problem for the Right Cauchy-Green deformation field can be posed as follows.

Problem: Let be a positive definite symmetric tensor field defined on the reference configuration. Under what conditions on does there exist a deformed configuration marked by the position field such that

Necessary conditions

Suppose that a field exists that satisfies condition (1). In terms of components with respect to a rectangular Cartesian basis

From finite strain theory we know that . Hence we can write

For two symmetric second-order tensor field that are mapped one-to-one we also have the relation

From the relation between of and that , we have

Then, from the relation

we have

From finite strain theory we also have

Therefore,

and we have

Again, using the commutative nature of the order of differentiation, we have

or

After collecting terms we get

From the definition of we observe that it is invertible and hence cannot be zero. Therefore,

We can show these are the mixed components of the Riemann-Christoffel curvature tensor. Therefore, the necessary conditions for -compatibility are that the Riemann-Christoffel curvature of the deformation is zero.

Sufficient conditions

The proof of sufficiency is a bit more involved. [5] [6] We start with the assumption that

We have to show that there exist and such that

From a theorem by T.Y.Thomas [7] we know that the system of equations

has unique solutions over simply connected domains if

The first of these is true from the defining of and the second is assumed. Hence the assumed condition gives us a unique that is continuous.

Next consider the system of equations

Since is and the body is simply connected there exists some solution to the above equations. We can show that the also satisfy the property that

We can also show that the relation

implies that

If we associate these quantities with tensor fields we can show that is invertible and the constructed tensor field satisfies the expression for .

See also

Related Research Articles

Navier–Stokes equations Equations describing the motion of viscous fluid substances

In physics, the Navier–Stokes equations are a set of partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician George Gabriel Stokes.

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

In mathematics, particularly in linear algebra, tensor analysis, and differential geometry, the Levi-Civita symbol represents a collection of numbers; defined from the sign of a permutation of the natural numbers 1, 2, …, n, for some positive integer n. It is named after the Italian mathematician and physicist Tullio Levi-Civita. Other names include the permutation symbol, antisymmetric symbol, or alternating symbol, which refer to its antisymmetric property and definition in terms of permutations.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

Onsager reciprocal relations Relations between flows and forces, or gradients, in thermodynamic systems

In thermodynamics, the Onsager reciprocal relations express the equality of certain ratios between flows and forces in thermodynamic systems out of equilibrium, but where a notion of local equilibrium exists.

In differential geometry, the Cotton tensor on a (pseudo)-Riemannian manifold of dimension n is a third-order tensor concomitant of the metric, like the Weyl tensor. The vanishing of the Cotton tensor for n = 3 is necessary and sufficient condition for the manifold to be conformally flat, as with the Weyl tensor for n ≥ 4. For n < 3 the Cotton tensor is identically zero. The concept is named after Émile Cotton.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

Electromagnetic tensor

In electromagnetism, the electromagnetic tensor or electromagnetic field tensor is a mathematical object that describes the electromagnetic field in spacetime. The field tensor was first used after the four-dimensional tensor formulation of special relativity was introduced by Hermann Minkowski. The tensor allows related physical laws to be written very concisely.

In continuum mechanics, the finite strain theory—also called large strain theory, or large deformation theory—deals with deformations in which strains and/or rotations are large enough to invalidate assumptions inherent in infinitesimal strain theory. In this case, the undeformed and deformed configurations of the continuum are significantly different, requiring a clear distinction between them. This is commonly the case with elastomers, plastically-deforming materials and other fluids and biological soft tissue.

Parameterized post-Newtonian formalism

In general relativity and many alternatives to it, the post-Newtonian formalism is a calculational tool that expresses Einstein's (nonlinear) equations of gravity in terms of the lowest-order deviations from Newton's law of universal gravitation. This allows approximations to Einstein's equations to be made in the case of weak fields. Higher-order terms can be added to increase accuracy, but for strong fields, it may be preferable to solve the complete equations numerically. Some of these post-Newtonian approximations are expansions in a small parameter, which is the ratio of the velocity of the matter forming the gravitational field to the speed of light, which in this case is better called the speed of gravity. In the limit, when the fundamental speed of gravity becomes infinite, the post-Newtonian expansion reduces to Newton's law of gravity.

Electromagnetic stress–energy tensor

In relativistic physics, the electromagnetic stress–energy tensor is the contribution to the stress–energy tensor due to the electromagnetic field. The stress–energy tensor describes the flow of energy and momentum in spacetime. The electromagnetic stress–energy tensor contains the negative of the classical Maxwell stress tensor that governs the electromagnetic interactions.

Covariant formulation of classical electromagnetism

The covariant formulation of classical electromagnetism refers to ways of writing the laws of classical electromagnetism in a form that is manifestly invariant under Lorentz transformations, in the formalism of special relativity using rectilinear inertial coordinate systems. These expressions both make it simple to prove that the laws of classical electromagnetism take the same form in any inertial coordinate system, and also provide a way to translate the fields and forces from one frame to another. However, this is not as general as Maxwell's equations in curved spacetime or non-rectilinear coordinate systems.

Maxwells equations in curved spacetime electromagnetism in general relativity

In physics, Maxwell's equations in curved spacetime govern the dynamics of the electromagnetic field in curved spacetime or where one uses an arbitrary coordinate system. These equations can be viewed as a generalization of the vacuum Maxwell's equations which are normally formulated in the local coordinates of flat spacetime. But because general relativity dictates that the presence of electromagnetic fields induce curvature in spacetime, Maxwell's equations in flat spacetime should be viewed as a convenient approximation.

Inhomogeneous electromagnetic wave equation wave equations describing the propagation of electromagnetic waves generated by nonzero source charges and currents

In electromagnetism and applications, an inhomogeneous electromagnetic wave equation, or nonhomogeneous electromagnetic wave equation, is one of a set of wave equations describing the propagation of electromagnetic waves generated by nonzero source charges and currents. The source terms in the wave equations make the partial differential equations inhomogeneous, if the source terms are zero the equations reduce to the homogeneous electromagnetic wave equations. The equations follow from Maxwell's equations.

Newman–Penrose formalism Notation in general relativity

The Newman–Penrose (NP) formalism is a set of notation developed by Ezra T. Newman and Roger Penrose for general relativity (GR). Their notation is an effort to treat general relativity in terms of spinor notation, which introduces complex forms of the usual variables used in GR. The NP formalism is itself a special case of the tetrad formalism, where the tensors of the theory are projected onto a complete vector basis at each point in spacetime. Usually this vector basis is chosen to reflect some symmetry of the spacetime, leading to simplified expressions for physical observables. In the case of the NP formalism, the vector basis chosen is a null tetrad: a set of four null vectors—two real, and a complex-conjugate pair. The two real members asymptotically point radially inward and radially outward, and the formalism is well adapted to treatment of the propagation of radiation in curved spacetime. The Weyl scalars, derived from the Weyl tensor, are often used. In particular, it can be shown that one of these scalars— in the appropriate frame—encodes the outgoing gravitational radiation of an asymptotically flat system.

Mathematical descriptions of the electromagnetic field Formulations of electromagnetism

There are various mathematical descriptions of the electromagnetic field that are used in the study of electromagnetism, one of the four fundamental interactions of nature. In this article, several approaches are discussed, although the equations are in terms of electric and magnetic fields, potentials, and charges with currents, generally speaking.

Classical electromagnetism and special relativity Relationship between relativity and pre-quantum electromagnetism

The theory of special relativity plays an important role in the modern theory of classical electromagnetism. First of all, it gives formulas for how electromagnetic objects, in particular the electric and magnetic fields, are altered under a Lorentz transformation from one inertial frame of reference to another. Secondly, it sheds light on the relationship between electricity and magnetism, showing that frame of reference determines if an observation follows electrostatic or magnetic laws. Third, it motivates a compact and convenient notation for the laws of electromagnetism, namely the "manifestly covariant" tensor form.

Kirchhoff–Love plate theory

The Kirchhoff–Love theory of plates is a two-dimensional mathematical model that is used to determine the stresses and deformations in thin plates subjected to forces and moments. This theory is an extension of Euler-Bernoulli beam theory and was developed in 1888 by Love using assumptions proposed by Kirchhoff. The theory assumes that a mid-surface plane can be used to represent a three-dimensional plate in two-dimensional form.

The Optical Metric was defined by German theoretical physicist Walter Gordon in 1923 to study the geometrical optics in curved space-time filled with moving dielectric materials. Let ua be the normalized (covariant) 4-velocity of the arbitrarily-moving dielectric medium filling the space-time, and assume that the fluid’s electromagnetic properties are linear, isotropic, transparent, nondispersive, and can be summarized by two scalar functions: a dielectric permittivity ε and a magnetic permeability μ. Then optical metric tensor is defined as

References

  1. C Amrouche, PG Ciarlet, L Gratie, S Kesavan, On Saint Venant's compatibility conditions and Poincaré's lemma, C. R. Acad. Sci. Paris, Ser. I, 342 (2006), 887-891. doi : 10.1016/j.crma.2006.03.026
  2. Barber, J. R., 2002, Elasticity - 2nd Ed., Kluwer Academic Publications.
  3. N.I. Muskhelishvili, Some Basic Problems of the Mathematical Theory of Elasticity. Leyden: Noordhoff Intern. Publ., 1975.
  4. Slaughter, W. S., 2003, The linearized theory of elasticity, Birkhauser
  5. 1 2 Acharya, A., 1999, On Compatibility Conditions for the Left Cauchy–Green Deformation Field in Three Dimensions, Journal of Elasticity, Volume 56, Number 2 , 95-105
  6. Blume, J. A., 1989, "Compatibility conditions for a left Cauchy-Green strain field", J. Elasticity, v. 21, p. 271-308.
  7. Thomas, T. Y., 1934, "Systems of total differential equations defined over simply connected domains", Annals of Mathematics, 35(4), p. 930-734