HOME

TheInfoList



OR:

In
continuum mechanics Continuum mechanics is a branch of mechanics that deals with the mechanical behavior of materials modeled as a continuous mass rather than as discrete particles. The French mathematician Augustin-Louis Cauchy was the first to formulate such m ...
, a compatible
deformation Deformation can refer to: * Deformation (engineering), changes in an object's shape or form due to the application of a force or forces. ** Deformation (physics), such changes considered and analyzed as displacements of continuum bodies. * Defor ...
(or
strain Strain may refer to: Science and technology * Strain (biology), variants of plants, viruses or bacteria; or an inbred animal used for experimental purposes * Strain (chemistry), a chemical stress of a molecule * Strain (injury), an injury to a mu ...
)
tensor field In mathematics and physics, a tensor field assigns a tensor to each point of a mathematical space (typically a Euclidean space or manifold). Tensor fields are used in differential geometry, algebraic geometry, general relativity, in the analysis ...
in a body is that ''unique'' tensor field that is obtained when the body is subjected to a
continuous Continuity or continuous may refer to: Mathematics * Continuity (mathematics), the opposing concept to discreteness; common examples include ** Continuous probability distribution or random variable in probability and statistics ** 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 condition In mathematics, certain systems of partial differential equations are usefully formulated, from the point of view of their underlying geometric and algebraic structure, in terms of a system of differential forms. The idea is to take advantage of the ...
s and were first derived for
linear elasticity 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 mec ...
by Barré de Saint-Venant in 1864 and proved rigorously by Beltrami in 1886.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. 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.Barber, J. R., 2002, Elasticity - 2nd Ed., Kluwer Academic Publications. In the context of
infinitesimal strain theory 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 (indeed, infinitesimal ...
, these conditions are equivalent to stating that the displacements in a body can be obtained by integrating the
strain Strain may refer to: Science and technology * Strain (biology), variants of plants, viruses or bacteria; or an inbred animal used for experimental purposes * Strain (chemistry), a chemical stress of a molecule * Strain (injury), an injury to a mu ...
s. Such an integration is possible if the Saint-Venant's tensor (or incompatibility tensor) \boldsymbol(\boldsymbol) vanishes in a simply-connected bodyN.I. Muskhelishvili, Some Basic Problems of the Mathematical Theory of Elasticity. Leyden: Noordhoff Intern. Publ., 1975. where \boldsymbol is the infinitesimal strain tensor and : \boldsymbol := \boldsymbol\times(\boldsymbol\times\boldsymbol)^T = \boldsymbol ~. For finite deformations the compatibility conditions take the form : \boldsymbol :=\boldsymbol\times\boldsymbol = \boldsymbol where \boldsymbol is the
deformation gradient 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 strai ...
.


Compatibility conditions for infinitesimal strains

The compatibility conditions in
linear elasticity 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 mec ...
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 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 (indeed, infinitesimal ...
problems the strain-displacement relations are : \varepsilon_ = \cfrac ~;~~ \varepsilon_ = \cfrac\left cfrac + \cfrac\right;~~ \varepsilon_ = \cfrac Repeated differentiation of these relations, in order to remove the displacements u_1 and u_2, gives us the two-dimensional compatibility condition for strains : \cfrac - 2\cfrac + \cfrac = 0 The only displacement field that is allowed by a compatible plane strain field is a plane displacement field, i.e., \mathbf = \mathbf(x_1, x_2) .


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 : \cfrac = \cfrac\left \cfrac + \cfrac - \cfrac\right 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 asSlaughter, W. S., 2003, ''The linearized theory of elasticity'', Birkhauser : e_~e_~\varepsilon_ = 0 where e_ is the
permutation symbol In mathematics, particularly in linear algebra, tensor analysis, and differential geometry, the Levi-Civita symbol or Levi-Civita epsilon represents a collection of numbers; defined from the sign of a permutation of the natural numbers , for s ...
. In direct tensor notation : \boldsymbol\times(\boldsymbol\times\boldsymbol)^T = \boldsymbol where the curl operator can be expressed in an orthonormal coordinate system as \boldsymbol\times\boldsymbol = e_\varepsilon_\mathbf_k\otimes\mathbf_r . The second-order tensor : \boldsymbol := \boldsymbol\times(\boldsymbol\times\boldsymbol)^T ~;~~ R_ := e_~e_~\varepsilon_ 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 : \boldsymbol\times\boldsymbol = \boldsymbol where \boldsymbol is the
deformation gradient 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 strai ...
. In terms of components with respect to a Cartesian coordinate system we can write these compatibility relations as : e_~\cfrac = 0 This condition is necessary if the deformation is to be continuous and derived from the mapping \mathbf = \boldsymbol(\mathbf,t) (see
Finite strain theory 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 strai ...
). 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 : R^\gamma_ := \frac Gamma^\gamma_- \frac Gamma^\gamma_+ \Gamma^\gamma_~\Gamma^\mu_ - \Gamma^\gamma_~\Gamma^\mu_ = 0 where \Gamma^k_ is the Christoffel symbol of the second kind. The quantity R^m_ represents the mixed components of the
Riemann-Christoffel curvature tensor In the mathematical field of differential geometry, the Riemann curvature tensor or Riemann–Christoffel tensor (after Bernhard Riemann and Elwin Bruno Christoffel) is the most common way used to express the curvature of Riemannian manifolds. I ...
.


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.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 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 : \mathbf = \mathbf - \mathbf ~;~~ u_i = x_i - X_i Also : \boldsymbol \mathbf = \frac ~;~~ \boldsymbol \mathbf = \frac What conditions on a given second-order tensor field \boldsymbol(\mathbf) on a body are necessary and sufficient so that there exists a unique vector field \mathbf(\mathbf) that satisfies : \boldsymbol \mathbf = \boldsymbol \quad \equiv \quad v_ = A_


Necessary conditions

For the necessary conditions we assume that the field \mathbf exists and satisfies v_ = A_. Then : v_ = A_ ~;~~ v_ = A_ Since changing the order of differentiation does not affect the result we have : v_ = v_ Hence : A_ = A_ From the well known identity for the curl of a tensor we get the necessary condition : \boldsymbol \times \boldsymbol = \boldsymbol


Sufficient conditions

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 \boldsymbol exists such that \boldsymbol \times \boldsymbol = \boldsymbol . We will integrate this field to find the vector field \mathbf along a line between points A and B (see Figure 2), i.e., : \mathbf(\mathbf_B) - \mathbf(\mathbf_A) = \int_^ \boldsymbol \mathbf\cdot~d\mathbf = \int_^ \boldsymbol(\mathbf)\cdot d\mathbf If the vector field \mathbf is to be single-valued then the value of the integral should be independent of the path taken to go from A to B. From
Stokes' theorem Stokes's theorem, also known as the Kelvin–Stokes theorem Nagayoshi Iwahori, et al.:"Bi-Bun-Seki-Bun-Gaku" Sho-Ka-Bou(jp) 1983/12Written in Japanese)Atsuo Fujimoto;"Vector-Kai-Seki Gendai su-gaku rekucha zu. C(1)" :ja:培風館, Bai-Fu-Kan( ...
, the integral of a second order tensor along a closed path is given by : \oint_ \boldsymbol\cdot d\mathbf = \int_ \mathbf\cdot(\boldsymbol \times \boldsymbol)~da Using the assumption that the curl of \boldsymbol is zero, we get : \oint_ \boldsymbol\cdot d\mathbf = 0 \quad \implies \quad \int_ \boldsymbol\cdot d\mathbf + \int_ \boldsymbol\cdot d\mathbf = 0 Hence the integral is path independent and the compatibility condition is sufficient to ensure a unique \mathbf 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 : \boldsymbol = \cfrac = \boldsymbol\mathbf Then the necessary and sufficient conditions for the existence of a compatible \boldsymbol field over a simply connected body are : \boldsymbol\times\boldsymbol = \boldsymbol


Compatibility of infinitesimal strains

The compatibility problem for small strains can be stated as follows. Given a symmetric second order tensor field \boldsymbol when is it possible to construct a vector field \mathbf such that : \boldsymbol = \frac boldsymbol\mathbf + (\boldsymbol\mathbf)^T


Necessary conditions

Suppose that there exists \mathbf such that the expression for \boldsymbol holds. Now : \boldsymbol\mathbf = \boldsymbol + \boldsymbol where : \boldsymbol := \frac boldsymbol\mathbf - (\boldsymbol\mathbf)^T Therefore, in index notation, : \boldsymbol \boldsymbol \equiv \omega_ = \frac (u_ - u_) = \frac (u_ + u_ - u_ - u_) = \varepsilon_ - \varepsilon_ If \boldsymbol is continuously differentiable we have \omega_ = \omega_. Hence, : \varepsilon_ - \varepsilon_ - \varepsilon_ + \varepsilon_ = 0 In direct tensor notation : \boldsymbol \times (\boldsymbol \times\boldsymbol)^T = \boldsymbol The above are necessary conditions. If \mathbf is the infinitesimal rotation vector then \boldsymbol \times \boldsymbol = \boldsymbol \mathbf+\boldsymbol \mathbf^T. Hence the necessary condition may also be written as \boldsymbol \times ( \boldsymbol \mathbf+\boldsymbol \mathbf^T)^T = \boldsymbol.


Sufficient conditions

Let us now assume that the condition \boldsymbol \times (\boldsymbol \times\boldsymbol)^T = \boldsymbol is satisfied in a portion of a body. Is this condition sufficient to guarantee the existence of a continuous, single-valued displacement field \mathbf? The first step in the process is to show that this condition implies that the infinitesimal rotation tensor \boldsymbol is uniquely defined. To do that we integrate \boldsymbol \mathbf along the path \mathbf_A to \mathbf_B, i.e., : \mathbf(\mathbf_B) - \mathbf(\mathbf_A) = \int_^ \boldsymbol \mathbf\cdot d\mathbf = \int_^ (\boldsymbol \times \boldsymbol)\cdot d\mathbf Note that we need to know a reference \mathbf(\mathbf_A) to fix the rigid body rotation. The field \mathbf(\mathbf) is uniquely determined only if the contour integral along a closed contour between \mathbf_A and \mathbf_b is zero, i.e., : \oint_^ (\boldsymbol \times \boldsymbol)\cdot d\mathbf = \boldsymbol But from Stokes' theorem for a simply-connected body and the necessary condition for compatibility : \oint_^ (\boldsymbol \times \boldsymbol)\cdot d\mathbf = \int_ \mathbf\cdot(\boldsymbol \times \boldsymbol\times\boldsymbol)~da = \boldsymbol Therefore, the field \mathbf is uniquely defined which implies that the infinitesimal rotation tensor \boldsymbol 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 \mathbf. As before we integrate the displacement gradient : \mathbf(\mathbf_B) - \mathbf(\mathbf_A) = \int_^ \boldsymbol \mathbf\cdot d\mathbf = \int_^ (\boldsymbol + \boldsymbol)\cdot d\mathbf From Stokes' theorem and using the relations \boldsymbol \times \boldsymbol = \boldsymbol \mathbf = -\boldsymbol \times \omega we have : \oint_^ (\boldsymbol + \boldsymbol)\cdot d\mathbf = \int_ \mathbf\cdot(\boldsymbol \times \boldsymbol+\boldsymbol \times \boldsymbol)~da = \boldsymbol Hence the displacement field \mathbf is also determined uniquely. Hence the compatibility conditions are sufficient to guarantee the existence of a unique displacement field \mathbf 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 \boldsymbol(\mathbf) be a positive definite symmetric tensor field defined on the reference configuration. Under what conditions on \boldsymbol does there exist a deformed configuration marked by the position field \mathbf(\mathbf) such that : (1)\quad\left(\frac\right)^T \left(\frac\right) = \boldsymbol


Necessary conditions

Suppose that a field \mathbf(\mathbf) exists that satisfies condition (1). In terms of components with respect to a rectangular Cartesian basis : \frac\frac = C_ From
finite strain theory 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 strai ...
we know that C_ = g_. Hence we can write : \delta_~\frac~\frac = g_ For two symmetric second-order tensor field that are mapped one-to-one we also have the
relation Relation or relations may refer to: General uses *International relations, the study of interconnection of politics, economics, and law on a global level *Interpersonal relationship, association or acquaintance between two or more people *Public ...
: G_ = \frac~\frac~g_ From the relation between of G_ and g_ that \delta_ = G_, we have : _\Gamma_^k = 0 Then, from the relation : \frac = \frac\,_\Gamma^\mu_ - \frac~\frac \,_\Gamma^m_ we have : \frac = F^m_\,_\Gamma^\mu_ \qquad; ~~ F^i_ := \frac From
finite strain theory 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 strai ...
we also have : _\Gamma_ = \frac\left(\frac + \frac - \frac\right) ~;~~ _\Gamma^\nu_ = g^ \,_\Gamma_ ~;~~ g_ = C_ ~;~~ g^ = C^ Therefore, : \,_\Gamma^\mu_ = \cfrac\left(\frac + \frac - \frac\right) and we have : \frac = F^m_~\cfrac\left(\frac + \frac - \frac\right) Again, using the commutative nature of the order of differentiation, we have : \frac = \frac \implies \frac\,_\Gamma^\mu_ + F^m_~\frac ,_\Gamma^\mu_= \frac\,_\Gamma^\mu_ + F^m_~\frac ,_\Gamma^\mu_ or : F^m_\,_\Gamma^\gamma_\,_\Gamma^\mu_ + F^m_~\frac ,_\Gamma^\mu_= F^m_\,_\Gamma^\gamma_\,_\Gamma^\mu_ + F^m_~\frac ,_\Gamma^\mu_ After collecting terms we get : F^m_\left(\,_\Gamma^\gamma_\,_\Gamma^\mu_ + \frac ,_\Gamma^\gamma_- \,_\Gamma^\gamma_\,_\Gamma^\mu_ - \frac ,_\Gamma^\gamma_right) = 0 From the definition of F^m_ we observe that it is invertible and hence cannot be zero. Therefore, : R^\gamma_ := \frac ,_\Gamma^\gamma_- \frac ,_\Gamma^\gamma_+ \,_\Gamma^\gamma_\,_\Gamma^\mu_ - \,_\Gamma^\gamma_\,_\Gamma^\mu_ = 0 We can show these are the mixed components of the
Riemann-Christoffel curvature tensor In the mathematical field of differential geometry, the Riemann curvature tensor or Riemann–Christoffel tensor (after Bernhard Riemann and Elwin Bruno Christoffel) is the most common way used to express the curvature of Riemannian manifolds. I ...
. Therefore, the necessary conditions for \boldsymbol-compatibility are that the Riemann-Christoffel curvature of the deformation is zero.


Sufficient conditions

The proof of sufficiency is a bit more involved.Blume, J. A., 1989, "Compatibility conditions for a left Cauchy-Green strain field", J. Elasticity, v. 21, p. 271-308. We start with the assumption that : R^\gamma_ = 0 ~;~~ g_ = C_ We have to show that there exist \mathbf and \mathbf such that : \frac\frac = C_ From a theorem by T.Y.Thomas Thomas, T. Y., 1934, "Systems of total differential equations defined over simply connected domains", Annals of Mathematics, 35(4), p. 930-734 we know that the system of equations : \frac = F^i_~\,_\Gamma^\gamma_ has unique solutions F^i_ over simply connected domains if : _\Gamma^\gamma_ = _\Gamma^\gamma_ ~;~~ R^\gamma_ = 0 The first of these is true from the defining of \Gamma^i_ and the second is assumed. Hence the assumed condition gives us a unique F^i_ that is C^2 continuous. Next consider the system of equations : \frac = F^i_ Since F^i_ is C^2 and the body is simply connected there exists some solution x^i(X^\alpha) to the above equations. We can show that the x^i also satisfy the property that : \det\left, \frac\ \ne 0 We can also show that the relation : \frac~g^~\frac = \delta^ implies that : g_ = C_ = \frac~\frac If we associate these quantities with tensor fields we can show that \frac is invertible and the constructed tensor field satisfies the expression for \boldsymbol{C}.


See also

*
Saint-Venant's compatibility condition In the mathematical theory of elasticity (mathematics), elasticity, Saint-Venant's compatibility condition defines the relationship between the Deformation (mechanics), strain \varepsilon and a displacement field (mechanics), displacement field \ u ...
*
Linear elasticity 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 mec ...
*
Deformation (mechanics) In physics, deformation is the continuum mechanics transformation of a body from a ''reference'' configuration to a ''current'' configuration. A configuration is a set containing the positions of all particles of the body. A deformation can ...
*
Infinitesimal strain theory 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 (indeed, infinitesimal ...
*
Finite strain theory 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 strai ...
*
Tensor derivative (continuum mechanics) The derivatives of scalars, vectors, and second-order tensors with respect to second-order tensors are of considerable use in continuum mechanics. These derivatives are used in the theories of nonlinear elasticity and plasticity, particularly i ...
*
Curvilinear coordinates In geometry, curvilinear coordinates are a coordinate system for Euclidean space in which the coordinate lines may be curved. These coordinates may be derived from a set of Cartesian coordinates by using a transformation that is locally inve ...


References


External links


Amit Acharya's notes on compatibility on iMechanicaPlasticity by J. Lubliner, sec. 1.2.4 p. 35
Continuum mechanics Elasticity (physics)