Elasticity tensor: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Create article from redirect page
Tag: Removed redirect
Line 1: Line 1:
The '''elasticity tensor''' is a fourth-rank [[tensor]] describing the [[Hooke's law|stress-strain relation]] in
#REDIRECT [[Hooke's law#Hooke's law for continuous media]]
a [[Linear elasticity|linear elastic]] material.{{sfn|Thorne|Blandford|2017|p=580}}{{sfn|Itin|Hehl|2013}} Other names are '''elastic modulus tensor''' and '''stiffness tensor'''. Common symbols include <math>\mathbf{C}</math> and <math>\mathbf{Y}</math>.

The defining equation can be written as
:<math>
T^{ij} = C^{ijkl} E_{kl}
</math>
where <math>T^{ij}</math> and <math>E_{kl}</math> are the compoments of the [[Cauchy stress tensor]] and [[infinitesimal strain tensor]], and <math>C^{ijkl}</math> are the components of the elasticity tensor. Summation over repeated indices is implied.<ref group=note name=n1>Here, upper and lower indices denote [[Covariance and contravariance of vectors|contravariant and covariant components]], respectively, though the distinction can be ignored for [[Cartesian coordinates]]. As a result, some references represent components using only lower indices.</ref> This relationship can be interpreted as a generalization of [[Hooke's law]] to a 3D [[Continuum mechanics|continuum]].

A general fourth-rank tensor <math>\mathbf{F}</math> in 3D has 3<sup>4</sup> = 81 independent components <math>F_{ijkl}</math>, but the elasticity tensor has at most 21 independent components.{{sfn|Thorne|Blandford|2017|p=581}} This fact follows from the symmetry of the stress and strain tensors, together with the requirement that the stress derives from an [[elastic energy]] potential. For [[isotropic]] materials, the elasticity tensor has just two independent components, which can be chosen to be the [[bulk modulus]] and [[shear modulus]].{{sfn|Thorne|Blandford|2017|p=581}}

== Definition ==
The most general linear relation between two second-rank tensors <math>\mathbf{T}, \mathbf{E}</math> is
:<math>
T^{ij} = C^{ijkl} E_{kl}
</math>
where <math>C^{ijkl}</math> are the components of a fourth-rank tensor <math>\mathbf{C}</math>.{{sfn|Thorne|Blandford|2017|p=580}}<ref group=note name=n1/> The elasticity tensor is defined as <math>\mathbf{C}</math> for the case where <math>\mathbf{T}</math> and <math>\mathbf{E}</math> are the stress and strain tensors, respectively.

The '''compliance tensor''' <math>\mathbf{K}</math> is defined from the inverse stress-strain relation:
:<math>
E^{ij} = K^{ijkl} T_{kl}
</math>
The two are related by
:<math>
K_{ijpq} C^{pqkl} = \frac{1}{2} \left(\delta_{i}^{k} \delta_{j}^{l} + \delta_{i}^{l} \delta_{j}^{k} \right)
</math>
where <math>\delta_{n}^{m}</math> is the [[Kronecker delta]].{{sfn|Hill|1965}}{{sfn|Cowin|1989}}<ref group=note>Combining the forward and inverse stress-strain relations gives {{math|''E''<sub>''ij''</sub> {{=}} ''K''<sub>''ijpq''</sub> ''C''<sup>''pqkl''</sup>''E''<sub>''kl''</sub>}}.
Due to the minor symmetries {{math|''C''<sup>''pqkl''</sup> {{=}} ''C''<sup>''qpkl''</sup>}} and {{math|''C''<sup>''pqkl''</sup> {{=}} ''C''<sup>''pqlk''</sup>}}, this equation does not uniquely determine {{math|''K''<sub>''ijpq''</sub> ''C''<sup>''pqkl''</sup>}}. In fact, {{math|''K''<sub>''ijpq''</sub> ''C''<sup>''pqkl'' </sup>{{=}}'' a'' δ<sup>''k''</sup><sub>''i''</sub>δ<sup>''l''</sup><sub>''j''</sub> + (1 − ''a'') δ<sup>''l''</sup><sub>''i''</sub>δ<sup>''k''</sup><sub>''j''</sub>}} is a solution for any {{math|0 ≤ ''a'' ≤ 1}}. However, only {{math|''a'' {{=}} 1/2}} preserves the minor symmetries of '''K''', so this is the correct solution from a physical standpoint.</ref>

Unless otherwise noted, this article assumes <math>\mathbf{C}</math> is defined from the stress-strain relation of a linear elastic material, in the limit of small strain.

== Special cases ==
=== Isotropic ===
For an isotropic material, <math>\mathbf{C}</math> simplifies to
:<math>
C^{ijkl} = \lambda \!\left( X \right) g^{ij} g^{kl} + \mu\!\left( X \right) \left(g^{ik} g^{jl} + g^{il} g^{kj} \right)
</math>
where <math>\lambda</math> and <math>\mu</math> are scalar functions of the material coordinates
<math>X</math>, and <math>\mathbf{g}</math> is the [[metric tensor]] in the reference frame of the material.{{sfn|Marsden|Hughes|1994|p=223}}{{sfn|Hehl|Itin|2002}} In an orthonormal Cartesian coordinate basis, there is no distinction between upper and lower indices, and the metric tensor can be replaced with the Kronecker delta:
:<math>
C_{ijkl} = \lambda \!\left( X \right) \delta_{ij} \delta_{kl} + \mu\!\left( X \right) \left(\delta_{ik} \delta_{jl} + \delta_{il} \delta_{kj} \right) \quad \text{[Cartesian coordinates]}</math>

Substituting the first equation into the stress-strain relation and summing over repeated indices gives
:<math>
T^{ij} = \lambda \!\left( X \right) \cdot \left(\mathrm{Tr} \, \mathbf{E} \right) g^{ij} + 2 \mu\!\left( X \right) E^{ij}
</math>
where <math>\mathrm{Tr}\, \mathbf{E} \equiv E^i_{\,i}</math> is the trace of <math>\mathbf{E}</math>.
In this form, <math>\mu</math> and <math>\lambda</math> can be identified with the first and second [[Lame coefficients]].
An equivalent expression is
:<math>
T^{ij} = K \!\left( X \right) \cdot \left(\mathrm{Tr}\, \mathbf{E} \right) g^{ij} + 2 \mu\!\left( X \right)\Sigma^{ij}
</math>
where <math>K = \lambda + (2/3) \mu</math> is the bulk modulus, and
:<math>
\Sigma^{ij} \equiv E^{ij} - (1/3)\left(\mathrm{Tr}\, \mathbf{E} \right) g^{ij}
</math>
are the components of the shear tensor <math>\mathbf{\Sigma}</math>.

===Cubic crystals===
The elasticity tensor of a [[cubic crystal]] has components
:<math>
\begin{align}
C^{ijkl} &= \lambda g^{ij} g^{kl} + \mu \left(g^{ik} g^{jl} + g^{il} g^{kj} \right) \\ &+ \alpha \left(a^i a^j a^k a^l + b^i b^j b^k b^l + c^i c^j c^k c^l\right)
\end{align}
</math>
where <math>\mathbf{a}</math>, <math>\mathbf{b}</math>, and <math>\mathbf{c}</math> are unit vectors corresponding to the three mutually perpendicular axes of the crystal [[unit cell]].{{sfn|Thomas|1966}} The coefficients <math>\lambda</math>, <math>\mu</math>, and <math>\alpha</math> are scalars; because they are coordinate-independent, they are intrinsic material constants. Thus, a crystal with cubic symmetry is described by three independent elastic constants.{{sfn|Landau|Lifshitz|1970}}

In an orthonormal Cartesian coordinate basis, there is no distinction between upper and lower indices, and <math>g^{ij}</math> is the Kronecker delta, so the expression simplifies to
:<math>
\begin{align}
C_{ijkl} &= \lambda \delta_{ij} \delta_{kl} + \mu \left(\delta_{ik} \delta_{jl} + \delta_{il} \delta_{kj} \right) \\ &+ \alpha \left(a_i a_j a_k a_l + b_i b_j b_k b_l + c_i c_j c_k c_l\right)
\end{align}
</math>

=== Other crystal classes ===
There are similar expressions for the components of <math>\mathbf{C}</math> in other crystal symmetry classes.{{sfn|Srinivasan|Nigam|1969}} The number of independent elastic constants for several of these is given in table 1.{{sfn|Landau|Lifshitz|1970}}
{| class="wikitable"
|+ Table 1: Number of independent elastic constants for various crystal symmetry classes.{{sfn|Landau|Lifshitz|1970}}
! Crystal family
! [[Crystallographic point group|Point group]]
! Independent components
|-
| [[Triclinic]]
|
| 21
|-
| [[Monoclinic]]
|
| 13
|-
| [[Orthorhombic]]
|
| 9
|-
| [[Tetragonal]]
| '''C'''<sub>4</sub>, '''S'''<sub>4</sub>, '''C'''<sub>4h</sub>
| 7
|-
| Tetragonal
| '''C'''<sub>4v</sub>, '''D'''<sub>2d</sub>, '''D'''<sub>4</sub>, '''D'''<sub>4h</sub>
| 6
|-
| [[Rhombohedral]]
| '''C'''<sub>3</sub>, '''S'''<sub>6</sub>
| 7
|-
| Rhombohedral
| '''C'''<sub>3v</sub>, '''D'''<sub>6</sub>, '''D'''<sub>3d</sub>
| 6
|-
| [[Hexagonal]]
|
| 5
|-
| [[Cubic crystal system|Cubic]]
|
| 3
|}

== Properties ==
=== Symmetries ===
The elasticity tensor has several symmetries that follow directly from its defining equation <math>T^{ij} = C^{ijkl} E_{kl}</math>.{{sfn|Thorne|Blandford|2017}}{{sfn|Itin|Hehl|2013|p=}} The symmetry of the stress and strain tensors implies that
:<math>
C_{ijkl} = C_{jikl} \qquad \text{and} \qquad C_{ijkl} = C_{jilk},
</math>
Usually, one also assumes that the stress derives from an elastic energy potential <math>U</math>:
:<math>
T^{ij} = \frac{\partial U}{\partial E_{ij}}
</math>
which implies
:<math>
C_{ijkl} = \frac{\partial^2 U}{\partial E_{ij}\partial E_{kl}}
</math>
Hence, <math>\mathbf{C}</math> must be symmetric under interchange of the first and second pairs of indices:
:<math>
C_{ijkl} = C_{klij}
</math>
The symmetries listed above reduce the number of independent components from 81 to 21. If a material has additional symmetries, then this number is further reduced.{{sfn|Landau|Lifshitz|1970}}

=== Transformations ===
Under rotation, the components <math>C^{ijkl}</math> transform as
:<math>
C'_{ijkl} = R_{ip} R_{jq} R_{kr} R_{ls} C^{pqrs}</math>
where <math>C'_{ijkl}</math> are the covariant components in the rotated basis, and <math>R_{ij}</math>
are the elements of the corresponding [[rotation matrix]]. A similar transformation rule holds for other linear transformations.

=== Invariants ===
The components of <math>\mathbf{C}</math> generally acquire different values under a change of basis. Nevertheless, for certain types of transformations,
there are specific combinations of components, called invariants, that remain unchanged. Invariants are defined with respect to a given set of transformations, formally known as a [[group operation]]. For example, an invariant with respect to the group of proper orthogonal transformations, called SO(3), is a quantity that remains constant under arbitrary 3D rotations.
<math>\mathbf{C}</math> possesses two linear invariants and seven quadratic invariants with respect to SO(3).{{sfn|Norris|2007}} The linear invariants are
:<math>
\begin{align}
L_1 &= C^{ij}_{\,\,\,ij} \\
L_2 &= C^{ii}_{\,\,\,jj}
\end{align}
</math>
and the quadratic invariants are
:<math>
\left\{ L_1^2, \, L_2^2, \, L_1 L_2, \, C_{ijkl} C^{ijkl}, \,C_{iikl} C^{jjkl}, \,C_{iikl} C^{jkjl}, \,C_{kiil} C^{kjjl} \right\}
</math>
These quantities are linearly independent, that is, none can be expressed as a linear combination of the others.
They are also complete, in the sense that there are no additional independent linear or quadratic invariants.{{sfn|Norris|2007}}

== Decompositions ==
A common strategy in tensor analysis is to decompose a tensor into simpler components that can be analyzed separately. For example, the
displacement gradient tensor <math>\mathbf{W} = \mathbf{\nabla} \mathbf{\xi}</math> can be decomposed as
:<math>
\mathbf{W} = \frac{1}{3} \Theta \mathbf{g} + \mathbf{\Sigma} + \mathbf{R}
</math>
where <math>\Theta</math> is a rank-0 tensor (a scalar), equal to the trace of <math>\mathbf{W}</math>;
<math>\mathbf{\Sigma}</math> is symmetric and trace-free; and <math>\mathbf{R}</math> is antisymmetric.{{sfn|Thorne|Blandford|2017|p=571}} Component-wise,
:<math>
\begin{align}
\Sigma^{ij} \equiv W^{(ij)} &= \frac{1}{2} \left(W^{ij} + W^{ji} \right) - \frac{1}{3} \left(\mathrm{Tr}\, \mathbf{W} \right) g^{ij} \\
R^{ij} \equiv W^{[ij]} &= \frac{1}{2} \left(W^{ij} - W^{ji} \right)
\end{align}
</math>
Here and later, symmeterization and antisymmeterization are denoted by <math>(ij)</math> and <math>[ij]</math>, respectively. This decomposition is irreducible, in the sense of being invariant under rotations, and is an important tool in the conceptual development of continuum mechanics.{{sfn|Thorne|Blandford|2017}}

The elasticity tensor has rank 4, and its decompositions are more complex and varied than those of a rank-2 tensor.{{sfn|Moakher|Norris|2006|pp=221–222}} Two examples are described below.

=== '''M''' and '''N''' tensors ===
This decomposition is obtained by symmeterization and antisymmeterization of the middle two indices:
:<math>
C^{ijkl} = M^{ijkl} + N^{ijkl}
</math>
where
:<math>
\begin{align}
M^{ijkl} \equiv C^{i(jk)l} = \frac{1}{2}\left(C^{ijkl} + C^{ikjl} \right) \\
N^{ijkl} \equiv C^{i[jk]l} = \frac{1}{2}\left(C^{ijkl} - C^{ikjl} \right)
\end{align}
</math>
A disadvantage of this decomposition is that <math>M^{ijkl}</math> and <math>N^{ijkl}</math> do not
obey all original symmetries of <math>C^{ijkl}</math>, as they are not symmetric under interchange of the first two indices. In addition, it is not irreducible, so it is not invariant under linear transformations such as rotations.{{sfn|Itin|Hehl|2013|p=}}

=== Irreducible representations ===
{{See also|Young tableau|Irreducible representation}}
An irreducible representation can be built by considering the notion of a totally symmetric tensor, which is invariant under the interchange of any two indices. A totally symmetric tensor <math>\mathbf{S}</math> can be constructed from
<math>\mathbf{C}</math> by summing over all <math>4! = 24</math> [[permutations]] of the indices
:<math>
\begin{align}
S^{ijkl} &= \frac{1}{4!}\sum_{(i,j,k,l)\in S_4} C^{ijkl} \\ &=\frac{1}{4!}\left(C^{ijkl} + C^{jikl}+ C^{ikjl} + \ldots \right)
\end{align}
</math>
where where <math>\mathbb{S}_4</math> is the set of all permutations of the four indices.{{sfn|Itin|Hehl|2013|p=}} Owing to the symmetries of <math>C^{ijkl}</math>, this sum reduces to
:<math>
S^{ijkl} = \frac{1}{3}\left(C^{ijkl} + C^{iklj} + C^{iljk} \right)
</math>
The difference
:<math>
A^{ijkl} \equiv C^{ijkl} - S^{ijkl} = \frac{1}{3}\left(2 C^{ijkl} - C^{ilkj} - C^{iklj} \right)
</math>
is an asymmetric tensor (''not'' antisymmetric). The decomposition <math>C^{ijkl} = S^{ijkl} + A^{ijkl}</math> can be shown to be unique and irreducible with respect to <math>\mathbb{S}_4</math>. In other words, any additional symmetrization operations on <math>\mathbf{S}</math> or <math>\mathbf{A}</math> will either leave it unchanged or evaluate to zero. It is also irreducible with respect to arbitrary linear transformations, that is, the [[general linear group]] <math>G(3,\mathbb{R})</math>.{{sfn|Itin|Hehl|2013|p=}}{{sfn|Itin|2020}}

However, this decomposition is not irreducible with respect to the group of rotations <math>SO(3)</math>. Instead, <math>\mathbf{S}</math> decomposes into three irreducible parts, and <math>\mathbf{A}</math> into two:
:<math>
\begin{align}
C^{ijkl} &= S^{ijkl} + A^{ijkl} \\
&= \left(^{(1)}\!S^{ijkl} + \, ^{(2)}\!S^{ijkl} + \, ^{(3)}\!S^{ijkl} \right) + \, \left(^{(1)}\!A^{ijkl} + ^{(2)}\!A^{ijkl} \right)
\end{align}
</math>
See Itin (2020){{sfn|Itin|2020}} for explicit expressions in terms of the components of <math>\mathbf{C}</math>.

This representation decomposes the space of elasticity tensors into a direct sum of subspaces:
:<math>
\mathcal{C} = \left(^{(1)}\!\mathcal{C} \oplus \, ^{(2)}\!\mathcal{C} \oplus \, ^{(3)}\!\mathcal{C} \right) \oplus \, \left(^{(4)}\!\mathcal{C} \oplus \, ^{(5)}\!\mathcal{C} \right)
</math>
with dimensions
:<math>
21 = (1 \oplus 5 \oplus 9) \oplus (1 \oplus 5)
</math>
These subspaces are each isomorphic to a [[harmonic tensor]] space <math>\mathbb{H}_n(\mathbb{R}^3)</math>.{{sfn|Itin|2020}}{{sfn | Olive | Kolev | Auffray | 2017}} Here, <math>\mathbb{H}_n(\mathbb{R}^3)</math> is the space of 3D, totally symmetric, traceless tensors of rank <math>n</math>. In particular, <math>^{(1)}\!\mathcal{C}</math> and <math>^{(4)}\!\mathcal{C}</math> correspond to <math>\mathbb{H}_1</math>, <math>^{(2)}\!\mathcal{C}</math> and <math>^{(5)}\!\mathcal{C}</math> correspond to <math>\mathbb{H}_2</math>, and <math>^{(3)}\!\mathcal{C}</math> corresponds to <math>\mathbb{H}_4</math>.

== See also ==
* [[Continuum mechanics]]
* [[Solid mechanics]]
* [[Constitutive equation]]
* [[Strength of materials]]
*{{annotated link|List of materials properties#Mechanical properties|List of materials properties § Mechanical properties}}
* [[Thermoelasticity]]
* [[Representation theory of finite groups]]

==Footnotes==
{{reflist|group=note}}

==References==
{{reflist}}

==Bibliography==
{{refbegin|2|indent=yes}}
*{{cite journal | last1 = Cowin | first1 = Stephen C. | title = Properties of the Anisotropic Elasticity Tensor | journal = The Quarterly Journal of Mechanics and Applied Mathematics | date = 1993 | volume = 46 | issue = 3 | pages = 539–539 | issn = 0033-5614 | eissn = 1464-3855 | doi = 10.1093/qjmam/46.3.539-a | pmid = | url = }}
*{{cite journal | last1 = Hehl | first1 = Friedrich W. | last2 = Itin | first2 = Yakov | title = The Cauchy Relations in Linear Elasticity Theory |journal = Journal of elasticity and the physical science of solids | date = 2002 | volume = 66 | issue = 2 | pages = 185–192 | issn = 0374-3535 | doi = 10.1023/A:1021225230036 | pmid = | url = https://arxiv.org/abs/cond-mat/0206175v1| arxiv = cond-mat/0206175}}
*{{cite journal | last1 = Hill | first1 = R. | title = Continuum micro-mechanics of elastoplastic polycrystals | journal = Journal of the Mechanics and Physics of Solids | date = April 1965 | volume = 13 | issue = 2 | pages = 89–101 | issn = 0022-5096 | doi = 10.1016/0022-5096(65)90023-2 | pmid = | url = https://hal.science/hal-03620021/document}}
*{{cite journal | last1 = Itin | first1 = Yakov | last2 = Hehl | first2 = Friedrich W. | title = The constitutive tensor of linear elasticity: Its decompositions, Cauchy relations, null Lagrangians, and wave propagation | journal = Journal of Mathematical Physics | date = April 2013 | volume = 54 | issue = 4 | page = 042903 | issn = 0022-2488 | eissn = 1089-7658 | doi = 10.1063/1.4801859 | pmid = | url = https://arxiv.org/abs/1208.1041v1 | arxiv = 1208.1041}}
*{{cite journal | last1 = Itin | first1 = Yakov | title = Irreducible matrix resolution for symmetry classes of elasticity tensors | journal = Mathematics and Mechanics of Solids | date = 20 April 2020 | volume = 25 | issue = 10 | pages = 1873–1895 | issn = 1081-2865 | eissn = 1741-3028 | doi = 10.1177/1081286520913596 | pmid = | url = | arxiv = 1812.03367}}
* {{cite book
|author-first1=Lev D. |author-last1=Landau |author-link1=Lev Landau
|author-first2=Evgeny M. |author-last2=Lifshitz |author-link2=Evgeny Lifshitz
|date=1970
|title=Theory of Elasticity
|edition=2nd |volume=7
|publisher=[[Pergamon Press]]
|isbn=978-0-08-006465-9
}}
* {{cite book | author1 = Jerrold E. Marsden | author2 = Thomas J. R. Hughes | year = 1994 | title = Mathematical Foundations of Elasticity | publisher = Dover Publications | pages = | isbn = 978-0-486-67865-8 | oclc = 1117171567 | url = https://books.google.com/books?id=STyKAAAAQBAJ&pg=PP1}}
*{{cite journal | last1 = Moakher | first1 = Maher | last2 = Norris | first2 = Andrew N. | title = The Closest Elastic Tensor of Arbitrary Symmetry to an Elasticity Tensor of Lower Symmetry | journal = Journal of Elasticity | date = 5 October 2006 | volume = 85 | issue = 3 | pages = 215–263 | issn = 0374-3535 | eissn = 1573-2681 | doi = 10.1007/s10659-006-9082-0 | pmid = | url = https://link.springer.com/content/pdf/10.1007/s10659-006-9082-0.pdf}}
*{{cite journal | last1 = Norris | first1 = A. N. | title = Quadratic invariants of elastic moduli | journal = The Quarterly Journal of Mechanics and Applied Mathematics | date = 22 May 2007 | volume = 60 | issue = 3 | pages = 367–389 | issn = 0033-5614 | eissn = 1464-3855 | doi = 10.1093/qjmam/hbm007 | pmid = | url = | arxiv = cond-mat/0612506 }}
* {{cite journal | last=Olive | first=M. | last2=Kolev | first2=B. | last3=Auffray | first3=N. | title=A Minimal Integrity Basis for the Elasticity Tensor | journal=Archive for Rational Mechanics and Analysis | publisher=Springer Science and Business Media LLC | volume=226 | issue=1 | date=2017-05-24 | issn=0003-9527 | doi=10.1007/s00205-017-1127-y | pages=1–31 | arxiv = 1605.09561}}
*{{cite journal | last1 = Srinivasan | first1 = T.P. | last2 = Nigam | first2 = S.D. | title = Invariant Elastic Constants for Crystals | journal = Journal of Mathematics and Mechanics | year = 1969 | volume = 19 | pages = 411–420 | issn = 1943-5274 | eissn = 0095-9057 | doi = 10.2307/24901866 | url = http://www.jstor.org/stable/24901866 | url-access= registration}}
*{{cite journal | last1 = Thomas | first1 = T. Y. | title = On the stress-strain relations for cubic crystals | journal = Proceedings of the National Academy of Sciences | date = February 1966 | volume = 55 | issue = 2 | pages = 235–239 | issn = 0027-8424 | eissn = 1091-6490 | doi = 10.1073/pnas.55.2.235 | pmid = 16591328 | pmc = 224128 | url = https://www.pnas.org/doi/pdf/10.1073/pnas.55.2.235}}
*{{cite book |first1=Kip S.|last1=Thorne|first2=Roger D.|last2=Blandford|author-link1=Kip Thorne|title=Modern Classical Physics: Optics, Fluids, Plasmas, Elasticity, Relativity, and Statistical Physics|publisher=Princeton University Press|year=2017|isbn=9780691159027}}



[[Category:Tensor physical quantities]]
[[Category:Tensor physical quantities]]

Revision as of 16:49, 25 February 2023

The elasticity tensor is a fourth-rank tensor describing the stress-strain relation in a linear elastic material.[1][2] Other names are elastic modulus tensor and stiffness tensor. Common symbols include and .

The defining equation can be written as

where and are the compoments of the Cauchy stress tensor and infinitesimal strain tensor, and are the components of the elasticity tensor. Summation over repeated indices is implied.[note 1] This relationship can be interpreted as a generalization of Hooke's law to a 3D continuum.

A general fourth-rank tensor in 3D has 34 = 81 independent components , but the elasticity tensor has at most 21 independent components.[3] This fact follows from the symmetry of the stress and strain tensors, together with the requirement that the stress derives from an elastic energy potential. For isotropic materials, the elasticity tensor has just two independent components, which can be chosen to be the bulk modulus and shear modulus.[3]

Definition

The most general linear relation between two second-rank tensors is

where are the components of a fourth-rank tensor .[1][note 1] The elasticity tensor is defined as for the case where and are the stress and strain tensors, respectively.

The compliance tensor is defined from the inverse stress-strain relation:

The two are related by

where is the Kronecker delta.[4][5][note 2]

Unless otherwise noted, this article assumes is defined from the stress-strain relation of a linear elastic material, in the limit of small strain.

Special cases

Isotropic

For an isotropic material, simplifies to

where and are scalar functions of the material coordinates , and is the metric tensor in the reference frame of the material.[6][7] In an orthonormal Cartesian coordinate basis, there is no distinction between upper and lower indices, and the metric tensor can be replaced with the Kronecker delta:

Substituting the first equation into the stress-strain relation and summing over repeated indices gives

where is the trace of . In this form, and can be identified with the first and second Lame coefficients. An equivalent expression is

where is the bulk modulus, and

are the components of the shear tensor .

Cubic crystals

The elasticity tensor of a cubic crystal has components

where , , and are unit vectors corresponding to the three mutually perpendicular axes of the crystal unit cell.[8] The coefficients , , and are scalars; because they are coordinate-independent, they are intrinsic material constants. Thus, a crystal with cubic symmetry is described by three independent elastic constants.[9]

In an orthonormal Cartesian coordinate basis, there is no distinction between upper and lower indices, and is the Kronecker delta, so the expression simplifies to

Other crystal classes

There are similar expressions for the components of in other crystal symmetry classes.[10] The number of independent elastic constants for several of these is given in table 1.[9]

Table 1: Number of independent elastic constants for various crystal symmetry classes.[9]
Crystal family Point group Independent components
Triclinic 21
Monoclinic 13
Orthorhombic 9
Tetragonal C4, S4, C4h 7
Tetragonal C4v, D2d, D4, D4h 6
Rhombohedral C3, S6 7
Rhombohedral C3v, D6, D3d 6
Hexagonal 5
Cubic 3

Properties

Symmetries

The elasticity tensor has several symmetries that follow directly from its defining equation .[11][2] The symmetry of the stress and strain tensors implies that

Usually, one also assumes that the stress derives from an elastic energy potential :

which implies

Hence, must be symmetric under interchange of the first and second pairs of indices:

The symmetries listed above reduce the number of independent components from 81 to 21. If a material has additional symmetries, then this number is further reduced.[9]

Transformations

Under rotation, the components transform as

where are the covariant components in the rotated basis, and are the elements of the corresponding rotation matrix. A similar transformation rule holds for other linear transformations.

Invariants

The components of generally acquire different values under a change of basis. Nevertheless, for certain types of transformations, there are specific combinations of components, called invariants, that remain unchanged. Invariants are defined with respect to a given set of transformations, formally known as a group operation. For example, an invariant with respect to the group of proper orthogonal transformations, called SO(3), is a quantity that remains constant under arbitrary 3D rotations.

possesses two linear invariants and seven quadratic invariants with respect to SO(3).[12] The linear invariants are

and the quadratic invariants are

These quantities are linearly independent, that is, none can be expressed as a linear combination of the others. They are also complete, in the sense that there are no additional independent linear or quadratic invariants.[12]

Decompositions

A common strategy in tensor analysis is to decompose a tensor into simpler components that can be analyzed separately. For example, the displacement gradient tensor can be decomposed as

where is a rank-0 tensor (a scalar), equal to the trace of ; is symmetric and trace-free; and is antisymmetric.[13] Component-wise,

Here and later, symmeterization and antisymmeterization are denoted by and , respectively. This decomposition is irreducible, in the sense of being invariant under rotations, and is an important tool in the conceptual development of continuum mechanics.[11]

The elasticity tensor has rank 4, and its decompositions are more complex and varied than those of a rank-2 tensor.[14] Two examples are described below.

M and N tensors

This decomposition is obtained by symmeterization and antisymmeterization of the middle two indices:

where

A disadvantage of this decomposition is that and do not obey all original symmetries of , as they are not symmetric under interchange of the first two indices. In addition, it is not irreducible, so it is not invariant under linear transformations such as rotations.[2]

Irreducible representations

An irreducible representation can be built by considering the notion of a totally symmetric tensor, which is invariant under the interchange of any two indices. A totally symmetric tensor can be constructed from by summing over all permutations of the indices

where where is the set of all permutations of the four indices.[2] Owing to the symmetries of , this sum reduces to

The difference

is an asymmetric tensor (not antisymmetric). The decomposition can be shown to be unique and irreducible with respect to . In other words, any additional symmetrization operations on or will either leave it unchanged or evaluate to zero. It is also irreducible with respect to arbitrary linear transformations, that is, the general linear group .[2][15]

However, this decomposition is not irreducible with respect to the group of rotations . Instead, decomposes into three irreducible parts, and into two:

See Itin (2020)[15] for explicit expressions in terms of the components of .

This representation decomposes the space of elasticity tensors into a direct sum of subspaces:

with dimensions

These subspaces are each isomorphic to a harmonic tensor space .[15][16] Here, is the space of 3D, totally symmetric, traceless tensors of rank . In particular, and correspond to , and correspond to , and corresponds to .

See also

Footnotes

  1. ^ a b Here, upper and lower indices denote contravariant and covariant components, respectively, though the distinction can be ignored for Cartesian coordinates. As a result, some references represent components using only lower indices.
  2. ^ Combining the forward and inverse stress-strain relations gives Eij = Kijpq CpqklEkl. Due to the minor symmetries Cpqkl = Cqpkl and Cpqkl = Cpqlk, this equation does not uniquely determine Kijpq Cpqkl. In fact, Kijpq Cpqkl = a δkiδlj + (1 − a) δliδkj is a solution for any 0 ≤ a ≤ 1. However, only a = 1/2 preserves the minor symmetries of K, so this is the correct solution from a physical standpoint.

References

  1. ^ a b Thorne & Blandford 2017, p. 580.
  2. ^ a b c d e Itin & Hehl 2013.
  3. ^ a b Thorne & Blandford 2017, p. 581.
  4. ^ Hill 1965.
  5. ^ Cowin 1989.
  6. ^ Marsden & Hughes 1994, p. 223.
  7. ^ Hehl & Itin 2002.
  8. ^ Thomas 1966.
  9. ^ a b c d Landau & Lifshitz 1970.
  10. ^ Srinivasan & Nigam 1969.
  11. ^ a b Thorne & Blandford 2017.
  12. ^ a b Norris 2007.
  13. ^ Thorne & Blandford 2017, p. 571.
  14. ^ Moakher & Norris 2006, pp. 221–222.
  15. ^ a b c Itin 2020.
  16. ^ Olive, Kolev & Auffray 2017.

Bibliography