In classical elasticity, it is assumed that the stress at each point depends only on the elastic strain components at that same point.1 In other words, the theory is local — the mechanical response at a point is determined solely by the state of deformation there: \[ \sigma_{ij} = F_{ij}(\epsilon_{11}, \epsilon_{22}, \dots, \epsilon_{12}) \quad (i,j=1,2,3) \] Additionally, we assume that there is no stress if all components of strain are zero (unless there are residual stresses).

Stiffness and Compliance Tensors

Expanding \(F_{ij}\) in power series: \[ \begin{aligned} \sigma_{11}&=C_{1111}\epsilon_{11}+C_{1112}\epsilon_{12}+C_{1113}\epsilon_{13}+\cdots+C_{1133}\epsilon_{33}\\ \sigma_{12}&=C_{1211}\epsilon_{12}+C_{1112}\epsilon_{12}+C_{1213}\epsilon_{13}+\cdots+C_{1233}\epsilon_{33}\\ \vdots\\ \sigma_{33}&=C_{3311}\epsilon_{11}+C_{3312}\epsilon_{12}+C_{3313}\epsilon_{13}+\cdots+C_{3333}\epsilon_{33} \end{aligned}\tag{*} \] or more concisely: \[ \sigma_{ij}=\sum_{k=1}^3\sum_{l=1}^3 C_{ijkl} \epsilon_{kl}. \] where \(C_{ijkl}\) is called the stiffness tensor or the tensor of elastic constants or elastic moduli of the material.2

In general, \(C_{ijkl}=C_{ijkl}(\mathbf{x})\); that is, the elastic coefficients vary from point to point. If \(C_{ijkl}\) is independent of the position, we say that the material is homogeneous.

The system of linear equations in (*) can be solved for \(\epsilon_{11}\), \(\epsilon_{12}\), …, \(\epsilon_{33}\), and we can express them in terms of \(\sigma_{11}\), \(\sigma_{12}\), …, \(\sigma_{33}\). The result will be another system of linear equations: \[ \begin{aligned} \epsilon_{11}&=S_{1111}\sigma_{11}+S_{1112}\sigma_{12}+\cdots+S_{1133}\sigma_{33}\\ \vdots\\ \epsilon_{33}&=S_{3311}\sigma_{11}+S_{3312}\sigma_{12}+\cdots+S_{3333}\sigma_{33} \end{aligned} \] or more concisely: \[ \epsilon_{ij}=\sum_{k=1}^3\sum_{l=1}^3S_{ijkl}\sigma_{kl}. \] Here we are dealing with another fourth rank tensor Sijkl, which is called compliance tensor.

Since \(i, j, k\), and \(l\), each vary between 1, 2, and 3, \(C_{ijkl}\) (or \(S_{ijkl}\)) has 81 components. However, not all of them are independent. Using some symmetry, we can reduce the number of independent components. In this section, we will see that, if the material is isotropic, meaning that its properties are the same in all directions, only two elastic moduli are enough to specify all stiffness components.

The first thing that we can use to reduce the number of independent components is the symmetry of stress and strain. Symmetry of stress \(\sigma_{ij}=\sigma_{ji}\) implies that is symmetric with respect to the first two indices: \[C_{{\color{red}{ij}}kl}=C_{{\color{red}{ji}}kl}.\]Similarly, symmetry of stain \(\epsilon_{kl}=\epsilon_{lk}\) implies that \(C_{ijkl}\) is symmetric with respect to the last two indices \[ C_{ij{\color{red}{kl}}}=C_{ij{\color{red}{lk}}}. \] Since each pair of \((ij)\) and \((kl)\) can take on six different value (1,1), (2,2), (3,3), (1, 2) = (2,1), (1,3) = (3,1), (2, 3)=(3,2), there are only 36 different elastic constants, at most.

Voigt Notation

Using double notation (e.g., σᵢⱼ) and dealing with a fourth-rank tensor Cijkl is cumbersome. Therefore, we introduce Voigt notation to simplify the representation of stress and strain.

The stress tensor components are arranged into a 6x1 column vector: \[ \begin{bmatrix} \sigma_{11} \\ \sigma_{22} \\ \sigma_{33} \\ \sigma_{23} \\ \sigma_{13} \\ \sigma_{12} \end{bmatrix} = \begin{bmatrix} \sigma_1 \\ \sigma_2 \\ \sigma_3 \\ \sigma_4 \\ \sigma_5 \\ \sigma_6 \end{bmatrix} \] That is, we number them according to the following: \[ \begin{bmatrix} ① & {\color{red}{⑥}} & {\color{red}{⑤}}\\ & ② & {\color{red}{④}}\\ & & ③ \end{bmatrix} \] Similarly, the corresponding strain components are arranged into a 6x1 vector. Note the factors of 2 for the shear strains, which are introduced to ensure work conjugacy between the stress and strain vectors.

\[ \begin{bmatrix} \epsilon_{11} \\ \epsilon_{22} \\ \epsilon_{33} \\ 2\epsilon_{23} \\ 2\epsilon_{13} \\ 2\epsilon_{12} \end{bmatrix} = \begin{bmatrix} \epsilon_1 \\ \epsilon_2 \\ \epsilon_3 \\ \epsilon_4 \\ \epsilon_5 \\ \epsilon_6 \end{bmatrix} = \begin{bmatrix} \epsilon_{11} \\ \epsilon_{22} \\ \epsilon_{33} \\ \gamma_{23} \\ \gamma_{13} \\ \gamma_{12} \end{bmatrix} \]

Note: Do not confuse σ₁, σ₂, σ₃ and ε₁, ε₂, ε₃ with the principal values of stresses and strains. Here, we are using them with a totally different meaning.

Elastic Coefficient Matrix

Using Voigt notation, the linear relationship between stress and strain can be written using a matrix [cmn]:

\[ \begin{bmatrix} \sigma_{11} \\ \sigma_{22} \\ \sigma_{33} \\ \sigma_{23} \\ \sigma_{13} \\ \sigma_{12} \end{bmatrix} = \begin{bmatrix} c_{11} & c_{12} & c_{13} & c_{14} & c_{15} & c_{16} \\ c_{21} & c_{22} & c_{23} & c_{24}& c_{25}& c_{26}\\ \vdots & & \ddots & & & \vdots \\ & & & & & \\ & & & & & \\ c_{61}& & \dots & & & c_{66} \end{bmatrix} \begin{bmatrix} \epsilon_{11} \\ \epsilon_{22} \\ \epsilon_{33} \\ \gamma_{23} \\ \gamma_{13} \\ \gamma_{12} \end{bmatrix} \]

From the above, it is clear that to express stress in terms of strain, we need 36 independent constants in the stiffness matrix [c].

Note: While Cᵢⱼₖₗ is a fourth-order tensor, the 6x6 matrix cₘₙ in Voigt notation is not a tensor. Therefore, we cannot use tensor transformation rules to find its components in a new coordinate system. We must transform the stress and strain components first and then deduce the new stiffness matrix.

Strain Energy

Previously we showed that the strain energy density, U₀,is given by \[ \begin{aligned} dU_0&=\sigma_{11} d\epsilon_{11}+\sigma_{12} d\epsilon_{12}+\cdots+\sigma_{33} d\epsilon_{33}\\ &=\sum_{i=1}^3\sum_{j=1}^3 \sigma_{ij}\ d\epsilon_{ij} \end{aligned} \] If there is a strain energy density function we can express stress as: \[ \sigma_{ij} = \frac{\partial U_0}{\partial \epsilon_{ij}} \] Using Voigt notation, we can write the stress energy density as \[ dU_0 = \sigma_1 d\epsilon_1 + \sigma_2 d\epsilon_2 + \sigma_3 d\epsilon_3 + \sigma_4 d\epsilon_4 + \dots + \sigma_6 d\epsilon_6 \] and \[ \sigma_{i} = \frac{\partial U_0}{\partial \epsilon_{i}} \] If the material is linear elastic, we have \(\sigma_i = \sum_{j=1}^{6} c_{ij} \epsilon_j\). Substituting this into the expression for dU₀: \[ dU_0 = \sum_{i=1}^{6} \sigma_i d\epsilon_i = \sum_{i=1}^{6} \sum_{k=1}^{6} c_{ik} \epsilon_k\, d\epsilon_i \] Integrating this expression gives the strain energy density: \[ U_0 = \sum_{i=1}^{6} \sum_{k=1}^{6} \frac{1}{2} c_{ik} \epsilon_k \epsilon_i \] By taking partial derivatives twice, we find: \[ \frac{\partial^2 U_0}{\partial \epsilon_i \partial \epsilon_j} = \frac{\partial}{\partial \epsilon_i}\left(\frac{\partial U_0}{\partial \epsilon_j}\right)=\frac{\partial \sigma_j}{\partial \epsilon_i} = c_{ji} \] \[ \frac{\partial^2 U_0}{\partial \epsilon_j \partial \epsilon_i} = \frac{\partial \sigma_i}{\partial \epsilon_j} = c_{ij} \] Since the order of differentiation does not matter for a function3: \[ \frac{\partial^2 U_0}{\partial \epsilon_i \partial \epsilon_j} = \frac{\partial^2 U_0}{\partial \epsilon_j \partial \epsilon_i} \] we must have:4 \[ c_{ji} = c_{ij} \] This means the stiffness matrix [c] must be symmetric. This symmetry reduces the number of different elastic constants to 21 (6 diagonal elements + 15 above diagonal elements).

The Effect of Material Symmetry

The number of independent elastic constants (21 for the most general anisotropic case) can be reduced if the material’s structure has some form of symmetry.

Material with One Plane of Symmetry

Let’s assume the material has a plane of material symmetry at z=0 (the xy-plane). This means the material’s response is identical if we reflect the coordinate system across this plane. Let’s define a new coordinate system (x’, y’, z’) such that: \[ x' = x, \quad y' = y, \quad z' = -z \] In the new coordinate system, due to the symmetry of the material, elastic coefficients remain invariant, but under this transformation, the stress and strain components change as follows: \[ \begin{aligned} \sigma_{x'x'} &= \sigma_{xx} & \sigma_{x'y'} &= \sigma_{xy} & \sigma_{x'z'} &= -\sigma_{xz} \\ \sigma_{y'y'} &= \sigma_{yy} & \sigma_{y'z'} &= -\sigma_{yz} & \sigma_{z'z'} &= \sigma_{zz} \end{aligned} \] \[ \begin{aligned} \epsilon_{x'x'} &= \epsilon_{xx} & \epsilon_{x'y'} &= \epsilon_{xy} & \epsilon_{x'z'} &= -\epsilon_{xz} \\ \epsilon_{y'y'} &= \epsilon_{yy} & \epsilon_{y'z'} &= -\epsilon_{yz} & \epsilon_{z'z'} &= \epsilon_{zz} \end{aligned} \] The constitutive relation for σₓₓ must be the same in both coordinate systems. In the original system: \[ \sigma_{xx} = c_{11}\epsilon_{xx} + c_{12}\epsilon_{yy} + c_{13}\epsilon_{zz} + c_{14}\gamma_{yz} + c_{15}\gamma_{xz} + c_{16}\gamma_{xy} \] In the new system: \[ \sigma_{x'x'} = c_{11}\epsilon_{x'x'} + c_{12}\epsilon_{y'y'} + c_{13}\epsilon_{z'z'} + c_{14}\gamma_{y'z'} + c_{15}\gamma_{x'z'} + c_{16}\gamma_{x'y'} \] Substituting the transformed components: \[ \sigma_{xx} = c_{11}\epsilon_{xx} + c_{12}\epsilon_{yy} + c_{13}\epsilon_{zz} - c_{14}\gamma_{yz} - c_{15}\gamma_{xz} + c_{16}\gamma_{xy} \] For the two equations to be identical for all possible strains, we must have: \[ c_{14} = -c_{14} \implies c_{14} = 0 \] \[ c_{15} = -c_{15} \implies c_{15} = 0 \] By applying the same logic to other stress components, we find that c₂₄ = c₂₅ = c₃₄ = c₃₅ = c₄₆ = c₅₆ = 0. The elastic matrix simplifies to: \[ [c_{mn}] = \begin{bmatrix} c_{11} & c_{12} & c_{13} & 0 & 0 & c_{16} \\ c_{12} & c_{22} & c_{23} & 0 & 0 & c_{26} \\ c_{13} & c_{23} & c_{33} & 0 & 0 & c_{36} \\ 0 & 0 & 0 & c_{44} & c_{45} & 0 \\ 0 & 0 & 0 & c_{45} & c_{55} & 0 \\ c_{16} & c_{26} & c_{36} & 0 & 0 & c_{66} \end{bmatrix} \] This reduces the number of elastic constants required to specify the stiffness tensor to 13.

Orthotropic Material

An orthotropic material has three mutually perpendicular planes of symmetry and three corresponding orthogonal axes. Many materials, such as wood, rolled metal sheet, and fiber-reinforced laminates, reinforced concrete can be treated as orthotropic.

If we align our coordinate axes with the normals of these planes, the matrix simplifies further. Applying a second symmetry plane (e.g., the xz-plane) will zero out additional coefficients. The resulting matrix for an orthotropic material is: \[ [c_{mn}] = \begin{bmatrix} c_{11} & c_{12} & c_{13} & 0 & 0 & 0 \\ c_{12} & c_{22} & c_{23} & 0 & 0 & 0 \\ c_{13} & c_{23} & c_{33} & 0 & 0 & 0 \\ 0 & 0 & 0 & c_{44} & 0 & 0 \\ 0 & 0 & 0 & 0 & c_{55} & 0 \\ 0 & 0 & 0 & 0 & 0 & c_{66} \end{bmatrix} \] The number of independent constants is now 9.

Cubic Material

A cubic material has the symmetry of a cube. In addition to being orthotropic, its properties are identical if we swap the coordinate axes (e.g., xy, yz, zx). This imposes further constraints: \[ \begin{aligned} c_{11} = c_{22} = c_{33} \\ c_{12} = c_{13} = c_{23} \\ c_{44} = c_{55} = c_{66} \end{aligned} \] The matrix of elastic constants for a cubic material has only 3 independent constants: \(c_{11}, c_{12}\) ,and \(c_{44}\): \[ [c_{mn}] = \begin{bmatrix} c_{11} & c_{12} & c_{12} & 0 & 0 & 0 \\ c_{12} & c_{11} & c_{12} & 0 & 0 & 0 \\ c_{12} & c_{12} & c_{11} & 0 & 0 & 0 \\ 0 & 0 & 0 & c_{44} & 0 & 0 \\ 0 & 0 & 0 & 0 & c_{44} & 0 \\ 0 & 0 & 0 & 0 & 0 & c_{44} \end{bmatrix} \] Many materials including those with FCC, BCC, and diamond cubic structures have cubic symmetry. Examples of such materials include aluminum, copper, nickel, silver, gold, iron, silicon, germanium.

Example:5 Both silicon (Si) and germanium (Ge) are crystals with a cubic unit cell. The edge length of the unit cell of Si is \(a_{\text{Si}} = 5.428\) Å, and that of Ge is \(a_{\text{Ge}} = 5.658\) Å. A Ge film, 10 nm thick, is grown epitaxially (i.e., with matching atomic positions) on the surface of a 100 µm thick Si substrate. Calculate the stress and strain components in the Ge film.

The respective elastic constants are given in GPa:

  • Si: \(c_{11} = 165.8\), \(c_{12} = 63.9\), \(c_{44} = 79.6\).
  • Ge: \(c_{11} = 128.5\), \(c_{12} = 48.2\), \(c_{44} = 66.7\).
Solution

The problem describes a thin film of germanium being grown on a much thicker substrate of silicon. This setup is common in the semiconductor industry.

  • Epitaxial Growth: This means the crystal lattice of the Ge film is forced to align with the crystal lattice of the Si substrate at the interface.
  • Thick Substrate Assumption: Because the Si substrate (100 µm) is vastly thicker than the Ge film (10 nm), we can assume the substrate is a rigid template. It does not bend or strain due to the presence of the thin film. The film must do all the deforming.
  • Coordinate System: We set up a coordinate system where axes 1 (x) and 2 (y) are in the plane of the film, parallel to the crystal cube edges, and axis 3 (z) is normal to the film surface.

Calculating the In-Plane Strain (\(\epsilon_{11}\) and \(\epsilon_{22}\))

The stress in the film originates from the lattice mismatch between Ge and Si. The natural lattice constant of Ge is larger than that of Si. To grow epitaxially, the Ge atoms must compress in the plane of the film to match the smaller spacing of the Si atoms.

The strain is defined as the change in length divided by the original length. * Original length (natural spacing of Ge) = \(a_{Ge}\) * Final length (forced spacing of Si) = \(a_{Si}\)

The in-plane strains, \(\epsilon_{11}\) and \(\epsilon_{22}\), are therefore equal due to the cubic symmetry of the materials. \[ \begin{aligned} \epsilon_{11} = \epsilon_{22} &= \frac{a_{\text{Si}} - a_{\text{Ge}}}{a_{Ge}} \\ &= \frac{5.428\,{\rm Å} - 5.658\,{\rm Å}}{5.658\, {\rm Å}} \approx -0.0406 \end{aligned} \]

This gives an in-plane strain of approximately -4.1%. The negative sign correctly indicates that the Ge film is under compression in the x and y directions. Because the crystal axes are aligned with our coordinate system, all shear strains are zero (\(\epsilon_{12} = \epsilon_{13} = \epsilon_{23} = 0\)).

Calculating the Out-of-Plane Strain (\(\epsilon_{33}\))

The film is compressed in the x-y plane, but it is free to expand or contract in the z-direction (normal to the surface). Physically, since the top surface of the film is free, there can be no force acting on it. This means the stress normal to the surface must be zero. \[ \sigma_{33} = 0 \] We can now use the generalized Hooke’s Law for a cubic crystal to relate this stress to the strains. The equation for \(\sigma_{33}\) is: \[ \sigma_{33} = c_{12}\epsilon_{11} + c_{12}\epsilon_{22}+c_{11}\epsilon_{33} . \] Using the conditions \(\sigma_{33} = 0\) and \(\epsilon_{11} = \epsilon_{22}\), we can solve for the unknown out-of-plane strain, \(\epsilon_{33}\). \[ 0 = c_{11}\epsilon_{33} + c_{12}\epsilon_{11} + c_{12}\epsilon_{11} = c_{11}\epsilon_{33} + 2c_{12}\epsilon_{11} \]

Rearranging for \(\epsilon_{33}\): \[ \epsilon_{33} = - \frac{2c_{12}}{c_{11}} \epsilon_{11} \] Now, we insert the numerical values for the germanium film:

\[ \epsilon_{33} = - \frac{2 \times 48.2 \, \text{GPa}}{128.5 \, \text{GPa}} \times (-0.0406) \approx +0.0304 \] This gives an out-of-plane strain of approximately +3.0%. The positive sign indicates an elongation in the z-direction.

Calculating the In-Plane Stress (\(\sigma_{11}\) and \(\sigma_{22}\))

Finally, we can calculate the in-plane stresses using Hooke’s Law and the strain values we’ve found. Due to the symmetry (\(\epsilon_{11}=\epsilon_{22}\)), the in-plane stresses will also be equal (\(\sigma_{11}=\sigma_{22}\)). This is known as a biaxial stress state.

The equation for \(\sigma_{11}\) is: \[ \sigma_{11} = c_{11}\epsilon_{11} + c_{12}\epsilon_{22} + c_{12}\epsilon_{33} \] Substitute the expression for \(\epsilon_{33}\) into this equation: \[ \sigma_{11} = c_{11}\epsilon_{11} + c_{12}\epsilon_{11} + c_{12} \left( - \frac{2c_{12}}{c_{11}} \epsilon_{11} \right) \] Now, we can factor out \(\epsilon_{11}\) to get the final expression: \[ \sigma_{11} = \sigma_{22} = \left[ c_{11} + c_{12} - \frac{2c_{12}^2}{c_{11}} \right] \epsilon_{11} \]

Inserting the numerical values for Ge: \[ \begin{aligned} \sigma_{11} &= \left[ 128.5 + 48.2 - \frac{2 \times (48.2)^2}{128.5} \right]\times 10^9 \times (-0.0406)\\ &\approx -5.71 \, \text{GPa} \end{aligned} \] The in-plane stress is -5.7 GPa. The negative sign confirms a large compressive stress. This stress is exceptionally high and stores a significant amount of strain energy in the film. In practice, if the film is grown beyond a certain “critical thickness,” this energy is often released through the formation of defects known as misfit dislocations, which would relieve some of the strain.

Isotropic Material

An isotropic material has the same properties in every direction. This is the highest level of material symmetry. For a material to be isotropic, its constitutive relations (and thus its elastic constants) must remain unchanged after any arbitrary rotation of the coordinate system.

We begin with the stiffness matrix for a cubic material, which has three independent constants (\(c_{11}\), \(c_{12}\), and \(c_{44}\)). An isotropic material is a special case of a cubic material, so we can find the additional constraint required for isotropy by enforcing rotational invariance.

By rotating the x and y axes by 45° around the z-axis and requiring the shear stress–strain relation to retain the same form in the rotated system, we find that for an isotropic material the elastic constants must satisfy \[\boxed{c_{44} = \tfrac{1}{2}(c_{11} - c_{12})}. \]

Let’s consider a new coordinate system (x’, y’, z’) that is rotated by 45° with respect to the original system (x, y, z) around the z-axis. The transformation relations for the stress and strain components can be derived.

For our purposes, we will need the following relations for the shear components in the x’-y’ plane: \[ \sigma_{1'2'} = \frac{1}{2}(\sigma_{22} - \sigma_{11})\tag{i} \] \[ \epsilon_{1'2'} = \frac{1}{2}(\epsilon_{22} - \epsilon_{11})\tag{ii} \] or equivalently \[ \gamma_{1'2'} = \epsilon_{22} - \epsilon_{11} \]

In the new (primed) coordinate system, the relationship between shear stress and engineering shear strain must have the same form as in the original system. In Voigt notation, this is \[\sigma_6 = c_{66} \epsilon_6\] or \[\sigma_{12}=c_{44}\gamma_{12}.\] For an isotropic material, the elastic constant \(c_{44}\) must be the same in both coordinate systems. Therefore, in the primed system: \[ \sigma_{1'2'} = c_{44}\gamma_{1'2'} \] Substituting (i) and (ii) into the above equation gives: \[ \frac{1}{2}(\sigma_{22} - \sigma_{11}) = c_{44}(\epsilon_{22} - \epsilon_{11}) \tag{iii} \] Now, let’s express the stresses \(\sigma_{11}\) and \(\sigma_{22}\) using the constitutive law for a cubic material in the original coordinate system: \[ \sigma_{11} = c_{11}\epsilon_{11} + c_{12}\epsilon_{22} \\ \sigma_{22} = c_{12}\epsilon_{11} + c_{11}\epsilon_{22} \] Now, we calculate the term \((\sigma_{22} - \sigma_{11})\): \[ \sigma_{22} - \sigma_{11} = (c_{12}\epsilon_{11} + c_{11}\epsilon_{22}) - (c_{11}\epsilon_{11} + c_{12}\epsilon_{22}) \\ = (c_{12} - c_{11})\epsilon_{11} + (c_{11} - c_{12})\epsilon_{22} \\ = (c_{11} - c_{12})(\epsilon_{22} - \epsilon_{11}) \] Finally, we substitute this result back into equation (iii): \[ \frac{1}{2}(c_{11} - c_{12})(\epsilon_{22} - \epsilon_{11}) = c_{44}(\epsilon_{22} - \epsilon_{11}) \] Since this relationship must hold for any arbitrary strain state, the strain terms \((\epsilon_{22} - \epsilon_{11})\) cancel out, leaving the required constraint for an isotropic material: \[ c_{44} = \frac{c_{11} - c_{12}}{2} \]

This leaves only 2 independent constants for an isotropic material. These are typically expressed as the Lamé parameters, λ and μ (where μ is the shear modulus, G). \[ \begin{aligned} c_{44} &= \mu = G \\ c_{12} &= \lambda \\ \end{aligned} \] \[ \Rightarrow c_{11} = \lambda + 2\mu \] Substituting these into the constitutive equations, for example σ₁₁: \[ \begin{aligned} \sigma_{11} &= c_{11}\epsilon_{11} + c_{12}\epsilon_{22} + c_{12}\epsilon_{33} \\ &= (\lambda+2\mu)\epsilon_{11} + \lambda\epsilon_{22} + \lambda\epsilon_{33} \\ &= \lambda(\epsilon_{11}+\epsilon_{22}+\epsilon_{33}) + 2\mu\epsilon_{11} \end{aligned} \] Similarly, for all components: \[ \begin{aligned} \sigma_{22} &= \lambda(\epsilon_{11}+\epsilon_{22}+\epsilon_{33}) + 2\mu\epsilon_{22} \\ \sigma_{33} &= \lambda(\epsilon_{11}+\epsilon_{22}+\epsilon_{33}) + 2\mu\epsilon_{33} \\ \sigma_{23} &= 2\mu\epsilon_{23} \\ \sigma_{13} &= 2\mu\epsilon_{13} \\ \sigma_{12} &= 2\mu\epsilon_{12} \end{aligned} \] These relationships can be written compactly using tensor notation: \[ \boxed{\sigma_{ij} = \lambda \left(\sum_{k=1}^{3} \epsilon_{kk}\right) \delta_{ij} + 2\mu\epsilon_{ij}} \] where δᵢⱼ is the Kronecker delta: \[ \delta_{ij} = \begin{cases} 1 & \text{if } i=j \\ 0 & \text{if } i \neq j \end{cases} \] Using the Einstein summation convention (where summation is implied over a repeated index), the equation becomes even more compact: \[ \bbox[5px,border:1px #f2f2f2;background-color:#f2f2f2]{\sigma_{ij} = \lambda \epsilon_{kk} \delta_{ij} + 2\mu\epsilon_{ij}} \]

Anisotropy Factor

In the section on isotropic materials, we derived the specific condition that the elastic constants of a cubic material must satisfy for it to be isotropic: \[ c_{44} = \frac{c_{11} - c_{12}}{2} \] This relationship provides a useful way to quantify the degree of elastic anisotropy in a cubic crystal. We can rearrange this expression into a ratio. This gives rise to the Zener anisotropy ratio (or anisotropy factor), denoted by A, which is defined as: \[ A = \frac{2c_{44}}{c_{11} - c_{12}} \] This dimensionless factor provides a measure of how anisotropic a material is:

Elastic Constants of Various Materials

MaterialCrystalc11 (GPa)c44 (GPa)c12 (GPa)E (GPa)\(\nu\)μ (GPa)A
Agfcc124.0046.1093.4043.750.4346.103.01
Alfcc107.3028.3060.9063.200.3628.301.22
Aufcc192.9041.50163.8042.460.4641.502.85
Cufcc168.4075.40121.4066.690.4275.403.21
Irfcc580.00256.00242.00437.510.29256.001.51
Nifcc246.50127.40147.30136.310.37127.402.57
Pbfcc49.5014.9042.3010.520.4614.904.14
Pdfcc227.1071.70176.0073.410.4471.702.81
Ptfcc346.7076.50250.70136.290.4276.501.59
Crbcc339.8099.0058.60322.560.1599.000.70
Febcc231.40116.40134.70132.280.37116.402.41
Kbcc4.142.632.212.600.352.632.73
Libcc13.508.7811.443.000.468.788.52
Mobcc440.80121.70172.40343.860.28121.700.91
Nabcc6.155.924.961.720.455.929.95
Nbbcc240.2028.20125.60153.950.3428.200.49
Tabcc260.2082.60154.50145.080.3782.601.56
Vbcc228.0042.60118.70146.720.3442.600.78
Wbcc522.40160.80204.40407.430.28160.801.01
Cdc949.00521.00151.00907.540.14521.001.31
Gedc128.4066.7048.20102.090.2766.701.66
Sidc166.2079.8064.40130.230.2879.801.57
GaAs118.8059.4053.7085.370.3159.401.82
GaP141.2070.5062.50102.850.3170.501.79
InP102.2046.0057.6060.680.3646.002.06
KCl39.506.304.9038.420.116.300.36
LiF114.0063.6047.7085.860.2963.601.92
MgO287.60151.4087.40246.860.23151.401.51
NaCl49.6012.9012.4044.640.2012.900.69
TiC500.00175.00113.00458.340.18175.000.9

Reference

References

  1. Nye, J. F. (1985). Physical properties of crystals: Their representation by tensors and matrices. Oxford University Press.
  2. Shames, I. H., & Cozzarelli, F. A. (1992). Elastic and inelastic stress analysis (2nd ed.). Prentice Hall.
  3. Sokolnikoff, I. S. (1956). Mathematical theory of elasticity (2nd ed.). McGraw-Hill.

  1. This assumption, however, is not essential. In nonlocal theories of elasticity, the stress at a point can depend not only on the strain at that point but also on its spatial variation (i.e., the gradient of strain) or, more generally, on the strain field in neighboring points. In the nonlocal elasticity theory proposed by C. Eringen, the stress at a point is related to the strain throughout the entire body through an integral relation. Specifically, the stress is expressed as a weighted average of the strain over the domain, where the weighting function assigns larger influence to strains at nearby points and gradually diminishes the contribution of points farther away. Thus, in nonlocal elasticity, the material response reflects long-range interactions, capturing effects that classical (local) elasticity cannot describe accurately, particularly at small length scales or in materials with microstructure.↩︎
  2. More specifically, \(C_{ijkl}\) is a fourth-rank tensor, meaning that its components transform between coordinate systems according to \[C_{m'n'o'p'} = \sum_{i=1}^{3}\sum_{j=1}^{3}\sum_{k=1}^{3}\sum_{l=1}^{3} l_{im'} \, l_{jn'} \, l_{ko'} \, l_{lp'} \, C_{ijkl}, \]where \(l_{im'} = \hat{\mathbf{e}}_{m'} \cdot \hat{\mathbf{e}}_i\) are the direction cosines between the new and old coordinate axes. Here, \(\hat{\mathbf{e}}_{m'}\) and \(\hat{\mathbf{e}}_i\) denote unit vectors along the \(m'\)th and \(i\)th coordinate axes in the new and old coordinate systems, respectively.↩︎
  3. Here we have assumed that the second partial derivatives are continuous.↩︎
  4. In terms of stiffness tensor \[ \frac{\partial^2 U_0}{\partial \epsilon_{kl}\partial \epsilon_{ij}}=\frac{\partial^2 U_0}{\partial \epsilon_{kl}\partial \epsilon_{ij}} \Rightarrow C_{{\color{red}{ij}}{\color{blue}{kl}}}=C_{{\color{blue}{kl}}{\color{red}{ij}}} \]↩︎
  5. This example is taken from Professor Zhigang Suo’s lecture notes (2008) for Engineering Science 240: Solid Mechanics.↩︎