Английская Википедия:Infinitesimal strain theory

Материал из Онлайн справочника
Перейти к навигацииПерейти к поиску

Шаблон:Short description Шаблон:More footnotes needed Шаблон:Continuum mechanics 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, infinitesimally smaller) than any relevant dimension of the body; so that its geometry and the constitutive properties of the material (such as density and stiffness) at each point of space can be assumed to be unchanged by the deformation.

With this assumption, the equations of continuum mechanics are considerably simplified. This approach may also be called small deformation theory, small displacement theory, or small displacement-gradient theory. It is contrasted with the finite strain theory where the opposite assumption is made.

The infinitesimal strain theory is commonly adopted in civil and mechanical engineering for the stress analysis of structures built from relatively stiff elastic materials like concrete and steel, since a common goal in the design of such structures is to minimize their deformation under typical loads. However, this approximation demands caution in the case of thin flexible bodies, such as rods, plates, and shells which are susceptible to significant rotations, thus making the results unreliable.[1]

Infinitesimal strain tensor

For infinitesimal deformations of a continuum body, in which the displacement gradient tensor (2nd order tensor) is small compared to unity, i.e. <math>\|\nabla \mathbf u\| \ll 1 </math>, it is possible to perform a geometric linearization of any one of the finite strain tensors used in finite strain theory, e.g. the Lagrangian finite strain tensor <math>\mathbf E</math>, and the Eulerian finite strain tensor <math>\mathbf e</math>. In such a linearization, the non-linear or second-order terms of the finite strain tensor are neglected. Thus we have

<math display="block">\mathbf E = \frac{1}{2} \left(\nabla_{\mathbf X}\mathbf u + (\nabla_{\mathbf X}\mathbf u)^T + (\nabla_{\mathbf X}\mathbf u)^T\nabla_{\mathbf X}\mathbf u\right)\approx \frac{1}{2}\left(\nabla_{\mathbf X}\mathbf u + (\nabla_{\mathbf X}\mathbf u)^T\right)</math> or <math display="block">E_{KL}= \frac{1}{2} \left(\frac{\partial U_K}{\partial X_L} +\frac{\partial U_L}{\partial X_K}+ \frac{\partial U_M}{\partial X_K} \frac{\partial U_M}{\partial X_L}\right)\approx \frac{1}{2}\left(\frac{\partial U_K}{\partial X_L}+\frac{\partial U_L}{\partial X_K}\right)</math> and <math display="block">\mathbf e =\frac{1}{2} \left(\nabla_{\mathbf x}\mathbf u + (\nabla_{\mathbf x}\mathbf u)^T - \nabla_{\mathbf x}\mathbf u(\nabla_{\mathbf x}\mathbf u)^T\right)\approx \frac{1}{2}\left(\nabla_{\mathbf x}\mathbf u + (\nabla_{\mathbf x}\mathbf u)^T\right)</math> or <math display="block">e_{rs}=\frac{1}{2} \left(\frac{\partial u_r}{\partial x_s} +\frac{\partial u_s}{\partial x_r} -\frac{\partial u_k}{\partial x_r} \frac{\partial u_k}{\partial x_s}\right)\approx \frac{1}{2}\left(\frac{\partial u_r}{\partial x_s} +\frac{\partial u_s}{\partial x_r}\right)</math>

This linearization implies that the Lagrangian description and the Eulerian description are approximately the same as there is little difference in the material and spatial coordinates of a given material point in the continuum. Therefore, the material displacement gradient tensor components and the spatial displacement gradient tensor components are approximately equal. Thus we have <math display="block">\mathbf E \approx \mathbf e \approx \boldsymbol \varepsilon = \frac{1}{2}\left((\nabla\mathbf u)^T + \nabla\mathbf u\right) </math> or <math display="block"> E_{KL}\approx e_{rs}\approx\varepsilon_{ij} = \frac{1}{2} \left(u_{i,j}+u_{j,i}\right)</math> where <math>\varepsilon_{ij}</math> are the components of the infinitesimal strain tensor <math>\boldsymbol \varepsilon</math>, also called Cauchy's strain tensor, linear strain tensor, or small strain tensor.

<math display="block">\begin{align} \varepsilon_{ij} &= \frac{1}{2}\left(u_{i,j}+u_{j,i}\right) \\ &= \begin{bmatrix} \varepsilon_{11} & \varepsilon_{12} & \varepsilon_{13} \\ \varepsilon_{21} & \varepsilon_{22} & \varepsilon_{23} \\ \varepsilon_{31} & \varepsilon_{32} & \varepsilon_{33} \\ \end{bmatrix} \\ &= \begin{bmatrix}

 \frac{\partial u_1}{\partial x_1} & \frac{1}{2} \left(\frac{\partial u_1}{\partial x_2}+\frac{\partial u_2}{\partial x_1}\right) & \frac{1}{2} \left(\frac{\partial u_1}{\partial x_3}+\frac{\partial u_3}{\partial x_1}\right) \\
  \frac{1}{2} \left(\frac{\partial u_2}{\partial x_1}+\frac{\partial u_1}{\partial x_2}\right) & \frac{\partial u_2}{\partial x_2} & \frac{1}{2} \left(\frac{\partial u_2}{\partial x_3}+\frac{\partial u_3}{\partial x_2}\right) \\
  \frac{1}{2} \left(\frac{\partial u_3}{\partial x_1}+\frac{\partial u_1}{\partial x_3}\right) & \frac{1}{2} \left(\frac{\partial u_3}{\partial x_2}+\frac{\partial u_2}{\partial x_3}\right) & \frac{\partial u_3}{\partial x_3} \\
 \end{bmatrix}

\end{align} </math> or using different notation: <math display="block">\begin{bmatrix} \varepsilon_{xx} & \varepsilon_{xy} & \varepsilon_{xz} \\

  \varepsilon_{yx} & \varepsilon_{yy} & \varepsilon_{yz} \\
  \varepsilon_{zx} & \varepsilon_{zy} & \varepsilon_{zz} \\
 \end{bmatrix}

= \begin{bmatrix}

 \frac{\partial u_x}{\partial x} & \frac{1}{2} \left(\frac{\partial u_x}{\partial y}+\frac{\partial u_y}{\partial x}\right) & \frac{1}{2} \left(\frac{\partial u_x}{\partial z}+\frac{\partial u_z}{\partial x}\right) \\
  \frac{1}{2} \left(\frac{\partial u_y}{\partial x}+\frac{\partial u_x}{\partial y}\right) & \frac{\partial u_y}{\partial y} & \frac{1}{2} \left(\frac{\partial u_y}{\partial z}+\frac{\partial u_z}{\partial y}\right) \\
  \frac{1}{2} \left(\frac{\partial u_z}{\partial x}+\frac{\partial u_x}{\partial z}\right) & \frac{1}{2} \left(\frac{\partial u_z}{\partial y}+\frac{\partial u_y}{\partial z}\right) & \frac{\partial u_z}{\partial z} \\
 \end{bmatrix} </math>

Furthermore, since the deformation gradient can be expressed as <math>\boldsymbol{F} = \boldsymbol{\nabla}\mathbf{u} + \boldsymbol{I}</math> where <math>\boldsymbol{I}</math> is the second-order identity tensor, we have <math display="block">\boldsymbol\varepsilon = \frac{1}{2} \left(\boldsymbol{F}^T+\boldsymbol{F}\right)-\boldsymbol{I}</math>

Also, from the general expression for the Lagrangian and Eulerian finite strain tensors we have <math display="block"> \begin{align} \mathbf E_{(m)}& =\frac{1}{2m} (\mathbf U^{2m}-\boldsymbol{I}) = \frac{1}{2m} [(\boldsymbol{F}^T\boldsymbol{F})^m - \boldsymbol{I}] \approx \frac{1}{2m} [\{\boldsymbol{\nabla}\mathbf{u}+(\boldsymbol{\nabla}\mathbf{u})^T + \boldsymbol{I}\}^m - \boldsymbol{I}]\approx \boldsymbol{\varepsilon}\\ \mathbf e_{(m)}& = \frac{1}{2m} (\mathbf V^{2m}-\boldsymbol{I})= \frac{1}{2m} [(\boldsymbol{F}\boldsymbol{F}^T)^m - \boldsymbol{I}]\approx \boldsymbol{\varepsilon} \end{align} </math>

Geometric derivation

Файл:2D geometric strain.svg
Figure 1. Two-dimensional geometric deformation of an infinitesimal material element.

Consider a two-dimensional deformation of an infinitesimal rectangular material element with dimensions <math>dx</math> by <math>dy</math> (Figure 1), which after deformation, takes the form of a rhombus. From the geometry of Figure 1 we have

<math display="block">\begin{align} \overline {ab} &= \sqrt{\left(dx+\frac{\partial u_x}{\partial x}dx \right)^2 + \left( \frac{\partial u_y}{\partial x}dx \right)^2} \\ &= dx\sqrt{1+2\frac{\partial u_x}{\partial x}+\left(\frac{\partial u_x}{\partial x}\right)^2 + \left(\frac{\partial u_y}{\partial x}\right)^2} \\ \end{align}</math>

For very small displacement gradients, i.e., <math>\|\nabla \mathbf u\| \ll 1 </math>, we have <math display="block">\overline {ab} \approx dx + \frac{\partial u_x}{\partial x} dx</math>

The normal strain in the <math>x</math>-direction of the rectangular element is defined by <math display="block">\varepsilon_x = \frac{\overline {ab}-\overline {AB}}{\overline {AB}}</math> and knowing that <math>\overline {AB}= dx</math>, we have <math display="block">\varepsilon_x = \frac{\partial u_x}{\partial x}</math>

Similarly, the normal strain in the Шаблон:Nowrap and Шаблон:Nowrap becomes <math display="block">\varepsilon_y = \frac{\partial u_y}{\partial y} \quad , \qquad \varepsilon_z = \frac{\partial u_z}{\partial z}</math>

The engineering shear strain, or the change in angle between two originally orthogonal material lines, in this case line <math>\overline {AC}</math> and <math>\overline {AB}</math>, is defined as <math display="block">\gamma_{xy}= \alpha + \beta</math>

From the geometry of Figure 1 we have <math display="block">\tan \alpha = \frac{\dfrac{\partial u_y}{\partial x}dx}{dx + \dfrac{\partial u_x}{\partial x} dx} = \frac{\dfrac{\partial u_y}{\partial x}}{1+\dfrac{\partial u_x}{\partial x}} \quad , \qquad \tan \beta=\frac{\dfrac{\partial u_x}{\partial y} dy}{dy+\dfrac{\partial u_y}{\partial y} dy}=\frac{\dfrac{\partial u_x}{\partial y}}{1+\dfrac{\partial u_y}{\partial y}}</math>

For small rotations, i.e., <math>\alpha</math> and <math>\beta</math> are <math>\ll 1</math> we have <math display="block">\tan \alpha \approx \alpha \quad , \qquad \tan \beta \approx \beta</math> and, again, for small displacement gradients, we have <math display="block">\alpha=\frac{\partial u_y}{\partial x} \quad , \qquad \beta=\frac{\partial u_x}{\partial y}</math> thus <math display="block">\gamma_{xy}= \alpha + \beta = \frac{\partial u_y}{\partial x} + \frac{\partial u_x}{\partial y}</math> By interchanging <math>x</math> and <math>y</math> and <math>u_x</math> and <math>u_y</math>, it can be shown that <math>\gamma_{xy} = \gamma_{yx}</math>.

Similarly, for the <math>y</math>-<math>z</math> and <math>x</math>-<math>z</math> planes, we have <math display="block">\gamma_{yz} = \gamma_{zy} = \frac{\partial u_y}{\partial z} + \frac{\partial u_z}{\partial y} \quad , \qquad \gamma_{zx} = \gamma_{xz} = \frac{\partial u_z}{\partial x} + \frac{\partial u_x}{\partial z}</math>

It can be seen that the tensorial shear strain components of the infinitesimal strain tensor can then be expressed using the engineering strain definition, Шаблон:Nowrap as <math display="block"> \begin{bmatrix} \varepsilon_{xx} & \varepsilon_{xy} & \varepsilon_{xz} \\ \varepsilon_{yx} & \varepsilon_{yy} & \varepsilon_{yz} \\ \varepsilon_{zx} & \varepsilon_{zy} & \varepsilon_{zz} \\ \end{bmatrix} = \begin{bmatrix} \varepsilon_{xx} & \gamma_{xy}/2 & \gamma_{xz}/2 \\ \gamma_{yx}/2 & \varepsilon_{yy} & \gamma_{yz}/2 \\ \gamma_{zx}/2 & \gamma_{zy}/2 & \varepsilon_{zz} \\ \end{bmatrix}</math>

Physical interpretation

From finite strain theory we have <math display="block">d\mathbf{x}^2 - d\mathbf{X}^2 = d\mathbf X \cdot 2\mathbf E \cdot d\mathbf X \quad\text{or}\quad (dx)^2 - (dX)^2 = 2E_{KL}\,dX_K\,dX_L</math>

For infinitesimal strains then we have <math display="block">d\mathbf{x}^2 - d\mathbf{X}^2 = d\mathbf X \cdot 2\mathbf{\boldsymbol \varepsilon} \cdot d\mathbf X \quad\text{or}\quad (dx)^2 - (dX)^2 = 2\varepsilon_{KL}\,dX_K\,dX_L</math>

Dividing by <math>(dX)^2</math> we have <math display="block">\frac{dx-dX}{dX}\frac{dx+dX}{dX}=2\varepsilon_{ij}\frac{dX_i}{dX}\frac{dX_j}{dX}</math>

For small deformations we assume that <math>dx \approx dX</math>, thus the second term of the left hand side becomes: <math>\frac{dx+dX}{dX} \approx 2</math>.

Then we have <math display="block">\frac{dx-dX}{dX} = \varepsilon_{ij}N_iN_j = \mathbf N \cdot \boldsymbol \varepsilon \cdot \mathbf N</math> where <math>N_i=\frac{dX_i}{dX}</math>, is the unit vector in the direction of <math>d\mathbf X</math>, and the left-hand-side expression is the normal strain <math>e_{(\mathbf N)}</math> in the direction of <math>\mathbf N</math>. For the particular case of <math>\mathbf N</math> in the <math>X_1</math> direction, i.e., <math>\mathbf N = \mathbf I_1</math>, we have <math display="block">e_{(\mathbf I_1)}=\mathbf I_1 \cdot \boldsymbol \varepsilon \cdot \mathbf I_1 = \varepsilon_{11}.</math>

Similarly, for <math>\mathbf N=\mathbf I_2</math> and <math>\mathbf N=\mathbf I_3</math> we can find the normal strains <math>\varepsilon_{22}</math> and <math>\varepsilon_{33}</math>, respectively. Therefore, the diagonal elements of the infinitesimal strain tensor are the normal strains in the coordinate directions.

Strain transformation rules

If we choose an orthonormal coordinate system (<math>\mathbf{e}_1,\mathbf{e}_2,\mathbf{e}_3</math>) we can write the tensor in terms of components with respect to those base vectors as <math display="block"> \boldsymbol{\varepsilon} = \sum_{i=1}^3 \sum_{j=1}^3 \varepsilon_{ij} \mathbf{e}_i\otimes\mathbf{e}_j </math> In matrix form, <math display="block">\underline{\underline{\boldsymbol{\varepsilon}}} = \begin{bmatrix}

 \varepsilon_{11} & \varepsilon_{12} & \varepsilon_{13} \\
 \varepsilon_{12} & \varepsilon_{22} & \varepsilon_{23} \\
 \varepsilon_{13} & \varepsilon_{23} & \varepsilon_{33} 

\end{bmatrix} </math> We can easily choose to use another orthonormal coordinate system (<math>\hat{\mathbf{e}}_1,\hat{\mathbf{e}}_2,\hat{\mathbf{e}}_3</math>) instead. In that case the components of the tensor are different, say <math display="block">

  \boldsymbol{\varepsilon} = \sum_{i=1}^3 \sum_{j=1}^3 \hat{\varepsilon}_{ij} \hat{\mathbf{e}}_i\otimes\hat{\mathbf{e}}_j \quad \implies \quad \underline{\underline{\hat{\boldsymbol{\varepsilon}}}} = \begin{bmatrix}
    \hat{\varepsilon}_{11} & \hat{\varepsilon}_{12} & \hat{\varepsilon}_{13} \\
    \hat{\varepsilon}_{12} & \hat{\varepsilon}_{22} & \hat{\varepsilon}_{23} \\
    \hat{\varepsilon}_{13} & \hat{\varepsilon}_{23} & \hat{\varepsilon}_{33}

\end{bmatrix} </math> The components of the strain in the two coordinate systems are related by <math display="block"> \hat{\varepsilon}_{ij} = \ell_{ip}~\ell_{jq}~\varepsilon_{pq} </math> where the Einstein summation convention for repeated indices has been used and <math>\ell_{ij} = \hat{\mathbf{e}}_i\cdot{\mathbf{e}}_j</math>. In matrix form <math display="block">

  \underline{\underline{\hat{\boldsymbol{\varepsilon}}}} = \underline{\underline{\mathbf{L}}} ~\underline{\underline{\boldsymbol{\varepsilon}}}~ \underline{\underline{\mathbf{L}}}^T
</math>

or <math display="block"> \begin{bmatrix}

 \hat{\varepsilon}_{11} & \hat{\varepsilon}_{12} & \hat{\varepsilon}_{13} \\
 \hat{\varepsilon}_{21} & \hat{\varepsilon}_{22} & \hat{\varepsilon}_{23} \\
 \hat{\varepsilon}_{31} & \hat{\varepsilon}_{32} & \hat{\varepsilon}_{33}

\end{bmatrix} = \begin{bmatrix}

 \ell_{11} & \ell_{12} & \ell_{13} \\
 \ell_{21} & \ell_{22} & \ell_{23} \\
 \ell_{31} & \ell_{32} & \ell_{33}

\end{bmatrix} \begin{bmatrix}

 \varepsilon_{11} & \varepsilon_{12} & \varepsilon_{13} \\
 \varepsilon_{21} & \varepsilon_{22} & \varepsilon_{23} \\
 \varepsilon_{31} & \varepsilon_{32} & \varepsilon_{33}

\end{bmatrix} \begin{bmatrix} \ell_{11} & \ell_{12} & \ell_{13} \\ \ell_{21} & \ell_{22} & \ell_{23} \\ \ell_{31} & \ell_{32} & \ell_{33} \end{bmatrix}^T </math>

Strain invariants

Certain operations on the strain tensor give the same result without regard to which orthonormal coordinate system is used to represent the components of strain. The results of these operations are called strain invariants. The most commonly used strain invariants are <math display="block"> \begin{align}

  I_1 & = \mathrm{tr}(\boldsymbol{\varepsilon}) \\
  I_2 & = \tfrac{1}{2}\{[\mathrm{tr}(\boldsymbol{\varepsilon})]^2 - \mathrm{tr}(\boldsymbol{\varepsilon}^2)\} \\
  I_3 & = \det(\boldsymbol{\varepsilon})

\end{align} </math> In terms of components <math display="block"> \begin{align}

  I_1 & = \varepsilon_{11} + \varepsilon_{22} + \varepsilon_{33} \\
  I_2 & = \varepsilon_{11}\varepsilon_{22} + \varepsilon_{22}\varepsilon_{33} + \varepsilon_{33}\varepsilon_{11} - \varepsilon_{12}^2 - \varepsilon_{23}^2 - \varepsilon_{31}^2 \\
  I_3 & = \varepsilon_{11}(\varepsilon_{22}\varepsilon_{33} - \varepsilon_{23}^2) - \varepsilon_{12}(\varepsilon_{21}\varepsilon_{33}-\varepsilon_{23}\varepsilon_{31}) + \varepsilon_{13}(\varepsilon_{21}\varepsilon_{32}-\varepsilon_{22}\varepsilon_{31})

\end{align} </math>

Principal strains

It can be shown that it is possible to find a coordinate system (<math>\mathbf{n}_1,\mathbf{n}_2,\mathbf{n}_3</math>) in which the components of the strain tensor are <math display="block">

  \underline{\underline{\boldsymbol{\varepsilon}}} = \begin{bmatrix} \varepsilon_{1} & 0 & 0 \\
    0 & \varepsilon_{2} & 0 \\
    0 & 0 & \varepsilon_{3} \end{bmatrix} \quad \implies \quad \boldsymbol{\varepsilon} = \varepsilon_{1} \mathbf{n}_1\otimes\mathbf{n}_1 + \varepsilon_{2} \mathbf{n}_2\otimes\mathbf{n}_2 + \varepsilon_{3} \mathbf{n}_3\otimes\mathbf{n}_3 
</math>

The components of the strain tensor in the (<math>\mathbf{n}_1,\mathbf{n}_2,\mathbf{n}_3</math>) coordinate system are called the principal strains and the directions <math>\mathbf{n}_i</math> are called the directions of principal strain. Since there are no shear strain components in this coordinate system, the principal strains represent the maximum and minimum stretches of an elemental volume.

If we are given the components of the strain tensor in an arbitrary orthonormal coordinate system, we can find the principal strains using an eigenvalue decomposition determined by solving the system of equations <math display="block"> (\underline{\underline{\boldsymbol{\varepsilon}}} - \varepsilon_i~\underline{\underline{\mathbf{I}}})~\mathbf{n}_i = \underline{\mathbf{0}} </math> This system of equations is equivalent to finding the vector <math>\mathbf{n}_i</math> along which the strain tensor becomes a pure stretch with no shear component.

Volumetric strain

The volumetric strain, also called bulk strain, is the relative variation of the volume, as arising from dilation or compression; it is the first strain invariant or trace of the tensor: <math display="block">\delta=\frac{\Delta V}{V_0} = I_1 = \varepsilon_{11} + \varepsilon_{22} + \varepsilon_{33}</math> Actually, if we consider a cube with an edge length a, it is a quasi-cube after the deformation (the variations of the angles do not change the volume) with the dimensions <math>a \cdot (1 + \varepsilon_{11}) \times a \cdot (1 + \varepsilon_{22}) \times a \cdot (1 + \varepsilon_{33})</math> and V0 = a3, thus <math display="block">\frac{\Delta V}{V_0} = \frac{\left ( 1 + \varepsilon_{11} + \varepsilon_{22} + \varepsilon_{33} + \varepsilon_{11} \cdot \varepsilon_{22} + \varepsilon_{11} \cdot \varepsilon_{33}+ \varepsilon_{22} \cdot \varepsilon_{33} + \varepsilon_{11} \cdot \varepsilon_{22} \cdot \varepsilon_{33} \right ) \cdot a^3 - a^3}{a^3}</math> as we consider small deformations, <math display="block">1 \gg \varepsilon_{ii} \gg \varepsilon_{ii} \cdot \varepsilon_{jj} \gg \varepsilon_{11} \cdot \varepsilon_{22} \cdot \varepsilon_{33} </math> therefore the formula.

Real variation of volume (top) and the approximated one (bottom): the green drawing shows the estimated volume and the orange drawing the neglected volume

In case of pure shear, we can see that there is no change of the volume.

Strain deviator tensor

The infinitesimal strain tensor <math>\varepsilon_{ij}</math>, similarly to the Cauchy stress tensor, can be expressed as the sum of two other tensors:

  1. a mean strain tensor or volumetric strain tensor or spherical strain tensor, <math>\varepsilon_M\delta_{ij}</math>, related to dilation or volume change; and
  2. a deviatoric component called the strain deviator tensor, <math>\varepsilon'_{ij}</math>, related to distortion.

<math display="block">\varepsilon_{ij}= \varepsilon'_{ij} + \varepsilon_M\delta_{ij}</math> where <math>\varepsilon_M</math> is the mean strain given by <math display="block">\varepsilon_M = \frac{\varepsilon_{kk}}{3} = \frac{\varepsilon_{11} + \varepsilon_{22} + \varepsilon_{33}}{3} = \tfrac{1}{3}I^e_1</math>

The deviatoric strain tensor can be obtained by subtracting the mean strain tensor from the infinitesimal strain tensor: <math display="block">\begin{align} \ \varepsilon'_{ij} &= \varepsilon_{ij} - \frac{\varepsilon_{kk}}{3}\delta_{ij} \\ \begin{bmatrix}

  \varepsilon'_{11} & \varepsilon'_{12} & \varepsilon'_{13} \\
  \varepsilon'_{21} & \varepsilon'_{22} & \varepsilon'_{23} \\
  \varepsilon'_{31} & \varepsilon'_{32} & \varepsilon'_{33} \\
 \end{bmatrix}

&=\begin{bmatrix}

  \varepsilon_{11} & \varepsilon_{12} & \varepsilon_{13} \\
  \varepsilon_{21} & \varepsilon_{22} & \varepsilon_{23} \\
  \varepsilon_{31} & \varepsilon_{32} & \varepsilon_{33} \\
 \end{bmatrix}

- \begin{bmatrix}

  \varepsilon_M & 0 & 0 \\
  0 & \varepsilon_M & 0 \\
  0 & 0 & \varepsilon_M \\
 \end{bmatrix} \\

&=\begin{bmatrix}

  \varepsilon_{11}-\varepsilon_M & \varepsilon_{12} & \varepsilon_{13} \\
  \varepsilon_{21} & \varepsilon_{22}-\varepsilon_M & \varepsilon_{23} \\
  \varepsilon_{31} & \varepsilon_{32} & \varepsilon_{33}-\varepsilon_M \\
 \end{bmatrix} \\

\end{align}</math>

Octahedral strains

Let (<math>\mathbf{n}_1, \mathbf{n}_2, \mathbf{n}_3</math>) be the directions of the three principal strains. An octahedral plane is one whose normal makes equal angles with the three principal directions. The engineering shear strain on an octahedral plane is called the octahedral shear strain and is given by <math display="block"> \gamma_{\mathrm{oct}} = \tfrac{2}{3}\sqrt{(\varepsilon_1-\varepsilon_2)^2 + (\varepsilon_2-\varepsilon_3)^2 + (\varepsilon_3-\varepsilon_1)^2} </math> where <math>\varepsilon_1, \varepsilon_2, \varepsilon_3</math> are the principal strains. Шаблон:Citation needed

The normal strain on an octahedral plane is given by <math display="block"> \varepsilon_{\mathrm{oct}} = \tfrac{1}{3}(\varepsilon_1 + \varepsilon_2 + \varepsilon_3) </math> Шаблон:Citation needed

Equivalent strain

A scalar quantity called the equivalent strain, or the von Mises equivalent strain, is often used to describe the state of strain in solids. Several definitions of equivalent strain can be found in the literature. A definition that is commonly used in the literature on plasticity is <math display="block">

  \varepsilon_{\mathrm{eq}} = \sqrt{\tfrac{2}{3} \boldsymbol{\varepsilon}^{\mathrm{dev}}:\boldsymbol{\varepsilon}^{\mathrm{dev}}}  = \sqrt{\tfrac{2}{3}\varepsilon_{ij}^{\mathrm{dev}}\varepsilon_{ij}^{\mathrm{dev}}} 
 ~;~~ \boldsymbol{\varepsilon}^{\mathrm{dev}} = \boldsymbol{\varepsilon} - \tfrac{1}{3}\mathrm{tr}(\boldsymbol{\varepsilon})~\boldsymbol{I} 
</math>

This quantity is work conjugate to the equivalent stress defined as <math display="block"> \sigma_{\mathrm{eq}} = \sqrt{\tfrac{3}{2} \boldsymbol{\sigma}^{\mathrm{dev}}:\boldsymbol{\sigma}^{\mathrm{dev}}} </math>

Compatibility equations

Шаблон:Main For prescribed strain components <math>\varepsilon_{ij}</math> the strain tensor equation <math>u_{i,j}+u_{j,i}= 2 \varepsilon_{ij}</math> represents a system of six differential equations for the determination of three displacements components <math>u_i</math>, giving an over-determined system. Thus, a solution does not generally exist for an arbitrary choice of strain components. Therefore, some restrictions, named compatibility equations, are imposed upon the strain components. With the addition of the three compatibility equations the number of independent equations are reduced to three, matching the number of unknown displacement components. These constraints on the strain tensor were discovered by Saint-Venant, and are called the "Saint Venant compatibility equations".

The compatibility functions serve to assure a single-valued continuous displacement function <math>u_i</math>. If the elastic medium is visualised as a set of infinitesimal cubes in the unstrained state, after the medium is strained, an arbitrary strain tensor may not yield a situation in which the distorted cubes still fit together without overlapping.

In index notation, the compatibility equations are expressed as <math display="block">\varepsilon_{ij,km}+\varepsilon_{km,ij}-\varepsilon_{ik,jm}-\varepsilon_{jm,ik}=0</math>

In engineering notation,

  • <math>\frac{\partial^2 \epsilon_x}{\partial y^2} + \frac{\partial^2 \epsilon_y}{\partial x^2} = 2 \frac{\partial^2 \epsilon_{xy}}{\partial x \partial y}</math>
  • <math>\frac{\partial^2 \epsilon_y}{\partial z^2} + \frac{\partial^2 \epsilon_z}{\partial y^2} = 2 \frac{\partial^2 \epsilon_{yz}}{\partial y \partial z}</math>
  • <math>\frac{\partial^2 \epsilon_x}{\partial z^2} + \frac{\partial^2 \epsilon_z}{\partial x^2} = 2 \frac{\partial^2 \epsilon_{zx}}{\partial z \partial x}</math>
  • <math>\frac{\partial^2 \epsilon_x}{\partial y \partial z} = \frac{\partial}{\partial x} \left ( -\frac{\partial \epsilon_{yz}}{\partial x} + \frac{\partial \epsilon_{zx}}{\partial y} + \frac{\partial \epsilon_{xy}}{\partial z}\right)</math>
  • <math>\frac{\partial^2 \epsilon_y}{\partial z \partial x} = \frac{\partial}{\partial y} \left ( \frac{\partial \epsilon_{yz}}{\partial x} - \frac{\partial \epsilon_{zx}}{\partial y} + \frac{\partial \epsilon_{xy}}{\partial z}\right)</math>
  • <math>\frac{\partial^2 \epsilon_z}{\partial x \partial y} = \frac{\partial}{\partial z} \left ( \frac{\partial \epsilon_{yz}}{\partial x} + \frac{\partial \epsilon_{zx}}{\partial y} - \frac{\partial \epsilon_{xy}}{\partial z}\right)</math>

Special cases

Plane strain

Файл:Plane strain.svg
Plane strain state in a continuum.

In real engineering components, stress (and strain) are 3-D tensors but in prismatic structures such as a long metal billet, the length of the structure is much greater than the other two dimensions. The strains associated with length, i.e., the normal strain <math>\varepsilon_{33}</math> and the shear strains <math>\varepsilon_{13}</math> and <math>\varepsilon_{23}</math> (if the length is the 3-direction) are constrained by nearby material and are small compared to the cross-sectional strains. Plane strain is then an acceptable approximation. The strain tensor for plane strain is written as: <math display="block">\underline{\underline{\boldsymbol{\varepsilon}}} = \begin{bmatrix} \varepsilon_{11} & \varepsilon_{12} & 0 \\ \varepsilon_{21} & \varepsilon_{22} & 0 \\

    0      &     0       & 0

\end{bmatrix}</math> in which the double underline indicates a second order tensor. This strain state is called plane strain. The corresponding stress tensor is: <math display="block">\underline{\underline{\boldsymbol{\sigma}}} = \begin{bmatrix} \sigma_{11} & \sigma_{12} & 0 \\ \sigma_{21} & \sigma_{22} & 0 \\

    0      &     0       & \sigma_{33}

\end{bmatrix}</math> in which the non-zero <math>\sigma_{33}</math> is needed to maintain the constraint <math>\epsilon_{33} = 0</math>. This stress term can be temporarily removed from the analysis to leave only the in-plane terms, effectively reducing the 3-D problem to a much simpler 2-D problem.

Antiplane strain

Шаблон:Main Antiplane strain is another special state of strain that can occur in a body, for instance in a region close to a screw dislocation. The strain tensor for antiplane strain is given by <math display="block">\underline{\underline{\boldsymbol{\varepsilon}}} = \begin{bmatrix} 0 & 0 & \varepsilon_{13} \\ 0 & 0 & \varepsilon_{23}\\ \varepsilon_{13} & \varepsilon_{23} & 0 \end{bmatrix}</math>

Relation to infinitesimal rotation tensor Шаблон:Anchor

Шаблон:See also

The infinitesimal strain tensor is defined as <math display="block"> \boldsymbol{\varepsilon} = \frac{1}{2} [\boldsymbol{\nabla}\mathbf{u} + (\boldsymbol{\nabla}\mathbf{u})^T]</math> Therefore the displacement gradient can be expressed as <math display="block"> \boldsymbol{\nabla}\mathbf{u} = \boldsymbol{\varepsilon} + \boldsymbol{W}</math> where <math display="block"> \boldsymbol{W} := \frac{1}{2} [\boldsymbol{\nabla}\mathbf{u} - (\boldsymbol{\nabla}\mathbf{u})^T]</math> The quantity <math>\boldsymbol{W}</math> is the infinitesimal rotation tensor or infinitesimal angular displacement tensor (related to the infinitesimal rotation matrix). This tensor is skew symmetric. For infinitesimal deformations the scalar components of <math>\boldsymbol{W}</math> satisfy the condition <math>|W_{ij}| \ll 1</math>. Note that the displacement gradient is small only if Шаблон:Em the strain tensor and the rotation tensor are infinitesimal.

The axial vector

A skew symmetric second-order tensor has three independent scalar components. These three components are used to define an axial vector, <math>\mathbf{w}</math>, as follows <math display="block"> W_{ij} = -\epsilon_{ijk}~w_k ~;~~ w_i = -\tfrac{1}{2}~\epsilon_{ijk}~W_{jk} </math> where <math>\epsilon_{ijk}</math> is the permutation symbol. In matrix form <math display="block"> \underline{\underline{\boldsymbol{W}}} = \begin{bmatrix} 0 & -w_3 & w_2 \\ w_3 & 0 & -w_1 \\ -w_2 & w_1 & 0\end{bmatrix} ~;~~ \underline{\mathbf{w}} = \begin{bmatrix} w_1 \\ w_2 \\ w_3 \end{bmatrix} </math> The axial vector is also called the infinitesimal rotation vector. The rotation vector is related to the displacement gradient by the relation <math display="block"> \mathbf{w} = \tfrac{1}{2}~ \boldsymbol{\nabla} \times \mathbf{u} </math> In index notation <math display="block"> w_i = \tfrac{1}{2}~\epsilon_{ijk}~u_{k,j} </math> If <math>\lVert\boldsymbol{W}\rVert \ll 1 </math> and <math>\boldsymbol{\varepsilon} = \boldsymbol{0}</math> then the material undergoes an approximate rigid body rotation of magnitude <math>|\mathbf{w}|</math> around the vector <math>\mathbf{w}</math>.

Relation between the strain tensor and the rotation vector

Given a continuous, single-valued displacement field <math>\mathbf{u}</math> and the corresponding infinitesimal strain tensor <math>\boldsymbol{\varepsilon}</math>, we have (see Tensor derivative (continuum mechanics)) <math display="block">\boldsymbol{\nabla}\times\boldsymbol{\varepsilon} = e_{ijk}~\varepsilon_{lj,i}~\mathbf{e}_k\otimes\mathbf{e}_l

= \tfrac{1}{2}~e_{ijk}~[u_{l,ji} + u_{j,li}]~\mathbf{e}_k\otimes\mathbf{e}_l </math>

Since a change in the order of differentiation does not change the result, <math>u_{l,ji} = u_{l,ij}</math>. Therefore <math display="block"> e_{ijk} u_{l,ji} = (e_{12k}+e_{21k}) u_{l,12} + (e_{13k}+e_{31k}) u_{l,13} + (e_{23k} + e_{32k}) u_{l,32} = 0 </math> Also <math display="block"> \tfrac{1}{2}~e_{ijk}~u_{j,li} = \left(\tfrac{1}{2}~e_{ijk}~u_{j,i}\right)_{,l} = \left(\tfrac{1}{2} ~ e_{kij}~u_{j,i}\right)_{,l} = w_{k,l} </math> Hence <math display="block"> \boldsymbol{\nabla} \times \boldsymbol{\varepsilon} = w_{k,l}~\mathbf{e}_k\otimes\mathbf{e}_l = \boldsymbol{\nabla}\mathbf{w} </math>

Relation between rotation tensor and rotation vector

From an important identity regarding the curl of a tensor we know that for a continuous, single-valued displacement field <math>\mathbf{u}</math>, <math display="block"> \boldsymbol{\nabla}\times(\boldsymbol{\nabla}\mathbf{u}) = \boldsymbol{0}. </math> Since <math>\boldsymbol{\nabla}\mathbf{u} = \boldsymbol{\varepsilon} + \boldsymbol{W}</math> we have <math display="block"> \boldsymbol{\nabla}\times\boldsymbol{W} = -\boldsymbol{\nabla}\times\boldsymbol{\varepsilon} = - \boldsymbol{\nabla} \mathbf{w}. </math>

Strain tensor in non-Cartesian coordinates

Strain tensor in cylindrical coordinates

In cylindrical polar coordinates (<math>r, \theta, z</math>), the displacement vector can be written as <math display="block"> \mathbf{u} = u_r~\mathbf{e}_r + u_\theta~\mathbf{e}_\theta + u_z~\mathbf{e}_z </math> The components of the strain tensor in a cylindrical coordinate system are given by:[2] <math display="block">\begin{align}

   \varepsilon_{rr} & = \cfrac{\partial u_r}{\partial r} \\
   \varepsilon_{\theta\theta} & = \cfrac{1}{r}\left(\cfrac{\partial u_\theta}{\partial \theta} + u_r\right) \\
   \varepsilon_{zz} & = \cfrac{\partial u_z}{\partial z} \\
   \varepsilon_{r\theta} & = \cfrac{1}{2} \left(\cfrac{1}{r}\cfrac{\partial u_r}{\partial \theta} + \cfrac{\partial u_\theta}{\partial r} - \cfrac{u_\theta}{r}\right) \\
   \varepsilon_{\theta z} & = \cfrac{1}{2} \left(\cfrac{\partial u_\theta}{\partial z} + \cfrac{1}{r} \cfrac{\partial u_z}{\partial \theta}\right) \\
   \varepsilon_{zr} & = \cfrac{1}{2} \left(\cfrac{\partial u_r}{\partial z} + \cfrac{\partial u_z}{\partial r}\right) 

\end{align}</math>

Strain tensor in spherical coordinates

Файл:3D Spherical.svg
Spherical coordinates (r, θ, φ) as commonly used in physics: radial distance r, polar angle θ (theta), and azimuthal angle φ (phi). The symbol ρ (rho) is often used instead of r.

In spherical coordinates (<math>r, \theta, \phi</math>), the displacement vector can be written as <math display="block">

  \mathbf{u} = u_r~\mathbf{e}_r + u_\theta~\mathbf{e}_\theta + u_\phi~\mathbf{e}_\phi
</math>

The components of the strain tensor in a spherical coordinate system are given by [2] <math display="block">\begin{align}

   \varepsilon_{rr} & = \cfrac{\partial u_r}{\partial r} \\
   \varepsilon_{\theta\theta} & = \cfrac{1}{r}\left(\cfrac{\partial u_\theta}{\partial \theta} + u_r\right) \\
   \varepsilon_{\phi\phi} & = \cfrac{1}{r\sin\theta}\left(\cfrac{\partial u_\phi}{\partial \phi} + u_r\sin\theta + u_\theta\cos\theta\right)\\
   \varepsilon_{r\theta} & = \cfrac{1}{2}\left(\cfrac{1}{r}\cfrac{\partial u_r}{\partial \theta} + \cfrac{\partial u_\theta}{\partial r}- \cfrac{u_\theta}{r}\right) \\
   \varepsilon_{\theta \phi} & = \cfrac{1}{2r}\left(\cfrac{1}{\sin\theta}\cfrac{\partial u_\theta}{\partial \phi} + \cfrac{\partial u_\phi}{\partial \theta} - u_\phi\cot\theta\right) \\
   \varepsilon_{\phi r} & = \cfrac{1}{2}\left(\cfrac{1}{r\sin\theta}\cfrac{\partial u_r}{\partial \phi} + \cfrac{\partial u_\phi}{\partial r} - \cfrac{u_\phi}{r}\right) 

\end{align} </math>

See also

References

Шаблон:Reflist

External links

Шаблон:Branches of physics Шаблон:Infinitesimals