Skip to contents

Fault offsets

The offset along a fault can be factorized into several components.

Fig. 1: Graphic illustration of displacement components along a fault. \mathbb{G} as the georeference frame with D = down, E = East, N = North.
Fig. 1: Graphic illustration of displacement components along a fault. 𝔾\mathbb{G} as the georeference frame with D = down, E = East, N = North.

Get different components with trigonometry

1. Input: Fault orientation (dip angle, dip direction), shorteninig direction, and horizontal throw

Knowing the horizontal throw (e.g.Β from plate motion parameters), the remaining components of the displacements along a given fault are as follows.

Ξ΄=|ΟƒHmaxβˆ’(dip direction+90∘)|\begin{equation}\delta = |\sigma_{\textrm{Hmax}} - (\textrm{dip direction}+90^{\circ})|\end{equation}

Slip components in the horizontal plane:

fstrike slip=|cosΞ΄*fhorizontal throw|fheave=fhorizontal throw2βˆ’fstrike slip2\begin{equation}\begin{split} f_\textrm{strike slip} & = |\cos{\delta} * f_\textrm{horizontal throw}|\\ f_\textrm{heave} & = \sqrt{f_\textrm{horizontal throw}^2 - f_\textrm{strike slip}^2} \end{split}\end{equation}

Slip components in the vertical plane perpendicular to the strike of the fault:

fdip slip=fheavecos(dip)fvertical throw=fdip slip2βˆ’fheave2\begin{equation}\begin{split} f_\textrm{dip slip} &= \frac{f_\textrm{heave}}{cos{(\textrm{dip})}}\\ f_\textrm{vertical throw} &= \sqrt{f_\textrm{dip slip}^2 - f_\textrm{heave}^2} \end{split}\end{equation}

Slip components in the fault plane plane:

fnet slip=fstrike slip2+fdip slip2rake=arctan(fdip slipfstrike slip)\begin{equation}\begin{split} f_\textrm{net slip} &= \sqrt{f_\textrm{strike slip}^2 + f_\textrm{dip slip}^2}\\ \textrm{rake} &= \arctan{\left(\frac{f_\textrm{dip slip}}{f_\textrm{strike slip}}\right)} \end{split}\end{equation}

Thus, the rake angle describes the ratio between the dip slip and the strike slip component.

Knowing the vertical throw (e.g.Β from thermochronology or petrology), the fault dip (?assumption), and the direction and amount of horizontal offset, the strike of the fault is as follows:

fheave=fvertical throw*tan(dip)Ξ΄=arcsin(fheavefhorizontal throw)strike=|ΟƒHmaxβˆ’Ξ΄|\begin{equation}\begin{split} f_\textrm{heave} &= f_\textrm{vertical throw} * \tan{(\textrm{dip})}\\ \delta &= \arcsin{\left(\frac{f_\textrm{heave}}{f_\textrm{horizontal throw}}\right)}\\ \textrm{strike} &= |\sigma_{\textrm{Hmax}} - \delta| \end{split}\end{equation}

Knowing the vertical throw (e.g.Β from thermochronology or petrology), the fault strike (geomorphology), and the direction and amount of horizontal offset, the dip of the fault is as follows:

Ξ΄=|ΟƒHmaxβˆ’strike|fheave=fhorizontal throw*sinΞ΄dip=arctan(fheavefvertical throw)\begin{equation}\begin{split} \delta &= |\sigma_{\textrm{Hmax}} - \textrm{strike}|\\ f_\textrm{heave} &= f_\textrm{horizontal throw} * \sin{\delta} \\ \textrm{dip} &= \arctan{\left(\frac{f_\textrm{heave}}{f_\textrm{vertical throw}}\right)} \end{split}\end{equation}

Knowing the vertical throw (e.g.Β from thermochronology or petrology) and the fault’s dip and rake, the horizontal offset, the horizontal throw, and the net-slip are as follows:

Fault displacement tensors

Each fault component is a vector describing its direction and length. For instance, the vector of the strike slip is:

fstrike slipβ†’=(βˆ₯fstrike-slipβˆ₯00)\begin{equation} \vec{f_\text{strike slip}} = \begin{pmatrix} \lVert f_\textrm{strike-slip}\rVert \\ 0 \\ 0 \end{pmatrix} \end{equation}

por <- data.frame("lat" = 45, "lon" = 0) # Point of origin in lat, lon coordinates
fault <- data.frame("dipdir" = 0, "dip" = 45) # fault given in dip direction and dip angle
strikeslip <- c("magnitude" = 2) # magnitude of strike slip component

# geographical_to_cartesian(por)
# pracma::deg2rad(strike)

Net slip vector

Net slip vector

fnetβ†’=(βˆ₯fstrike-slipβˆ₯βˆ₯fheaveβˆ₯βˆ₯fvertical throwβˆ₯)\begin{equation} \vec{f_\text{net}} = \begin{pmatrix} \lVert f_\textrm{strike-slip}\rVert \\ \lVert f_\textrm{heave}\rVert \\ \lVert f_\textrm{vertical throw}\rVert \end{pmatrix} \end{equation}

location <- data.frame("lat" = 45, "lon" = 0) # position of fault in lat, lon coordinates
fault <- data.frame("dipdir" = 0, "dip" = 45) # fault givin in dip direction and dip angle
horizontalthrow <- data.frame("direction" = 0, "magnitude" = 1) # direction and magnitude of horizontal throw
netslip_vector(location = location, fault = fault, horizontalthrow = horizontalthrow)

Principal displacement tensor

The Eigen values of π–₯ij\mathsf{F}_{ij}; represented as {f1β†’,f2β†’,f3β†’}\{ \vec{f_1}, \vec{f_2}, \vec{f_3} \} are referred to as the strike slip, heave, and vertical throw component, respectively. These orthonormal vectors define a orthogonal matrix, i.e.Β the principal displacement tensor π–₯𝔽\mathsf{F}_\mathbb{F}:

π–₯𝔽=[βˆ₯fstrike-slipβˆ₯000βˆ₯fheaveβˆ₯000βˆ₯fvertical throwβˆ₯]\begin{equation}\mathsf{F}_{\mathbb{F}} = {\begin{bmatrix} \lVert f_\textrm{strike-slip}\rVert & 0 & 0\\ 0 & \lVert f_\textrm{heave}\rVert & 0\\ 0 & 0 & \lVert f_\textrm{vertical throw}\rVert\end{bmatrix}}\end{equation}

The Eigen values comprise the netslip vector.

# F.f <- eigen(ORT)
# F.f

The tensor π–₯\mathsf{F} can also be defined by the magnitudes of the fault displacements:

F.f <- displacement_tensor(g)
F.f

Orientation tensor

Fault orientation tensor is defined by the fault plane’s location, orientation (dip direction and dip angle), and the fault’s slip (direction and magnitude):

π–₯ij=[f11f12f13f21f22f23f31f32f33]\begin{equation}\mathsf{F}_{ij} = {\begin{bmatrix}f_{11} & f_{12} & f_{13}\\ f_{21} & f_{22} & f_{23}\\ f_{31} & f_{32} & f_{33}\end{bmatrix}}\end{equation}

From Principal displacement tensor to Orientation tensor

Translation point of origin in π–₯𝔽\mathsf{F_\mathbb{F}} into point of measurement and rotate into fault orientation π–₯𝔽𝔾\mathsf{F}_\mathbb{FG}

Stress tensor

In continuum mechanics, the Cauchy stress tensor Οƒ\sigma, true stress tensor, or simply called the stress tensor is a second order tensor named after Augustin-Louis Cauchy.

The tensor consists of nine components Οƒij\sigma_{ij} that completely define the state of stress at a point inside a material in the deformed state, placement, or configuration. The tensor relates a unit-length direction vector n to the traction vector T(n)T^{(n)} across an imaginary surface perpendicular to n:

T(n)=n*Οƒ\begin{equation} T^{(n)} = n * \sigma \end{equation} or Tj(n)=ni*Οƒij\begin{equation} T^{(n)}_j = n_{i}*\sigma_{ij}\end{equation}

Transformation of the stress tensor

It can be shown that the stress tensor is a contravariant second order tensor, which is a statement of how it transforms under a change of the coordinate system. From an xix_i-system to an xiβ€²x_i'-system, the components Οƒij\sigma_{ij} in the initial system are transformed into the components Οƒijβ€²\sigma_{ij}' in the new system according to the tensor transformation rule (see Figure 2):

Οƒ=aimajnΟƒmn\sigma = a_{im} a_{jn} \sigma_{mn} or Οƒβ€²=𝖠σ𝖠T\sigma' = \mathsf{A} \sigma \mathsf{A}^{T},

where 𝖠\mathsf{A} is a rotation matrix with components aija_{ij}. In matrix form this is [Οƒβ€²11Οƒβ€²12Οƒβ€²13Οƒβ€²21Οƒβ€²22Οƒβ€²23Οƒβ€²31Οƒβ€²32Οƒβ€²33]=[a11a12a13a21a22a23a31a32a33][Οƒ11Οƒ12Οƒ13Οƒ21Οƒ22Οƒ23Οƒ31Οƒ32Οƒ33][a11a12a13a21a22a23a31a32a33]\begin{equation} \begin{bmatrix} \sigma'_{11} & \sigma'_{12} & \sigma'_{13} \\ \sigma'_{21} & \sigma'_{22} & \sigma'_{23} \\ \sigma'_{31} & \sigma'_{32} & \sigma'_{33} \end{bmatrix} = \begin{bmatrix} a_{11} & a_{12} & a_{13} \\ a_{21} & a_{22} & a_{23} \\ a_{31} & a_{32} & a_{33} \end{bmatrix} \begin{bmatrix} \sigma_{11} & \sigma_{12} & \sigma_{13} \\ \sigma_{21} & \sigma_{22} & \sigma_{23} \\ \sigma_{31} & \sigma_{32} & \sigma_{33} \end{bmatrix} \begin{bmatrix} a_{11} & a_{12} & a_{13} \\ a_{21} & a_{22} & a_{23} \\ a_{31} & a_{32} & a_{33} \end{bmatrix} \end{equation}

Fig. 2: Components of stress in three dimensions amd transformation of the Cauchy tensor. Source: https://commons.wikimedia.org/wiki/File:Stress_transformation_3D.svg
Fig. 2: Components of stress in three dimensions amd transformation of the Cauchy tensor. Source: https://commons.wikimedia.org/wiki/File:Stress_transformation_3D.svg

Expanding the matrix operation, and simplifying terms using the symmetry of the stress tensor, gives

Οƒ11β€²=a112Οƒ11+a122Οƒ22+a132Οƒ33+2a11a12Οƒ12+2a11a13Οƒ13+2a12a13Οƒ23\sigma_{11}' = a_{11}^{2}\sigma_{11} + a_{12}^{2}\sigma_{22} + a_{13}^{2}\sigma_{33} + 2a_{11}a_{12}\sigma_{12} + 2a_{11}a_{13}\sigma_{13} + 2a_{12}a_{13}\sigma_{23},

Οƒ22β€²=a212Οƒ11+a222Οƒ22+a232Οƒ33+2a21a22Οƒ12+2a21a23Οƒ13+2a22a23Οƒ23\sigma _{22}'=a_{21}^{2}\sigma _{11}+a_{22}^{2}\sigma _{22}+a_{23}^{2}\sigma _{33}+2a_{21}a_{22}\sigma _{12}+2a_{21}a_{23}\sigma _{13}+2a_{22}a_{23}\sigma _{23},

Οƒ33β€²=a312Οƒ11+a322Οƒ22+a332Οƒ33+2a31a32Οƒ12+2a31a33Οƒ13+2a32a33Οƒ23\sigma _{33}'=a_{31}^{2}\sigma _{11}+a_{32}^{2}\sigma _{22}+a_{33}^{2}\sigma _{33}+2a_{31}a_{32}\sigma _{12}+2a_{31}a_{33}\sigma _{13}+2a_{32}a_{33}\sigma _{23},

Οƒ12β€²=a11a21Οƒ11+a12a22Οƒ22+a13a23Οƒ33+(a11a22+a12a21)Οƒ12+(a12a23+a13a22)Οƒ23+(a11a23+a13a21)Οƒ13\sigma _{12}'=a_{11}a_{21}\sigma _{11}+a_{12}a_{22}\sigma _{22}+a_{13}a_{23}\sigma _{33}+(a_{11}a_{22}+a_{12}a_{21})\sigma _{12}+(a_{12}a_{23}+a_{13}a_{22})\sigma _{23}+(a_{11}a_{23}+a_{13}a_{21})\sigma _{13},

Οƒ23β€²=a21a31Οƒ11+a22a32Οƒ22+a23a33Οƒ33+(a21a32+a22a31)Οƒ12+(a22a33+a23a32)Οƒ23+(a21a33+a23a31)Οƒ13\sigma _{23}'=a_{21}a_{31}\sigma _{11}+a_{22}a_{32}\sigma _{22}+a_{23}a_{33}\sigma _{33}+(a_{21}a_{32}+a_{22}a_{31})\sigma _{12}+(a_{22}a_{33}+a_{23}a_{32})\sigma _{23}+(a_{21}a_{33}+a_{23}a_{31})\sigma _{13},

Οƒ13β€²=a11a31Οƒ11+a12a32Οƒ22+a13a33Οƒ33+(a11a32+a12a31)Οƒ12+(a12a33+a13a32)Οƒ23+(a11a33+a13a31)Οƒ13\sigma _{13}'=a_{11}a_{31}\sigma _{11}+a_{12}a_{32}\sigma _{22}+a_{13}a_{33}\sigma _{33}+(a_{11}a_{32}+a_{12}a_{31})\sigma _{12}+(a_{12}a_{33}+a_{13}a_{32})\sigma _{23}+(a_{11}a_{33}+a_{13}a_{31})\sigma _{13}.

The Mohr circle for stress is a graphical representation of this transformation of stresses.

Normal and shear stresses

The magnitude of the normal stress component Οƒn of any stress vector T(n)T(n) acting on an arbitrary plane with normal unit vector n at a given point, in terms of the components Οƒij\sigma_{ij} of the stress tensor Οƒ\sigma, is the dot product of the stress vector and the normal unit vector:

Οƒnβ†’=𝐓(n)β‹…nβ†’=Ti(n)ni=Οƒijninj.\begin{equation} \begin{split} \sigma_\vec{n} & = \mathbf{T}^{(n)}\cdot \vec{n} \\ & = T^{(n)}_i n_i \\ & =\sigma_{ij} n_i n_j. \end{split} \end{equation}

The magnitude of the shear stress component Ο„n\tau_n, acting orthogonal to the vector n, can then be found using the Pythagorean theorem:

Ο„n=(T(𝐧))2βˆ’Οƒn2=Ti(𝐧)Ti(𝐧)βˆ’Οƒn2,\begin{equation} \begin{split} \tau_{\mathrm {n} }&={\sqrt{\left(T^{(\mathbf{n})}\right)^{2} - \sigma_{\mathrm{n}}^{2}}}\\ &={\sqrt{T_{i}^{(\mathbf{n})}T_{i}^{(\mathbf{n})} - \sigma_{\mathrm{n}}^{2}}}, \end{split} \end{equation}

where

(T(𝐧))2=Ti(𝐧)Ti(𝐧)=(Οƒijnj)(Οƒiknk)=ΟƒijΟƒiknjnk.\begin{equation}{\displaystyle \left(T^{(\mathbf {n} )}\right)^{2}=T_{i}^{(\mathbf {n} )}T_{i}^{(\mathbf {n} )}=\left(\sigma _{ij}n_{j}\right)\left(\sigma _{ik}n_{k}\right)=\sigma _{ij}\sigma _{ik}n_{j}n_{k}}.\end{equation}

Principal stress

At every point in a stressed body there are at least three planes, called principal planes, with normal vectors n→\vec{n}, called principal directions, where the corresponding stress vector is perpendicular to the plane, i.e., parallel or in the same direction as the normal vector n→\vec{n}, and where there are no normal shear stressesτn→{\displaystyle \tau _{\vec{n}}}. The three stresses normal to these principal planes are called principal stresses.

The components Οƒij\sigma _{ij} of the stress tensor depend on the orientation of the coordinate system at the point under consideration. However, the stress tensor itself is a physical quantity and as such, it is independent of the coordinate system chosen to represent it. There are certain invariants associated with every tensor which are also independent of the coordinate system. For example, a vector is a simple tensor of rank one. In three dimensions, it has three components. The value of these components will depend on the coordinate system chosen to represent the vector, but the magnitude of the vector is a physical quantity (a scalar) and is independent of the Cartesian coordinate system chosen to represent the vector (so long as it is normal). Similarly, every second rank tensor (such as the stress and the strain tensors) has three independent invariant quantities associated with it. One set of such invariants are the principal stresses of the stress tensor, which are just the eigenvalues of the stress tensor. Their direction vectors are the principal directions or eigenvectors.

A stress vector parallel to the normal unit vector n→\vec{n} is given by:

𝐓(n)=Ξ»nβ†’=𝛔nnβ†’\begin{equation} \mathbf{T}^{(n)} = \lambda \vec{n} = \mathbf{\sigma}_{n}\vec{n} \end{equation}

where Ξ»\lambda is a constant of proportionality, and in this particular case corresponds to the magnitudes Οƒn\sigma_{n} of the normal stress vectors or principal stresses.

A coordinate system with axes oriented to the principal directions implies that the normal stresses are the principal stresses and the stress tensor is represented by a diagonal matrix:

Οƒij=[Οƒ1000Οƒ2000Οƒ3]\begin{equation} \sigma_{ij}={\begin{bmatrix}\sigma_{1}&0&0\\0&\sigma_{2}&0\\0&0&\sigma_{3}\end{bmatrix}} \end{equation}

The principal stresses can be combined to form the stress invariants, I1I_{1}, I2I_{2}, and I3I_{3}. The first and third invariant are the trace and determinant respectively, of the stress tensor. Thus,

I1=Οƒ1+Οƒ2+Οƒ3I2=Οƒ1Οƒ2+Οƒ2Οƒ3+Οƒ3Οƒ1I3=Οƒ1Οƒ2Οƒ3\begin{equation} \begin{split} I_{1} & = \sigma_{1}+\sigma_{2}+\sigma_{3}\\ I_{2} & = \sigma_{1}\sigma_{2}+\sigma_{2}\sigma_{3}+\sigma_{3}\sigma_{1}\\ I_{3} & = \sigma_{1}\sigma_{2}\sigma_{3} \end{split} \end{equation}

Because of its simplicity, the principal coordinate system is often useful when considering the state of the elastic medium at a particular point. Principal stresses are often expressed in the following equation for evaluating stresses in the x and y directions or axial and bending stresses on a part. The principal normal stresses can then be used to calculate the von Mises stress and ultimately the safety factor and margin of safety.

Οƒ1,Οƒ2=Οƒx+Οƒy2Β±(Οƒxβˆ’Οƒy2)2+Ο„xy2\begin{equation} \sigma_{1},\sigma_{2}={\frac {\sigma_{x}+\sigma_{y}}{2}}\pm {\sqrt {\left({\frac {\sigma_{x}-\sigma_{y}}{2}}\right)^{2}+\tau_{xy}^{2}}} \end{equation}

Using just the part of the equation under the square root is equal to the maximum and minimum shear stress for plus and minus. This is shown as:

Ο„max,Ο„min=Β±(Οƒxβˆ’Οƒy2)2+Ο„xy2\begin{equation} \tau_{\max },\tau_{\min }=\pm {\sqrt {\left({\frac {\sigma_{x}-\sigma_{y}}{2}}\right)^{2}+\tau_{xy}^{2}}} \end{equation}

Stress deviator tensor

The stress tensor Οƒij\sigma_{ij} can be expressed as the sum of two other stress tensors:

  • a mean hydrostatic stress tensor or volumetric stress tensor or mean normal stress tensor, πδij\pi \delta_{ij}, which tends to change the volume of the stressed body; and
  • a deviatoric component called the stress deviator tensor, sijs_{ij}, which tends to distort it.

So

Οƒij=sij+πδij,\begin{equation} \sigma_{ij} = s_{ij} + \pi \delta_{ij}, \end{equation}

where Ο€\pi is the mean stress given by

Ο€=Οƒkk3=Οƒ11+Οƒ22+Οƒ333=13I1,\begin{equation} \pi ={\frac {\sigma_{kk}}{3}}={\frac{\sigma_{11}+\sigma _{22}+\sigma_{33}}{3}}={\frac {1}{3}}I_{1}, \end{equation}

Pressure (pp) is generally defined as negative one-third the trace of the stress tensor minus any stress the divergence of the velocity contributes with, i.e.

p=Ξ»βˆ‡β‹…uβ†’βˆ’Ο€=Ξ»βˆ‚ukβˆ‚xkβˆ’Ο€=βˆ‘kΞ»βˆ‚ukβˆ‚xkβˆ’Ο€,\begin{equation} p=\lambda \,\nabla \cdot {\vec {u}}-\pi =\lambda \,{\frac {\partial u_{k}}{\partial x_{k}}}-\pi =\sum_{k}\lambda \,{\frac {\partial u_{k}}{\partial x_{k}}}-\pi, \end{equation}

where Ξ»\lambda is a proportionality constant, βˆ‡\nabla is the divergence operator, xkx_{k} is the k:th Cartesian coordinate, uβ†’\vec{u} is the velocity and uku_{k} is the k:th Cartesian component of uβ†’\vec{u}.

The deviatoric stress tensor can be obtained by subtracting the hydrostatic stress tensor from the Cauchy stress tensor:

sij=Οƒijβˆ’Οƒkk3Ξ΄ij,[s11s12s13s21s22s23s31s32s33]=[Οƒ11Οƒ12Οƒ13Οƒ21Οƒ22Οƒ23Οƒ31Οƒ32Οƒ33]βˆ’[Ο€000Ο€000Ο€]=[Οƒ11βˆ’Ο€Οƒ12Οƒ13Οƒ21Οƒ22βˆ’Ο€Οƒ23Οƒ31Οƒ32Οƒ33βˆ’Ο€].\begin{equation} \begin{split} s_{ij} & = \sigma_{ij} - {\frac {\sigma_{kk}}{3}}\delta_{ij},\\ \begin{bmatrix}s_{11}&s_{12}&s_{13}\\s_{21}&s_{22}&s_{23}\\s_{31}&s_{32}&s_{33}\end{bmatrix} & = \begin{bmatrix}\sigma_{11}&\sigma_{12}&\sigma_{13}\\\sigma_{21}&\sigma_{22}&\sigma_{23}\\\sigma_{31}&\sigma_{32}&\sigma_{33}\end{bmatrix} - \begin{bmatrix}\pi &0&0\\0&\pi &0\\0&0&\pi \end{bmatrix}\\ & = \begin{bmatrix}\sigma_{11}-\pi &\sigma_{12}&\sigma_{13}\\\sigma_{21}&\sigma_{22}-\pi &\sigma_{23}\\\sigma_{31}&\sigma_{32}&\sigma_{33}-\pi \end{bmatrix}. \end{split} \end{equation}

Maximum horizontal stress1

We de fine a vertical plane in the geographic coordinate system, 𝔾\mathbb{G}, using its unit normal vector, n→𝔾T=(nN,nE,nD)=(cosΞ±,sinΞ±,0)\vec{n}^T_\mathbb{G} = (n_N, n_E, n_D) = (\cos\alpha, \sin\alpha, 0), where Ξ±\alpha is the normal’s trend angle measured clockwise from north and the plane’s strike is Ξ±+Ο€/2\alpha+\pi/2. We represent nβ†’\vec{n} with respect to the principal stress coordinate system using the transformation matrix

π– π•Šπ”Ύ=(Οƒβ†’1gβ†’1Οƒβ†’1gβ†’2Οƒβ†’1gβ†’3Οƒβ†’2gβ†’1Οƒβ†’2gβ†’2Οƒβ†’2gβ†’3Οƒβ†’3gβ†’1Οƒβ†’3gβ†’2Οƒβ†’3gβ†’3)\begin{equation} \mathsf{A}_{\mathbb{S}\mathbb{G}} = \begin{pmatrix} \vec{\sigma}_1 \vec{g}_1 & \vec{\sigma}_1 \vec{g}_2 & \vec{\sigma}_1 \vec{g}_3 \\ \vec{\sigma}_2 \vec{g}_1 & \vec{\sigma}_2 \vec{g}_2 & \vec{\sigma}_2 \vec{g}_3 \\ \vec{\sigma}_3 \vec{g}_1 & \vec{\sigma}_3 \vec{g}_2 & \vec{\sigma}_3 \vec{g}_3 \\ \end{pmatrix} \end{equation}

(i.e.Β aij=Οƒigja_{ij} = \sigma_i g_j). Here Οƒi=(ΟƒiN,ΟƒiE,ΟƒiD)\sigma_i = (\sigma_{iN}, \sigma_{iE}, \sigma_{iD}) are unit vectors in the principal stress directions (principal stress coordinate system π•Š\mathbb{S} with unit vectors {π›”πŸβ†’,π›”πŸβ†’,π›”πŸ‘β†’}\{ \mathbf{\vec{\sigma_1}}, \mathbf{\vec{\sigma_2}}, \mathbf{\vec{\sigma_3}} \}) with respect to the geographic coordinate system (𝔾\mathbb{G}), and g1β†’=(1,0,0)\vec{g_1} = (1,0,0), g2β†’=(0,1,0)\vec{g_2} =(0,1,0) and g3β†’=(0,0,1)\vec{g_3} = (0,0,1) are the basis vectors in the geographic coordinate system. The normal vector expressed with respect to the principal stress system is then

π§β†’π•Š=π– π•Šπ”Ύπ§β†’π”Ύ=(Οƒ1NnN+Οƒ1EnEΟƒ2NnN+Οƒ2EnEΟƒ3NnN+Οƒ3EnE),\begin{equation} \vec{\mathbf{n}}_\mathbb{S} = \mathsf{A}_{\mathbb{S}\mathbb{G}}\vec{\mathbf{n}}_\mathbb{G} = \begin{pmatrix} \sigma_{1N}n_N + \sigma_{1E}n_E\\ \sigma_{2N}n_N + \sigma_{2E}n_E\\ \sigma_{3N}n_N + \sigma_{3E}n_E \end{pmatrix}, \end{equation}

where, for example, Οƒ1N\sigma_{1N} is the north component of the Οƒβ†’1\vec{\sigma}_1 unit vector. In the principal coordinate system, the stress tensor is diagonal

π–²π•Š=(βˆ₯Οƒ1βˆ₯000βˆ₯Οƒ2βˆ₯000βˆ₯Οƒ3βˆ₯)\begin{equation} \mathsf{S}_\mathbb{S} = \begin{pmatrix} \lVert \sigma_1 \rVert & 0 & 0 \\ 0 & \lVert\sigma_2\rVert & 0 \\ 0 & 0 & \lVert\sigma_3\rVert \end{pmatrix} \end{equation}

and the normal stress acting on the vertical plane of interest is

Οƒn=(π§β†’π•ŠTπ–²π•Šπ§β†’π•Š)=[βˆ₯Οƒ1βˆ₯(Οƒ1NnN+Οƒ1EnE)2+βˆ₯Οƒ2βˆ₯(Οƒ2NnN+Οƒ2EnE)2+βˆ₯Οƒ3βˆ₯(Οƒ3NnN+Οƒ3EnE)2]π§β†’π•Š=βˆ₯Οƒnβˆ₯π§β†’π•Š.\begin{equation} \begin{split} \sigma_n &= \left( \mathbf{\vec{n}}^T_\mathbb{S} \mathsf{S}_\mathbb{S} \mathbf{\vec{n}}_\mathbb{S} \right)\\ &= \left[ \lVert\sigma_1\rVert ( \sigma_{1N}n_N + \sigma_{1E}n_E )^2 + \lVert\sigma_2\rVert ( \sigma_{2N}n_N + \sigma_{2E}n_E )^2 + \lVert\sigma_3\rVert (\sigma_{3N}n_N + \sigma_{3E}n_E)^2 \right] \mathbf{\vec{n}}_{\mathbb{S}}\\ &= \lVert\sigma_n\rVert \mathbf{\vec{n}}_\mathbb{S}. \end{split} \end{equation}

This normal stress corresponds to the horizontal stress in the direction of 𝐧→\mathbf{\vec{n}}.

We can find the direction of the maximum horizontal stress analytically by differentiating the last equation with respect to Ξ±\alpha.

dΟƒndΞ±=[βˆ₯Οƒ1βˆ₯(Οƒ1E2βˆ’Οƒ1N2)+βˆ₯Οƒ2βˆ₯(Οƒ2E2βˆ’Οƒ2N2)+βˆ₯Οƒ3βˆ₯(Οƒ3E2βˆ’Οƒ3N2)]sin2Ξ±+2[βˆ₯Οƒ1βˆ₯Οƒ1NΟƒ1E+βˆ₯Οƒ2βˆ₯Οƒ2NΟƒ2E+βˆ₯Οƒ3βˆ₯Οƒ3NΟƒ3E]cos2Ξ±.\begin{equation} \frac{d\sigma_n}{d\alpha} = \left[ \lVert\sigma_1\rVert ( \sigma_{1E}^2 - \sigma_{1N}^2 ) + \lVert\sigma_2\rVert ( \sigma_{2E}^2 - \sigma_{2N}^2 ) + \lVert\sigma_3\rVert ( \sigma_{3E}^2 - \sigma_{3N}^2 ) \right] \sin{2\alpha} + 2 \left[\lVert\sigma_1\rVert \sigma_{1N} \sigma_{1E} + \lVert\sigma_2\rVert \sigma_{2N} \sigma_{2E} + \lVert\sigma_3\rVert \sigma_{3N} \sigma_{3E} \right] \cos{2\alpha}. \end{equation}

Table 1 Summary of the conditions under which the denominator of eq. (11) equals zero. Ξ±H\alpha_H and Ξ±1\alpha_1 denote the trends of the axis of maximum horizontal stress, ΟƒH\sigma_H, and the principal axis of maximum compressive stress, π›”πŸβ†’\mathbf{\vec{\sigma_1}}, respectively.
Condition Interpretation
Οƒ1N=Β±Οƒ1E\sigma_{1N} = \pm \sigma_{1E} and R=1R = 1 π•ŠH\mathbb{S}_H undefined if π›”πŸβ†’\mathbf{\vec{\sigma_1}} vertical, else Ξ±H=Ξ±1\alpha_H = \alpha_1 (45∘^{\circ} or 135∘^{\circ})
Οƒ1N=Β±Οƒ1E\sigma_{1N} = \pm \sigma_{1E} and Οƒ2N=Β±Οƒ2E\sigma_{2N} = \pm \sigma_{2E} π•ŠH\mathbb{S}_H undefined if R=0R = 0, else Ξ±H=Ξ±1\alpha_H = \alpha_1 (45∘^{\circ} or 135$^{})
(Οƒ1N2βˆ’Οƒ1E2)+(1βˆ’R)(Οƒ2N2βˆ’Οƒ2E2)=0(\sigma_{1N}^2 - \sigma_{1E}^2) + (1 - R)(\sigma_{2N}^2 - \sigma_{2E}^2) = 0 π•ŠH\mathbb{S}_H undefined if R=0R = 0, else Ξ±H=Ξ±1\alpha_H = \alpha_1 (45∘^{\circ} or 135$^{})

In general (cf.Β Table 1), Οƒn\sigma_n has one maximum and one minimum in the interval 0≀α≀π0 \leq \alpha \leq \pi and setting the derivative in the last equation to zero, we find these stationary points:

tan2Ξ±=2(βˆ₯Οƒ1βˆ₯Οƒ1NΟƒ1E+βˆ₯Οƒ2βˆ₯Οƒ2NΟƒ2E+βˆ₯Οƒ3βˆ₯Οƒ3NΟƒ3E)βˆ₯Οƒ1βˆ₯(Οƒ1E2βˆ’Οƒ1N2)+βˆ₯Οƒ2βˆ₯(Οƒ2E2βˆ’Οƒ2N2)+βˆ₯Οƒ3βˆ₯(Οƒ3E2βˆ’Οƒ3N2).\begin{equation} \tan{2\alpha} = \frac{ 2 ( \lVert\sigma_1\rVert \sigma_{1N} \sigma_{1E} + \lVert\sigma_2\rVert \sigma_{2N} \sigma_{2E} + \lVert\sigma_3\rVert \sigma_{3N} \sigma_{3E} )}{\lVert\sigma_1\rVert ( \sigma_{1E}^2 - \sigma_{1N}^2 ) + \lVert\sigma_2\rVert ( \sigma_{2E}^2 - \sigma_{2N}^2 ) + \lVert\sigma_3\rVert ( \sigma_{3E}^2 - \sigma_{3N}^2 )}. \end{equation}

Using the second derivative of Οƒn\sigma_n with respect to Ξ±\alpha we can determine whether Οƒn(Ξ±)\sigma_n (\alpha) yields a maximum or a minimum value, and identify the ΟƒH\sigma_H trend (Ξ±H\alpha_H) accordingly:

Ξ±H={Ξ±if βˆ₯Οƒn(Ξ±)βˆ₯ is a maximumΞ±+Ο€/2if βˆ₯Οƒn(Ξ±)βˆ₯ is a minimum.\begin{equation} \alpha_H = \begin{cases} \alpha & \text{if $\lVert\sigma_n (\alpha)\rVert$ is a maximum} \\ \alpha + \pi/2 & \text{if $\lVert\sigma_n (\alpha)\rVert$ is a minimum} \end{cases}. \end{equation}

Mountain building and earthquakes: Some simple geometrical calculations

Suppose we wish to compute how long it will take for 3000 meters (roughly 10,000 feet) of uplift to occur. Clearly, the time it requires is equal to the uplift (v=fvertical throwv=f_{\text{vertical throw}}) divided by the rate at which uplift occurs v̇=δfvertical throw)δt\dot{v} = \frac{\delta f_{\text{vertical throw})}}{\delta t}.

tv=|v|v̇\begin{equation} t_{\text{v}} = \frac{|v|}{\dot{v}} \end{equation}

To compute how horizontal strain Ρ\varepsilon that accumulates at a known rate (Ρ̇=5m/200ys\dot{\varepsilon} = 5\,\text{m} / 200\,\text{ys} in this example) translates into vertical uplift rate v̇\dot{v}, we note that it is related to the tangent of the dip of the fault (β\beta) as follows

v̇=ḣ*tanβ\begin{equation} \dot{v} = \dot{h} * \tan{\beta} \end{equation} where

ḣ=dearthquakeΔtearthquake\begin{equation} \dot{h} = \frac{d_\text{earthquake}}{\Delta t_\text{earthquake}} \end{equation}

dearthquaked_\text{earthquake} is the slip during average earthquake; Ξ”tearthquake\Delta t_\text{earthquake} is the earthquake recurrence interval.

Suppose that uplift occurs along a shallow fault dipping 10 degrees from horizontal. Then,

tv=3000m5m/200yr*tan10∘=680000yr\begin{equation} t_{v} = \frac{3\,000\,\text{m}}{5\,\text{m} / 200\,\text{yr}} * \tan{10^{\circ}} = 680\,000\,\text{yr} \end{equation}

For a steeper fault (30∘30^{\circ}), uplift of 3000m3\,000 \text{m} occurs even more rapidly, in only 210000yr210\,000\,\text{yr}.

Thus, for rapid convergence rates and even shallow fault dips, significant topographic relief can appear in only several hundred thousand years - a geological instant! Of course, these calculations ignore the important effects of erosion and downward flexure of the lithosphere under the load of a growing mountain, both of which slow the rate at which a mountain can grow. We shall consider these effects more in future lectures.


Abbreviations and symbols

Two coordinate systems are of importance here: a principal stress coordinate system, π•Š\mathbb{S}, with unit vectors {Οƒ1β†’,Οƒ2β†’,Οƒ3β†’}\{ \vec{\sigma_1},\vec{\sigma_2},\vec{\sigma_3} \} aligned along the eigenvectors whose eigenvalues are {βˆ₯Οƒ1βˆ₯,βˆ₯Οƒ2βˆ₯,βˆ₯Οƒ3βˆ₯}\{ \lVert\sigma_1\rVert,\lVert\sigma_2\rVert, \lVert\sigma_3\rVert \}, respectively; and a geographic coordinate system, 𝔾\mathbb{G}, with unit vectors {g1β†’,g2β†’,g3β†’}\{\vec{g_1}, \vec{g_2}, \vec{g_3} \} aligned with the north, east and down directions, respectively. In this paper, matrices are represented in sans serif upper-case letters (e.g. 𝖲\mathsf{S}), vectors in bold lower-case letters (e.g. 𝐧\mathbf{n}), and vector magnitudes in upper-case italics (e.g.Β Sn=βˆ₯𝐬nβˆ₯S_n = \lVert \mathbf{s}_n \rVert; Table 2). The components of matrices and vectors are represented, for example, as sijs_{ij} and nin_i, respectively, and a caret and subscript (e.g. 𝐬𝐒→\mathbf{\vec{s_i}} ) are used together to denote the ith unit basis vector of a coordinate system (i=1,2,3i = 1, 2, 3). Vector orientations are expressed in terms of trend and plunge angles as trend/plunge (e.g.Β 135/25). For convenience, we use upper-case bold symbols for the two most frequently used parameters, namely the vectors of maximum horizontal stress SHS_H and its proxy SPS_P.

Table 2 Mathematical symbols. Indices ii and jj each span the range 1, 2, 3, and an arrow (.β†’\vec{.}) denotes a unit vector.
Symbol Explanation
Ξ±\alpha Arbitrary trend angle measured clockwise from north
Ξ±H\alpha_H Trend of the maximum horizontal compressive stress
Ξ±P\alpha_P Trend of the maximum horizontal stress proxy
𝖠𝕏𝕐\mathsf{A}_{\mathbb{XY}} Rotation matrix between coordinate systems 𝕐\mathbb{Y} and 𝕏\mathbb{X}
Ξ²\beta fault dip angle
𝖣\mathsf{D} Deviatoric stress tensor (with respect to S3𝖨S_3\mathsf{I})
𝐝n\mathbf{d}_n Deviatoric normal stress vector
DnD_n Magnitude of the deviatoric normal stress vector
π–₯\mathsf{F} Fault displacement tensor
fif_i Fault displacement component
fi̇=Δfi/Δt\dot{f_i} = {\Delta f_i}/{\Delta t} Rate of fault displacement, i.e. displacement over time
gβ†’i\vec{g}_i Axes of the north-eastβ€”down coordinate system 𝔾\mathbb{G}
𝖨\mathsf{I} Identity matrix
𝐧→𝕏\mathbf{\vec{n}}_{\mathbb{X}} Unit normal vector represented in coordinate system 𝕏\mathbb{X}
nin_i Components of the vector 𝐧\mathbf{n}
RR Stress ratio equal to (S1βˆ’S2)/(S1βˆ’S3)(S_1 βˆ’S_2)/(S_1 βˆ’S_3)
𝖲\mathsf{S} Stress tensor
𝐒𝐇\mathbf{S_H} Maximum horizontal compressive stress vector
𝛔→𝐒\mathbf{\vec{\sigma}_i} Axes of the principal stress coordinate system π•Š\mathbb{S}
Οƒij\sigma_{ij} Components of matrix 𝖲\mathsf{S} (i.e.Β the jjth component of siβ†’\vec{s_i} )
𝛔𝐧\mathbf{\sigma_n} Normal stress vector
Sn=βˆ₯Οƒnβˆ₯S_n = \lVert \sigma_n \rVert Magnitude of the normal stress vector
𝐒P=𝛔P\mathbf{S}_P = \mathbf{\sigma}_P Maximum horizontal stress proxy vector
Si=βˆ₯Οƒiβˆ₯S_i = \lVert \sigma_i \rVert Magnitudes of the principal stresses (S1β‰₯S2β‰₯S3)(S_1 \geq S_2 \geq S_3)
tt Time
Ξ”t=|t2βˆ’t1|\Delta t = |t_2 - t_1| Time interval
𝕏\mathbb{X} Arbitrary coordinate system

References