Curvilinear coordinates








Curvilinear (top), affine (right), and Cartesian (left) coordinates in two-dimensional space


In geometry, curvilinear coordinates are a coordinate system for Euclidean space in which the coordinate lines may be curved. Commonly used curvilinear coordinate systems include: rectangular, spherical, and cylindrical coordinate systems. These coordinates may be derived from a set of Cartesian coordinates by using a transformation that is locally invertible (a one-to-one map) at each point. This means that one can convert a point given in a Cartesian coordinate system to its curvilinear coordinates and back. The name curvilinear coordinates, coined by the French mathematician Lamé, derives from the fact that the coordinate surfaces of the curvilinear systems are curved.


Well-known examples of curvilinear coordinate systems in three-dimensional Euclidean space (R3) are Cartesian, cylindrical and spherical polar coordinates. A Cartesian coordinate surface in this space is a coordinate plane; for example z = 0 defines the x-y plane. In the same space, the coordinate surface r = 1 in spherical polar coordinates is the surface of a unit sphere, which is curved. The formalism of curvilinear coordinates provides a unified and general description of the standard coordinate systems.


Curvilinear coordinates are often used to define the location or distribution of physical quantities which may be, for example, scalars, vectors, or tensors. Mathematical expressions involving these quantities in vector calculus and tensor analysis (such as the gradient, divergence, curl, and Laplacian) can be transformed from one coordinate system to another, according to transformation rules for scalars, vectors, and tensors. Such expressions then become valid for any curvilinear coordinate system.


Depending on the application, a curvilinear coordinate system may be simpler to use than the Cartesian coordinate system. For instance, a physical problem with spherical symmetry defined in R3 (for example, motion of particles under the influence of central forces) is usually easier to solve in spherical polar coordinates than in Cartesian coordinates. Equations with boundary conditions that follow coordinate surfaces for a particular curvilinear coordinate system may be easier to solve in that system. One would for instance describe the motion of a particle in a rectangular box in Cartesian coordinates, whereas one would prefer spherical coordinates for a particle in a sphere. Spherical coordinates are one of the most used curvilinear coordinate systems in such fields as Earth sciences, cartography, and physics (in particular quantum mechanics, relativity), and engineering.




Contents






  • 1 Orthogonal curvilinear coordinates in 3 Dimensions


    • 1.1 Coordinates, basis, and vectors




  • 2 Vector calculus


    • 2.1 Differential elements




  • 3 Covariant and contravariant bases


  • 4 Covariant basis


    • 4.1 Constructing a covariant basis in one dimension


    • 4.2 Constructing a covariant basis in three dimensions


    • 4.3 Jacobian of the transformation




  • 5 Generalization to n dimensions


  • 6 Transformation of coordinates


  • 7 Vector and tensor algebra in three-dimensional curvilinear coordinates


  • 8 Tensors in curvilinear coordinates


    • 8.1 The metric tensor in orthogonal curvilinear coordinates


      • 8.1.1 Relation to Lamé coefficients


      • 8.1.2 Example: Polar coordinates




    • 8.2 The alternating tensor


    • 8.3 Christoffel symbols


    • 8.4 Vector operations




  • 9 Vector and tensor calculus in three-dimensional curvilinear coordinates


    • 9.1 Geometric elements


    • 9.2 Integration


    • 9.3 Differentiation




  • 10 Fictitious forces in general curvilinear coordinates


  • 11 See also


  • 12 References


  • 13 Further reading


  • 14 External links





Orthogonal curvilinear coordinates in 3 Dimensions



Coordinates, basis, and vectors




Fig. 1 - Coordinate surfaces, coordinate lines, and coordinate axes of general curvilinear coordinates.




Fig. 2 - Coordinate surfaces, coordinate lines, and coordinate axes of spherical coordinates. Surfaces: r - spheres, θ - cones, φ - half-planes; Lines: r - straight beams, θ - vertical semicircles, φ - horizontal circles; Axes: r - straight beams, θ - tangents to vertical semicircles, φ - tangents to horizontal circles


For now, consider 3d space. A point P in 3d space (or its position vector r) can be defined using Cartesian coordinates (x, y, z) [equivalently written (x1, x2, x3)], by r=xex+yey+zez{displaystyle mathbf {r} =xmathbf {e} _{x}+ymathbf {e} _{y}+zmathbf {e} _{z}}mathbf{r} = x mathbf{e}_x + ymathbf{e}_y + zmathbf{e}_z, where ex, ey, ez are the standard basis vectors.


It can also be defined by its curvilinear coordinates (q1, q2, q3) if this triplet of numbers defines a single point in an unambiguous way. The relation between the coordinates is then given by the invertible transformation functions:



x=f1(q1,q2,q3),y=f2(q1,q2,q3),z=f3(q1,q2,q3){displaystyle x=f^{1}(q^{1},q^{2},q^{3}),,y=f^{2}(q^{1},q^{2},q^{3}),,z=f^{3}(q^{1},q^{2},q^{3})}{displaystyle x=f^{1}(q^{1},q^{2},q^{3}),,y=f^{2}(q^{1},q^{2},q^{3}),,z=f^{3}(q^{1},q^{2},q^{3})}

q1=g1(x,y,z),q2=g2(x,y,z),q3=g3(x,y,z){displaystyle q^{1}=g^{1}(x,y,z),,q^{2}=g^{2}(x,y,z),,q^{3}=g^{3}(x,y,z)}{displaystyle q^{1}=g^{1}(x,y,z),,q^{2}=g^{2}(x,y,z),,q^{3}=g^{3}(x,y,z)}


The surfaces q1 = constant, q2 = constant, q3 = constant are called the coordinate surfaces; and the space curves formed by their intersection in pairs are called the coordinate curves. The coordinate axes are determined by the tangents to the coordinate curves at the intersection of three surfaces. They are not in general fixed directions in space, which happens to be the case for simple Cartesian coordinates, and thus there is generally no natural global basis for curvilinear coordinates.


In the Cartesian system, the standard basis vectors can be derived from the derivative of the location of point P with respect to the local coordinate


ex=∂r∂x;ey=∂r∂y;ez=∂r∂z.{displaystyle mathbf {e} _{x}={dfrac {partial mathbf {r} }{partial x}};;mathbf {e} _{y}={dfrac {partial mathbf {r} }{partial y}};;mathbf {e} _{z}={dfrac {partial mathbf {r} }{partial z}}.}mathbf{e}_x = dfrac{partialmathbf{r}}{partial x}; ;<br />
mathbf{e}_y = dfrac{partialmathbf{r}}{partial y}; ;<br />
mathbf{e}_z = dfrac{partialmathbf{r}}{partial z}.

Applying the same derivatives to the curvilinear system locally at point P defines the natural basis vectors:


h1=∂r∂q1;h2=∂r∂q2;h3=∂r∂q3.{displaystyle mathbf {h} _{1}={dfrac {partial mathbf {r} }{partial q_{1}}};;mathbf {h} _{2}={dfrac {partial mathbf {r} }{partial q_{2}}};;mathbf {h} _{3}={dfrac {partial mathbf {r} }{partial q_{3}}}.}mathbf{h}_1 = dfrac{partialmathbf{r}}{partial q_1}; ;<br />
mathbf{h}_2 = dfrac{partialmathbf{r}}{partial q_2}; ;<br />
mathbf{h}_3 = dfrac{partialmathbf{r}}{partial q_3}.

Such a basis, whose vectors change their direction and/or magnitude from point to point is called a local basis. All bases associated with curvilinear coordinates are necessarily local. Basis vectors that are the same at all points are global bases, and can be associated only with linear or affine coordinate systems.


Note: for this article e is reserved for the standard basis (Cartesian) and h or b is for the curvilinear basis.


These may not have unit length, and may also not be orthogonal. In the case that they are orthogonal at all points where the derivatives are well-defined, we define the Lamé coefficients (after Gabriel Lamé) by


h1=|h1|;h2=|h2|;h3=|h3|{displaystyle h_{1}=|mathbf {h} _{1}|;;h_{2}=|mathbf {h} _{2}|;;h_{3}=|mathbf {h} _{3}|}h_1 = |mathbf{h}_1|; ; h_2 = |mathbf{h}_2|; ; h_3 = |mathbf{h}_3|

and the curvilinear orthonormal basis vectors by


b1=h1h1;b2=h2h2;b3=h3h3.{displaystyle mathbf {b} _{1}={dfrac {mathbf {h} _{1}}{h_{1}}};;mathbf {b} _{2}={dfrac {mathbf {h} _{2}}{h_{2}}};;mathbf {b} _{3}={dfrac {mathbf {h} _{3}}{h_{3}}}.}mathbf{b}_1 = dfrac{mathbf{h}_1}{h_1}; ;<br />
mathbf{b}_2 = dfrac{mathbf{h}_2}{h_2}; ;<br />
mathbf{b}_3 = dfrac{mathbf{h}_3}{h_3}.

It is important to note that these basis vectors may well depend upon the position of P; it is therefore necessary that they are not assumed to be constant over a region. (They technically form a basis for the tangent bundle of R3{displaystyle mathbb {R} ^{3}}mathbb {R} ^{3} at P, and so are local to P.)


In general, curvilinear coordinates allow the natural basis vectors hi not all mutually perpendicular to each other, and not required to be of unit length: they can be of arbitrary magnitude and direction. The use of an orthogonal basis makes vector manipulations simpler than for non-orthogonal. However, some areas of physics and engineering, particularly fluid mechanics and continuum mechanics, require non-orthogonal bases to describe deformations and fluid transport to account for complicated directional dependences of physical quantities. A discussion of the general case appears later on this page.



Vector calculus




Differential elements


In orthogonal curvilinear coordinates, since the total differential change in r is


dr=∂r∂q1dq1+∂r∂q2dq2+∂r∂q3dq3=h1dq1b1+h2dq2b2+h3dq3b3{displaystyle dmathbf {r} ={dfrac {partial mathbf {r} }{partial q^{1}}}dq^{1}+{dfrac {partial mathbf {r} }{partial q^{2}}}dq^{2}+{dfrac {partial mathbf {r} }{partial q^{3}}}dq^{3}=h_{1}dq^{1}mathbf {b} _{1}+h_{2}dq^{2}mathbf {b} _{2}+h_{3}dq^{3}mathbf {b} _{3}}{displaystyle dmathbf {r} ={dfrac {partial mathbf {r} }{partial q^{1}}}dq^{1}+{dfrac {partial mathbf {r} }{partial q^{2}}}dq^{2}+{dfrac {partial mathbf {r} }{partial q^{3}}}dq^{3}=h_{1}dq^{1}mathbf {b} _{1}+h_{2}dq^{2}mathbf {b} _{2}+h_{3}dq^{3}mathbf {b} _{3}}

so scale factors are hi=|∂r∂qi|{displaystyle h_{i}=left|{frac {partial mathbf {r} }{partial q^{i}}}right|}{displaystyle h_{i}=left|{frac {partial mathbf {r} }{partial q^{i}}}right|}


In non-orthogonal coordinates the length of dr=dq1h1+dq2h2+dq3h3{displaystyle dmathbf {r} =dq^{1}mathbf {h} _{1}+dq^{2}mathbf {h} _{2}+dq^{3}mathbf {h} _{3}}{displaystyle dmathbf {r} =dq^{1}mathbf {h} _{1}+dq^{2}mathbf {h} _{2}+dq^{3}mathbf {h} _{3}} is the positive square root of dr⋅dr=dqidqjhi⋅hj{displaystyle dmathbf {r} cdot dmathbf {r} =dq^{i}dq^{j}mathbf {h} _{i}cdot mathbf {h} _{j}}{displaystyle dmathbf {r} cdot dmathbf {r} =dq^{i}dq^{j}mathbf {h} _{i}cdot mathbf {h} _{j}} (with Einstein summation convention). The six independent scalar products gij=hi.hj of the natural basis vectors generalize the three scale factors defined above for orthogonal coordinates. The nine gij are the components of the metric tensor, which has only three non zero components in orthogonal coordinates: g11=h1h1, g22=h2h2, g33=h3h3.



Covariant and contravariant bases





A vector v (red) represented by • a vector basis (yellow, left: e1, e2, e3), tangent vectors to coordinate curves (black) and • a covector basis or cobasis (blue, right: e1, e2, e3), normal vectors to coordinate surfaces (grey) in general (not necessarily orthogonal) curvilinear coordinates (q1, q2, q3). Note the basis and cobasis do not coincide unless the coordinate system is orthogonal.[1]


Spatial gradients, distances, time derivatives and scale factors are interrelated within a coordinate system by two groups of basis vectors:



  1. basis vectors that are locally tangent to their associated coordinate pathline:
    bi=∂r∂qi{displaystyle mathbf {b} _{i}={dfrac {partial mathbf {r} }{partial q^{i}}}}{displaystyle mathbf {b} _{i}={dfrac {partial mathbf {r} }{partial q^{i}}}}

    which transforms like covariant vectors (denoted by lowered indices), or

  2. basis vectors that are locally normal to the isosurface created by the other coordinates:

    bi=∇qi{displaystyle mathbf {b} ^{i}=nabla q^{i}}{displaystyle mathbf {b} ^{i}=nabla q^{i}}


    which transforms like contravariant vectors (denoted by raised indices), ∇ is the del operator.


Consequently, a general curvilinear coordinate system has two sets of basis vectors for every point: {b1, b2, b3} is the covariant basis, and {b1, b2, b3} is the contravariant (a.k.a. reciprocal) basis. The covariant and contravariant basis vectors types have identical direction for orthogonal curvilinear coordinate systems, but as usual have inverted units with respect to each other.


Note the following important equality:


bi⋅bj=δji{displaystyle mathbf {b} _{i}cdot mathbf {b} ^{j}=delta _{j}^{i}}{displaystyle mathbf {b} _{i}cdot mathbf {b} ^{j}=delta _{j}^{i}}

wherein δji{displaystyle delta _{j}^{i}} delta^i_j denotes the generalized Kronecker delta.






A vector v can be specified in terms either basis, i.e.,


v=v1b1+v2b2+v3b3=v1b1+v2b2+v3b3{displaystyle mathbf {v} =v^{1}mathbf {b} _{1}+v^{2}mathbf {b} _{2}+v^{3}mathbf {b} _{3}=v_{1}mathbf {b} ^{1}+v_{2}mathbf {b} ^{2}+v_{3}mathbf {b} ^{3}} mathbf{v} = v^1mathbf{b}_1 + v^2mathbf{b}_2 + v^3mathbf{b}_3 = v_1mathbf{b}^1 + v_2mathbf{b}^2 + v_3mathbf{b}^3

Using the Einstein summation convention, the basis vectors relate to the components by[2](pp30–32)



v⋅bi=vkbk⋅bi=vkδki=vi{displaystyle mathbf {v} cdot mathbf {b} ^{i}=v^{k}mathbf {b} _{k}cdot mathbf {b} ^{i}=v^{k}delta _{k}^{i}=v^{i}} mathbf{v}cdotmathbf{b}^i = v^kmathbf{b}_kcdotmathbf{b}^i = v^kdelta^i_k = v^i

v⋅bi=vkbk⋅bi=vkδik=vi{displaystyle mathbf {v} cdot mathbf {b} _{i}=v_{k}mathbf {b} ^{k}cdot mathbf {b} _{i}=v_{k}delta _{i}^{k}=v_{i}} mathbf{v}cdotmathbf{b}_i = v_kmathbf{b}^kcdotmathbf{b}_i = v_kdelta_i^k = v_i


and



v⋅bi=vkbk⋅bi=gkivk{displaystyle mathbf {v} cdot mathbf {b} _{i}=v^{k}mathbf {b} _{k}cdot mathbf {b} _{i}=g_{ki}v^{k}} mathbf{v}cdotmathbf{b}_i = v^kmathbf{b}_kcdotmathbf{b}_i = g_{ki}v^k

v⋅bi=vkbk⋅bi=gkivk{displaystyle mathbf {v} cdot mathbf {b} ^{i}=v_{k}mathbf {b} ^{k}cdot mathbf {b} ^{i}=g^{ki}v_{k}} mathbf{v}cdotmathbf{b}^i = v_kmathbf{b}^kcdotmathbf{b}^i = g^{ki}v_k


where g is the metric tensor (see below).


A vector can be specified with covariant coordinates (lowered indices, written vk) or contravariant coordinates (raised indices, written vk). From the above vector sums, it can be seen that contravariant coordinates are associated with covariant basis vectors, and covariant coordinates are associated with contravariant basis vectors.


A key feature of the representation of vectors and tensors in terms of indexed components and basis vectors is invariance in the sense that vector components which transform in a covariant manner (or contravariant manner) are paired with basis vectors that transform in a contravariant manner (or covariant manner).



Covariant basis




Constructing a covariant basis in one dimension




Fig. 3 – Transformation of local covariant basis in the case of general curvilinear coordinates


Consider the one-dimensional curve shown in Fig. 3. At point P, taken as an origin, x is one of the Cartesian coordinates, and q1 is one of the curvilinear coordinates. The local (non-unit) basis vector is b1 (notated h1 above, with b reserved for unit vectors) and it is built on the q1 axis which is a tangent to that coordinate line at the point P. The axis q1 and thus the vector b1 form an angle α{displaystyle alpha }alpha with the Cartesian x axis and the Cartesian basis vector e1.


It can be seen from triangle PAB that


cos⁡α=|e1||b1|⇒|e1|=|b1|cos⁡α{displaystyle cos alpha ={cfrac {|mathbf {e} _{1}|}{|mathbf {b} _{1}|}}quad Rightarrow quad |mathbf {e} _{1}|=|mathbf {b} _{1}|cos alpha } cos alpha = cfrac{|mathbf{e}_1|}{|mathbf{b}_1|} quad Rightarrow quad |mathbf{e}_1| = |mathbf{b}_1|cos alpha

where |e1|, |b1| are the magnitudes of the two basis vectors, i.e., the scalar intercepts PB and PA. Note that PA is also the projection of b1 on the x axis.


However, this method for basis vector transformations using directional cosines is inapplicable to curvilinear coordinates for the following reasons:



  1. By increasing the distance from P, the angle between the curved line q1 and Cartesian axis x increasingly deviates from α{displaystyle alpha }alpha .

  2. At the distance PB the true angle is that which the tangent at point C forms with the x axis and the latter angle is clearly different from α{displaystyle alpha }alpha .


The angles that the q1 line and that axis form with the x axis become closer in value the closer one moves towards point P and become exactly equal at P.


Let point E be located very close to P, so close that the distance PE is infinitesimally small. Then PE measured on the q1 axis almost coincides with PE measured on the q1 line. At the same time, the ratio PD/PE (PD being the projection of PE on the x axis) becomes almost exactly equal to cos⁡α{displaystyle cos alpha }{displaystyle cos alpha }.


Let the infinitesimally small intercepts PD and PE be labelled, respectively, as dx and dq1. Then



cos⁡α=dxdq1=|e1||b1|{displaystyle cos alpha ={cfrac {dx}{dq^{1}}}={frac {|mathbf {e} _{1}|}{|mathbf {b} _{1}|}}}cos alpha = cfrac{dx}{dq^1} = frac{|mathbf{e}_1|}{|mathbf{b}_1|}.

Thus, the directional cosines can be substituted in transformations with the more exact ratios between infinitesimally small coordinate intercepts. It follows that the component (projection) of b1 on the x axis is



p1=b1⋅e1|e1|=|b1||e1||e1|cos⁡α=|b1|dxdq1⇒p1|b1|=dxdq1{displaystyle p^{1}=mathbf {b} _{1}cdot {cfrac {mathbf {e} _{1}}{|mathbf {e} _{1}|}}=|mathbf {b} _{1}|{cfrac {|mathbf {e} _{1}|}{|mathbf {e} _{1}|}}cos alpha =|mathbf {b} _{1}|{cfrac {dx}{dq^{1}}}quad Rightarrow quad {cfrac {p^{1}}{|mathbf {b} _{1}|}}={cfrac {dx}{dq^{1}}}}p^1 = mathbf{b}_1cdotcfrac{mathbf{e}_1}{|mathbf{e}_1|} = |mathbf{b}_1|cfrac{|mathbf{e}_1|}{|mathbf{e}_1|}cosalpha = |mathbf{b}_1|cfrac{dx}{dq^1} quad Rightarrow quad cfrac{p^1}{|mathbf{b}_1|} = cfrac{dx}{dq^1}.

If qi = qi(x1, x2, x3) and xi = xi(q1, q2, q3) are smooth (continuously differentiable) functions the transformation ratios can be written as qi∂xj{displaystyle {cfrac {partial q^{i}}{partial x_{j}}}}cfrac{partial q^i}{partial x_j} and xi∂qj{displaystyle {cfrac {partial x_{i}}{partial q^{j}}}}cfrac{partial x_i}{partial q^j}. That is, those ratios are partial derivatives of coordinates belonging to one system with respect to coordinates belonging to the other system.



Constructing a covariant basis in three dimensions


Doing the same for the coordinates in the other 2 dimensions, b1 can be expressed as:


b1=p1e1+p2e2+p3e3=∂x1∂q1e1+∂x2∂q1e2+∂x3∂q1e3{displaystyle mathbf {b} _{1}=p^{1}mathbf {e} _{1}+p^{2}mathbf {e} _{2}+p^{3}mathbf {e} _{3}={cfrac {partial x_{1}}{partial q^{1}}}mathbf {e} _{1}+{cfrac {partial x_{2}}{partial q^{1}}}mathbf {e} _{2}+{cfrac {partial x_{3}}{partial q^{1}}}mathbf {e} _{3}}<br />
mathbf{b}_1 = p^1mathbf{e}_1 + p^2mathbf{e}_2 + p^3mathbf{e}_3 = cfrac{partial x_1}{partial q^1} mathbf{e}_1 + cfrac{partial x_2}{partial q^1} mathbf{e}_2 + cfrac{partial x_3}{partial q^1} mathbf{e}_3<br />

Similar equations hold for b2 and b3 so that the standard basis {e1, e2, e3} is transformed to a local (ordered and normalised) basis {b1, b2, b3} by the following system of equations:


b1=∂x1∂q1e1+∂x2∂q1e2+∂x3∂q1e3b2=∂x1∂q2e1+∂x2∂q2e2+∂x3∂q2e3b3=∂x1∂q3e1+∂x2∂q3e2+∂x3∂q3e3{displaystyle {begin{aligned}mathbf {b} _{1}&={cfrac {partial x_{1}}{partial q^{1}}}mathbf {e} _{1}+{cfrac {partial x_{2}}{partial q^{1}}}mathbf {e} _{2}+{cfrac {partial x_{3}}{partial q^{1}}}mathbf {e} _{3}\mathbf {b} _{2}&={cfrac {partial x_{1}}{partial q^{2}}}mathbf {e} _{1}+{cfrac {partial x_{2}}{partial q^{2}}}mathbf {e} _{2}+{cfrac {partial x_{3}}{partial q^{2}}}mathbf {e} _{3}\mathbf {b} _{3}&={cfrac {partial x_{1}}{partial q^{3}}}mathbf {e} _{1}+{cfrac {partial x_{2}}{partial q^{3}}}mathbf {e} _{2}+{cfrac {partial x_{3}}{partial q^{3}}}mathbf {e} _{3}end{aligned}}}begin{align}<br />
   mathbf{b}_1 & = cfrac{partial x_1}{partial q^1} mathbf{e}_1 + cfrac{partial x_2}{partial q^1} mathbf{e}_2 + cfrac{partial x_3}{partial q^1} mathbf{e}_3 \<br />
   mathbf{b}_2 & = cfrac{partial x_1}{partial q^2} mathbf{e}_1 + cfrac{partial x_2}{partial q^2} mathbf{e}_2 + cfrac{partial x_3}{partial q^2} mathbf{e}_3 \<br />
   mathbf{b}_3 & = cfrac{partial x_1}{partial q^3} mathbf{e}_1 + cfrac{partial x_2}{partial q^3} mathbf{e}_2 + cfrac{partial x_3}{partial q^3} mathbf{e}_3<br />
end{align}

By analogous reasoning, one can obtain the inverse transformation from local basis to standard basis:


e1=∂q1∂x1b1+∂q2∂x1b2+∂q3∂x1b3e2=∂q1∂x2b1+∂q2∂x2b2+∂q3∂x2b3e3=∂q1∂x3b1+∂q2∂x3b2+∂q3∂x3b3{displaystyle {begin{aligned}mathbf {e} _{1}&={cfrac {partial q^{1}}{partial x_{1}}}mathbf {b} _{1}+{cfrac {partial q^{2}}{partial x_{1}}}mathbf {b} _{2}+{cfrac {partial q^{3}}{partial x_{1}}}mathbf {b} _{3}\mathbf {e} _{2}&={cfrac {partial q^{1}}{partial x_{2}}}mathbf {b} _{1}+{cfrac {partial q^{2}}{partial x_{2}}}mathbf {b} _{2}+{cfrac {partial q^{3}}{partial x_{2}}}mathbf {b} _{3}\mathbf {e} _{3}&={cfrac {partial q^{1}}{partial x_{3}}}mathbf {b} _{1}+{cfrac {partial q^{2}}{partial x_{3}}}mathbf {b} _{2}+{cfrac {partial q^{3}}{partial x_{3}}}mathbf {b} _{3}end{aligned}}}begin{align}<br />
   mathbf{e}_1 & = cfrac{partial q^1}{partial x_1} mathbf{b}_1 + cfrac{partial q^2}{partial x_1} mathbf{b}_2 + cfrac{partial q^3}{partial x_1} mathbf{b}_3 \<br />
   mathbf{e}_2 & = cfrac{partial q^1}{partial x_2} mathbf{b}_1 + cfrac{partial q^2}{partial x_2} mathbf{b}_2 + cfrac{partial q^3}{partial x_2} mathbf{b}_3 \<br />
   mathbf{e}_3 & = cfrac{partial q^1}{partial x_3} mathbf{b}_1 + cfrac{partial q^2}{partial x_3} mathbf{b}_2 + cfrac{partial q^3}{partial x_3} mathbf{b}_3<br />
end{align}


Jacobian of the transformation


The above systems of linear equations can be written in matrix form using the Einstein summation convention as



xi∂qkei=bk,∂qi∂xkbi=ek{displaystyle {cfrac {partial x_{i}}{partial q^{k}}}mathbf {e} _{i}=mathbf {b} _{k},quad {cfrac {partial q^{i}}{partial x_{k}}}mathbf {b} _{i}=mathbf {e} _{k}}cfrac{partial x_i}{partial q^k} mathbf{e}_i = mathbf{b}_k, quad cfrac{partial q^i}{partial x_k} mathbf{b}_i = mathbf{e}_k.

This coefficient matrix of the linear system is the Jacobian matrix (and its inverse) of the transformation. These are the equations that can be used to transform a Cartesian basis into a curvilinear basis, and vice versa.


In three dimensions, the expanded forms of these matrices are


J=[∂x1∂q1∂x1∂q2∂x1∂q3∂x2∂q1∂x2∂q2∂x2∂q3∂x3∂q1∂x3∂q2∂x3∂q3],J−1=[∂q1∂x1∂q1∂x2∂q1∂x3∂q2∂x1∂q2∂x2∂q2∂x3∂q3∂x1∂q3∂x2∂q3∂x3]{displaystyle mathbf {J} ={begin{bmatrix}{cfrac {partial x_{1}}{partial q^{1}}}&{cfrac {partial x_{1}}{partial q^{2}}}&{cfrac {partial x_{1}}{partial q^{3}}}\{cfrac {partial x_{2}}{partial q^{1}}}&{cfrac {partial x_{2}}{partial q^{2}}}&{cfrac {partial x_{2}}{partial q^{3}}}\{cfrac {partial x_{3}}{partial q^{1}}}&{cfrac {partial x_{3}}{partial q^{2}}}&{cfrac {partial x_{3}}{partial q^{3}}}\end{bmatrix}},quad mathbf {J} ^{-1}={begin{bmatrix}{cfrac {partial q^{1}}{partial x_{1}}}&{cfrac {partial q^{1}}{partial x_{2}}}&{cfrac {partial q^{1}}{partial x_{3}}}\{cfrac {partial q^{2}}{partial x_{1}}}&{cfrac {partial q^{2}}{partial x_{2}}}&{cfrac {partial q^{2}}{partial x_{3}}}\{cfrac {partial q^{3}}{partial x_{1}}}&{cfrac {partial q^{3}}{partial x_{2}}}&{cfrac {partial q^{3}}{partial x_{3}}}\end{bmatrix}}}<br />
mathbf{J} = begin{bmatrix}<br />
       cfrac{partial x_1}{partial q^1} & cfrac{partial x_1}{partial q^2} & cfrac{partial x_1}{partial q^3} \<br />
       cfrac{partial x_2}{partial q^1} & cfrac{partial x_2}{partial q^2} & cfrac{partial x_2}{partial q^3} \<br />
       cfrac{partial x_3}{partial q^1} & cfrac{partial x_3}{partial q^2} & cfrac{partial x_3}{partial q^3} \<br />
     end{bmatrix},quad<br />
mathbf{J}^{-1} = begin{bmatrix}<br />
       cfrac{partial q^1}{partial x_1} & cfrac{partial q^1}{partial x_2} & cfrac{partial q^1}{partial x_3} \<br />
       cfrac{partial q^2}{partial x_1} & cfrac{partial q^2}{partial x_2} & cfrac{partial q^2}{partial x_3} \<br />
       cfrac{partial q^3}{partial x_1} & cfrac{partial q^3}{partial x_2} & cfrac{partial q^3}{partial x_3} \<br />
     end{bmatrix}<br />

In the inverse transformation (second equation system), the unknowns are the curvilinear basis vectors. For any specific location there can only exist one and only one set of basis vectors (else the basis is not well defined at that point). This condition is satisfied if and only if the equation system has a single solution, from linear algebra, a linear equation system has a single solution (non-trivial) only if the determinant of its system matrix is non-zero:


det(J−1)≠0{displaystyle det(mathbf {J} ^{-1})neq 0} det(mathbf{J}^{-1})  neq 0

which shows the rationale behind the above requirement concerning the inverse Jacobian determinant.



Generalization to n dimensions


The formalism extends to any finite dimension as follows.


Consider the real Euclidean n-dimensional space, that is Rn = R × R × ... × R (n times) where R is the set of real numbers and × denotes the Cartesian product, which is a vector space.


The coordinates of this space can be denoted by: x = (x1, x2,...,xn). Since this is a vector (an element of the vector space), it can be written as:


x=∑i=1nxiei{displaystyle mathbf {x} =sum _{i=1}^{n}x_{i}mathbf {e} ^{i}} mathbf{x} = sum_{i=1}^n x_imathbf{e}^i

where e1 = (1,0,0...,0), e2 = (0,1,0...,0), e3 = (0,0,1...,0),...,en = (0,0,0...,1) is the standard basis set of vectors for the space Rn, and i = 1, 2,...n is an index labelling components. Each vector has exactly one component in each dimension (or "axis") and they are mutually orthogonal (perpendicular) and normalized (has unit magnitude).


More generally, we can define basis vectors bi so that they depend on q = (q1, q2,...,qn), i.e. they change from point to point: bi = bi(q). In which case to define the same point x in terms of this alternative basis: the coordinates with respect to this basis vi also necessarily depend on x also, that is vi = vi(x). Then a vector v in this space, with respect to these alternative coordinates and basis vectors, can be expanded as a linear combination in this basis (which simply means to multiply each basis vector ei by a number vi – scalar multiplication):


v=∑j=1nv¯jbj=∑j=1nv¯j(q)bj(q){displaystyle mathbf {v} =sum _{j=1}^{n}{bar {v}}^{j}mathbf {b} _{j}=sum _{j=1}^{n}{bar {v}}^{j}(mathbf {q} )mathbf {b} _{j}(mathbf {q} )} mathbf{v} = sum_{j=1}^n bar{v}^jmathbf{b}_j = sum_{j=1}^n bar{v}^j(mathbf{q})mathbf{b}_j(mathbf{q})

The vector sum that describes v in the new basis is composed of different vectors, although the sum itself remains the same.



Transformation of coordinates


From a more general and abstract perspective, a curvilinear coordinate system is simply a coordinate patch on the differentiable manifold En (n-dimensional Euclidean space) that is diffeomorphic to the Cartesian coordinate patch on the manifold.[3] Note that two diffeomorphic coordinate patches on a differential manifold need not overlap differentiably. With this simple definition of a curvilinear coordinate system, all the results that follow below are simply applications of standard theorems in differential topology.


The transformation functions are such that there's a one-to-one relationship between points in the "old" and "new" coordinates, that is, those functions are bijections, and fulfil the following requirements within their domains:



  1. They are smooth functions: qi = qi(x)

  2. The inverse Jacobian determinant
    J−1=|∂q1∂x1∂q1∂x2⋯q1∂xn∂q2∂x1∂q2∂x2⋯q2∂xn⋮qn∂x1∂qn∂x2⋯qn∂xn|≠0{displaystyle J^{-1}={begin{vmatrix}{dfrac {partial q^{1}}{partial x_{1}}}&{dfrac {partial q^{1}}{partial x_{2}}}&cdots &{dfrac {partial q^{1}}{partial x_{n}}}\{dfrac {partial q^{2}}{partial x_{1}}}&{dfrac {partial q^{2}}{partial x_{2}}}&cdots &{dfrac {partial q^{2}}{partial x_{n}}}\vdots &vdots &ddots &vdots \{dfrac {partial q^{n}}{partial x_{1}}}&{dfrac {partial q^{n}}{partial x_{2}}}&cdots &{dfrac {partial q^{n}}{partial x_{n}}}end{vmatrix}}neq 0} J^{-1}=begin{vmatrix}<br />
dfrac{partial q^1}{partial x_1} & dfrac{partial q^1}{partial x_2} & cdots & dfrac{partial q^1}{partial x_n} \<br />
dfrac{partial q^2}{partial x_1} & dfrac{partial q^2}{partial x_2} & cdots & dfrac{partial q^2}{partial x_n} \<br />
vdots & vdots & ddots & vdots \<br />
dfrac{partial q^n}{partial x_1} & dfrac{partial q^n}{partial x_2} & cdots & dfrac{partial q^n}{partial x_n}<br />
end{vmatrix} neq 0

    is not zero; meaning the transformation is invertible: xi(q).


    according to the inverse function theorem. The condition that the Jacobian determinant is not zero reflects the fact that three surfaces from different families intersect in one and only one point and thus determine the position of this point in a unique way.[4]



Vector and tensor algebra in three-dimensional curvilinear coordinates


Note: the Einstein summation convention of summing on repeated indices is used below.

Elementary vector and tensor algebra in curvilinear coordinates is used in some of the older scientific literature in mechanics and physics and can be indispensable to understanding work from the early and mid-1900s, for example the text by Green and Zerna.[5] Some useful relations in the algebra of vectors and second-order tensors in curvilinear coordinates are given in this section. The notation and contents are primarily from Ogden,[6] Naghdi,[7] Simmonds,[2] Green and Zerna,[5] Basar and Weichert,[8] and Ciarlet.[9]



Tensors in curvilinear coordinates



A second-order tensor can be expressed as


S=Sijbi⊗bj=Sijbi⊗bj=Sijbi⊗bj=Sijbi⊗bj{displaystyle {boldsymbol {S}}=S^{ij}mathbf {b} _{i}otimes mathbf {b} _{j}=S^{i}{}_{j}mathbf {b} _{i}otimes mathbf {b} ^{j}=S_{i}{}^{j}mathbf {b} ^{i}otimes mathbf {b} _{j}=S_{ij}mathbf {b} ^{i}otimes mathbf {b} ^{j}}<br />
   boldsymbol{S} = S^{ij}mathbf{b}_iotimesmathbf{b}_j = S^i{}_jmathbf{b}_iotimesmathbf{b}^j = S_i{}^jmathbf{b}^iotimesmathbf{b}_j = S_{ij}mathbf{b}^iotimesmathbf{b}^j<br />

where {displaystyle scriptstyle otimes }scriptstyleotimes denotes the tensor product. The components Sij are called the contravariant components, Sij the mixed right-covariant components, Sij the mixed left-covariant components, and Sij the covariant components of the second-order tensor. The components of the second-order tensor are related by


Sij=gikSkj=gjkSik=gikgjℓSkℓ{displaystyle S^{ij}=g^{ik}S_{k}{}^{j}=g^{jk}S^{i}{}_{k}=g^{ik}g^{jell }S_{kell }} S^{ij} = g^{ik}S_k{}^j = g^{jk}S^i{}_k = g^{ik}g^{jell}S_{kell}


The metric tensor in orthogonal curvilinear coordinates



At each point, one can construct a small line element dx, so the square of the length of the line element is the scalar product dx • dx and is called the metric of the space, given by:



dx⋅dx=∂xi∂qj∂xi∂qkdqjdqk{displaystyle dmathbf {x} cdot dmathbf {x} ={cfrac {partial x_{i}}{partial q^{j}}}{cfrac {partial x_{i}}{partial q^{k}}}dq^{j}dq^{k}}dmathbf{x}cdot dmathbf{x} = cfrac{partial x_i}{partial q^j}cfrac{partial x_i}{partial q^k}dq^jdq^k<br />
 .

The following portion of the above equation


xk∂qi∂xk∂qj=gij(qi,qj)=bi⋅bj{displaystyle {cfrac {partial x_{k}}{partial q^{i}}}{cfrac {partial x_{k}}{partial q^{j}}}=g_{ij}(q^{i},q^{j})=mathbf {b} _{i}cdot mathbf {b} _{j}}{cfrac  {partial x_{k}}{partial q^{i}}}{cfrac  {partial x_{k}}{partial q^{j}}}=g_{{ij}}(q^{i},q^{j})={mathbf  {b}}_{i}cdot {mathbf  {b}}_{j}

is a symmetric tensor called the fundamental (or metric) tensor of the Euclidean space in curvilinear coordinates.


Indices can be raised and lowered by the metric:


vi=gikvk{displaystyle v^{i}=g^{ik}v_{k}} v^i = g^{ik}v_k


Relation to Lamé coefficients


Defining the scale factors hi by


hihj=gij=bi⋅bj⇒hi=gii=|bi|=|∂x∂qi|{displaystyle h_{i}h_{j}=g_{ij}=mathbf {b} _{i}cdot mathbf {b} _{j}quad Rightarrow quad h_{i}={sqrt {g_{ii}}}=left|mathbf {b} _{i}right|=left|{cfrac {partial mathbf {x} }{partial q^{i}}}right|} h_ih_j = g_{ij} = mathbf{b}_icdotmathbf{b}_j quad Rightarrow quad h_i =sqrt{g_{ii}}= left|mathbf{b}_iright|=left|cfrac{partialmathbf{x}}{partial q^i}right|

gives a relation between the metric tensor and the Lamé coefficients. Note also that


gij=∂x∂qi⋅x∂qj=(hkiek)⋅(hmjem)=hkihkj{displaystyle g_{ij}={cfrac {partial mathbf {x} }{partial q^{i}}}cdot {cfrac {partial mathbf {x} }{partial q^{j}}}=left(h_{ki}mathbf {e} _{k}right)cdot left(h_{mj}mathbf {e} _{m}right)=h_{ki}h_{kj}} g_{ij} = cfrac{partialmathbf{x}}{partial q^i}cdotcfrac{partialmathbf{x}}{partial q^j}<br />
= left( h_{ki}mathbf{e}_kright)cdotleft( h_{mj}mathbf{e}_mright)<br />
= h_{ki}h_{kj}

where hij are the Lamé coefficients. For an orthogonal basis we also have:


g=g11g22g33=h12h22h32⇒g=h1h2h3=J{displaystyle g=g_{11}g_{22}g_{33}=h_{1}^{2}h_{2}^{2}h_{3}^{2}quad Rightarrow quad {sqrt {g}}=h_{1}h_{2}h_{3}=J} g = g_{11}g_{22}g_{33} = h_1^2h_2^2h_3^2 quad Rightarrow quad sqrt{g} = h_1h_2h_3 = J


Example: Polar coordinates


If we consider polar coordinates for R2, note that


(x,y)=(rcos⁡θ,rsin⁡θ){displaystyle (x,y)=(rcos theta ,rsin theta )}{displaystyle (x,y)=(rcos theta ,rsin theta )}

(r, θ) are the curvilinear coordinates, and the Jacobian determinant of the transformation (r,θ) → (r cos θ, r sin θ) is r.


The orthogonal basis vectors are br = (cos θ, sin θ), bθ = (−sin θ, cos θ). The scale factors are hr = 1 and hθ= r. The fundamental tensor is g11 =1, g22 =r2, g12 = g21 =0.



The alternating tensor


In an orthonormal right-handed basis, the third-order alternating tensor is defined as


E=εijkei⊗ej⊗ek{displaystyle {boldsymbol {mathcal {E}}}=varepsilon _{ijk}mathbf {e} ^{i}otimes mathbf {e} ^{j}otimes mathbf {e} ^{k}}{displaystyle {boldsymbol {mathcal {E}}}=varepsilon _{ijk}mathbf {e} ^{i}otimes mathbf {e} ^{j}otimes mathbf {e} ^{k}}

In a general curvilinear basis the same tensor may be expressed as


E=Eijkbi⊗bj⊗bk=Eijkbi⊗bj⊗bk{displaystyle {boldsymbol {mathcal {E}}}={mathcal {E}}_{ijk}mathbf {b} ^{i}otimes mathbf {b} ^{j}otimes mathbf {b} ^{k}={mathcal {E}}^{ijk}mathbf {b} _{i}otimes mathbf {b} _{j}otimes mathbf {b} _{k}}<br />
  boldsymbol{mathcal{E}} = mathcal{E}_{ijk}mathbf{b}^iotimesmathbf{b}^jotimesmathbf{b}^k<br />
   = mathcal{E}^{ijk}mathbf{b}_iotimesmathbf{b}_jotimesmathbf{b}_k<br />

It can also be shown that


Eijk=1Jεijk=1+gεijk{displaystyle {mathcal {E}}^{ijk}={cfrac {1}{J}}varepsilon _{ijk}={cfrac {1}{+{sqrt {g}}}}varepsilon _{ijk}}{displaystyle {mathcal {E}}^{ijk}={cfrac {1}{J}}varepsilon _{ijk}={cfrac {1}{+{sqrt {g}}}}varepsilon _{ijk}}


Christoffel symbols




Christoffel symbols of the first kind



bi,j=∂bi∂qj=Γijkbk⇒bi,j⋅bk=Γijk{displaystyle mathbf {b} _{i,j}={frac {partial mathbf {b} _{i}}{partial q^{j}}}=Gamma _{ijk}mathbf {b} ^{k}quad Rightarrow quad mathbf {b} _{i,j}cdot mathbf {b} _{k}=Gamma _{ijk}}<br />
mathbf{b}_{i,j} = frac{partial mathbf{b}_i}{partial q^j} = Gamma_{ijk}mathbf{b}^k quad Rightarrow quad<br />
mathbf{b}_{i,j} cdot mathbf{b}_k = Gamma_{ijk}<br />

where the comma denotes a partial derivative (see Ricci calculus). To express Γijk in terms of gij we note that


gij,k=(bi⋅bj),k=bi,k⋅bj+bi⋅bj,k=Γikj+Γjkigik,j=(bi⋅bk),j=bi,j⋅bk+bi⋅bk,j=Γijk+Γkjigjk,i=(bj⋅bk),i=bj,i⋅bk+bj⋅bk,i=Γjik+Γkij{displaystyle {begin{aligned}g_{ij,k}&=(mathbf {b} _{i}cdot mathbf {b} _{j})_{,k}=mathbf {b} _{i,k}cdot mathbf {b} _{j}+mathbf {b} _{i}cdot mathbf {b} _{j,k}=Gamma _{ikj}+Gamma _{jki}\g_{ik,j}&=(mathbf {b} _{i}cdot mathbf {b} _{k})_{,j}=mathbf {b} _{i,j}cdot mathbf {b} _{k}+mathbf {b} _{i}cdot mathbf {b} _{k,j}=Gamma _{ijk}+Gamma _{kji}\g_{jk,i}&=(mathbf {b} _{j}cdot mathbf {b} _{k})_{,i}=mathbf {b} _{j,i}cdot mathbf {b} _{k}+mathbf {b} _{j}cdot mathbf {b} _{k,i}=Gamma _{jik}+Gamma _{kij}end{aligned}}}<br />
begin{align}<br />
g_{ij,k} & = (mathbf{b}_icdotmathbf{b}_j)_{,k} = mathbf{b}_{i,k}cdotmathbf{b}_j + mathbf{b}_icdotmathbf{b}_{j,k}<br />
= Gamma_{ikj} + Gamma_{jki}\<br />
g_{ik,j} & = (mathbf{b}_icdotmathbf{b}_k)_{,j} = mathbf{b}_{i,j}cdotmathbf{b}_k + mathbf{b}_icdotmathbf{b}_{k,j}<br />
= Gamma_{ijk} + Gamma_{kji}\<br />
g_{jk,i} & = (mathbf{b}_jcdotmathbf{b}_k)_{,i} = mathbf{b}_{j,i}cdotmathbf{b}_k + mathbf{b}_jcdotmathbf{b}_{k,i}<br />
= Gamma_{jik} + Gamma_{kij}<br />
end{align}<br />

Since


bi,j=bj,i⇒Γijk=Γjik{displaystyle mathbf {b} _{i,j}=mathbf {b} _{j,i}quad Rightarrow quad Gamma _{ijk}=Gamma _{jik}}mathbf{b}_{i,j} = mathbf{b}_{j,i}quadRightarrowquadGamma_{ijk} = Gamma_{jik}

using these to rearrange the above relations gives


Γijk=12(gik,j+gjk,i−gij,k)=12[(bi⋅bk),j+(bj⋅bk),i−(bi⋅bj),k]{displaystyle Gamma _{ijk}={frac {1}{2}}(g_{ik,j}+g_{jk,i}-g_{ij,k})={frac {1}{2}}[(mathbf {b} _{i}cdot mathbf {b} _{k})_{,j}+(mathbf {b} _{j}cdot mathbf {b} _{k})_{,i}-(mathbf {b} _{i}cdot mathbf {b} _{j})_{,k}]}Gamma_{ijk} = frac{1}{2}(g_{ik,j} + g_{jk,i} - g_{ij,k}) = frac{1}{2}[(mathbf{b}_icdotmathbf{b}_k)_{,j} + (mathbf{b}_jcdotmathbf{b}_k)_{,i} - (mathbf{b}_icdotmathbf{b}_j)_{,k}]<br />



Christoffel symbols of the second kind



Γijk=Γjik,∂bi∂qj=Γijkbk{displaystyle Gamma _{ij}{}^{k}=Gamma _{ji}{}^{k},quad {cfrac {partial mathbf {b} _{i}}{partial q^{j}}}=Gamma _{ij}{}^{k}mathbf {b} _{k}}Gamma_{ij}{}^k = Gamma_{ji}{}^k,quad cfrac{partial mathbf{b}_i}{partial q^j} = Gamma_{ij}{}^kmathbf{b}_k

This implies that


Γijk=∂bi∂qj⋅bk=−bi⋅bk∂qj{displaystyle Gamma _{ij}{}^{k}={cfrac {partial mathbf {b} _{i}}{partial q^{j}}}cdot mathbf {b} ^{k}=-mathbf {b} _{i}cdot {cfrac {partial mathbf {b} ^{k}}{partial q^{j}}}} Gamma_{ij}{}^k = cfrac{partial mathbf{b}_i}{partial q^j}cdotmathbf{b}^k = -mathbf{b}_icdotcfrac{partial mathbf{b}^k}{partial q^j}

Other relations that follow are


bi∂qj=−Γijkbk,∇bi=Γijkbk⊗bj,∇bi=−Γjkibk⊗bj{displaystyle {cfrac {partial mathbf {b} ^{i}}{partial q^{j}}}=-Gamma ^{i}{}_{jk}mathbf {b} ^{k},quad {boldsymbol {nabla }}mathbf {b} _{i}=Gamma _{ij}{}^{k}mathbf {b} _{k}otimes mathbf {b} ^{j},quad {boldsymbol {nabla }}mathbf {b} ^{i}=-Gamma _{jk}{}^{i}mathbf {b} ^{k}otimes mathbf {b} ^{j}}<br />
cfrac{partial mathbf{b}^i}{partial q^j} = -Gamma^i{}_{jk}mathbf{b}^k,quad<br />
boldsymbol{nabla}mathbf{b}_i = Gamma_{ij}{}^kmathbf{b}_kotimesmathbf{b}^j,quad<br />
boldsymbol{nabla}mathbf{b}^i = -Gamma_{jk}{}^imathbf{b}^kotimesmathbf{b}^j<br />


Vector operations




  1. Dot product:

    The scalar product of two vectors in curvilinear coordinates is[2](p32)


    u⋅v=uivi=uivi=gijuivj=gijuivj{displaystyle mathbf {u} cdot mathbf {v} =u^{i}v_{i}=u_{i}v^{i}=g_{ij}u^{i}v^{j}=g^{ij}u_{i}v_{j}}<br />
   mathbf{u}cdotmathbf{v} = u^iv_i = u_iv^i = g_{ij}u^iv^j = g^{ij}u_iv_j<br />



  2. Cross product:

    The cross product of two vectors is given by[2](pp32–34)


    v=ϵijkujvkei{displaystyle mathbf {u} times mathbf {v} =epsilon _{ijk}{u}_{j}{v}_{k}mathbf {e} _{i}}<br />
  mathbf{u}timesmathbf{v} = epsilon_{ijk}{u}_j{v}_kmathbf{e}_i<br />

    where ϵijk{displaystyle epsilon _{ijk}}epsilon _{ijk} is the permutation symbol and ei{displaystyle mathbf {e} _{i}}mathbf {e} _{i} is a Cartesian basis vector. In curvilinear coordinates, the equivalent expression is


    v=[(bm×bn)⋅bs]umvnbs=Esmnumvnbs{displaystyle mathbf {u} times mathbf {v} =[(mathbf {b} _{m}times mathbf {b} _{n})cdot mathbf {b} _{s}]u^{m}v^{n}mathbf {b} ^{s}={mathcal {E}}_{smn}u^{m}v^{n}mathbf {b} ^{s}}<br />
  mathbf{u}timesmathbf{v} = [(mathbf{b}_mtimesmathbf{b}_n)cdotmathbf{b}_s]u^mv^nmathbf{b}^s<br />
    = mathcal{E}_{smn}u^mv^nmathbf{b}^s<br />

    where Eijk{displaystyle {mathcal {E}}_{ijk}}mathcal{E}_{ijk} is the third-order alternating tensor.



Vector and tensor calculus in three-dimensional curvilinear coordinates


Note: the Einstein summation convention of summing on repeated indices is used below.

Adjustments need to be made in the calculation of line, surface and volume integrals. For simplicity, the following restricts to three dimensions and orthogonal curvilinear coordinates. However, the same arguments apply for n-dimensional spaces. When the coordinate system is not orthogonal, there are some additional terms in the expressions.


Simmonds,[2] in his book on tensor analysis, quotes Albert Einstein saying[10]



The magic of this theory will hardly fail to impose itself on anybody who has truly understood it; it represents a genuine triumph of the method of absolute differential calculus, founded by Gauss, Riemann, Ricci, and Levi-Civita.



Vector and tensor calculus in general curvilinear coordinates is used in tensor analysis on four-dimensional curvilinear manifolds in general relativity,[11] in the mechanics of curved shells,[9] in examining the invariance properties of Maxwell's equations which has been of interest in metamaterials[12][13] and in many other fields.


Some useful relations in the calculus of vectors and second-order tensors in curvilinear coordinates are given in this section. The notation and contents are primarily from Ogden,[14] Simmonds,[2] Green and Zerna,[5] Basar and Weichert,[8] and Ciarlet.[9]


Let φ = φ(x) be a well defined scalar field and v = v(x) a well-defined vector field, and λ1, λ2... be parameters of the coordinates



Geometric elements




  1. Tangent vector: If x(λ) parametrizes a curve C in Cartesian coordinates, then
    x∂λ=∂x∂qi∂qi∂λ=(hki∂qi∂λ)bk{displaystyle {partial mathbf {x} over partial lambda }={partial mathbf {x} over partial q^{i}}{partial q^{i} over partial lambda }=left(h_{ki}{cfrac {partial q^{i}}{partial lambda }}right)mathbf {b} _{k}} {partial mathbf{x} over partial lambda} = {partial mathbf{x} over partial q^i}{partial q^i over partial lambda} = left( h_{ki}cfrac{partial q^i}{partial lambda}right)mathbf{b}_k

    is a tangent vector to C in curvilinear coordinates (using the chain rule). Using the definition of the Lamé coefficients, and that for the metric gij = 0 when ij, the magnitude is:


    |∂x∂λ|=hkihkj∂qi∂λqj∂λ=gij∂qi∂λqj∂λ=hi2(∂qi∂λ)2{displaystyle left|{partial mathbf {x} over partial lambda }right|={sqrt {h_{ki}h_{kj}{cfrac {partial q^{i}}{partial lambda }}{cfrac {partial q^{j}}{partial lambda }}}}={sqrt {g_{ij}{cfrac {partial q^{i}}{partial lambda }}{cfrac {partial q^{j}}{partial lambda }}}}={sqrt {h_{i}^{2}left({cfrac {partial q^{i}}{partial lambda }}right)^{2}}}} left|{partial mathbf{x} over partial lambda} right| = sqrt{h_{ki}h_{kj}cfrac{partial q^i}{partial lambda}cfrac{partial q^j}{partial lambda}} = sqrt{ g_{ij}cfrac{partial q^i}{partial lambda}cfrac{partial q^j}{partial lambda}} = sqrt{h_{i}^2left(cfrac{partial q^i}{partial lambda}right)^2}



  2. Tangent plane element: If x(λ1, λ2) parametrizes a surface S in Cartesian coordinates, then the following cross product of tangent vectors is a normal vector to S with the magnitude of infinitesimal plane element, in curvilinear coordinates. Using the above result,
    x∂λx∂λ2=(∂x∂qi∂qi∂λ1)×(∂x∂qj∂qj∂λ2)=Ekmp(hki∂qi∂λ1)(hmj∂qj∂λ2)bp{displaystyle {partial mathbf {x} over partial lambda _{1}}times {partial mathbf {x} over partial lambda _{2}}=left({partial mathbf {x} over partial q^{i}}{partial q^{i} over partial lambda _{1}}right)times left({partial mathbf {x} over partial q^{j}}{partial q^{j} over partial lambda _{2}}right)={mathcal {E}}_{kmp}left(h_{ki}{partial q^{i} over partial lambda _{1}}right)left(h_{mj}{partial q^{j} over partial lambda _{2}}right)mathbf {b} _{p}} {partial mathbf{x} over partial lambda_1}times {partial mathbf{x} over partial lambda_2} =left({partial mathbf{x} over partial q^i}{partial q^i over partial lambda_1}right) times left({partial mathbf{x} over partial q^j}{partial q^j over partial lambda_2}right) = mathcal{E}_{kmp}left( h_{ki}{partial q^i over partial lambda_1}right)left(h_{mj}{partial q^j over partial lambda_2}right) mathbf{b}_p

    where E{displaystyle {mathcal {E}}}{mathcal {E}} is the permutation symbol. In determinant form:


    x∂λx∂λ2=|e1e2e3h1i∂qi∂λ1h2i∂qi∂λ1h3i∂qi∂λ1h1j∂qj∂λ2h2j∂qj∂λ2h3j∂qj∂λ2|{displaystyle {partial mathbf {x} over partial lambda _{1}}times {partial mathbf {x} over partial lambda _{2}}={begin{vmatrix}mathbf {e} _{1}&mathbf {e} _{2}&mathbf {e} _{3}\h_{1i}{dfrac {partial q^{i}}{partial lambda _{1}}}&h_{2i}{dfrac {partial q^{i}}{partial lambda _{1}}}&h_{3i}{dfrac {partial q^{i}}{partial lambda _{1}}}\h_{1j}{dfrac {partial q^{j}}{partial lambda _{2}}}&h_{2j}{dfrac {partial q^{j}}{partial lambda _{2}}}&h_{3j}{dfrac {partial q^{j}}{partial lambda _{2}}}end{vmatrix}}}{partial mathbf{x} over partial lambda_1}times {partial mathbf{x} over partial lambda_2}<br />
=begin{vmatrix}<br />
mathbf{e}_1 & mathbf{e}_2 & mathbf{e}_3 \<br />
h_{1i} dfrac{partial q^i}{partial lambda_1} & h_{2i} dfrac{partial q^i}{partial lambda_1} & h_{3i} dfrac{partial q^i }{partial lambda_1} \<br />
h_{1j} dfrac{partial q^j}{partial lambda_2} & h_{2j} dfrac{partial q^j}{partial lambda_2} & h_{3j} dfrac{partial q^j }{partial lambda_2}<br />
end{vmatrix}




Integration























Operator
Scalar field
Vector field

Line integral

(x)ds=∫abφ(x(λ))|∂x∂λ|dλ{displaystyle int _{C}varphi (mathbf {x} )ds=int _{a}^{b}varphi (mathbf {x} (lambda ))left|{partial mathbf {x} over partial lambda }right|dlambda } int_C varphi(mathbf{x}) ds = int_a^b varphi(mathbf{x}(lambda))left|{partial mathbf{x} over partial lambda}right| dlambda

Cv(x)⋅ds=∫abv(x(λ))⋅(∂x∂λ)dλ{displaystyle int _{C}mathbf {v} (mathbf {x} )cdot dmathbf {s} =int _{a}^{b}mathbf {v} (mathbf {x} (lambda ))cdot left({partial mathbf {x} over partial lambda }right)dlambda } int_C mathbf{v}(mathbf{x}) cdot dmathbf{s} = int_a^b mathbf{v}(mathbf{x}(lambda))cdotleft({partial mathbf{x} over partial lambda}right) dlambda

Surface integral

(x)dS=∬(x(λ1,λ2))|∂x∂λx∂λ2|dλ1dλ2{displaystyle int _{S}varphi (mathbf {x} )dS=iint _{T}varphi (mathbf {x} (lambda _{1},lambda _{2}))left|{partial mathbf {x} over partial lambda _{1}}times {partial mathbf {x} over partial lambda _{2}}right|dlambda _{1}dlambda _{2}}int_S varphi(mathbf{x}) dS = iint_T varphi(mathbf{x}(lambda_1, lambda_2)) left|{partial mathbf{x} over partial lambda_1}times {partial mathbf{x} over partial lambda_2}right| dlambda_1 dlambda_2

Sv(x)⋅dS=∬Tv(x(λ1,λ2))⋅(∂x∂λx∂λ2)dλ1dλ2{displaystyle int _{S}mathbf {v} (mathbf {x} )cdot dS=iint _{T}mathbf {v} (mathbf {x} (lambda _{1},lambda _{2}))cdot left({partial mathbf {x} over partial lambda _{1}}times {partial mathbf {x} over partial lambda _{2}}right)dlambda _{1}dlambda _{2}}int_S mathbf{v}(mathbf{x}) cdot dS = iint_T mathbf{v}(mathbf{x}(lambda_1, lambda_2)) cdotleft({partial mathbf{x} over partial lambda_1}times {partial mathbf{x} over partial lambda_2}right) dlambda_1 dlambda_2

Volume integral

(x,y,z)dV=∭(q1,q2,q3)Jdq1dq2dq3{displaystyle iiint _{V}varphi (x,y,z)dV=iiint _{V}chi (q_{1},q_{2},q_{3})Jdq_{1}dq_{2}dq_{3}}iiint_V varphi(x,y,z) dV = iiint_V chi(q_1,q_2,q_3) Jdq_1dq_2dq_3

Vu(x,y,z)dV=∭Vv(q1,q2,q3)Jdq1dq2dq3{displaystyle iiint _{V}mathbf {u} (x,y,z)dV=iiint _{V}mathbf {v} (q_{1},q_{2},q_{3})Jdq_{1}dq_{2}dq_{3}}iiint_V mathbf{u}(x,y,z) dV = iiint_V mathbf{v}(q_1,q_2,q_3) Jdq_1dq_2dq_3


Differentiation


The expressions for the gradient, divergence, and Laplacian can be directly extended to n-dimensions, however the curl is only defined in 3d.


The vector field bi is tangent to the qi coordinate curve and forms a natural basis at each point on the curve. This basis, as discussed at the beginning of this article, is also called the covariant curvilinear basis. We can also define a reciprocal basis, or contravariant curvilinear basis, bi. All the algebraic relations between the basis vectors, as discussed in the section on tensor algebra, apply for the natural basis and its reciprocal at each point x.

































Operator
Scalar field
Vector field
2nd order tensor field

Gradient

φ=1hi∂φqibi{displaystyle nabla varphi ={cfrac {1}{h_{i}}}{partial varphi over partial q^{i}}mathbf {b} ^{i}} nablavarphi = cfrac{1}{h_i}{partialvarphi over partial q^i} mathbf{b}^i

v=1hi2∂v∂qi⊗bi{displaystyle nabla mathbf {v} ={cfrac {1}{h_{i}^{2}}}{partial mathbf {v} over partial q^{i}}otimes mathbf {b} _{i}}nablamathbf{v} = cfrac{1}{h_i^2}{partial mathbf{v} over partial q^i}otimesmathbf{b}_i

S=∂S∂qi⊗bi{displaystyle {boldsymbol {nabla }}{boldsymbol {S}}={cfrac {partial {boldsymbol {S}}}{partial q^{i}}}otimes mathbf {b} ^{i}}boldsymbol{nabla}boldsymbol{S} = cfrac{partial boldsymbol{S}}{partial q^i}otimesmathbf{b}^i

Divergence
N/A

v=1∏jhj∂qi(vi∏j≠ihj){displaystyle nabla cdot mathbf {v} ={cfrac {1}{prod _{j}h_{j}}}{frac {partial }{partial q^{i}}}(v^{i}prod _{jneq i}h_{j})} nabla cdot mathbf{v} = cfrac{1}{prod_j h_j} frac{partial }{partial q^i}(v^iprod_{jne i} h_j)

(∇S)⋅a=∇(S⋅a){displaystyle ({boldsymbol {nabla }}cdot {boldsymbol {S}})cdot mathbf {a} ={boldsymbol {nabla }}cdot ({boldsymbol {S}}cdot mathbf {a} )}<br />
   (boldsymbol{nabla}cdotboldsymbol{S})cdotmathbf{a} = boldsymbol{nabla}cdot(boldsymbol{S}cdotmathbf{a})<br />

where a is an arbitrary constant vector.
In curvilinear coordinates,


S=[∂Sij∂qk−ΓkilSlj−ΓkjlSil]gikbj{displaystyle {boldsymbol {nabla }}cdot {boldsymbol {S}}=left[{cfrac {partial S_{ij}}{partial q^{k}}}-Gamma _{ki}^{l}S_{lj}-Gamma _{kj}^{l}S_{il}right]g^{ik}mathbf {b} ^{j}}boldsymbol{nabla}cdotboldsymbol{S} = left[cfrac{partial S_{ij}}{partial q^k} - Gamma^l_{ki}S_{lj} - Gamma^l_{kj}S_{il}right]g^{ik}mathbf{b}^j



Laplacian

=1∏jhj∂qi(∏jhjhi2∂φqi){displaystyle nabla ^{2}varphi ={cfrac {1}{prod _{j}h_{j}}}{frac {partial }{partial q^{i}}}left({cfrac {prod _{j}h_{j}}{h_{i}^{2}}}{frac {partial varphi }{partial q^{i}}}right)}<br />
nabla^2 varphi =  cfrac{1}{prod _j h_j}frac{partial }{partial q^i}left(cfrac{prod _j h_j}{h_i^2}frac{partial varphi}{partial q^i}right)<br />



Curl
N/A
For vector fields in 3d only,

×v=1h1h2h3eiϵijkhi∂(hkvk)∂qj{displaystyle nabla times mathbf {v} ={frac {1}{h_{1}h_{2}h_{3}}}mathbf {e} _{i}epsilon _{ijk}h_{i}{frac {partial (h_{k}v_{k})}{partial q^{j}}}} nablatimesmathbf{v} = frac{1}{h_1h_2h_3} mathbf{e}_i epsilon_{ijk} h_i frac{partial (h_k v_k)}{partial q^j}


where ϵijk{displaystyle epsilon _{ijk}}epsilon _{ijk} is the Levi-Civita symbol.



See Curl of a tensor field


Fictitious forces in general curvilinear coordinates


By definition, if a particle with no forces acting on it has its position expressed in an inertial coordinate system, (x1x2x3t), then there it will have no acceleration (d2xj/dt2 = 0).[15] In this context, a coordinate system can fail to be “inertial” either due to non-straight time axis or non-straight space axes (or both). In other words, the basis vectors of the coordinates may vary in time at fixed positions, or they may vary with position at fixed times, or both. When equations of motion are expressed in terms of any non-inertial coordinate system (in this sense), extra terms appear, called Christoffel symbols. Strictly speaking, these terms represent components of the absolute acceleration (in classical mechanics), but we may also choose to continue to regard d2xj/dt2 as the acceleration (as if the coordinates were inertial) and treat the extra terms as if they were forces, in which case they are called fictitious forces.[16] The component of any such fictitious force normal to the path of the particle and in the plane of the path’s curvature is then called centrifugal force.[17]


This more general context makes clear the correspondence between the concepts of centrifugal force in rotating coordinate systems and in stationary curvilinear coordinate systems. (Both of these concepts appear frequently in the literature.[18][19][20]) For a simple example, consider a particle of mass m moving in a circle of radius r with angular speed w relative to a system of polar coordinates rotating with angular speed W. The radial equation of motion is mr” = Fr + mr(w + W)2. Thus the centrifugal force is mr times the square of the absolute rotational speed A = w + W of the particle. If we choose a coordinate system rotating at the speed of the particle, then W = A and w = 0, in which case the centrifugal force is mrA2, whereas if we choose a stationary coordinate system we have W = 0 and w = A, in which case the centrifugal force is again mrA2. The reason for this equality of results is that in both cases the basis vectors at the particle’s location are changing in time in exactly the same way. Hence these are really just two different ways of describing exactly the same thing, one description being in terms of rotating coordinates and the other being in terms of stationary curvilinear coordinates, both of which are non-inertial according to the more abstract meaning of that term.


When describing general motion, the actual forces acting on a particle are often referred to the instantaneous osculating circle tangent to the path of motion, and this circle in the general case is not centered at a fixed location, and so the decomposition into centrifugal and Coriolis components is constantly changing. This is true regardless of whether the motion is described in terms of stationary or rotating coordinates.



See also



  • Covariance and contravariance

  • Introduction to the mathematics of general relativity

  • Orthogonal coordinates

  • Frenet–Serret formulas

  • Covariant derivative

  • Tensor derivative (continuum mechanics)

  • Curvilinear perspective

  • Del in cylindrical and spherical coordinates



References





  1. ^ J.A. Wheeler; C. Misner; K.S. Thorne (1973). Gravitation. W.H. Freeman & Co. ISBN 0-7167-0344-0..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output q{quotes:"""""""'""'"}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  2. ^ abcdef Simmonds, J. G. (1994). A brief on tensor analysis. Springer. ISBN 0-387-90639-8.


  3. ^ Boothby, W. M. (2002). An Introduction to Differential Manifolds and Riemannian Geometry (revised ed.). New York, NY: Academic Press.


  4. ^ McConnell, A. J. (1957). Application of Tensor Analysis. New York, NY: Dover Publications, Inc. Ch. 9, sec. 1. ISBN 0-486-60373-3.


  5. ^ abc Green, A. E.; Zerna, W. (1968). Theoretical Elasticity. Oxford University Press. ISBN 0-19-853486-8.


  6. ^ Ogden, R. W. (2000). Nonlinear elastic deformations. Dover.


  7. ^ Naghdi, P. M. (1972). "Theory of shells and plates". In S. Flügge. Handbook of Physics. VIa/2. pp. 425–640.


  8. ^ ab Basar, Y.; Weichert, D. (2000). Numerical continuum mechanics of solids: fundamental concepts and perspectives. Springer.


  9. ^ abc Ciarlet, P. G. (2000). Theory of Shells. 1. Elsevier Science.


  10. ^ Einstein, A. (1915). "Contribution to the Theory of General Relativity". In Laczos, C. The Einstein Decade. p. 213. ISBN 0-521-38105-3.


  11. ^ Misner, C. W.; Thorne, K. S.; Wheeler, J. A. (1973). Gravitation. W. H. Freeman and Co. ISBN 0-7167-0344-0.


  12. ^ Greenleaf, A.; Lassas, M.; Uhlmann, G. (2003). "Anisotropic conductivities that cannot be detected by EIT". Physiological measurement. 24 (2): 413–419. doi:10.1088/0967-3334/24/2/353. PMID 12812426.


  13. ^ Leonhardt, U.; Philbin, T.G. (2006). "General relativity in electrical engineering". New Journal of Physics. 8 (10): 247. doi:10.1088/1367-2630/8/10/247.


  14. ^ Ogden


  15. ^ Friedman, Michael (1989). The Foundations of Space–Time Theories. Princeton University Press. ISBN 0-691-07239-6.


  16. ^ Stommel, Henry M.; Moore, Dennis W. (1989). An Introduction to the Coriolis Force. Columbia University Press. ISBN 0-231-06636-8.


  17. ^ Beer; Johnston (1972). Statics and Dynamics (2nd ed.). McGraw–Hill. p. 485. ISBN 0-07-736650-6.


  18. ^ Hildebrand, Francis B. (1992). Methods of Applied Mathematics. Dover. p. 156. ISBN 0-13-579201-0.


  19. ^ McQuarrie, Donald Allan (2000). Statistical Mechanics. University Science Books. ISBN 0-06-044366-9.


  20. ^ Weber, Hans-Jurgen; Arfken, George Brown (2004). Essential Mathematical Methods for Physicists. Academic Press. p. 843. ISBN 0-12-059877-9.




Further reading


.mw-parser-output .refbegin{font-size:90%;margin-bottom:0.5em}.mw-parser-output .refbegin-hanging-indents>ul{list-style-type:none;margin-left:0}.mw-parser-output .refbegin-hanging-indents>ul>li,.mw-parser-output .refbegin-hanging-indents>dl>dd{margin-left:0;padding-left:3.2em;text-indent:-3.2em;list-style:none}.mw-parser-output .refbegin-100{font-size:100%}



  • Spiegel, M. R. (1959). Vector Analysis. New York: Schaum's Outline Series. ISBN 0-07-084378-3.


  • Arfken, George (1995). Mathematical Methods for Physicists. Academic Press. ISBN 0-12-059877-9.




External links



  • Planetmath.org Derivation of Unit vectors in curvilinear coordinates

  • MathWorld's page on Curvilinear Coordinates

  • Prof. R. Brannon's E-Book on Curvilinear Coordinates


  • Wikiversity:Introduction to Elasticity/Tensors#The divergence of a tensor field – Wikiversity, Introduction to Elasticity/Tensors.









Popular posts from this blog

Xamarin.iOS Cant Deploy on Iphone

Glorious Revolution

Dulmage-Mendelsohn matrix decomposition in Python