Fact-checked by Grok 2 weeks ago

Curvilinear coordinates

Curvilinear coordinates are a class of coordinate systems in Euclidean space where the coordinate curves are not necessarily straight lines, unlike Cartesian coordinates, and instead consist of intersecting families of curved surfaces that define the position of points through the values of constants on those surfaces. These systems are particularly useful for exploiting symmetries in physical or mathematical problems, such as rotational invariance, by aligning the coordinates with the geometry of the domain. Common examples include cylindrical coordinates, which extend polar coordinates into three dimensions using radial distance r, azimuthal angle \phi, and height z, and spherical coordinates, which use radial distance r, polar angle \theta, and azimuthal angle \phi. A key subclass is orthogonal curvilinear coordinates, where the coordinate directions are mutually perpendicular at every point, forming an orthonormal basis that simplifies vector operations and differential equations. In these systems, scale factors h_i account for the variation in metric along each coordinate direction, defined as the ratio of infinitesimal arc length to coordinate differential (h_i = ds/du_i), and are essential for expressing gradients, divergences, and curls in a form analogous to Cartesian coordinates. Rectangular coordinates serve as a special case where scale factors are unity. Curvilinear coordinates find widespread application in fields like electromagnetism, fluid dynamics, and quantum mechanics, where they facilitate the solution of partial differential equations by matching boundary conditions to coordinate surfaces, such as spheres or cylinders. For instance, in cylindrical coordinates, the position is given by x = r \cos \phi, y = r \sin \phi, z = z, while in spherical coordinates, x = r \sin \theta \cos \phi, y = r \sin \theta \sin \phi, z = r \cos \theta, with conventions varying slightly between physics and mathematics contexts. These transformations enable more intuitive formulations of problems with inherent symmetries, reducing computational complexity.

Introduction and Fundamentals

Definition and Motivation

Curvilinear coordinates provide a generalization of the standard Cartesian coordinate system, where the loci of constant coordinate values form families of curved surfaces rather than flat planes. In three dimensions, these coordinates are typically denoted by variables u, v, w, which parameterize points in space through a smooth position vector \mathbf{r}(u, v, w) that maps to the corresponding Cartesian coordinates x, y, z. This mapping assumes an injective and differentiable transformation from the coordinate domain to Euclidean space, allowing for flexible descriptions of geometric configurations that deviate from rectilinear grids. Readers are presumed to be acquainted with Cartesian coordinates and basic vector notation, as curvilinear systems build directly upon these foundations. The primary motivation for employing curvilinear coordinates arises in physical and mathematical problems exhibiting inherent symmetries, such as cylindrical or spherical geometries, where Cartesian coordinates lead to cumbersome expressions and inefficient computations. By aligning the coordinate surfaces with the problem's symmetry, curvilinear systems simplify the formulation of governing equations, particularly partial differential equations (PDEs) like those in electrostatics, fluid dynamics, and heat conduction. For instance, in domains with rotational invariance, these coordinates facilitate the separation of variables technique, transforming multidimensional PDEs into a set of ordinary differential equations (ODEs) that are solvable along individual coordinate directions, thereby reducing computational complexity and enhancing analytical tractability. Common examples include polar coordinates in two dimensions and spherical coordinates in three dimensions, which exploit such symmetries effectively. Historically, the of curvilinear coordinates emerged in the amid advances in , where solving \nabla^2 \phi = 0 for gravitational and electrostatic potentials required tools beyond Cartesian frameworks. introduced the equation in his 1822 treatise on to model fluid equilibrium, but it was Lamé who, in the , coined the "curvilinear coordinates" and applied them to transform into separable forms, such as , for and elasticity problems. Building on earlier work by on surface , Lamé's contributions provided a powerful to handle curved geometries in , influencing subsequent developments in vector calculus and differential . These innovations remain foundational for modern applications in physics and engineering.

Basic Coordinate Transformations

Curvilinear coordinates provide a framework for describing points in space that aligns with the symmetries inherent in many physical problems, such as those involving cylindrical or spherical geometries. The fundamental setup for transforming from Cartesian coordinates (x, y, z) to general curvilinear coordinates (u, v, w) involves expressing the position vector \mathbf{r} as a function of the new coordinates. Specifically, the position vector is given by \mathbf{r}(u, v, w) = x(u, v, w) \mathbf{i} + y(u, v, w) \mathbf{j} + z(u, v, w) \mathbf{k}, where \mathbf{i}, \mathbf{j}, and \mathbf{k} are the standard Cartesian unit vectors. This mapping defines the curvilinear system, with coordinate surfaces corresponding to level sets such as u = \text{constant}, v = \text{constant}, and w = \text{constant}, which intersect to form a curvilinear grid in space. At any point in space, the tangent vectors to the coordinate curves provide a natural basis for the tangent space. These are the partial derivatives of the position vector: \mathbf{e}_u = \frac{\partial \mathbf{r}}{\partial u}, \quad \mathbf{e}_v = \frac{\partial \mathbf{r}}{\partial v}, \quad \mathbf{e}_w = \frac{\partial \mathbf{r}}{\partial w}. These vectors point in the directions of increasing u, v, and w, respectively, while holding the other coordinates fixed, and they span the tangent space at that point. For the transformation to be well-defined and useful, the mapping from (u, v, w) to (x, y, z) must be a diffeomorphism, meaning it is smooth, bijective, and has a smooth inverse, ensuring a one-to-one correspondence between points in the coordinate domains locally. This invertibility requires that the Jacobian matrix of the transformation has a non-zero determinant at each point, guaranteeing the local existence of the inverse function. A simple illustration of this setup occurs in one dimension, where a curve is parameterized by arc length s along its path. The position vector \mathbf{r}(s) satisfies |\frac{d\mathbf{r}}{ds}| = 1, so the tangent vector \frac{d\mathbf{r}}{ds} is a unit vector pointing along the curve, demonstrating how the coordinate s directly measures distance without scaling factors in this basic case.

Orthogonal Curvilinear Coordinates

Systems in Two Dimensions

In two dimensions, orthogonal curvilinear coordinate systems provide a framework for describing points in the plane using curves that intersect at right angles, facilitating the solution of problems with specific geometric symmetries. One of the most fundamental systems is the polar coordinate system, which employs radial distance r \geq 0 and azimuthal angle \theta \in [0, 2\pi). The transformation to Cartesian coordinates is given by
x = r \cos \theta, \quad y = r \sin \theta,
with the inverse relations
r = \sqrt{x^2 + y^2}, \quad \theta = \atan2(y, x). The scale factors for this system are h_r = 1 and h_\theta = r. Geometrically, curves of constant r form concentric circles centered at the origin, while curves of constant \theta are straight rays emanating from the origin. This system is particularly useful for problems exhibiting rotational symmetry, such as wave propagation in circular domains or fluid flow around cylindrical objects.
Another important two-dimensional orthogonal system is the parabolic coordinate system, using parameters \sigma \geq 0 and \tau \geq 0. The transformation equations are
x = \sigma \tau, \quad y = \frac{1}{2} (\sigma^2 - \tau^2),
with inverse forms
\sigma = \sqrt{\sqrt{x^2 + y^2} + y}, \quad \tau = \sqrt{\sqrt{x^2 + y^2} - y}. The scale factors are identical for both coordinates: h_\sigma = h_\tau = \sqrt{\sigma^2 + \tau^2}. The coordinate curves consist of confocal parabolas: constant \sigma traces parabolas opening to the right with focus at the origin, while constant \tau traces parabolas opening to the left. These coordinates are applied in quantum mechanics to solve the Schrödinger equation for the hydrogen atom, where they separate variables in the presence of electric fields, as in the Stark effect.
The elliptic coordinate system, denoted by \mu \geq 0 and \nu \in [0, 2\pi), is defined relative to two foci separated by distance $2a > 0. The transformation is
x = a \cosh \mu \cos \nu, \quad y = a \sinh \mu \sin \nu.
The scale factors are h_\mu = h_\nu = a \sqrt{\sinh^2 \mu + \sin^2 \nu}. Constant \mu curves are confocal ellipses with foci at (\pm a, 0), and constant \nu curves are confocal hyperbolas sharing the same foci. This system aids in analyzing boundary value problems with elliptical geometries, such as electrostatic potentials around elliptic cylinders or Stokes flow in elliptic domains.
The following table compares these three common two-dimensional orthogonal curvilinear systems, highlighting their transformation equations and scale factors:
SystemCoordinatesTransformation to CartesianScale Factors
Polarr \geq 0, \theta \in [0, 2\pi)x = r \cos \theta, y = r \sin \thetah_r = 1, h_\theta = r
Parabolic\sigma \geq 0, \tau \geq 0x = \sigma \tau, y = \frac{1}{2} (\sigma^2 - \tau^2)h_\sigma = h_\tau = \sqrt{\sigma^2 + \tau^2}
Elliptic\mu \geq 0, \nu \in [0, 2\pi)x = a \cosh \mu \cos \nu, y = a \sinh \mu \sin \nuh_\mu = h_\nu = a \sqrt{\sinh^2 \mu + \sin^2 \nu}

Systems in Three Dimensions

In three-dimensional space, orthogonal curvilinear coordinate systems extend the utility of two-dimensional systems by incorporating a third coordinate to describe volumes, particularly those exhibiting cylindrical, spherical, or toroidal symmetries. These systems simplify the mathematical description of physical phenomena with inherent rotational or radial invariance, such as fluid flows around axes or gravitational fields from point sources. Building on polar coordinates in the plane, which use radial distance and azimuthal angle, three-dimensional extensions add a longitudinal or height coordinate to fill space comprehensively. Cylindrical coordinates (\rho, \phi, z) provide a natural framework for problems with axial symmetry, where \rho is the radial distance from the z-axis, \phi is the azimuthal angle in the xy-plane, and z is the height along the axis. The transformation to Cartesian coordinates is given by x = \rho \cos \phi, \quad y = \rho \sin \phi, \quad z = z, with inverse relations \rho = \sqrt{x^2 + y^2}, \phi = \atan2(y, x), and z = z. The scale factors, which determine the metric in these coordinates, are h_\rho = 1, h_\phi = \rho, and h_z = 1. These coordinates are particularly useful in fluid dynamics for modeling axisymmetric flows, such as accretion onto a central object or pipe flows, where the azimuthal uniformity reduces the problem to two effective dimensions. Spherical coordinates (r, \theta, \phi) are suited to scenarios with radial symmetry from a point origin, where r is the radial distance, \theta is the polar angle from the positive z-axis, and \phi is the azimuthal angle. The relations to Cartesian coordinates are x = r \sin \theta \cos \phi, \quad y = r \sin \theta \sin \phi, \quad z = r \cos \theta, with inverses r = \sqrt{x^2 + y^2 + z^2}, \theta = \arccos(z/r), and \phi = \atan2(y, x). The corresponding scale factors are h_r = 1, h_\theta = r, and h_\phi = r \sin \theta. In gravitation, these coordinates facilitate the analysis of radial fields, as exemplified by Newton's theorem, which equates the attraction of a uniform-density sphere to that of a point mass at its center, simplifying potential calculations for planetary or stellar models. Toroidal coordinates (\xi, \eta, \phi) describe geometries resembling rings or doughnuts, generated by rotating bipolar coordinates around an axis, with \xi as the toroidal (poloidal) angle, \eta as the hyperbolic parameter controlling distance from the ring, and \phi as the azimuthal angle; a scale parameter a sets the ring radius. The transformation equations are x = \frac{a \sinh \eta \cos \phi}{\cosh \eta - \cos \xi}, \quad y = \frac{a \sinh \eta \sin \phi}{\cosh \eta - \cos \xi}, \quad z = \frac{a \sin \xi}{\cosh \eta - \cos \xi}, where $0 \leq \xi < 2\pi, $0 < \eta < \infty, and $0 \leq \phi < 2\pi. The scale factors are h_\xi = a / (\cosh \eta - \cos \xi), h_\eta = a / (\cosh \eta - \cos \xi), and h_\phi = a \sinh \eta / (\cosh \eta - \cos \xi). These coordinates are applied in problems involving ring-like structures, such as electromagnetic fields around toroidal inductors or plasma equilibria in tokamaks.

Basis Vectors and Metric Tensor

Covariant and Contravariant Bases

In curvilinear coordinates, the covariant basis vectors are defined as the partial derivatives of the position vector \mathbf{r} with respect to the coordinate variables u^i, given by \mathbf{e}_i = \frac{\partial \mathbf{r}}{\partial u^i}. These vectors are tangent to the coordinate curves and are generally not normalized, spanning the tangent space at each point in the manifold. The inner product of the covariant basis vectors yields the components of the metric tensor, \mathbf{e}_i \cdot \mathbf{e}_j = g_{ij}, where g_{ij} describes the geometry induced by the coordinate system. The contravariant basis vectors \mathbf{e}^i are defined such that they form the dual basis, satisfying the reciprocity relation \mathbf{e}^i \cdot \mathbf{e}_j = \delta^i_j, where \delta^i_j is the Kronecker delta. These contravariant vectors can be expressed using the inverse metric as \mathbf{e}^i = g^{ik} \mathbf{e}_k, where g^{ik} is the inverse of the metric tensor g_{ij}. In one dimension, the covariant basis reduces to a single vector along the coordinate curve parameterized by arc length s, where \mathbf{e}_1 = \frac{d\mathbf{r}}{ds} represents the unit tangent vector to the curve. In three dimensions, the contravariant basis vectors are constructed from the volume spanned by the covariant basis, using the reciprocity relation \mathbf{e}^i = \frac{\mathbf{e}_j \times \mathbf{e}_k}{\mathbf{e}_i \cdot (\mathbf{e}_j \times \mathbf{e}_k)} for cyclic permutations of indices i, j, k = 1, 2, 3, where the denominator is the scalar triple product defining the local volume element. Unlike the constant orthonormal basis vectors in Cartesian coordinates, the covariant and contravariant bases in curvilinear coordinates vary with position, reflecting the changing geometry of the coordinate system. In the special case of orthogonal curvilinear coordinates, the bases align with the unit vectors along the coordinate directions.

Scale Factors and Lamé Coefficients

In orthogonal curvilinear coordinates, the scale factors h_i are defined as the magnitudes of the partial derivatives of the position vector \mathbf{r} with respect to the coordinate variables u^i, i.e., h_i = \left| \frac{\partial \mathbf{r}}{\partial u^i} \right| for i = 1, 2, 3. These factors account for the local stretching of the coordinate lines relative to Cartesian systems, arising from the orthogonality condition where the basis vectors are mutually perpendicular. The scale factors are equivalently known as Lamé coefficients H_i, named after the mathematician Gabriel Lamé, who introduced them in the context of curvilinear systems for elasticity and geometry problems. In this notation, H_i = h_i, and they directly relate to the infinitesimal arc length element via the line element squared: ds^2 = h_1^2 (du^1)^2 + h_2^2 (du^2)^2 + h_3^2 (du^3)^2, which follows from the differential position vector d\mathbf{r} = \sum_i h_i du^i \hat{\mathbf{e}}_i, where \hat{\mathbf{e}}_i are the orthonormal unit vectors along the coordinate directions. The unit vectors are obtained by normalizing the covariant basis vectors, such that \hat{\mathbf{e}}_i = \frac{1}{h_i} \frac{\partial \mathbf{r}}{\partial u^i}, providing a physical interpretation where the scale factors convert coordinate differentials into physical displacements. For computation, the h_i can be found using dot products of the partial derivatives; in orthogonal systems, the off-diagonal terms vanish, yielding h_i^2 = \frac{\partial \mathbf{r}}{\partial u^i} \cdot \frac{\partial \mathbf{r}}{\partial u^i}. As an example in spherical coordinates (r, \theta, \phi), where \mathbf{r} = r \hat{\mathbf{r}} with \hat{\mathbf{r}} = \sin\theta \cos\phi \, \hat{\mathbf{x}} + \sin\theta \sin\phi \, \hat{\mathbf{y}} + \cos\theta \, \hat{\mathbf{z}}, the scale factors are h_r = 1, h_\theta = r, and h_\phi = r \sin\theta, derived from the respective magnitudes \left| \frac{\partial \mathbf{r}}{\partial r} \right| = 1, \left| \frac{\partial \mathbf{r}}{\partial \theta} \right| = r, and \left| \frac{\partial \mathbf{r}}{\partial \phi} \right| = r \sin\theta. In the metric tensor for orthogonal coordinates, the diagonal components are g_{ii} = h_i^2 (no summation over i), with g_{ij} = 0 for i \neq j, linking the scale factors directly to the geometry of the coordinate system. This diagonal form simplifies vector operations while capturing the curvature effects inherent to the coordinates.

Vector Operations in Orthogonal Coordinates

Gradient and Directional Derivatives

In orthogonal curvilinear coordinates, the gradient of a scalar function f(u^1, u^2, u^3) is a vector that points in the direction of the steepest increase of f and whose magnitude equals the rate of change of f in that direction. This operator generalizes the Cartesian gradient to systems where the coordinate lines are curved but mutually orthogonal. The derivation begins with the differential change in f, given by the total differential: df = \sum_{i=1}^3 \frac{\partial f}{\partial u^i} \, du^i. By definition, this equals the dot product of the gradient with the infinitesimal displacement vector: df = \nabla f \cdot d\mathbf{r}. In orthogonal curvilinear coordinates, the displacement is d\mathbf{r} = \sum_{i=1}^3 h_i \, du^i \, \hat{\mathbf{e}}_i, where h_i are the scale factors and \hat{\mathbf{e}}_i are the unit basis vectors along the coordinate directions. Substituting and using the orthogonality of the basis vectors (\hat{\mathbf{e}}_i \cdot \hat{\mathbf{e}}_j = \delta_{ij}) yields the components of \nabla f by equating coefficients: \nabla f = \sum_{i=1}^3 \frac{1}{h_i} \frac{\partial f}{\partial u^i} \, \hat{\mathbf{e}}_i. The directional derivative of f along an arbitrary unit vector \hat{\mathbf{a}} is then the projection of the gradient onto that direction: \nabla f \cdot \hat{\mathbf{a}} = \sum_{i=1}^3 a_i \frac{1}{h_i} \frac{\partial f}{\partial u^i}, where a_i = \hat{\mathbf{a}} \cdot \hat{\mathbf{e}}_i are the components of \hat{\mathbf{a}} in the curvilinear basis; this gives the rate of change of f per unit distance along \hat{\mathbf{a}}. For example, in two-dimensional polar coordinates (r, \theta) with scale factors h_r = 1 and h_\theta = r, the gradient simplifies to \nabla f = \frac{\partial f}{\partial r} \, \hat{\mathbf{e}}_r + \frac{1}{r} \frac{\partial f}{\partial \theta} \, \hat{\mathbf{e}}_\theta.

Divergence and Curl

In orthogonal curvilinear coordinates (u^1, u^2, u^3) with scale factors h_1, h_2, h_3, the divergence of a vector field \mathbf{A} with physical components A_i = \mathbf{A} \cdot \hat{\mathbf{e}}_i (where \hat{\mathbf{e}}_i are the unit basis vectors) is given by \nabla \cdot \mathbf{A} = \frac{1}{h_1 h_2 h_3} \left[ \frac{\partial (A_1 h_2 h_3)}{\partial u^1} + \frac{\partial (A_2 h_1 h_3)}{\partial u^2} + \frac{\partial (A_3 h_1 h_2)}{\partial u^3} \right]. This expression arises from applying Gauss's divergence theorem to an infinitesimal volume element, where the net flux through the coordinate surfaces yields the partial derivatives scaled by the appropriate products of scale factors. The curl of \mathbf{A} has components (\nabla \times \mathbf{A})_1 = \frac{1}{h_2 h_3} \left[ \frac{\partial (A_3 h_3)}{\partial u^2} - \frac{\partial (A_2 h_2)}{\partial u^3} \right], with the other components obtained by cyclic permutation of the indices. These forms derive from Stokes's theorem applied to an infinitesimal area element, capturing the circulation around coordinate faces and incorporating the scale factors to account for the geometry. These operators are fundamental in fields like fluid dynamics, where divergence describes source terms in continuity equations, and electromagnetism, where curl governs rotational aspects of fields such as Faraday's law. A representative application appears in computing the magnetic field of an electric current dipole, modeled via the vector potential in spherical coordinates (r, \theta, \phi) with scale factors h_r = 1, h_\theta = r, h_\phi = r \sin \theta. The azimuthal vector potential is A_\phi = \frac{\mu_0 m \sin \theta}{4\pi r^2} for magnetic moment \mathbf{m} = m \hat{\mathbf{z}}, and the curl yields the dipole field components B_r = \frac{2\mu_0 m \cos \theta}{4\pi r^3} and B_\theta = \frac{\mu_0 m \sin \theta}{4\pi r^3}, with B_\phi = 0. The Laplacian of a scalar f, expressible as the divergence of the gradient, takes the form \nabla^2 f = \frac{1}{h_1 h_2 h_3} \sum_{i=1}^3 \frac{\partial}{\partial u^i} \left( \frac{h_1 h_2 h_3}{h_i^2} \frac{\partial f}{\partial u^i} \right).

Tensor Framework in Curvilinear Coordinates

Metric Tensor Properties

In curvilinear coordinates, the metric tensor g_{ij} is defined as the inner product of the covariant basis vectors \mathbf{e}_i = \frac{\partial \mathbf{r}}{\partial u^i} and \mathbf{e}_j = \frac{\partial \mathbf{r}}{\partial u^j}, yielding g_{ij} = \mathbf{e}_i \cdot \mathbf{e}_j = \frac{\partial \mathbf{r}}{\partial u^i} \cdot \frac{\partial \mathbf{r}}{\partial u^j}. This symmetric positive-definite tensor encapsulates the geometry of the coordinate system by determining infinitesimal distances through the line element ds^2 = g_{ij} \, du^i \, du^j, where summation over repeated indices is implied. The determinant of the metric tensor, g = \det(g_{ij}), equals the square of the scalar triple product of the basis vectors, g = \left[ \mathbf{e}_1 \cdot (\mathbf{e}_2 \times \mathbf{e}_3) \right]^2, which represents the squared volume of the parallelepiped spanned by \mathbf{e}_1, \mathbf{e}_2, \mathbf{e}_3. For orientable coordinate systems in Euclidean space, g > 0, ensuring the basis vectors form a right-handed frame and the coordinate transformation is locally invertible. The inverse metric tensor g^{ij} satisfies g^{ik} g_{kj} = \delta^i_j, where \delta^i_j is the Kronecker delta, and serves to raise indices on vectors and tensors; for instance, the contravariant components of a vector \mathbf{v} are obtained via v^i = g^{ij} v_j. In the special case of orthogonal curvilinear coordinates, the metric tensor is diagonal, with components g_{ii} = h_i^2 (no summation), where h_i are the scale factors along each coordinate direction. The metric tensor is invariant under changes of coordinates, as its components transform in a manner that preserves the underlying geometric structure of distances and angles, forming the foundation of Riemannian geometry for describing manifolds.

Christoffel Symbols

In the tensor framework of curvilinear coordinates, Christoffel symbols of the second kind, denoted \Gamma^k_{ij}, serve as the connection coefficients that account for the variation of basis vectors across coordinate patches, enabling covariant differentiation on manifolds equipped with a metric tensor. They are defined in terms of the metric tensor g_{ij} and its inverse g^{kl} as \Gamma^k_{ij} = \frac{1}{2} g^{kl} \left( \frac{\partial g_{lj}}{\partial u^i} + \frac{\partial g_{li}}{\partial u^j} - \frac{\partial g_{ij}}{\partial u^l} \right), where u^i are the curvilinear coordinates; this expression arises from requiring the connection to be metric-compatible and torsion-free, ensuring parallel transport preserves lengths and angles. A key property of these symbols is their symmetry in the lower indices, \Gamma^k_{ij} = \Gamma^k_{ji}, which follows directly from the symmetry of the metric tensor and the partial derivatives in the definition, reflecting the absence of torsion in the Levi-Civita connection. In Cartesian coordinates on flat Euclidean space, where the basis vectors are constant and the metric is independent of position, all Christoffel symbols vanish, \Gamma^k_{ij} = 0, simplifying tensor operations to ordinary partial derivatives. The Christoffel symbols facilitate the covariant derivative, which extends directional derivatives to curved spaces while maintaining tensorial character. For a contravariant vector field v^i, the covariant derivative along direction u^j is \nabla_j v^i = \frac{\partial v^i}{\partial u^j} + \Gamma^i_{jk} v^k, incorporating the connection to correct for basis changes; similarly, for a covariant vector (covector) w_i, \nabla_j w_i = \frac{\partial w_i}{\partial u^j} - \Gamma^k_{ji} w_k, where the minus sign arises from the transformation rules for lowered indices. In orthogonal curvilinear coordinates, where the metric tensor is diagonal, the Christoffel symbols simplify significantly, with many components vanishing; only terms involving derivatives of the diagonal scale factors remain non-zero. For example, in polar coordinates (r, \theta), the non-zero symbols include \Gamma^\theta_{r\theta} = \Gamma^\theta_{\theta r} = \frac{1}{r} and \Gamma^r_{\theta\theta} = -r, reflecting the radial dependence of the angular basis vector. Christoffel symbols play a central role in the geodesic equation, which describes the shortest paths (or extremal curves) on the manifold: for a curve parameterized by affine parameter s, \frac{d^2 u^k}{ds^2} + \Gamma^k_{ij} \frac{du^i}{ds} \frac{du^j}{ds} = 0, where the second term encodes the intrinsic geometry via the connection, generalizing straight-line motion in Euclidean space.

Calculus and Integration

Line, Surface, and Volume Elements

In orthogonal curvilinear coordinates (u^1, u^2, u^3), the infinitesimal displacement along the coordinate curve where only u^i varies is given by dl_i = h_i \, du^i, where h_i is the scale factor associated with the i-th coordinate. The general line element ds for an arbitrary displacement is then ds = \sqrt{\sum_{i=1}^3 (h_i \, du^i)^2} = \sqrt{g_{ij} \, du^i \, du^j}, where g_{ij} = h_i^2 \delta_{ij} is the diagonal metric tensor for the orthogonal system. For surface elements, consider a surface of constant u^3. The infinitesimal area vector is d\mathbf{S} = h_1 h_2 \, du^1 \, du^2 \, \mathbf{n}_3, where \mathbf{n}_3 is the unit normal in the direction of increasing u^3. Analogous expressions hold for surfaces of constant u^1 or u^2, with the appropriate scale factors and normals. This form arises from the magnitude of the cross product of the tangent vectors along the surface coordinates. The volume element in orthogonal curvilinear coordinates is dV = h_1 h_2 h_3 \, du^1 \, du^2 \, du^3, obtained as the scalar triple product of the infinitesimal displacement vectors along each coordinate direction. Consequently, the integral of a scalar function f over a volume becomes \int f \, dV = \int f(u^1, u^2, u^3) \, h_1 h_2 h_3 \, du^1 \, du^2 \, du^3. For flux integrals, \int \mathbf{A} \cdot d\mathbf{S} uses the vector form of the surface element. As an example, consider the flux of a radial vector field \mathbf{A} = A_r \hat{r} through a sphere of radius r in spherical coordinates, where the scale factors are h_r = 1, h_\theta = r, and h_\phi = r \sin \theta. The surface element is d\mathbf{S} = r^2 \sin \theta \, d\theta \, d\phi \, \hat{r}, so the flux is \int_0^{2\pi} \int_0^\pi A_r r^2 \sin \theta \, d\theta \, d\phi = 4\pi r^2 A_r.

Jacobian Determinant and Change of Variables

In the context of curvilinear coordinates, the Jacobian matrix facilitates the transformation between Cartesian coordinates \mathbf{x} = (x^1, x^2, x^3) and generalized curvilinear coordinates \mathbf{u} = (u^1, u^2, u^3), where \mathbf{x} = \mathbf{x}(\mathbf{u}). The components of the Jacobian matrix are given by J^i_j = \frac{\partial x^i}{\partial u^j}, representing the partial derivatives of the position vector components with respect to the curvilinear parameters. The determinant of this matrix, denoted J = \det(J^i_j) = \left| \frac{\partial(x^1, x^2, x^3)}{\partial(u^1, u^2, u^3)} \right|, quantifies the local scaling of volumes under the coordinate change and is essential for preserving the integrity of integrals. The absolute value of the Jacobian determinant accounts for orientation and ensures positive scaling factors in integration. In general curvilinear systems, this determinant equals \sqrt{|g|}, where g = \det(g_{ij}) is the determinant of the metric tensor with components g_{ij} = \frac{\partial \mathbf{r}}{\partial u^i} \cdot \frac{\partial \mathbf{r}}{\partial u^j}, linking the transformation to the geometry of the coordinate system. Computationally, for a position vector \mathbf{r}(u^1, u^2, u^3), the Jacobian determinant is the absolute value of the scalar triple product: J = \left| \frac{\partial \mathbf{r}}{\partial u^1} \cdot \left( \frac{\partial \mathbf{r}}{\partial u^2} \times \frac{\partial \mathbf{r}}{\partial u^3} \right) \right|, which directly evaluates the oriented volume spanned by the tangent vectors to the coordinate curves. The change of variables theorem in multiple integrals relies on the Jacobian to transform domains and integrands accordingly. For a region R in Cartesian coordinates and corresponding region S in curvilinear coordinates, the integral transforms as \int_R f(\mathbf{x}) \, d\mathbf{x} = \int_S f(\mathbf{x}(\mathbf{u})) \left| \det \left( \frac{\partial \mathbf{x}}{\partial \mathbf{u}} \right) \right| d\mathbf{u}, where the absolute value preserves the measure regardless of orientation. In orthogonal curvilinear coordinates, where the coordinate surfaces are mutually perpendicular, the Jacobian simplifies to the product of the scale factors: |J| = h_1 h_2 h_3, with h_i = \left| \frac{\partial \mathbf{r}}{\partial u^i} \right|. Lower-dimensional cases illustrate the Jacobian's role in scaling. In one dimension, the arc length element scales as ds = \left| \frac{dx}{du} \right| du, where \left| \frac{dx}{du} \right| is the 1D Jacobian. In two dimensions, the area element transforms via the 2D Jacobian determinant \left| \frac{\partial(x,y)}{\partial(u,v)} \right| du \, dv, which measures the parallelogram area formed by the partial derivatives \frac{\partial \mathbf{r}}{\partial u} and \frac{\partial \mathbf{r}}{\partial v}.

Generalizations

Non-Orthogonal Systems

In non-orthogonal curvilinear coordinate systems, also referred to as oblique systems, the coordinate axes are not mutually perpendicular, resulting in a metric tensor g_{ij} with non-zero off-diagonal elements that reflect the angles between the axes being unequal to 90 degrees. This generality allows for more flexible descriptions of complex geometries compared to orthogonal systems, though it introduces additional computational complexity in vector and tensor operations. The metric tensor is defined as g_{ij} = \mathbf{e}_i \cdot \mathbf{e}_j, where \mathbf{e}_i = \partial \mathbf{r} / \partial q^i are the covariant basis vectors tangent to the coordinate curves. The covariant basis vectors \mathbf{e}_i in non-orthogonal systems are not perpendicular, so vector decomposition requires both covariant components A_i (projections onto \mathbf{e}_i) and contravariant components A^i (projections onto the dual basis \mathbf{e}^i, satisfying \mathbf{e}^i \cdot \mathbf{e}_j = \delta^i_j). To interpret physical quantities intuitively, physical components A_{(i)} are often defined as A_{(i)} = A_i / \sqrt{g_{ii}} (no summation over i), representing the projection onto the unit vector tangent to the i-th coordinate line. This adjustment accounts for the varying lengths of the basis vectors but does not fully resolve the obliqueness, making cross terms in g_{ij} essential for accurate transformations. Challenges arise in expressing vector operations, as the non-perpendicularity complicates projections and requires the full inverse metric g^{ij} for raising indices. An illustrative example appears in geophysical modeling of Earth's gravity field within a rotating Earth model, where non-orthogonal curvilinear coordinates (such as (q, \chi, \nu)) are employed to capture the oblate spheroid shape and rotational effects, introducing cross terms in the metric to describe the reference model's parameters varying along one coordinate. These systems facilitate the analysis of normal modes and perturbations in the gravitational potential without assuming orthogonality. For integration, the volume element is dV = \sqrt{|g|} \, dq^1 dq^2 dq^3, where g = \det(g_{ij}), differing from the orthogonal case's simple product of scale factors. Christoffel symbols become more involved due to the off-diagonal metric components, affecting geodesic and derivative calculations.

Extension to n Dimensions

In n-dimensional Euclidean space \mathbb{R}^n, curvilinear coordinates are introduced via a differentiable position vector \mathbf{r}(u^1, \dots, u^n) that parametrizes points in the space, where u^i are the coordinate functions. The geometry is encoded in the metric tensor g_{ij}, defined by the inner product of partial derivatives: g_{ij} = \frac{\partial \mathbf{r}}{\partial u^i} \cdot \frac{\partial \mathbf{r}}{\partial u^j}, which determines distances, angles, and the overall structure of the coordinate system. This metric is symmetric and positive definite, allowing the coordinate system to describe arbitrary curved manifolds embedded in \mathbb{R}^n. The covariant basis vectors are given by \mathbf{e}_i = \frac{\partial \mathbf{r}}{\partial u^i}, which are tangent to the coordinate curves and span the tangent space at each point, forming a complete local frame for the n-dimensional space. These basis vectors may not be orthogonal or normalized, reflecting the potential non-orthogonality of the coordinates. The volume element in these coordinates, essential for integration, is dV = \sqrt{|\det g|} \, du^1 \dots du^n, where \det g is the determinant of the metric tensor; this generalizes the Jacobian factor from lower dimensions and ensures invariant measure under coordinate transformations. The gradient of a scalar function f(u^1, \dots, u^n) in this framework is expressed as \nabla f = g^{ij} \left( \frac{\partial f}{\partial u^j} \right) \mathbf{e}_i, where g^{ij} is the inverse metric tensor (satisfying g^{ik} g_{kj} = \delta^i_j) and Einstein summation convention is used over repeated indices. This contravariant form points in the direction of steepest ascent, weighted by the inverse metric to account for the coordinate distortion. In Riemannian manifolds, this extends naturally to curved spaces beyond flat Euclidean geometry. Such generalizations find application in higher-dimensional physics, notably in Kaluza-Klein theory, where curvilinear coordinates parametrize extra compact dimensions (e.g., a fifth dimension as a circle) to unify gravity and electromagnetism through the metric components of a (4+1)-dimensional spacetime. This approach, originally proposed by Theodor Kaluza in 1921, relies on the n-dimensional metric to derive lower-dimensional field equations, influencing modern string theory and extra-dimensional models.

Applications in Dynamics

Fictitious Forces

In non-inertial reference frames described by curvilinear coordinates, the equations of motion for a particle include additional terms known as fictitious forces, which account for the acceleration of the frame itself relative to an inertial frame. These forces arise when transforming Newton's second law from an inertial Cartesian system to a curvilinear system, particularly in cases like rotating frames where the coordinate basis vectors vary with position or time. For a particle of mass m, the effective force in the non-inertial frame can be expressed as \mathbf{F}_{\text{eff}} = \mathbf{F} - m \left( \mathbf{a}_{\text{frame}} + 2 \boldsymbol{\omega} \times \mathbf{v}_{\text{rel}} + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \dot{\boldsymbol{\omega}} \times \mathbf{r} \right), where \mathbf{a}_{\text{frame}} is the acceleration of the frame's origin, \boldsymbol{\omega} is the angular velocity, \mathbf{v}_{\text{rel}} is the relative velocity, \mathbf{r} is the position vector relative to the origin, and the dot denotes time derivative; this form holds generally but is evaluated in the curvilinear basis. In curvilinear coordinates, these fictitious forces can alternatively be derived using the Christoffel symbols of the second kind, \Gamma^i_{jk}, which capture the geometry of the coordinate system. The acceleration in coordinate components becomes \ddot{x}^i + \Gamma^i_{jk} \dot{x}^j \dot{x}^k, so the equation of motion is m (\ddot{x}^i + \Gamma^i_{jk} \dot{x}^j \dot{x}^k) = F^i, where the term m \Gamma^i_{jk} \dot{x}^j \dot{x}^k represents the fictitious force components arising from the curvature of the coordinates. Among these, the Coriolis force in a rotating frame is given by -2m \boldsymbol{\omega} \times \mathbf{v}_{\text{rel}}, which deflects moving particles perpendicular to both their velocity and the rotation axis, and the centrifugal force by -m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), which acts outward from the axis of rotation. These terms emerge naturally when the curvilinear coordinates align with the rotating geometry, such as in systems with azimuthal dependence. A representative example occurs in cylindrical coordinates (r, \theta, z) for a frame rotating with constant angular velocity \omega about the z-axis. The radial equation of motion includes terms like -r \dot{\theta}^2 + 2 \omega r \dot{\theta} + \omega^2 r, where -r \dot{\theta}^2 is the coordinate-curvature term, $2 \omega r \dot{\theta} corresponds to the Coriolis contribution (proportional to -2m \omega \times v), and \omega^2 r to the centrifugal term; the azimuthal equation features r \ddot{\theta} + 2 \dot{r} \dot{\theta}, incorporating Coriolis effects on angular motion. These azimuthal acceleration terms highlight how rotation couples the coordinate velocities, producing observable deflections in the rotating frame. Fictitious forces in curvilinear coordinates relate directly to the geodesic equation, which describes the straight-line paths (geodesics) in the coordinate manifold: \ddot{x}^i + \Gamma^i_{jk} \dot{x}^j \dot{x}^k = 0. In this context, the fictitious forces represent deviations from these geodesics when external forces are absent, interpreting the apparent acceleration as arising from the "curvature" of the coordinate space rather than true physical forces.

Rotating Coordinate Systems

In curvilinear coordinates, rotating coordinate systems introduce time-dependent transformations that account for the rotation of the reference frame relative to an inertial one, commonly used in mechanics to describe motion on rotating bodies like Earth. The position vector \mathbf{r} is expressed in both inertial coordinates \mathbf{u} and rotating coordinates \mathbf{u}' via a time-dependent orthogonal rotation matrix R(t), such that \mathbf{u}' = R(t) \mathbf{u}, where R(t) satisfies R^T R = I and \det R = 1. The angular velocity vector \boldsymbol{\omega} characterizes the rotation, defined through the relation \frac{dR}{dt} = [\boldsymbol{\omega}] R, where [\boldsymbol{\omega}] is the skew-symmetric matrix representing the cross product with \boldsymbol{\omega}. The key to analyzing motion in such systems lies in the transformation of time derivatives between frames. For any vector \mathbf{A}, the inertial time derivative is related to the rotating frame derivative by the operator equation: \left( \frac{d\mathbf{A}}{dt} \right)_I = \left( \frac{d\mathbf{A}}{dt} \right)_R + \boldsymbol{\omega} \times \mathbf{A}. This follows from the rotation of the basis vectors in the rotating frame. Applying this to the position vector \mathbf{r} (which is the same in both frames) yields the velocity transformation: \mathbf{v}_I = \mathbf{v}_R + \boldsymbol{\omega} \times \mathbf{r}, where \mathbf{v}_R = \left( \frac{d\mathbf{r}}{dt} \right)_R. To obtain the acceleration, apply the operator again to \mathbf{v}_I: \mathbf{a}_I = \left( \frac{d\mathbf{v}_I}{dt} \right)_I = \left( \frac{d\mathbf{v}_R}{dt} \right)_R + \boldsymbol{\omega} \times \mathbf{v}_R + \frac{d\boldsymbol{\omega}}{dt} \times \mathbf{r} + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \boldsymbol{\omega} \times \mathbf{v}_R. Simplifying, the inertial acceleration is: \mathbf{a}_I = \mathbf{a}_R + 2 \boldsymbol{\omega} \times \mathbf{v}_R + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \frac{d\boldsymbol{\omega}}{dt} \times \mathbf{r}, where \mathbf{a}_R = \left( \frac{d\mathbf{v}_R}{dt} \right)_R. Newton's second law in the inertial frame, m \mathbf{a}_I = \mathbf{F}, thus becomes in the rotating frame: m \mathbf{a}_R = \mathbf{F} - m \left[ 2 \boldsymbol{\omega} \times \mathbf{v}_R + \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}) + \frac{d\boldsymbol{\omega}}{dt} \times \mathbf{r} \right]. The additional terms represent fictitious forces: the Coriolis force -2m \boldsymbol{\omega} \times \mathbf{v}_R, the centrifugal force -m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}), and the Euler force -m \frac{d\boldsymbol{\omega}}{dt} \times \mathbf{r}. The Euler force arises specifically when the angular velocity \boldsymbol{\omega} is not constant, accounting for the tangential acceleration due to changes in rotation rate; for constant \boldsymbol{\omega}, this term vanishes, leaving only Coriolis and centrifugal effects. This derivation highlights the role of time-varying basis vectors in rotating curvilinear systems, extending beyond static transformations. Unlike fictitious forces in static curvilinear coordinates, which stem from spatial metric variations, those in rotating systems originate directly from the explicit time dependence of the coordinate rotation. A prominent application is the Foucault pendulum, analyzed in spherical coordinates fixed to the rotating Earth. In these coordinates (\theta, \phi), where \theta is the colatitude and \phi the azimuthal angle, the pendulum's position is \mathbf{r} = l (\sin\theta \cos\phi, \sin\theta \sin\phi, \cos\theta), with Earth's rotation introducing a Coriolis term that causes the plane of oscillation to precess. The precession angular rate is \boldsymbol{\omega} \sin\lambda, where \lambda is the latitude and \boldsymbol{\omega} is Earth's angular velocity ($7.29 \times 10^{-5} rad/s), opposite to Earth's rotation; for example, at the North Pole (\lambda = 90^\circ), the full precession period is one sidereal day. This effect demonstrates the Coriolis influence in curvilinear coordinates adapted to a rotating frame, providing empirical evidence of Earth's rotation without relying on celestial observations.

References

  1. [1]
    Curvilinear Coordinates - math - Oregon State University
    Dec 1, 2012 · The whole point of using curvilinear coordinates is that they are better adapted to the symmetries of the given problem. Ideally, this means ...
  2. [2]
    [PDF] A.7 ORTHOGONAL CURVILINEAR COORDINATES
    Orthogonal curvilinear coordinates are systems where vectors aligned with coordinate directions are mutually perpendicular, and varying one coordinate ...
  3. [3]
    [PDF] Untitled
    (1) Define a vector function r, that encodes the geometry. r = x(u₁, U2, 43) ex + y (U, M2, M3) Sy. + Z(U,, U2, Uz) Ez. U's are like r, Q, z. variables for ...
  4. [4]
    None
    Below is a merged response that consolidates all the information from the provided summaries into a single, comprehensive overview of curvilinear coordinates. To maximize detail and clarity, I will use a table in CSV format to organize the key aspects (Definition, Motivation/Historical Context, Sections, and Useful URLs) across the different segments of the document. Following the table, I will provide a narrative summary that integrates the details for a cohesive understanding.
  5. [5]
    1.3 Curvilinear Coordinates - BOOKS
    A coordinate system can be thought of as a collection of such “constant coordinate” surfaces, and the coordinates of a given point are just the values of those ...
  6. [6]
    [PDF] Coordinate Systems and Separation of Variables
    We are now faced with a spherical polar coordinate system, with the motivation that we might employ the separation of variables technique to solve.
  7. [7]
    None
    ### Summary of Laplace's Equation and Curvilinear Coordinates
  8. [8]
    Gabriel Lamé - Biography - MacTutor - University of St Andrews
    ... curvilinear coordinates. Curvilinear coordinates proved a very powerful tool in Lamé's hands. He used them to transform Laplace's equation into ellipsoidal ...
  9. [9]
    Curvilinear Coordinates - BOOKS
    round problems should be done in round coordinates. The ...
  10. [10]
    [PDF] Curvilinear Analysis in a Euclidean Space
    Jun 17, 2004 · This manuscript is a student's introduction on the mathematics of curvilinear coordinates, but can also serve as an information resource for ...
  11. [11]
    [PDF] Introduction to Differential Geometry
    A map F : V → W is called a diffeomorphism provided it is 1-1 and onto, and both F and F−1 are smooth. Conditions (1) and (2) in the IFT say that F : V → W is a ...
  12. [12]
    Parabolic Cylindrical Coordinates -- from Wolfram MathWorld
    . The scale factors are. h_1, = sqrt(u^2+v^2). (7). h_2, = sqrt(u^2+v^2). (8). h_3 ... See also. Confocal Paraboloidal Coordinates, Helmholtz Differential ...Missing: polar | Show results with:polar
  13. [13]
    [PDF] COORDINATE SYSTEMS , ox ox - Math@LSU
    Cartesian coordinates use perpendicular planes. Curvilinear coordinates use surfaces. The 1/-, v-, z-system is used in electrostatics and hydrodynamics.  ...
  14. [14]
    Elliptic Cylindrical Coordinates -- from Wolfram MathWorld
    ... coordinates by (x^2)/(a^2cosh^2u)+(y^2)/(a^2sinh^2u) = 1 (4) (x^2)/(a^2cos^2v)-(y^2)/(a^2sin^2v) = 1. (5) The scale factors are h_u = asqrt(cosh^2usin^2v+ ...Missing: polar | Show results with:polar
  15. [15]
    Solution of Two-Dimensional Stokes Flow With Elliptical ...
    Apr 22, 2021 · In this paper, analytical developments of the biharmonic equation representing two-dimensional Stokes flow are performed with elliptical coordinates.
  16. [16]
    Cylindrical Coordinates - Richard Fitzpatrick
    In the cylindrical coordinate system, , , and , where , , and , , are standard Cartesian coordinates. Thus, is the perpendicular distance from the -axis, and ...
  17. [17]
    [PDF] AN INTRODUCTION TO CURVILINEAR ORTHOGONAL ...
    Orthogonal coordinates have unit vectors where the dot product of any two unit vectors is 0 unless they are the same vector. The distance between points is the ...
  18. [18]
    A Proper Discretization of Hydrodynamic Equations in Cylindrical ...
    Cylindrical coordinates are often used in computational fluid dynamics, particularly when one is considering gas flow accreting onto a central object.
  19. [19]
    [PDF] 18.02 Multivariable Calculus - MIT OpenCourseWare
    Example: Newton's theorem: the gravitational attraction of a spherical planet with uniform density δ is the same as that of the equivalent point mass at its ...
  20. [20]
    [PDF] SIESTA: A scalable iterative equilibrium solver for toroidal applications
    SIESTA is a solver for rapidly obtaining MHD equilibria in toroidal systems, using an iterative approach and a physics-based preconditioner.
  21. [21]
    [PDF] An Introduction to Vectors and Tensors from a Computational ... - UTC
    Nonorthogonal coordinates introduce a dual basis: one basis is parallel to the coordinate lines (the contravariant basis) and a reciprocal basis (the covariant ...
  22. [22]
    [PDF] An Introduction to Tensors for Students of Physics and Engineering
    Joseph C. Kolecki. Glenn Research Center, Cleveland, Ohio. An Introduction to Tensors for Students of Physics and Engineering. NASA/TM—2002-211716. September ...
  23. [23]
    [PDF] Ae/APh 101 - Joseph Shepherd
    1.4.2 Curvilinear Coordinates. Scale factors Consider an orthogonal curvilinear coordinate system (x1,x2,x3) defined by a triad of unit vectors (e1,e2,e3) ...Missing: mathematical | Show results with:mathematical
  24. [24]
    [PDF] Appendix A Curvilinear coordinates
    we restrict ourselves to orthogonal curvilinear coordinates systems eξi ... hi Lamé coefficients. Those coefficients (might) depend on the coordinates ...
  25. [25]
    Div, Grad and Curl in Orthogonal Curvilinear Coordinates - Galileo
    Putting this together with the expression for the gradient gives immediately the expression for the Laplacian operator in curvilinear coordinates: ∇2ψ=1h1h ...
  26. [26]
    [PDF] Orthogonal Curvilinear Coordinates - Research
    Gradient, Divergence and Curl. The gradient operator ∇ takes the derivative of a quantity with respect to a change in position. In orthogonal curvilinear ...
  27. [27]
    [PDF] Vector operators in curvilinear coordinate systems - Physics
    Gradient in curvilinear (orthogonal) coordinate system. • Most generally, we ... • Derivation following problem 2, start with ∇ x1 = ê1 h1. , and then.
  28. [28]
    [PDF] Jackson J.D. Classical electrodynamics (Wiley, 1962)
    3.1 Laplace's equation in spherical coordinates, 54. 3.2 Legendre polynomials ... 9.3 Magnetic dipole and quadrupole fields, 273. 9.4 Center-fed linear ...
  29. [29]
    [PDF] Physics 504, Lecture 4 Feb. 1, 2010 1 Curvilinear Coordinates
    Feb 1, 2010 · This real symmetric matrix is known as the metric tensor. It is in general a nontrivial function of the position, gij(q). To repeat: the ...
  30. [30]
    [PDF] Curvilinear coordinate systems
    A curvilinear coordinate system is defined relative to a Cartesian coordinate system. ... The quantities gij form a matrix called the metric tensor G = {gij}.
  31. [31]
    [PDF] The Riemann Curvature Tensor - Louisiana Tech Digital Commons
    May 6, 2019 · The metric tensor is of particular importance, allowing for the concept of distances and maintaing the invariance of distance in different ...
  32. [32]
    Lecture Notes on General Relativity - S. Carroll
    The Riemann tensor, with four indices, naively has n4 independent components in an n-dimensional space. In fact the antisymmetry property (3.64) means that ...
  33. [33]
  34. [34]
  35. [35]
    [PDF] 18.022: Multivariable calculus — The change of variables theorem
    We note that the matrix DF(x) is invertible if and only if the determinant det DF(x) is non-zero. This determinant is called the Jacobian of F at x. The change- ...
  36. [36]
    [PDF] Chapter 5 - University of California, Berkeley
    ... volume element To understand the appearance of the factor√g in the volume element, let us consider a two-dimensional Euclidean plane covered by curvilinear ...
  37. [37]
    [PDF] Change of Variables and the Jacobian - Elsevier
    In these instances, we need to compute the determinant of the Jacobian matrix in order to include the proper scaling factor when we change coordinates. jJj = @x.
  38. [38]
    5.7 Change of Variables in Multiple Integrals - Calculus Volume 3
    Mar 30, 2016 · Now we need to define the Jacobian for three variables. Definition. The Jacobian determinant J ( u , v , w ) J ( u , v , w ) in three ...
  39. [39]
    [PDF] FW Math 321, 10/01/2003 Curvilinear Coordinates Let x, y and z be ...
    In curvilinear coordinates, the unit vectors qi depend on the coordinates. We need to know their. derivatives with respect to the qj, ∂qi/∂qj, for various ...
  40. [40]
    Normal-mode theory of a rotating earth model using a Lagrangian ...
    In this section, we define a set of non-orthogonal curvilinear coordinates (q, χ, ν) such that the parameters of the reference earth model depend on q only.
  41. [41]
    [PDF] 1.3 special tensors
    Mar 1, 2013 · ... position vector which is a function of the generalized coordinates and r = r(u, v, w) = x(u, v, w) e1 + y(u, v, w) e2 + z(u, v, w) e3. Page 3 ...
  42. [42]
    [PDF] 1.18 Curvilinear Coordinates: Tensor Calculus
    Jan 18, 2010 · From the definition 1.18.23 for the gradient of a scalar field, and Eqn. 1.18.30, one has for an orthogonal curvilinear coordinate system,. 3.
  43. [43]
    [PDF] Kaluza-Klein Theory
    Dec 3, 2022 · We require the fifth dimension to be compactified, rolled-up in a small circle, explaining the fact that an extra spatial dimension cannot be ...
  44. [44]
    Connection dynamics of reduced five-dimensional Kaluza-Klein ...
    The connection dynamics of the five-dimensional Kaluza-Klein theory reduced on four-dimensional spacetime is obtained by performing the Hamiltonian analysis ...
  45. [45]
    6.4 Fictitious Forces and Non-inertial Frames: The Coriolis Force
    Jul 13, 2022 · This fictitious force is called the centrifugal force—it explains the rider's motion in the rotating frame of reference. (b) In an inertial ...
  46. [46]
    Curved Coordinate Systems and Fictitious Forces - MathPages
    According to many reference texts, centrifugal and Coriolis forces (and fictitious forces in general) arise only from accelerated motion of the coordinate ...
  47. [47]
    [PDF] Unit 3 Coordinates and Transformations WG Harter
    Feb 3, 2018 · Christoffel expansion of fictitious forces: Rotating frames. 35 ... The intrinsic derivative takes account of all the "fictitious forces" due to ...
  48. [48]
    [PDF] Rotating Coordinate Systems
    Rotating Coordinate Systems. 4.1 Time Derivatives in a Rotating Frame. First recall the result that, for a vector A of fixed length, rotating about the origin.
  49. [49]
  50. [50]
    [PDF] 7.1 Equations of Motion in Rotating Coordinates - MIT
    Nov 20, 2002 · In the coordinate system rotating at the constant angular velocity, the momentum equation reads, after dropping subscripts R. ρÃdq dt+ 2~Ω ...
  51. [51]
    [PDF] Euler Force
    rotating frames to stationary frames​​ Then, by taking time derivatives, formulas are e velocity of the particle as seen in the two frames, and the acceleratio e ...
  52. [52]
    Rotating reference frames
    The last two terms in this equation are so-called fictitious forces; such forces are always needed to account for motion observed in non-inertial reference ...Missing: thornton marion classical<|control11|><|separator|>
  53. [53]
    [PDF] Foucault pendulum (with a twist)
    The goal of this paper is to study the motion of a nonlinear, spherical pendulum on a rotating planet. It turns out that the problem on a fixed energy level ...
  54. [54]
    [PDF] Chapter VII. Rotating Coordinate Systems - twister.ou.edu
    ... systems. Our goal is to find the equations of motion in the absolute and rotating coordinates. Now, we just showed that a. V. V r. = +Ω× r. r r r. (7.9) where V.