Fact-checked by Grok 2 weeks ago

Line element

In , the line element, commonly denoted as ds^2, is a quadratic that defines the squared or interval between two infinitesimally close points on a manifold, given by ds^2 = g_{ij} \, dx^i \, dx^j, where g_{ij} are the components of the and dx^i represent coordinate differentials. This expression encapsulates the intrinsic of the space, allowing for the measurement of distances, angles, and curvatures independent of any embedding in a higher-dimensional . In , the g_{ij} is a smooth, symmetric, and positive definite (0,2)- on the manifold, providing the coefficients for the line element and inducing an inner product on the tangent spaces at each point, enabling the computation of lengths of tangent s as \sqrt{g_{ij} v^i v^j} for a v. In , the along a is obtained by integrating the line element: L = \int_a^b \sqrt{g_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt}} \, dt, generalizing the to curved spaces. For surfaces embedded in \mathbb{R}^3, the line element corresponds to the first fundamental form, expressed as ds^2 = E \, du^2 + 2F \, du \, dv + G \, dv^2, where E = \mathbf{r}_u \cdot \mathbf{r}_u, F = \mathbf{r}_u \cdot \mathbf{r}_v, and G = \mathbf{r}_v \cdot \mathbf{r}_v for a parametrization \mathbf{r}(u,v), capturing the surface's intrinsic metric properties. In pseudo-Riemannian geometry, such as , the has an indefinite signature, and the line element defines spacetime intervals, with for timelike paths given by \tau = \int \sqrt{ - ds^2 }. Beyond , the line element plays a central role in physics, particularly in , where it describes the of via metrics like the Minkowski line element ds^2 = -dt^2 + dx^2 + dy^2 + dz^2 (in units where c = 1) in flat spacetime or the ds^2 = -(1 - 2m/r) dt^2 + (1 - 2m/r)^{-1} dr^2 + r^2 d\theta^2 + r^2 \sin^2\theta d\phi^2 (in units where G = c = 1) for black holes, determining geodesics as paths of extremal or distance. In , it facilitates the study of geodesics, , and isometries, with examples in such as spherical systems yielding ds^2 = dr^2 + r^2 d\theta^2 + r^2 \sin^2\theta d\phi^2, highlighting how the adapts to the coordinate while preserving geometric invariants.

Fundamentals

Definition and arc length

The line element emerged from the calculus of variations in the 18th century, where it served as a foundational tool for determining —the curves of extremal length on surfaces. Leonhard Euler laid the groundwork in his 1744 treatise Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, applying variational principles to minimize path lengths and deriving equations now known as the Euler-Lagrange equations for such problems. advanced these ideas in 1760, refining the analytical framework to eliminate purely geometric constructions and extending applications to minimal surfaces and . further developed the concept in 1827, introducing the line element as a measure intrinsic to curved surfaces, independent of embedding in , which enabled the study of geodesic properties without reference to external coordinates. In modern , the line element ds quantifies the infinitesimal along a smooth \gamma in a , parameterized by local coordinates x^i(t), where t varies over an interval [t_1, t_2]. It is defined as ds = \sqrt{ds^2}, with ds^2 denoting the squared line element, a positive that captures the local ; the encodes this form, ensuring coordinate independence. Infinitesimally, ds arises from the norm of the \dot{\gamma}(t) = \frac{dx^i}{dt} \frac{\partial}{\partial x^i}, representing the instantaneous direction and magnitude of motion along the curve at each point. This interpretation aligns with first principles: as the parameter increment dt approaches zero, ds approximates the between nearby points on the curve, scaled by the local metric structure. The finite arc length s of the curve from t_1 to t_2 is obtained by integrating the line element: s = \int_{t_1}^{t_2} ds = \int_{t_1}^{t_2} \sqrt{ds^2} \, dt, where the integrand is the speed \sqrt{ds^2 / dt^2}, ensuring reparameterization invariance when the curve is (nonzero speed). For proper setup with curves, one assumes the parameterization is smooth and the nowhere vanishes, allowing the to accumulate the total along the path. As a representative example, consider a simple parametric curve in the plane, such as the graph of a parabola traced by \gamma(t) = (t, t^2 / 2) for t \in [0, 1]. The arc length is computed as s = \int_0^1 \sqrt{1 + t^2} \, dt = \frac{1}{2} \left[ t \sqrt{1 + t^2} + \sinh^{-1} t \right]_0^1 \approx 1.148, illustrating how the integral aggregates infinitesimal segments to yield the total curved distance, exceeding the straight-line separation of endpoints.

Relation to the metric tensor

The g_{ij} is defined as a , symmetric, non-degenerate on the of a manifold, assigning an inner product to each T_p M at every point p \in M. In local coordinates, its components g_{ij} form a that varies across the manifold. The squared line element ds^2 is expressed using the via the equation ds^2 = g_{ij} \, dx^i \, dx^j, where the Einstein is employed, implying summation over repeated indices i and j from 1 to the dimension of the manifold. This captures the squared distance between nearby points. Under a coordinate x^i = x^i(\tilde{x}^k), the components of the transform as \tilde{g}_{kl} = \frac{\partial x^i}{\partial \tilde{x}^k} \frac{\partial x^j}{\partial \tilde{x}^l} g_{ij}, ensuring that ds^2 remains , thereby defining distances independently of the choice of coordinates. For Riemannian manifolds, the metric is positive-definite, meaning g_{ij} v^i v^j > 0 for all non-zero vectors v, with signature (n, 0) in n-dimensions. In pseudo-Riemannian cases, the metric has an indefinite signature (p, q) with p + q = n and both p, q > 0, allowing for both positive and negative eigenvalues, as in Lorentzian metrics with signature (1, n-1). In cases where g_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt} > 0 (e.g., in Riemannian manifolds or spacelike curves in pseudo-Riemannian ones), the arc length s of a curve \gamma(t) parameterized by t \in [a, b] is s = \int_a^b \sqrt{g_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt}} \, dt, which is independent of the parameterization. For unit speed curves, the parameterization satisfies g_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt} = 1, normalizing the speed to unity along the curve. In pseudo-Riemannian manifolds, for timelike curves where the form is negative, proper time is defined using \int_a^b \sqrt{ - g_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt} } \, dt (assuming signature with negative time component); further details are covered in subsequent sections on spacetime applications.

Euclidean spaces

Cartesian coordinates

In flat Euclidean space, the line element in Cartesian coordinates provides the simplest expression for measuring infinitesimal distances. For an n-dimensional space with coordinates x^1, x^2, \dots, x^n, the line element is given by ds^2 = \delta_{ij} \, dx^i \, dx^j = \sum_{i=1}^n (dx^i)^2, where \delta_{ij} is the Kronecker delta, which equals 1 if i = j and 0 otherwise, representing the flat metric tensor in this coordinate system. This form arises as a special case of the general line element ds^2 = g_{ij} \, dx^i \, dx^j, where the metric components g_{ij} are constant and diagonal. The s along a parameterized by t, with position \mathbf{r}(t) = (x^1(t), \dots, x^n(t)), is computed by integrating the : s = \int_a^b \sqrt{\sum_{i=1}^n \left( \frac{dx^i}{dt} \right)^2} \, dt = \int_a^b \|\mathbf{r}'(t)\| \, dt. For a straight line between two points, this integral simplifies directly to the \sqrt{\sum_{i=1}^n (\Delta x^i)^2}, generalizing the from two dimensions—where ds^2 = dx^2 + dy^2 yields the hypotenuse \sqrt{(\Delta x)^2 + (\Delta y)^2}—to higher dimensions. Under orthogonal transformations, such as rotations, the Cartesian coordinates can be changed via a R with R^T R = I, preserving the flat : the line element remains ds^2 = \sum_{i=1}^n (dx'^i)^2 in the new coordinates x'^i = R^i_j x^j, since the transforms as \delta_{ij} under such bases. This invariance ensures that distances and angles are unchanged, reflecting the of . In applications, the line element primarily defines distances between points along paths, essential for optimization problems like shortest paths (geodesics, which are straight lines here) and in vector calculus for line integrals. While it extends to higher-dimensional forms like the volume element dV = dx^1 \wedge \cdots \wedge dx^n for integration over regions, the focus remains on one-dimensional arc lengths. Historically, the conceptual foundation traces to Euclidean geometry in Euclid's Elements (circa 300 BCE), where the Pythagorean theorem underpins finite distances, evolving into the infinitesimal line element through 17th- and 18th-century calculus developments in arc length by Leibniz and Euler, and formalized in n-dimensional vector spaces by the 19th century.

Curvilinear coordinates

In Euclidean space, orthogonal curvilinear coordinates provide a system where the coordinate curves intersect at right angles, allowing for a simplified representation of distances and geometry that aligns with the inherent symmetries of certain problems. Common examples include two-dimensional polar coordinates (r, \theta), where r is the radial distance and \theta is the azimuthal angle, and three-dimensional cylindrical coordinates (r, \theta, z) or spherical coordinates (r, \theta, \phi), with \theta as the polar angle and \phi as the azimuthal angle. These systems are particularly advantageous for problems exhibiting rotational or spherical symmetry, such as calculating fields around circular or spherical objects, where the coordinate choice reduces the complexity of equations compared to Cartesian coordinates—for instance, describing a circle requires a constant r in polar coordinates rather than a varying relation like x^2 + y^2 = r^2. The line element in an n-dimensional orthogonal curvilinear coordinate system (u^1, u^2, \dots, u^n) takes the diagonal form ds^2 = h_1^2 (du^1)^2 + h_2^2 (du^2)^2 + \dots + h_n^2 (du^n)^2, where the h_i are the scale factors, also known as Lamé coefficients, which account for the stretching or compression along each coordinate direction. These scale factors are derived from the position vector \mathbf{r}(u^1, u^2, \dots, u^n) in Euclidean space, with h_i = \left| \frac{\partial \mathbf{r}}{\partial u^i} \right|, representing the magnitude of the infinitesimal displacement vector along the u^i-direction. In two-dimensional polar coordinates, the scale factors are h_r = 1 and h_\theta = r, yielding the line element ds^2 = dr^2 + r^2 d\theta^2. The arc length along a is found by integrating ds = \sqrt{dr^2 + r^2 d\theta^2}; for example, the of a circle at fixed r is \int_0^{2\pi} r \, d\theta = 2\pi r. For three-dimensional spherical coordinates, the scale factors are h_r = 1, h_\theta = r, and h_\phi = r \sin \theta, giving ds^2 = dr^2 + r^2 d\theta^2 + r^2 \sin^2 \theta \, d\phi^2. Here, the arc length integral along, say, a great circle (fixed r, \theta = \pi/2, varying \phi) simplifies to \int_0^{2\pi} r \, d\phi = 2\pi r, highlighting the coordinate system's efficiency in symmetric geometries. This form emerges as a special case of the Cartesian line element when scale factors are unity, but curvilinear systems introduce position-dependent factors to better capture curved symmetries.

General manifolds

Riemannian manifolds

In a Riemannian manifold, the line element generalizes the notion of infinitesimal distance to curved spaces, providing a way to measure lengths and angles intrinsically without reference to an . A is a smooth manifold equipped with a positive-definite g_{ij}(x) at each point, which varies smoothly with position x. The line element is expressed in local coordinates as ds^2 = g_{ij}(x) \, dx^i \, dx^j, where summation over repeated indices i, j = 1, \dots, n is implied (), and the components g_{ij} form a symmetric positive-definite that may include off-diagonal terms, reflecting the possible non-orthogonality of the coordinate basis. This form allows the metric to capture the local geometry, including curvature, by how g_{ij} changes across the manifold. Under a change of coordinates from x^i to x'^k, the line element remains , ensuring that distances are independent of the coordinate choice. The transforms as a tensor via the of the : g'_{kl}(x') = \frac{\partial x^m}{\partial x'^k} \frac{\partial x^n}{\partial x'^l} g_{mn}(x), where the partial derivatives account for how displacements dx^i map to dx'^k. This law preserves the positive-definiteness and the geometric structure, confirming that the defines an intrinsic geometry on the manifold. Geodesics on a are the curves that locally minimize the , analogous to straight lines in . The L of a \gamma(t) parameterized by t \in [a, b] is given by L = \int_a^b \sqrt{g_{ij}(\gamma(t)) \frac{dx^i}{dt} \frac{dx^j}{dt}} \, dt. To find the extremal paths, one applies the : the Euler-Lagrange equations for the \mathcal{L} = \sqrt{g_{ij} \dot{x}^i \dot{x}^j} (where \dot{x} = dx/dt) yield the equation \frac{d^2 x^k}{dt^2} + \Gamma^k_{ij} \frac{dx^i}{dt} \frac{dx^j}{dt} = 0, with the of the second kind defined as \Gamma^k_{ij} = \frac{1}{2} g^{kl} \left( \partial_i g_{jl} + \partial_j g_{il} - \partial_l g_{ij} \right), where g^{kl} is the inverse (g^{kl} g_{lm} = \delta^k_m) and \partial_i = \partial / \partial x^i. These symbols encode the effects through the of the . A concrete example is the on a two-dimensional , obtained by rotating a in the rz-plane around the z-axis in \mathbb{R}^3. If the generating is parameterized as r = r(z) with z as the meridional coordinate and \theta the azimuthal , the induced Riemannian from the is ds^2 = \left(1 + \left(\frac{dr}{dz}\right)^2 \right) dz^2 + r(z)^2 \, d\theta^2, which is diagonal but position-dependent, illustrating how in higher dimensions determines the intrinsic . For instance, a of radius R arises when r(z) = \sqrt{R^2 - z^2} for |z| < R, yielding the ds^2 = \frac{R^2}{R^2 - z^2} dz^2 + (R^2 - z^2) d\theta^2, though geodesics (great circles) can be computed intrinsically without the embedding. The invariant nature of ds^2 under diffeomorphisms underscores its role in defining the manifold's geometry solely through the metric, independent of any external coordinates or embeddings. This tensorial object encapsulates all information about lengths, angles, and volumes, forming the foundation for studying curvature via the Riemann tensor derived from second derivatives of g_{ij}. In contrast to curvilinear coordinates in Euclidean space, where the metric is derived from a flat background and often diagonalizes to orthogonal form, the general Riemannian case admits arbitrary variation and mixing of components, enabling the description of truly curved geometries.

Pseudo-Riemannian manifolds

In pseudo-Riemannian manifolds, the line element generalizes the Riemannian case by incorporating a metric tensor with an indefinite signature, allowing for both positive and negative eigenvalues in its quadratic form. The signature is denoted as (p, q), where p is the number of positive eigenvalues and q the number of negative ones, with the total dimension n = p + q. This contrasts with the positive-definite (n, 0) signature of , enabling structures that distinguish between different types of directions in the tangent space. A prominent example is the (3, 1) or (1, 3), commonly used in spacetime contexts, where the metric has three spatial positive directions and one temporal negative direction (or vice versa, depending on convention). The line element on a pseudo-Riemannian manifold is expressed in local coordinates as ds^2 = g_{ij} \, dx^i \, dx^j, where g_{ij} is the indefinite metric tensor, symmetric and non-degenerate but not positive definite. Unlike in the Riemannian setting, the sign of ds^2 classifies infinitesimal displacements: spacelike if ds^2 > 0, timelike if ds^2 < 0, and null if ds^2 = 0. This classification arises directly from the , partitioning the into orthogonal subspaces corresponding to positive, negative, and degenerate directions. Under smooth coordinate transformations, the line element transforms as a scalar, preserving its form and the metric's signature locally, just as in the Riemannian case. The indefinite nature introduces a , geometrically manifested through at each point. These cones are the sets of null directions where ds^2 = 0, forming double cones that separate timelike interiors (where causal influences can propagate) from spacelike exteriors (where they cannot). In a manifold of Lorentzian signature, the divides the into future and past timelike regions, with spacelike vectors lying outside. This structure is intrinsic to the and holds generally, without reference to flat models. For an illustrative example, consider a two-dimensional with (1, 1) , such as a curved surface equipped with a ds^2 = -f(u,v)^2 \, du^2 + g(u,v)^2 \, dv^2, where f and g are positive smooth functions ensuring non-degeneracy. Here, curves with ds^2 < 0 are timelike, tracing paths within the light cones, while the 's affects behavior without altering the fundamental classification. Such manifolds demonstrate how the line element adapts to indefinite metrics while retaining tensorial under diffeomorphisms.

Applications in spacetime

Minkowski spacetime

In Minkowski spacetime, the line element describes the invariant spacetime interval in flat four-dimensional spacetime, as formulated in . The metric takes the form ds^2 = -c^2 dt^2 + dx^2 + dy^2 + dz^2, where c is the , t is the , and x, y, z are spatial coordinates, employing the (-, +, +, +). Equivalently, in abstract index notation, it is expressed as ds^2 = \eta_{\mu\nu} dx^\mu dx^\nu, with \eta_{\mu\nu} = \operatorname{diag}(-1, 1, 1, 1) (in units where c=1). This form was introduced by to geometrize Einstein's , unifying space and time into a single manifold where the line element remains invariant under coordinate transformations. For timelike paths, where ds^2 < 0, the line element defines the \tau experienced by an observer along their worldline. The infinitesimal proper time is given by d\tau = \frac{|ds|}{c} = \sqrt{-ds^2}/c, and the total proper time for a path is the \tau = \frac{1}{c} \int \sqrt{-ds^2}. This quantity is Lorentz invariant, representing the time measured by a clock moving with the observer, distinct from in any inertial frame. In , this leads to effects, where moving clocks tick slower relative to stationary ones. The line element is preserved under Lorentz transformations, which include spatial rotations and boosts between inertial frames. Boosts along the x-direction, for instance, mix time and space coordinates while leaving ds^2 unchanged, embodying the principle of that physical laws are the same in all inertial frames. These transformations form the , ensuring the invariance of the spacetime interval as a fundamental postulate. Null geodesics correspond to paths where ds^2 = 0, defining the worldlines of rays propagating at speed c. These form cones at each , separating timelike (inside the cone) and spacelike (outside) intervals, with the cone's generators tracing null directions. In inertial coordinates, such paths satisfy dx = \pm c dt (with dy = dz = 0 for radial propagation), illustrating the of . Minkowski spacetime uses Cartesian-like coordinates tied to inertial frames, where observers are at rest relative to the axes. Boosts can be parameterized using , defined by v = c \tanh \phi, which adds hyperbolically under velocity : \phi_{\text{total}} = \phi_1 + \phi_2. This parameterization simplifies Lorentz transformations, expressing them via hyperbolic functions and highlighting the geometric analogy to rotations in .

Curved spacetimes

In , the line element describing curved spacetimes takes the form ds^2 = g_{\mu\nu}(x) \, dx^\mu \, dx^\nu, where the g_{\mu\nu} varies with position and is determined by the G_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}, which relate to the distribution of matter and energy. These equations, formulated in 1915, provide the foundation for modeling gravitational effects as geometric distortions of . A seminal exact solution is the , derived by in 1916 shortly after the field equations, which applies to the vacuum exterior of a static, spherically symmetric mass M and is given by ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\Omega^2, where d\Omega^2 = d\theta^2 + \sin^2\theta \, d\phi^2. This metric is interpreted in modern as describing the around a non-rotating , featuring an at r_s = 2GM/c^2, the boundary beyond which radial null geodesics—paths of light rays—cannot escape. For timelike observers, proper time \tau along geodesics is measured by c^2 d\tau^2 = -ds^2, which slows relative to coordinate time t near the horizon due to . Radial null geodesics in this geometry, satisfying ds^2 = 0 with d\theta = d\phi = 0, approach the horizon in finite proper distance but infinite coordinate time for distant observers, highlighting the of the . In , the Friedmann–Lemaître–Robertson–Walker (FLRW) models homogeneous and isotropic expanding universes as ds^2 = -c^2 dt^2 + a(t)^2 \left[ \frac{dr^2}{1 - k r^2} + r^2 d\Omega^2 \right], where a(t) is the time-dependent scale factor governing expansion and k parameterizes spatial (k = 0, +1, -1 for flat, closed, or open geometries, respectively). This form, first proposed in 1922, arises as a solution to the Einstein equations with a perfect fluid source and underpins the standard Big Bang model. For rotating masses, the extends the Schwarzschild solution to include , describing the around rotating black holes while preserving asymptotic flatness. motion in these curved follows the equation \frac{d^2 x^\mu}{d\tau^2} + \Gamma^\mu_{\nu\sigma} \frac{dx^\nu}{d\tau} \frac{dx^\sigma}{d\tau} = 0, where the \Gamma^\mu_{\nu\sigma} = \frac{1}{2} g^{\mu\lambda} (\partial_\nu g_{\sigma\lambda} + \partial_\sigma g_{\nu\lambda} - \partial_\lambda g_{\nu\sigma}) encode the connection and cause paths to deviate from straight lines in flat space, reflecting gravitational deflection. This formulation, central to , predicts phenomena like orbital and light bending observed in astrophysical tests.

References

  1. [1]
    [PDF] An Introduction to Differential Geometry through Computation
    This is not a book on classical differential geometry or tensor analysis, ... Another name for a metric tensor is a “line-element field”. (written ds2) ...
  2. [2]
    [PDF] dr bob's elementary differential geometry - Villanova University
    Mar 4, 2018 · form,” usually called the “line element” in differential geometry. The diagonal matrix g = JT J consists of the coefficients of this ...
  3. [3]
    [PDF] A Brief Survey of the Calculus of Variations - arXiv
    Abstract. In this paper, we trace the development of the theory of the calculus of variations. From its roots in the work of Greek thinkers and continuing ...
  4. [4]
    [PDF] Outline of a History of Differential Geometry (II)
    As EULER and LAGRANGE, GAUSS uses " Gaussian coordinates," taking x, y, z, equal ... of surfaces, linear element, geodesics, Gaussian curvature, curvatura.
  5. [5]
    [PDF] Notes on Differential Geometry
    May 3, 2004 · These notes are an attempt to summarize some of the key mathe- matical aspects of differential geometry, as they apply in particular.
  6. [6]
    [PDF] Differential Geometry in Physics - UNCW
    In terms of these forms, the line element becomes ds2 = gαβθαθβ = 2θ0θ1 − (θ2)2 − (θ3)2, where g01 = g10 = −g22 = −g33 = 1, while all the other gαβ = 0 ...
  7. [7]
    [PDF] Differential Geometry of Curves and Surfaces by Do Carmo.
    Arc length parametrization. (1) Assume regular curve α. Arc length. Consider starting point α(t0). Arc length is distance traveled from t0 ...
  8. [8]
    [PDF] lee-smooth-manifolds.pdf - MIT Mathematics
    ... Lee, John M., 1950- p. cm. - (Graduate texts in mathematics; 218). Includes bibliographical references and index. K.A. Ribet. Mathematics Department.
  9. [9]
  10. [10]
    [PDF] math 147a: introduction to differential geometry
    Arc-length. By the Pythagorean theorem, we know the length of a vector v = (v1,...,vn) is given by kvk = p(v1)2 + · + (vn)2. To compute the length of a ...
  11. [11]
    [PDF] DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces
    If we think of H as modeling non-Euclidean geometry, with lines in our geometry being the geodesics, ... (The last is the hyperbolic Pythagorean Theorem.) ...
  12. [12]
    [PDF] Lectures on Differential Geometry Math 240C
    Jun 6, 2011 · These lecture notes cover Riemannian geometry, including tangent/co-tangent spaces, metrics, geodesics, and the Levi-Civita connection.
  13. [13]
    [PDF] PHZ 6607 Special and General Relativity I Handout #1 - UF Physics
    Handout #6: Infinitesimal Line Elements. Consider the two dimensional metric for flat space in cartesian coordinates: ds2 = dx2+ dy2, and impose the ...
  14. [14]
    Curvilinear Coordinates -- from Wolfram MathWorld
    Orthogonal curvilinear coordinates therefore have a simple line element. ds^2 ... (6). where the scale factors are. h_i=|(partialr)/(partialu_i)|. (7). and ...
  15. [15]
    [PDF] A.7 ORTHOGONAL CURVILINEAR COORDINATES
    To illustrate the derivation of scale factors and base vectors, consider the 0 quanti ties in cylindrical coordinates. From Eqs. (A.7-3) and (A.7-25) it is ...
  16. [16]
    [PDF] THE GEOMETRY OF MATHEMATICAL METHODS - BOOKS
    May 3, 2025 · Polar coordinates are useful for situations with circular symmetry in the plane. ... This is the general line element in spherical coordinates.
  17. [17]
    3 Introducing Riemannian Geometry‣ General Relativity ... - DAMTP
    A general Riemannian metric gives us a way to measure the length of a vector X at each point, It also allows us to measure the angle between any two vectors X ...
  18. [18]
    [PDF] John M. Lee - Introduction to Riemannian Manifolds
    This book is designed as a textbook for a graduate course on Riemannian geometry for students who are familiar with the basic theory of smooth manifolds. It ...
  19. [19]
    [PDF] An Introduction to Riemannian Geometry - UPenn CIS
    The study of Riemannian Geometry is rather meaningless without some basic knowledge on Gaussian Geometry that is the differential geometry of curves and ...
  20. [20]
    [PDF] Chapter 7 Geodesics on Riemannian Manifolds - UPenn CIS
    Definition 7.1.1 Given any Riemannian manifold, M, a smooth parametric curve (for short, curve) on M is a map, γ:I → M, where I is some open interval of R.
  21. [21]
    [PDF] 11. Geodesics and Classical Mechanics
    In terms of covariant differentiation on a Riemannian manifold, this equation. (11 : 13) says that the covariant derivative with respect to t of the velocity ...<|control11|><|separator|>
  22. [22]
    [PDF] Riemannian Geometry
    (5) (Surfaces of revolution). Let γ(t)=(φ(t),ψ(t)) be a curve into R2 and let M ⊂ R3 be the surface of revolution for γ with metric tensor ( ˙φ2 + ˙ψ2)dt2 ...
  23. [23]
    [PDF] Riemannian Geometry - Dr Hovhannes Khudaverdian
    vature K(p) depends only on Riemannian metric on the surface C. Indeed ... The ”free” Lagrangian of the surface of revolution is. Lfree = 1 + f02(h) ˙h2 ...
  24. [24]
    [PDF] Introduction to differential and Riemannian geometry - Hal-Inria
    May 2, 2020 · The addition of a Riemannian metric enables length and angle measurements on tangent spaces giving rise to the notions of curve length,.
  25. [25]
    Semi-Riemannian Geometry With Applications to Relativity
    In stock Free delivery1st Edition, Volume 103 - June 28, 1983 · Latest edition · Imprint: Academic Press · Author: Barrett O'Neill · Language: English · Hardback ISBN: 9780125267403. 9 7 ...
  26. [26]
    The Large Scale Structure of Space-Time
    Hawking, University of Cambridge, George F. R. Ellis, University of Cape Town. Publisher: Cambridge University Press. Online publication date: February 2023.Missing: URL | Show results with:URL
  27. [27]
    [PDF] Space and Time - UCSD Math
    It was Hermann Minkowski (Einstein's mathematics professor) who announced the new four- dimensional (spacetime) view of the world in 1908, which he deduced from ...
  28. [28]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES
    It is known that Maxwell's electrodynamics—as usually understood at the present time—when applied to moving bodies, leads to asymmetries which do.
  29. [29]
    The Field Equations of Gravitation - Wikisource, the free online library
    Aug 9, 2025 · We obtain the ten general covariant equations of the gravitational field in spaces, in which matter is absent.
  30. [30]
    On the gravitational field of a mass point according to Einstein's theory
    May 12, 1999 · View a PDF of the paper ... Loinger of the fundamental memoir, that contains the ORIGINAL form of the solution of Schwarzschild's problem.
  31. [31]
    Revisiting timelike and null geodesics in the Schwarzschild spacetime
    Oct 18, 2022 · The main result presented in this paper is a concise formula describing all types of timelike and null trajectories in the Schwarzschild metric ...
  32. [32]
    Gravitational Field of a Spinning Mass as an Example of ...
    Feb 21, 2014 · A novel solution to Einstein's gravitational equations, discovered in 1963, turned out to describe the curvature of space around every astrophysical black hole.
  33. [33]
    [PDF] General Relativity - DAMTP - University of Cambridge
    If we wished, we could substitute the Christoffel symbols into the geodesic equation (1.7) to determine the equations of motion. However, it's typically ...