First fundamental form
In differential geometry, the first fundamental form of a surface is a quadratic differential form that defines the intrinsic metric on the tangent space at each point of the surface, enabling the measurement of lengths, angles, and areas without reference to the ambient Euclidean space. Introduced by Carl Friedrich Gauss in his seminal 1827 work Disquisitiones generales circa superficies curvas, it is expressed for a parametrized surface \mathbf{r}(u, v) as ds^2 = E\, du^2 + 2F\, du\, dv + G\, dv^2, where the coefficients are the dot products 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.[1][2] These coefficients form a symmetric positive definite matrix \begin{pmatrix} E & F \\ F & G \end{pmatrix} for a regular surface, ensuring the form is positive definite and zero only for the trivial tangent vector.[3] The first fundamental form captures the Riemannian metric structure of the surface, determining the arc length of a curve \gamma(t) = (u(t), v(t)) along the surface as \int \sqrt{E (u')^2 + 2F u' v' + G (v')^2} \, dt.[4] It also defines the angle \theta between two tangent vectors via the cosine formula \cos \theta = \frac{E u_1' u_2' + F (u_1' v_2' + v_1' u_2') + G v_1' v_2'}{|\gamma_1'| |\gamma_2'|}, and the surface area element as dA = \sqrt{EG - F^2} \, du \, dv, where EG - F^2 > 0 for orientable surfaces.[2][3] This intrinsic nature means two surfaces are locally isometric—and thus share the same geometry—if and only if their first fundamental forms coincide under corresponding parametrizations, a key insight from Gauss's theorema egregium relating it to Gaussian curvature.[4] Gauss's formulation revolutionized the study of curved spaces by distinguishing intrinsic properties, measurable by inhabitants on the surface, from extrinsic ones dependent on embedding, paving the way for developments in general relativity and modern geometry.[1] For example, on a sphere of radius a parametrized by spherical coordinates, the first fundamental form simplifies to ds^2 = a^2 \sin^2 v \, du^2 + a^2 \, dv^2 when F = 0, reflecting orthogonal coordinate lines.[4] In computational applications, such as computer-aided design, it facilitates precise surface interrogation and representation.[2]Historical Background
Carl Friedrich Gauss's Contribution
Carl Friedrich Gauss introduced the first fundamental form in his seminal 1827 publication, Disquisitiones generales circa superficies curvas, presented to the Royal Society of Göttingen on October 8, 1827.[1] This work laid the foundation for the intrinsic theory of curved surfaces embedded in three-dimensional Euclidean space, marking a pivotal advancement in differential geometry. Gauss's primary motivation was to investigate properties of surfaces that could be determined solely from measurements within the surface itself, without reliance on the surrounding ambient space. He sought to identify geometric invariants preserved under isometric deformations, such as bending without stretching or tearing, thereby distinguishing intrinsic characteristics—like distances along the surface—from extrinsic ones dependent on embedding. This approach was influenced by practical problems in geodesy, where accurate surface measurements on the Earth were essential, but Gauss emphasized its broader theoretical implications for understanding surface geometry.[1] In the paper, Gauss defined the first fundamental form using the line element ds^2 = E \, dp^2 + 2F \, dp \, dq + G \, dq^2, where p and q are parameters parameterizing the surface, and the coefficients are given by \begin{align*} E &= \left( \frac{\partial x}{\partial p} \right)^2 + \left( \frac{\partial y}{\partial p} \right)^2 + \left( \frac{\partial z}{\partial p} \right)^2, \\ F &= \frac{\partial x}{\partial p} \frac{\partial x}{\partial q} + \frac{\partial y}{\partial p} \frac{\partial y}{\partial q} + \frac{\partial z}{\partial p} \frac{\partial z}{\partial q}, \\ G &= \left( \frac{\partial x}{\partial q} \right)^2 + \left( \frac{\partial y}{\partial q} \right)^2 + \left( \frac{\partial z}{\partial q} \right)^2, \end{align*} with (x, y, z) denoting the position coordinates of points on the surface. These coefficients, derived from the dot products of partial derivatives, encapsulate the metric structure induced by the embedding.[1] A key insight from Gauss's analysis is that the first fundamental form uniquely determines the lengths of curves and areas of regions on the surface, independent of how the surface is embedded in the ambient space. This intrinsic nature allows for the computation of such quantities using only the form's coefficients, without reference to extrinsic coordinates, as Gauss demonstrated through its application to infinitesimal displacements along the surface. This principle underpins his famous Theorema Egregium, which further highlights the form's role in preserving Gaussian curvature under isometries.[1]Evolution in Modern Differential Geometry
In his 1854 habilitation lecture, Bernhard Riemann extended the concept of the first fundamental form beyond surfaces to n-dimensional manifolds, introducing the metric tensor g_{ij} as a generalization that defines distances and angles intrinsically within the manifold.[5] This framework allowed for the study of geometry without embedding in Euclidean space, laying the groundwork for Riemannian geometry by treating the metric as a quadratic form on tangent spaces.[6] Building on Riemann's ideas, Gregorio Ricci-Curbastro and Tullio Levi-Civita developed tensor calculus in the early 20th century, formalizing the first fundamental form within absolute differential calculus, which provided tools for coordinate-independent manipulations of the metric tensor.[7] Their work, particularly in the 1901 publication Méthodes de calcul différentiel absolu et leurs applications, integrated the metric into a broader system of tensor analysis, enabling rigorous computations on curved spaces.[8] The metric tensor, derived from the principles of the first fundamental form, played a pivotal role in Albert Einstein's formulation of general relativity in 1915, where it describes the geometry of spacetime influenced by mass and energy.[9] In Einstein's field equations, the metric g_{\mu\nu} governs gravitational effects as curvature, marking a profound application of intrinsic geometry to physics.[10] In contemporary differential geometry, the first fundamental form is understood as the pullback of the Euclidean metric to the tangent space of a submanifold, facilitating the analysis of intrinsic properties such as geodesics and curvature without reference to the ambient space.[11] This perspective underscores its role in studying Riemannian manifolds and submanifold geometry, emphasizing the form's capacity to encode all measurable distances and angles on the surface itself.[12]Mathematical Definition
For Parametric Surfaces
For a smooth parametric surface given by a parameterization \mathbf{X}: U \subset \mathbb{R}^2 \to \mathbb{R}^3, where U is an open set and \mathbf{X}(u,v) = (x(u,v), y(u,v), z(u,v)), the first fundamental form is defined as the symmetric bilinear form on the tangent space induced by the Euclidean inner product in \mathbb{R}^3.[13] Specifically, for tangent vectors \mathbf{w} and \mathbf{z} at a point p = \mathbf{X}(u,v), it is given by I_p(\mathbf{w}, \mathbf{z}) = \mathbf{w} \cdot \mathbf{z}, where \cdot denotes the standard dot product.[14] In terms of the parameterization, this yields the quadratic form ds^2 = I\left( \frac{\partial \mathbf{X}}{\partial u}, \frac{\partial \mathbf{X}}{\partial u} \right) du^2 + 2 I\left( \frac{\partial \mathbf{X}}{\partial u}, \frac{\partial \mathbf{X}}{\partial v} \right) du \, dv + I\left( \frac{\partial \mathbf{X}}{\partial v}, \frac{\partial \mathbf{X}}{\partial v} \right) dv^2, which measures infinitesimal distances on the surface.[13] The coefficients of this form, often denoted E, F, and G, are computed from the partial derivatives \mathbf{X}_u = \frac{\partial \mathbf{X}}{\partial u} and \mathbf{X}_v = \frac{\partial \mathbf{X}}{\partial v}, which span the tangent plane at each point. Thus, E = |\mathbf{X}_u|^2 = \mathbf{X}_u \cdot \mathbf{X}_u, \quad F = \mathbf{X}_u \cdot \mathbf{X}_v, \quad G = |\mathbf{X}_v|^2 = \mathbf{X}_v \cdot \mathbf{X}_v, resulting in the line element ds^2 = E \, du^2 + 2F \, du \, dv + G \, dv^2.[14][13] The quadratic form defined by these coefficients is positive definite, with E > 0, G > 0, and EG - F^2 > 0, ensuring the form defines a valid Riemannian metric.[14][13] The parameterization is assumed to be regular, meaning the surface is immersed without singularities, which requires \mathbf{X}_u \times \mathbf{X}_v \neq \mathbf{0} everywhere in U.[13] This cross product being nonzero guarantees that \{\mathbf{X}_u, \mathbf{X}_v\} forms a basis for the two-dimensional tangent space and that the parameterization is locally injective.[14] Moreover, the first fundamental form is unique up to coordinate changes; under a reparameterization (u,v) = (u(s,t), v(s,t)), the form transforms via the Jacobian matrix but remains invariant as an intrinsic object on the surface.[13] This independence ensures that metric properties derived from it, such as lengths and angles, are well-defined regardless of the chosen coordinates.[14]General Formulation as an Induced Metric
In differential geometry, the first fundamental form provides a general framework for the induced metric on a submanifold M \subset \mathbb{R}^n. At each point p \in M, it is defined as the symmetric bilinear form g_p: T_p M \times T_p M \to \mathbb{R} given by g_p(v, w) = \langle v, w \rangle_{\mathbb{R}^n} for tangent vectors v, w \in T_p M, where \langle \cdot, \cdot \rangle_{\mathbb{R}^n} is the standard Euclidean inner product on the ambient space \mathbb{R}^n.[15][16] This construction arises as the pullback of the Euclidean metric via an immersion or embedding of M, restricting the ambient inner product to the tangent spaces of the submanifold.[16] As such, it equips M with a natural Riemannian metric, enabling measurements of lengths, angles, and areas intrinsically on M.[15] A key feature of this formulation is its intrinsic nature: the first fundamental form depends only on the geometry within M itself, defined exclusively on its tangent bundle, and remains unchanged under isometric re-embeddings of M into Euclidean space.[16] Unlike extrinsic descriptions that rely on the codimension or normal spaces, g_p captures the geometry observable from within M, independent of how M is positioned or oriented in \mathbb{R}^n.[15] This independence ensures that properties derived from the first fundamental form, such as geodesic distances, are preserved under rigid motions of the ambient space.[16] The bilinear form g_p exhibits essential algebraic properties that underpin its role as a metric tensor. It is symmetric, satisfying g_p(v, w) = g_p(w, v) for all v, w \in T_p M, and positive definite, meaning g_p(v, v) > 0 for all nonzero v \in T_p M, which induces a norm \|v\|_p = \sqrt{g_p(v, v)} and guarantees that M inherits a Riemannian structure from the embedding.[15][16] These properties allow g to define inner products on tangent vectors, facilitating the computation of angles via \cos \theta = \frac{g_p(v, w)}{\|v\|_p \|w\|_p} and ensuring the metric is non-degenerate.[15] To express this metric in local coordinates \{x^i\}_{i=1}^m on M, where m = \dim M, the first fundamental form takes the form g = g_{ij} \, dx^i \, dx^j, with components g_{ij} = g_p \left( \frac{\partial}{\partial x^i}, \frac{\partial}{\partial x^j} \right).[16] In terms of a local parametrization or immersion x: U \subset \mathbb{R}^m \to \mathbb{R}^n, these coefficients reduce to g_{ij} = \left\langle \frac{\partial x}{\partial u^i}, \frac{\partial x}{\partial u^j} \right\rangle_{\mathbb{R}^n}, linking the abstract definition to coordinate-based computations while preserving the induced structure.[15] The matrix (g_{ij}) is symmetric and positive definite, reflecting the bilinear properties at each point.[16]Notation and Properties
Coefficients and Line Element
The first fundamental form of a surface, in a parametric representation with coordinates u and v, is expressed in standard notation asI = E \, du^2 + 2F \, du \, dv + G \, dv^2,
where the coefficients E, F, and G are smooth functions of the parameters (u, v). This quadratic form was originally introduced by Carl Friedrich Gauss in his 1827 work Disquisitiones generales circa superficies curvas, using parameters p and q and the expression E \, dp^2 + 2F \, dp \, dq + G \, dq^2 to describe infinitesimal distances on curved surfaces.[1][17] The line element ds, representing the infinitesimal arc length along a curve on the surface, is given by
ds = \sqrt{E \, du^2 + 2F \, du \, dv + G \, dv^2},
which quantifies distances within the tangent plane at each point and serves as the basis for intrinsic measurements independent of the embedding space. This formulation arises as the induced metric from the Euclidean inner product on the partial derivatives of the parametrization.[17] The coefficient F vanishes, i.e., F = 0, when the parametrization is orthogonal, meaning the coordinate curves (lines of constant u or v) are perpendicular in the tangent plane, which simplifies computations of lengths and angles. In such cases, the form reduces to I = E \, du^2 + G \, dv^2.[17] Under a change of coordinates from (u, v) to (u', v'), the coefficients transform according to the tensorial law
g'_{ij} = \frac{\partial u^k}{\partial u'^i} \frac{\partial u^l}{\partial u'^j} g_{kl},
where g_{ij} denotes the metric components (with g_{11} = E, g_{12} = g_{21} = F, g_{22} = G), ensuring the line element ds^2 remains invariant and the intrinsic geometry is preserved. This contravariant transformation property underscores the first fundamental form's role as a Riemannian metric on the surface.[17]
Matrix and Tensor Representation
In local coordinates (u, v) on a parametrized surface \mathbf{r}(u, v), the first fundamental form is represented by the symmetric $2 \times 2 matrix G = \begin{pmatrix} E & F \\ F & G \end{pmatrix}, where E = \langle \mathbf{r}_u, \mathbf{r}_u \rangle, F = \langle \mathbf{r}_u, \mathbf{r}_v \rangle, and G = \langle \mathbf{r}_v, \mathbf{r}_v \rangle are the coefficients derived from the induced Euclidean inner product on the tangent space.[18] This matrix formulation allows for algebraic computations of lengths and angles via the quadratic form \mathbf{X}^T G \mathbf{Y}, where \mathbf{X} and \mathbf{Y} are tangent vectors expressed in the coordinate basis.[19] The matrix G is symmetric by construction, since F = \langle \mathbf{r}_u, \mathbf{r}_v \rangle = \langle \mathbf{r}_v, \mathbf{r}_u \rangle, and positive definite provided E > 0, G > 0, and \det G = EG - F^2 > 0.[18] The positive definiteness ensures that the inner product defines a proper Riemannian metric on the surface, while the determinant condition \det G > 0 corresponds to |\mathbf{r}_u \times \mathbf{r}_v|^2 > 0, guaranteeing that the parametrization is immersive and regular at each point.[19] This determinant plays a key role in surface area computations, as the infinitesimal area element is dA = \sqrt{\det G} \, du \, dv.[18] In abstract tensor notation, the first fundamental form is the covariant metric tensor g = g_{ij} \, dx^i \otimes dx^j, where the components g_{ij} are precisely the entries of G (with repeated indices summed over i, j = 1, 2) and dx^1 = du, dx^2 = dv.[20] The tensor g is symmetric (g_{ij} = g_{ji}) and positive definite, inducing the geometry on the tangent bundle. The contravariant inverse metric g^{ij} is given by the adjugate matrix G^{-1}, scaled by $1/\det G, and serves to raise tensor indices, for example, transforming a covariant vector to contravariant form via v^i = g^{ij} v_j.[20] As a symmetric positive definite matrix, G admits an orthogonal diagonalization G = P D P^T, where P is an orthogonal matrix whose columns are the eigenvectors, defining the principal directions in the parameter space, and D = \operatorname{diag}(\lambda_1, \lambda_2) has positive eigenvalues \lambda_1, \lambda_2 > 0. These eigenvalues quantify the stretch or scaling of the metric along the corresponding principal directions, where an orthogonal coordinate system aligns with the eigenvectors, simplifying the form to \lambda_1 du'^2 + \lambda_2 dv'^2 (with F' = 0).[21]Applications to Surface Measurements
Arc Lengths of Curves
The first fundamental form provides the intrinsic metric for computing the lengths of curves lying on a parametrized surface \mathbf{r}(u, v). For a smooth curve \gamma(t) = (u(t), v(t)) on the surface, where t ranges from a to b and the curve is traced with non-zero speed, the arc length L is given by the line integral L = \int_a^b \sqrt{E \left( \frac{du}{dt} \right)^2 + 2F \frac{du}{dt} \frac{dv}{dt} + G \left( \frac{dv}{dt} \right)^2 } \, dt, where E = \langle \mathbf{r}_u, \mathbf{r}_u \rangle, F = \langle \mathbf{r}_u, \mathbf{r}_v \rangle, and G = \langle \mathbf{r}_v, \mathbf{r}_v \rangle are the coefficients of the first fundamental form evaluated along the curve.[13] This formula arises from the induced Euclidean inner product on the tangent space, integrating the infinitesimal distances ds along the curve's tangent vector (u', v').[1] Special cases simplify the expression for coordinate curves. For a u-curve, where v is constant and thus dv/dt = 0, the arc length reduces to L = \int \sqrt{E} \, du, measuring distance along lines of constant v. Similarly, for a v-curve with u constant and du/dt = 0, it becomes L = \int \sqrt{G} \, dv, quantifying length along lines of constant u. These specializations highlight how the metric coefficients E and G determine the scaling of distances in the principal parameter directions.[13] Geometrically, this integral captures the intrinsic distance traveled along the curve as measured solely by the surface's metric, without reference to the ambient Euclidean space. It reflects the surface's local geometry through the first fundamental form, enabling the definition of shortest paths (geodesics) and other metric properties that are preserved under isometries of the surface.[13][1] The arc length is invariant under reparameterizations of the curve, as substituting a new parameter s = s(t) with s' > 0 yields the same value due to the chain rule in the integrand, ensuring the total length depends only on the curve's image, not its traversal speed. Furthermore, this length is independent of the surface's embedding in \mathbb{R}^3, depending exclusively on the first fundamental form, which encodes all intrinsic metric information.[13]Areas of Regions
The area of a region on a parametrized surface is determined using the first fundamental form via an integral over the corresponding domain in the parameter space. For a surface patch parametrized by \mathbf{X}(u,v) and a region D in the (u,v)-plane, the surface area A is given by A = \iint_D \sqrt{EG - F^2} \, du \, dv, where E = \langle \mathbf{X}_u, \mathbf{X}_u \rangle, F = \langle \mathbf{X}_u, \mathbf{X}_v \rangle, and G = \langle \mathbf{X}_v, \mathbf{X}_v \rangle are the coefficients of the first fundamental form.[13] This formula arises from the geometry of the tangent plane, as \sqrt{EG - F^2} = \| \mathbf{X}_u \times \mathbf{X}_v \|, the magnitude of the cross product of the partial derivatives, which measures the area of the parallelogram they span.[13] The factor \sqrt{EG - F^2} corresponds to the square root of the determinant of the metric tensor associated with the first fundamental form, scaling the infinitesimal area element du \, dv in the parameter domain to reflect the actual surface measure.[22] In the special case of an orthogonal parametrization, where the coordinate curves are perpendicular and thus F = 0, the area integral simplifies to A = \iint_D \sqrt{EG} \, du \, dv. [13] This form highlights the product of the scaling factors along the orthogonal directions. The computed area is independent of the choice of parametrization, as changes in coordinates preserve the value of the integral through the Jacobian determinant, underscoring the intrinsic nature of the first fundamental form in defining surface areas.[13]Illustrative Example: The Unit Sphere
A standard parameterization of the unit sphere S^2 is given by \mathbf{X}(u,v) = (\sin v \cos u, \sin v \sin u, \cos v), where $0 \leq u \leq 2\pi is the azimuthal angle and $0 \leq v \leq \pi is the colatitude.$$] [23] The coefficients of the first fundamental form for this parameterization are E = \sin^2 v, F = 0, and G = 1, yielding the line element[ ds^2 = \sin^2 v , du^2 + dv^2. The [determinant](/page/Determinant) of the [metric tensor](/page/Metric_tensor) $\mathbf{G}$ is $\det \mathbf{G} = EG - F^2 = \sin^2 v$.$$\] [](https://math.etsu.edu/multicalc/prealpha/Chap3/Chap3-8/printversion.pdf) To compute the length of a curve on the sphere using the first fundamental form, consider the equator, given by the path with constant $v = \pi/2$ and $u$ varying from 0 to $2\pi$. Substituting into the line element gives $ds = \sin(\pi/2) \, du = du$, so the [arc length](/page/Arc_length) is \[ L = \int_0^{2\pi} du = 2\pi. $$$$\] [](https://math.etsu.edu/multicalc/prealpha/Chap3/Chap3-8/printversion.pdf) The first fundamental form also enables computation of surface areas via the area element $\sqrt{\det \mathbf{G}} = \sin v$. For the entire unit sphere, the surface area is \[ A = \int_0^{2\pi} \int_0^\pi \sin v \, dv \, du = \left[ \int_0^{2\pi} du \right] \left[ -\cos v \Big|_0^\pi \right] = (2\pi)(2) = 4\pi, confirming the classical result for the surface area of the unit sphere.$$] [23]