Fact-checked by Grok 2 weeks ago

Second fundamental form

In , the second fundamental form is a defined on the of a smooth surface embedded in , quantifying the extrinsic by measuring how the surface deviates from its tangent plane in the normal direction. For a parametrized surface given by a position \mathbf{r}(u,v), it is expressed in coordinates as II = L\, du^2 + 2M\, du\, dv + N\, dv^2, where the coefficients L = \mathbf{r}_{uu} \cdot \mathbf{n}, M = \mathbf{r}_{uv} \cdot \mathbf{n}, and N = \mathbf{r}_{vv} \cdot \mathbf{n} involve the second partial derivatives of \mathbf{r} dotted with the unit normal vector \mathbf{n}. This form, introduced by in his 1827 work Disquisitiones generales circa superficies curvas, contrasts with the intrinsic , which captures metric properties like lengths and angles on the surface itself. The second fundamental form plays a central role in determining key curvature invariants, such as the normal curvature \kappa_n = II / I along a direction, where I is the first fundamental form, providing a signed measure of bending perpendicular to the tangent plane. Its eigenvalues correspond to the principal curvatures, which describe the maximum and minimum curvatures at a point, while the Gaussian curvature K, the product of these principal curvatures, emerges as the determinant of the matrix relating the second and first fundamental forms, linking extrinsic and intrinsic geometry via Gauss's Theorema Egregium. For a surface where the second fundamental form vanishes identically, the surface must be planar, highlighting its role in classifying flat versus curved embeddings. In higher dimensions, the second fundamental form generalizes to in Riemannian manifolds, serving as the shape operator that encodes how the curves relative to the ambient space, with applications in and . Its symmetry ensures it defines a , facilitating spectral decompositions essential for studying surface evolution, minimal surfaces, and variational problems.

Surfaces in Euclidean three-space

Historical context and motivation

The study of curved surfaces in three-dimensional space traces its roots to the eighteenth century, with significant precursors in the works of mathematicians like Leonhard Euler and Gaspard Monge. Euler explored the curvature of surfaces through osculating planes and developable surfaces in his 1760 paper "Recherches sur la courbure des surfaces," laying early groundwork for understanding how surfaces deviate from flatness. Monge, often regarded as the father of differential geometry, advanced the analysis of curved surfaces in his 1785 work "Mémoire sur la théorie des déblais et des remblais," where he introduced concepts like lines of curvature and partial differential equations for surface properties, influencing the geometric study of embeddings in Euclidean space. The second fundamental form was formally introduced by in his seminal 1827 paper "Disquisitiones generales circa superficies curvas," presented to the Royal Society of . In this work, Gauss developed a to quantify the extrinsic of surfaces, building on but distinctly advancing the earlier French school of led by Monge. Gauss's approach emphasized the local embedding of surfaces, distinguishing it from prior efforts by providing a systematic tool for measuring bending independent of coordinate choices. Conceptually, the second fundamental form arises from the need to capture how a surface bends away from its tangent in the ambient space, contrasting with the , which encodes the intrinsic metric structure measurable solely from within the surface. For instance, a exhibits pronounced extrinsic as points nearby deviate sharply from the tangent , while a shows none, highlighting the form's role in distinguishing embedded shapes geometrically. This extrinsic perspective motivates its study by revealing how local deviations influence global form, such as in the of abstract manifolds. In differential geometry, the second fundamental form serves as a crucial bridge between the local properties of a surface's embedding and its overall shape, enabling the derivation of key invariants like principal curvatures. Its applications extend to physics, particularly in the theory of minimal surfaces where zero mean curvature—derived from this form—describes equilibrium shapes like soap films. In engineering, it underpins shell theory for analyzing thin curved structures under load, informing designs in architecture and aerospace by quantifying bending stresses.

Definition via parametrization

Consider a oriented surface S in three-space \mathbb{R}^3 given by a parametrization \mathbf{r}: U \to \mathbb{R}^3, where U \subset \mathbb{R}^2 is an and \mathbf{r}(u,v) is , meaning the partial derivatives \mathbf{r}_u = \frac{\partial \mathbf{r}}{\partial u} and \mathbf{r}_v = \frac{\partial \mathbf{r}}{\partial v} satisfy \mathbf{r}_u \times \mathbf{r}_v \neq \mathbf{0}. These partial derivatives form a basis for the tangent plane T_p S at each point p = \mathbf{r}(u,v) \in S. The unit normal vector \mathbf{n} to S at p is then defined as \mathbf{n} = \frac{\mathbf{r}_u \times \mathbf{r}_v}{|\mathbf{r}_u \times \mathbf{r}_v|}, providing an orientation for the surface. The second fundamental form II at p is a symmetric bilinear form on the tangent plane T_p S, measuring the extrinsic curvature of S in the normal direction. For tangent vectors X, Y \in T_p S, it is defined as II(X, Y) = -\langle d\mathbf{n}(X), Y \rangle, where d\mathbf{n}: T_p S \to T_{\mathbf{n}(p)} S^2 is the differential of the Gauss map \mathbf{n}: S \to S^2, and \langle \cdot, \cdot \rangle denotes the Euclidean inner product in \mathbb{R}^3. Equivalently, since \mathbb{R}^3 is equipped with its flat Levi-Civita connection \nabla, we have II(X, Y) = \langle \nabla_X Y, \mathbf{n} \rangle, where \nabla_X Y is the directional derivative of the vector field Y (extended suitably off the surface) along a curve \sigma: (-\epsilon, \epsilon) \to S with \sigma(0) = p and \sigma'(0) = X, projected onto the normal \mathbf{n}. This equivalence follows from the fact that the tangential component of \nabla_X Y lies in T_p S and is orthogonal to \mathbf{n}, while the full decomposition \nabla_X Y = (\nabla_X Y)^\top + II(X,Y) \mathbf{n} isolates the normal part. In the local coordinates (u,v), the second fundamental form is expressed as the quadratic form II = L \, du^2 + 2M \, du \, dv + N \, dv^2, with coefficients \begin{align*} L &= \langle \mathbf{r}{uu}, \mathbf{n} \rangle, \ M &= \langle \mathbf{r}{uv}, \mathbf{n} \rangle, \ N &= \langle \mathbf{r}_{vv}, \mathbf{n} \rangle, \end{align*} where \mathbf{r}_{uu} = \frac{\partial^2 \mathbf{r}}{\partial u^2}, \mathbf{r}_{uv} = \frac{\partial^2 \mathbf{r}}{\partial u \partial v}, and \mathbf{r}_{vv} = \frac{\partial^2 \mathbf{r}}{\partial v^2} are the second partial derivatives of the parametrization. These coefficients, introduced by Gauss in his classical treatment of surface curvature, represent the matrix \begin{pmatrix} L & M \\ M & N \end{pmatrix} of II with respect to the basis \{\mathbf{r}_u, \mathbf{r}_v\}, defining a quadratic form on T_p S. The second fundamental form is independent of the choice of parametrization in the tensorial sense: under a change of coordinates (u,v) \to (\tilde{u}, \tilde{v}), the coefficients transform via the Jacobian matrix of the coordinate change, ensuring that II behaves as a covariant (0,2)-tensor on S, contravariantly weighted by the of the first form's . This tensorial nature guarantees that geometric quantities derived from II are well-defined on the surface.

Notations and conventions

The second fundamental form, often denoted as II, can be expressed in classical notation as the quadratic form II = L \, du^2 + 2M \, du \, dv + N \, dv^2, where the coefficients L = \mathbf{r}_{uu} \cdot \mathbf{n}, M = \mathbf{r}_{uv} \cdot \mathbf{n}, and N = \mathbf{r}_{vv} \cdot \mathbf{n} are computed from the second partial derivatives of the position \mathbf{r}(u,v) and the unit \mathbf{n}. This form, originating from early texts, emphasizes the bilinear nature of the form along coordinate directions. In contrast, modern tensor notation represents II as II = h_{ij} \, du^i \, du^j, with components h_{ij} = \langle \partial^2 \mathbf{r} / \partial u^i \partial u^j, \mathbf{n} \rangle for indices i,j = 1,2 corresponding to coordinates u^1 = u, u^2 = v, providing a more general framework for abstract manifolds. In physicist's conventions, particularly , adopts index notation II = b_{\alpha\beta} \, du^\alpha \, du^\beta, where b_{\alpha\beta} = \partial^2 r^\gamma / \partial u^\alpha \partial u^\beta \, n_\gamma employs Einstein summation over repeated indices \gamma in the embedding space, facilitating computations in curved spacetimes or material deformations. This notation aligns with tensor analysis in physics, where the form describes extrinsic of hypersurfaces, such as spacelike slices in , and surface stresses in . The choice of normal direction introduces a sign convention for II, as replacing \mathbf{n} with -\mathbf{n} negates all components, thereby flipping the signs of derived mean curvature while leaving Gaussian curvature invariant. For closed surfaces, the outward normal typically yields positive mean curvature for convex shapes like spheres, whereas the inward normal reverses this, impacting interpretations in applications such as fluid interfaces or gravitational horizons.
NotationFormCoefficients
ClassicalL \, du^2 + 2M \, du \, dv + N \, dv^2L = \mathbf{r}_{uu} \cdot \mathbf{n}, etc.
Tensorh_{ij} \, du^i \, du^jh_{ij} = \langle \partial^2 \mathbf{r} / \partial u^i \partial u^j, \mathbf{n} \rangle[](https://www.cmu.edu/biolphys/deserno/pdf/diff_geom.pdf)
Physicistb_{\alpha\beta} \, du^\alpha \, du^\betab_{\alpha\beta} = r^\gamma_{,\alpha\beta} \, n_\gamma[](https://faculty.etsu.edu/gardnerr/5310/5310pdf/dg1-5.pdf)
In a coordinate-free , serves as a symmetric , or section of \mathrm{Hom}([TS](/page/TS) \otimes [TS](/page/TS), \mathbb{R}), intrinsically tied to the choice of for scalar valuation in one embeddings.

Geometric interpretation and properties

The shape operator

The Gauss map of an oriented surface S embedded in three-space is the smooth mapping n: S \to S^2 that sends each point p \in S to its unit vector n(p) at that point, providing a way to associate the local of the surface with points on the unit . This map captures the extrinsic geometry of the surface by tracking how the direction varies across S. The differential of the Gauss map, dn_p: T_p S \to T_{n(p)} S^2, takes values in the to the at n(p), which can be naturally identified with T_p S via the orthogonal onto the (since n is perpendicular to T_p S). The shape operator, also known as the Weingarten map, is the linear S_p: T_p S \to T_p S defined at each point p \in S by S_p(v) = -\nabla_v n, where \nabla_v n denotes the of the normal field n in the direction of the v \in T_p S (in the ambient , this is simply the v \cdot n). This operator was first introduced by Julius Weingarten in his study of developable surfaces. Note that the for S_p may vary across texts; some define it with the positive sign \nabla_v n, but the negative sign ensures consistency with the standard definition of the second fundamental form as a measure of . The image of S_p lies in T_p S because the unit length condition \langle n, n \rangle = 1 implies \langle \nabla_v n, n \rangle = 0, so \nabla_v n is orthogonal to n and thus tangent to S. The shape operator relates directly to the second fundamental form \mathrm{II} via the first fundamental form \mathrm{I}, the induced metric on S: specifically, \mathrm{II}(v, w) = \mathrm{I}(S_p(v), w) for all v, w \in T_p S. Since \mathrm{II} is symmetric as a bilinear form, it follows that S_p is self-adjoint with respect to \mathrm{I}, meaning \mathrm{I}(S_p(v), w) = \mathrm{I}(v, S_p(w)) for all tangent vectors v, w. In local coordinates \{u^i\} on S, if the matrix of \mathrm{I} is (g_{ij}) and the matrix of \mathrm{II} is (h_{ij}), then the matrix representation of S_p is given by S^j_i = g^{jk} h_{ki}, where g^{jk} is the inverse metric. The eigenvalues of S_p correspond to the principal curvatures of the surface at p, though their explicit computation and interpretation are addressed elsewhere. Geometrically, the shape operator quantifies how tangent vectors induce rotations in the normal direction, thereby describing the bending of the surface away from its tangent plane. For a v, S_p(v) indicates the direction and magnitude of the change in when moving along v, providing a of the surface's in the ambient space.

Curvatures derived from the second fundamental form

The principal curvatures of a surface at a point are the eigenvalues \kappa_1 and \kappa_2 (with \kappa_1 \geq \kappa_2) of the shape operator S, which encodes the relative to the . These eigenvalues represent the maximum and minimum values of the in principal directions, where the normal curvature \kappa_n(v) for a v is given by \kappa_n(v) = \mathrm{II}(v, v), and the principal directions are the corresponding eigenvectors of S. In matrix form, if the has matrix g = \begin{pmatrix} E & F \\ F & G \end{pmatrix} and the second fundamental form has matrix b = \begin{pmatrix} L & M \\ M & N \end{pmatrix}, then the matrix of S is b g^{-1}, and the principal curvatures satisfy the \det(b g^{-1} - \kappa I) = 0, or equivalently, \kappa^2 - \operatorname{tr}(b g^{-1}) \kappa + \det(b g^{-1}) = 0. The solutions are \kappa_{1,2} = \frac{\operatorname{tr}(b g^{-1}) \pm \sqrt{ [\operatorname{tr}(b g^{-1})]^2 - 4 \det(b g^{-1}) }}{2}, where \operatorname{tr}(b g^{-1}) = (EG + NL - 2FM)/(EG - F^2) and \det(b g^{-1}) = (LN - M^2)/(EG - F^2). The Gaussian curvature K is the product of the principal curvatures, K = \kappa_1 \kappa_2 = \det(b g^{-1}) = \frac{LN - M^2}{EG - F^2}, providing a measure of the intrinsic product of bending in orthogonal principal directions; this quantity was originally defined by Gauss as the "measure of curvature" using analogous second-order surface quantities. The mean curvature H is the average of the principal curvatures, H = \frac{\kappa_1 + \kappa_2}{2} = \frac{1}{2} \operatorname{tr}(b g^{-1}) = \frac{EN + GL - 2FM}{2(EG - F^2)}, capturing the average extrinsic bending of the surface. For a of R, parametrized in spherical coordinates, the is I = R^2 (d\theta^2 + \sin^2\theta \, d\phi^2) and the second fundamental form is \mathrm{II} = R (d\theta^2 + \sin^2\theta \, d\phi^2), so S = (1/R) I (up to the metric), yielding \kappa_1 = \kappa_2 = 1/R, K = 1/R^2, and H = 1/R. In contrast, for an infinite of R, parametrized by axial coordinate u and angular coordinate v, the forms are I = du^2 + R^2 dv^2 and \mathrm{II} = R dv^2, so the matrix of S is diagonal with entries 0 and $1/R, giving principal curvatures \kappa_1 = 1/R, \kappa_2 = 0, K = 0, and H = 1/(2R).

Gauss's theorema egregium

Gauss's , established in Carl Friedrich Gauss's 1827 paper Disquisitiones generales circa superficies curvas, asserts that the K of an embedded surface in Euclidean three-space is an intrinsic invariant, determined exclusively by the I and independent of the specific embedding. Specifically, the extrinsic expression K = \frac{\det(\mathrm{II})}{\det(\mathrm{I})} = \frac{LN - M^2}{EG - F^2}, where \mathrm{II} denotes the with coefficients L, M, N and I has coefficients E, F, G, equals the of the induced Riemannian on the surface, which can be computed using only the derived from I. The proof utilizes the Gauss-Codazzi equations, which connect the intrinsic ( \Gamma^k_{ij} from I) to \mathrm{II}. When substituted into the components of the R^i_{jkl} = \partial_k \Gamma^i_{lj} - \partial_l \Gamma^i_{kj} + \Gamma^i_{km} \Gamma^m_{lj} - \Gamma^i_{lm} \Gamma^m_{kj}, the extrinsic contributions from \mathrm{II} cancel in the formula K = \frac{R_{1212}}{\det(I)}, yielding an intrinsic expression in terms of E, F, G and their first partial derivatives. For instance, K \sqrt{EG - F^2} = \frac{\partial}{\partial u} \left( \frac{\partial_u G - \partial_v F}{2\sqrt{EG - F^2}} \right) - \frac{\partial}{\partial v} \left( \frac{\partial_v E - \partial_u F}{2\sqrt{EG - F^2}} \right) + quadratic terms in the , confirming the embedding independence. This result implies that surfaces sharing the same intrinsic metric possess identical ; notably, both a and a exhibit K = 0, enabling development of the onto the without metric distortion. Gauss's 1827 demonstration underscored the separation of intrinsic surface from extrinsic properties, fundamentally influencing the development of .

Generalizations to submanifolds

Hypersurfaces in Riemannian manifolds

In a Riemannian manifold (\tilde{M}^n, g) of dimension n, a hypersurface M^{n-1} is a codimension-one immersed submanifold equipped with the induced metric \bar{g} on its tangent bundle TM. The hypersurface is assumed to be oriented, allowing the choice of a smooth unit normal vector field \nu: M \to T\tilde{M} such that g(\nu, \nu) = 1 and \nu is orthogonal to TM at each point. This setup generalizes the classical notion of surfaces in Euclidean space to curved ambient geometries, where the second fundamental form captures the extrinsic curvature relative to the ambient connection. The second fundamental form II on M is defined as the symmetric bilinear form II: TM \times TM \to \mathbb{R} given by II(X, Y) = g(\nabla_X Y, \nu) for tangent vectors X, Y \in TM, where \nabla denotes the on \tilde{M}. Equivalently, since \nabla_X \nu is tangent to M (as g(\nabla_X \nu, \nu) = 0), one has II(X, Y) = -g(\nabla_X \nu, Y). In local coordinates \{ \partial_i \} on M, the components are h_{ij} = II(\partial_i, \partial_j) = g(\nabla_{\partial_i} \partial_j, \nu), forming a symmetric (0,2)-tensor with respect to \bar{g}. This measures the normal component of the ambient , distinguishing how M bends within \tilde{M}. The shape operator (or Weingarten map) S: TM \to TM is the tangent vector field defined by S(X) = -\nabla_X \nu, which is the tangential projection of the of . It is a with respect to \bar{g}, satisfying II(X, Y) = \bar{g}(S(X), Y). In matrix form with respect to an , the eigenvalues of S are the principal curvatures \kappa_1, \dots, \kappa_{n-1}. For the classical case of a in \mathbb{R}^n, this recovers the standard second fundamental form of surfaces in three-space when n=3. From the shape operator, the mean curvature scalar is H = \frac{1}{n-1} \operatorname{trace}_{\bar{g}} S = \frac{1}{n-1} \sum \kappa_i, representing the average extrinsic , while a Gaussian curvature analog is \det S = \prod \kappa_i (for n=3, this coincides with the intrinsic via the Gauss equation). In curved spaces, such as a in S^n, II quantifies deviations from great hyperspheres, which are totally (where II = 0). These quantities highlight how embedding interacts with the ambient .

Arbitrary codimension and normal bundle

Consider a \Sigma^k isometrically immersed in a M^n with m = n - k > 1. The NM over \Sigma is the to the T\Sigma in TM|_{\Sigma}, equipped with the induced normal . A local orthonormal frame for the normal bundle is denoted \{\nu_a\}_{a=1}^m, where each \nu_a is a unit normal vector field along \Sigma. The second fundamental form extends to a vector-valued II: T\Sigma \times T\Sigma \to NM, defined by II(X, Y) = (\nabla_X Y)^\perp for tangent vectors X, Y \in T\Sigma, where \nabla is the on M and (\cdot)^\perp denotes the orthogonal projection onto NM. This form is symmetric, II(X, Y) = II(Y, X), and captures the extrinsic of \Sigma in directions normal to the . As a tensor, it belongs to \mathrm{Hom}(T\Sigma \otimes T\Sigma, NM). The scalar components with respect to the normal frame are given by II^a(X, Y) = \langle \nabla_X Y, \nu_a \rangle = \langle II(X, Y), \nu_a \rangle for a = 1, \dots, m. Associated to each normal direction \nu_a, there is a shape operator S_a: T\Sigma \to T\Sigma defined by S_a(X) = -\mathrm{proj}_{T\Sigma} (\nabla_X \nu_a), which is the tangential projection of the of \nu_a. This operator is , and the components of the second fundamental form relate to it via II^a(X, Y) = \langle S_a(X), Y \rangle. The collection \{S_a\}_{a=1}^m generalizes the single shape operator from the case, reflecting the higher-dimensional freedom in the normal space. In local coordinates \{x^i\}_{i=1}^k on \Sigma, with tangent basis \{\partial_i\} and the normal frame \{\nu_a\}, the components take the form h^a_{ij} = \langle \nabla_{\partial_i} \partial_j, \nu_a \rangle = II^a(\partial_i, \partial_j). These h^a_{ij} are the coefficients of the vector-valued form in the normal basis, and curvatures generalize accordingly: for instance, the mean curvature vector involves traces \mathrm{tr}(S_a) \nu_a over the tangent indices for each a. This structure allows the second fundamental form to encode the full spectrum of normal curvatures in higher codimension. A representative example occurs for a curve \gamma: I \to \mathbb{R}^3, which has k=1 and m=2. Here, the normal bundle is spanned by the principal N and binormal B of the Frenet frame \{T, N, B\}, where T = \dot{\gamma}/|\dot{\gamma}|. The second fundamental form II(T, T) projects to \kappa N in the N-direction (with \kappa), while the binormal component relates to the torsion \tau through the normal connection on the frame; together, these components capture both the normal curvature and binormal twisting of the in \mathbb{R}^3.

Gauss-Codazzi equations

The Gauss-Codazzi equations constitute a system of partial differential equations that serve as compatibility conditions for the immersion of a Riemannian submanifold \Sigma into an ambient Riemannian manifold M. These equations relate the intrinsic Riemann curvature tensor R^\Sigma of \Sigma, induced by its metric, to the Riemann curvature tensor R^M of the ambient space and the second fundamental form II of the immersion, which captures the extrinsic geometry. For a submanifold of arbitrary codimension, the second fundamental form II: T\Sigma \times T\Sigma \to NM takes values in the normal bundle NM, and in an orthonormal frame \{ \nu^a \}_{a=1}^m for NM, it decomposes as II(X,Y) = \sum_{a=1}^m II^a(X,Y) \nu^a, where II^a are the scalar components. The Gauss equation expresses the difference between the intrinsic and ambient curvatures in terms of the second form. For tangent vectors X, Y, Z, W \in T\Sigma, it states \langle R^\Sigma(X,Y)Z, W \rangle = \langle R^M(X,Y)Z, W \rangle + \sum_{a=1}^m \left[ II^a(X,W) II^a(Y,Z) - II^a(X,Z) II^a(Y,W) \right], where \langle \cdot, \cdot \rangle denotes the on M. This equation projects the ambient onto the tangent space of \Sigma, with the second term arising from the normal components of the covariant derivatives. In the special case of a hypersurface (codimension m=1), where NM is trivialized by a unit normal \nu and II(X,Y) = h(X,Y) \nu with h the scalar second form, the equation simplifies to \langle R^\Sigma(X,Y)Z, W \rangle = \langle R^M(X,Y)Z, W \rangle + h(X,W) h(Y,Z) - h(X,Z) h(Y,W). For ambient space (R^M = 0), the Gauss equation reduces to the intrinsic being expressible via the determinant of the second form relative to the . The Codazzi equation governs the covariant derivative of the second fundamental form along the submanifold, ensuring its symmetry with respect to the induced connection. It is given by (\nabla_X II)(Y,Z) - (\nabla_Y II)(X,Z) = (R^M(X,Y)Z)^\perp, where \nabla denotes the induced Levi-Civita connection on \Sigma extended to the normal bundle via the second fundamental form, and (\cdot)^\perp projects onto the normal bundle NM. In local coordinates (u^i) on \Sigma, with components h^a_{ij} = II^a(\partial_i, \partial_j), the a-component is (\nabla_k h^a_{ij}) - (\nabla_j h^a_{ik}) = \langle R^M(\partial_k, \partial_j) \partial_i, \nu^a \rangle, where \nabla_k h^a_{ij} = \partial_k h^a_{ij} - \Gamma^l_{k i} h^a_{l j} - \Gamma^l_{k j} h^a_{i l}. For hypersurfaces in Euclidean space (R^M = 0), the equation simplifies to the Mainardi-Codazzi form, such as \partial_v e - \partial_u f = \Gamma^1_{12} e + \cdots in terms of the second fundamental form coefficients e, f, g. This equation captures the normal component of the ambient curvature's action on tangent vectors. These equations have profound implications for the geometry of submanifolds. They provide necessary and sufficient conditions for the local isometric embeddability of a into a higher-dimensional with prescribed and second fundamental form, as guaranteed by the fundamental for submanifolds. In the classical setting of surfaces in \mathbb{R}^3 (codimension 1, flat ambient), the tangential projection of the Gauss equation recovers Gauss's , expressing the intrinsically via the alone. The full system ensures the consistency of intrinsic and extrinsic structures, with violations implying non-embeddability. A general proof of the Gauss-Codazzi equations derives from the integrability of the immersion, viewed through the flat connection on the Whitney sum bundle T\Sigma \oplus N\Sigma. The ambient Levi-Civita connection \tilde{\nabla} on TM induces a connection on T\Sigma \oplus N\Sigma that splits into the induced connection \nabla on T\Sigma, the normal connection D on N\Sigma, and the second fundamental form II as the cross terms: \tilde{\nabla}_X (Y + \xi) = \nabla_X Y + D_X \xi + II(X,Y) for Y \in T\Sigma, \xi \in N\Sigma. The curvature of the induced connection on T\Sigma \oplus N\Sigma coincides with the restriction of the ambient curvature \tilde{R} to sections of this bundle. Computing the components of this curvature with respect to the splitting T\Sigma \oplus N\Sigma yields the Gauss equation (tangential-tangential component), Codazzi equation (tangent-normal component), and Ricci equation (normal-normal component) as the integrability conditions.

References

  1. [1]
    [PDF] Chapter 5. The Second Fundamental Form
    2. The second fundamental form. This is a certain bilinear form L : TpM ×TpM → IR associated in a natural way with the Weingarten map. We now describe the ...
  2. [2]
    3.3 Second fundamental form II (curvature) - MIT
    In order to quantify the curvatures of a surface $ S$ , we consider a curve $ C$ on $ S$ which passes through point $ P$ as shown in Fig. 3.6.
  3. [3]
    [PDF] DIFFERENTIAL GEOMETRY OF CURVES AND SURFACES 5. The ...
    The quadratic form corresponding to5 −d(N)P is the second fundamental form. ... Gauss, in his paper entitled Disquisitiones generales circa superficies curvas, ...
  4. [4]
    [PDF] General Investigations of Curved Surfaces - Project Gutenberg
    Disquisitiones generales circa superficies curvas auctore Carolo Friderico Gauss. Gottingæ. Typis Dieterichianis. 1828. II. In Monge's Application de l'analyse ...
  5. [5]
    [PDF] m435: introduction to differential geometry
    The idea of the second fundamental form is to measure, in R3, how Σ curves away from its tangent plane at a given point. The first fundamental form is an.
  6. [6]
    Gaspard Monge - Biography - MacTutor - University of St Andrews
    Gaspard Monge is considered the father of differential geometry because of his work Application de l'analyse à la géométrie where he introduced the concept ...
  7. [7]
    [PDF] disquisitiones generales circa superficies curvas - Numdam
    Carl Friedrich Gauss Carl Friedrich Gauss. Disquisitiones generales circa superficies curvas generales circa superficies curvas superficies c by. Peter ...
  8. [8]
    [PDF] Basics of the Differential Geometry of Surfaces - CIS UPenn
    The second fundamental form was introduced by Gauss in 1827. Unlike the first fundamental form, the second fundamental form is not necessarily positive or ...
  9. [9]
    [PDF] A method of shell theory in determination of the surface from ...
    Feb 17, 2007 · We introduce the tensor which maps the Cartesian plane into the tangent plane of the surface. Then by analogy to the polar.
  10. [10]
    None
    Below is a merged summary of the second fundamental form from Manfredo P. do Carmo's book, "Differential Geometry of Curves and Surfaces," based on the provided segments. To retain all information in a dense and organized manner, I will use a combination of narrative text and a table in CSV format to capture details such as definitions, setups, coefficients, expressions, page references, and URLs. The narrative will provide an overview, while the table will consolidate specific details for clarity and completeness.
  11. [11]
    [PDF] DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces
    Compute the second fundamental form IIP of the following parametrized surfaces. Then calculate the matrix of the shape operator, and determine H and K. a ...
  12. [12]
    [PDF] Notes on Differential Geometry
    May 3, 2004 · From this we now get the second fundamental form bij and its determinant b: bij := ei,j · ٌ = أ. −2 cos(2φ) 2r sin(2φ). 2r sin(2φ) 2r2 cos(2φ) !
  13. [13]
    [PDF] Differential Geometry in Physics - UNCW
    of the second fundamental form as we show below. The second fundamental form measures the distance from a point on the surface to the tangent plane at a ...
  14. [14]
    [PDF] Differential Geometry in Physics
    There are many excellent texts in differential geometry but very few have an early introduction to differential forms and their applications to physics. It ...
  15. [15]
    Ueber eine Klasse auf einander abwickelbarer Flächen. - EuDML
    Weingarten, J.. "Ueber eine Klasse auf einander abwickelbarer Flächen.." Journal für die reine und angewandte Mathematik 59 (1861): 382-393. <http://eudml ...Missing: pdf | Show results with:pdf
  16. [16]
    [PDF] Gauß Curvature in Terms of the First Fundamental Form - Math
    Mar 26, 2004 · We prove Gauß's Theorem and compute some examples for a general metric. ... We find the Codazzi Equation and show Gauß's Theorema Egregium.
  17. [17]
    [PDF] Unit 14: Theorema Egregium - Harvard Mathematics Department
    14.1. In 1827, Karl Friedrich Gauss proved the “theorema egregium”. Is curvature determined by distance measurements within the geometry alone, without ...Missing: sketch | Show results with:sketch
  18. [18]
    [PDF] curvature and the theorema egregium of gauss
    Dec 31, 2015 · The second fundamental form at p is defined to be H. The uniqueness of H implies that it is an extrinsic geometric invariant tensor of S. The ...Missing: history | Show results with:history
  19. [19]
    [PDF] Riemannian Manifolds: An Introduction to Curvature
    After introducing some basic definitions and terminology concerning sub- manifolds, we define a tensor field called the second fundamental form, which ...
  20. [20]
    [PDF] Introduction to RIEMANNIAN GEOMETRY - Mathematics
    is called the second fundamental form of the oriented hypersurface M in the given local coordinates. Fix a point y = y(x) ∈ M and a nonzero tangent vector ...<|control11|><|separator|>
  21. [21]
    [PDF] mean curvature of riemannian hypersurface under the exponential ...
    Sep 28, 2020 · Define the second fundamental form A of Σ to be the vectored-valued bilinear form on TΣ. For X, Y ∈ TΣ. (2.1). A(X, Y )=(∇XY ). ⊥. Here the ...
  22. [22]
    [PDF] riemannian submanifolds
    Jul 7, 2013 · called the second fundamental form of the submanifold (or of the immersion). ... do Carmo, Stability of minimal surfaces in a 3-dimensional.
  23. [23]
    [PDF] Submanifold geometry - IME-USP
    Jun 8, 2016 · It follows that all principal curvatures have the same sign, namely, the second fundamental form is definite as a symmetric bilinear form, for ...<|control11|><|separator|>
  24. [24]
    [PDF] minimal submanifolds in higher codimension
    where h∂Ω denotes the second fundamental form of ∂Ω with respect to the inner unit normal. Since P n i=1 |D⊥Vi|2 and P n i=1 |DT Vi|2 are independent of.
  25. [25]
    [PDF] INTRODUCTION TO DIFFERENTIAL GEOMETRY - ETH Zürich
    Jan 11, 2011 · These are notes for the lecture course “Differential Geometry I” held by the second author at ETH Zürich in the fall semester 2010.