Fact-checked by Grok 2 weeks ago

Covariant derivative

In , the covariant derivative is a fundamental operator that extends the to fields and s on a smooth manifold, incorporating the manifold's intrinsic through an to ensure coordinate-independent differentiation. It measures the rate of change of a along a direction while accounting for the "twisting" of the manifold, defined abstractly by axioms including in the argument, the for tensors, and compatibility with the manifold's structure in local coordinates via . For a V along a c(t), the covariant derivative \frac{DV}{dt} satisfies \frac{D}{dt}(fV) = f'V + f\frac{DV}{dt} for scalar functions f and reduces to the ordinary \nabla_{\dot{c}}X when V is the restriction of an ambient field X. The concept was pioneered by Tullio Levi-Civita in 1917, who introduced it in the context of parallel transport on Riemannian manifolds to define a notion of "straightness" (geodesics) and specify the Riemann curvature tensor geometrically. In a pseudo-Riemannian manifold (M, g), the unique torsion-free, metric-compatible connection—the Levi-Civita connection—provides the standard covariant derivative, satisfying \nabla g = 0 (metric compatibility) and T(X, Y) = \nabla_X Y - \nabla_Y X - [X, Y] = 0 (torsion-freeness), with explicit components given by Christoffel symbols \Gamma^k_{ij} = \frac{1}{2} g^{k\ell} (\partial_i g_{j\ell} + \partial_j g_{i\ell} - \partial_\ell g_{ij}). This structure enables parallel transport, where a vector field V along a curve satisfies \frac{DV}{dt} = 0, preserving lengths and angles isometrically. Beyond , the covariant derivative is indispensable in , particularly in , where it governs the motion of particles and fields on curved , with the derived from the to describe gravitational effects without additional structures. It extends naturally to higher-rank tensors via the Leibniz rule, \nabla_X (T \otimes S) = (\nabla_X T) \otimes S + T \otimes (\nabla_X S), facilitating computations of and other geometric invariants essential for understanding manifold and dynamics.

Historical Development

Origins in Riemannian Geometry

The foundations of the covariant derivative emerged from efforts to develop an intrinsic geometry for manifolds, independent of any embedding in . In his 1854 habilitation lecture at the , titled "On the Hypotheses Which Lie at the Foundations of Geometry," introduced the concept of an n-dimensional manifold equipped with a , allowing for the study of and distances solely through measurements within the manifold itself. This framework generalized to higher dimensions, emphasizing that geometric properties could be defined intrinsically without reference to an ambient space, thereby setting the stage for differentiation rules that respect the manifold's structure. Building on Riemann's ideas, advanced the differentiation of s on curved manifolds in his 1869 paper published in the Journal für die reine und angewandte Mathematik. introduced symbols, now known as \Gamma^k_{ij}, which serve as coefficients to account for the variation of basis vectors along curved paths, enabling a consistent way to differentiate vectors while remaining to the manifold. These symbols appear in Christoffel's formula for the difference between the directional derivative D_X Y and the covariant derivative \nabla_X Y of a Y along a direction X, expressed in components as (\nabla_X Y)^k - (D_X Y)^k = \Gamma^k_{ij} X^i Y^j, where D_X Y = X^i \partial_i Y^k \partial_k represents the ordinary projected onto the coordinate basis. This correction term encapsulates the geometry's , ensuring the result stays within the . Gregorio Ricci-Curbastro further formalized these concepts within his development of absolute differential calculus, beginning around and culminating in key publications through 1900. In works such as his studies on hypersurfaces in Riemannian manifolds, Ricci introduced the operation of as a means to generalize tensor manipulations invariant under coordinate changes, integrating Christoffel's symbols into a systematic for tensor on curved spaces. This calculus provided the tools for handling derivatives that transform covariantly, laying the groundwork for broader applications in .

Key Contributions and Evolution

In 1917, Tullio Levi-Civita introduced the concept of a torsion-free affine connection compatible with a given metric tensor, which uniquely determines the Levi-Civita covariant derivative on Riemannian manifolds. This formulation provided a rigorous geometric interpretation of parallel transport and differentiation, resolving ambiguities in earlier uses of Christoffel symbols for general coordinate systems. Building on this foundation in the 1920s, Élie Cartan integrated the covariant derivative into his method of moving frames, where he developed connection forms to describe affine connections in terms of differential forms on frame bundles. Cartan's approach generalized the Levi-Civita connection to spaces with arbitrary torsion and curvature, enabling a coordinate-free treatment that influenced subsequent developments in differential geometry. His key works from this period, such as those on generalized spaces of affine connection, emphasized the role of these forms in defining covariant differentiation for tensorial objects. The covariant derivative also played a pivotal role in physics during this era. In 1918, proposed a of and , where he extended the metric-compatible connection to include , introducing a precursor to modern gauge covariant derivatives that transformed under local symmetries. This idea, though initially unsuccessful for unifying forces, laid groundwork for later . Meanwhile, Albert Einstein's 1915 formulation of relied on the covariant derivative to express the field equations in a generally covariant manner, ensuring physical laws hold under arbitrary coordinate transformations. Post-World War II advancements, beginning in the 1940s, saw the covariant derivative integrated into and theory. and developed the Chern-Weil theory of characteristic classes, using connections and their curvatures to define topological invariants of vector bundles, bridging with . Concurrently, and others incorporated connections into the study of coherent sheaves and algebraic vector bundles, facilitating the computation of groups and advancing the Grothendieck reformulation of . These efforts established the covariant derivative as a central tool in modern bundle theory, with applications extending to complex manifolds and holomorphic structures.

Motivation and Intuitive Understanding

Limitations of Partial Derivatives on Manifolds

In the context of manifolds, partial derivatives of tensor fields fail to yield tensorial objects because they implicitly assume that the basis vectors associated with the remain constant across the space. On a curved manifold, however, the basis vectors vary with position, leading to changes in the components of tensor fields that are not captured by ordinary differentiation. This results in expressions that transform non-tensorially under coordinate changes, meaning they depend on the specific choice of coordinates rather than intrinsic geometric properties. A concrete illustration of this limitation occurs on the sphere, where coordinates such as are used. Differentiating a , such as one representing to the surface, using partial derivatives with respect to these angular coordinates produces results that are not under rotations of the sphere. For instance, the apparent rate of change of the vector's components alters if the coordinate grid is rotated, reflecting the varying orientation of the coordinate basis rather than a true geometric . The issue arises explicitly in the transformation law for the partial derivative of a vector field under a coordinate change x' = x'(x). The components transform as \partial_i' V'^j = \frac{\partial x^k}{\partial x'^i} \frac{\partial x'^j}{\partial x^l} \partial_k V^l + V^l \frac{\partial^2 x'^j}{\partial x'^i \partial x^l}, where the second term is an extraneous contribution that violates the tensor transformation rule, introducing coordinate-dependent artifacts. This non-tensorial behavior motivated the development of Ricci's absolute differential calculus in the late 19th and early 20th centuries, which introduced covariant differentiation to ensure invariance and facilitate calculations on manifolds independent of coordinate choices.

Geometric Interpretation via Parallel Transport

In , parallel transport provides a geometric means to move vectors along on a manifold while preserving their "direction" relative to the manifold's structure. Specifically, a V along a c: (a, b) \to M is said to be parallel if its components remain constant in a natural adapted to the , meaning the transport does not introduce any or beyond the manifold's intrinsic . This process is defined such that the covariant derivative along the vanishes, ensuring the vector is transported without deviation from this constant-component condition. The covariant derivative \nabla_X Y of a vector field Y in the direction of X geometrically measures the extent to which Y fails to be ed along the integral curves of X. In other words, it quantifies the infinitesimal change in Y relative to the parallel transport rule, capturing how the manifold's causes vectors to evolve as they are moved. Visually, this is analogous to keeping a vector "level" while traversing a hilly surface: unlike simply dragging the vector (which would ignore the terrain's slope), parallel transport adjusts the vector to stay aligned with the local tangent plane, and the covariant derivative tracks any necessary correction due to the hills' twists. A striking example occurs on the unit S^2, where parallel transporting a around a closed loop, such as a circle at \theta_0 < \pi/2, results in —a net of the upon return to the starting point. This angle is $2\pi (1 - \cos \theta_0), directly quantifying the 's through the path dependence of the transport. In the infinitesimal along a parameterized by s, parallel transport satisfies \frac{dV}{ds} \big|_{\parallel} = 0, so the covariant derivative \nabla_V V = 0 precisely characterizes geodesics as curves where the tangent vector is parallel to itself.

Informal Approaches

Embedding into Euclidean Space

One informal approach to constructing the covariant derivative on a M involves embedding M isometrically as a into a higher-dimensional \mathbb{R}^n via a map F: M \to \mathbb{R}^n. This embedding allows the use of standard in the ambient flat space, followed by back to the of M. To define the covariant derivative \nabla_X Y of a Y on M in the direction of another X at a point p \in M, first extend Y to a Y^\text{ext} on an open neighborhood of F(p) in \mathbb{R}^n. The covariant derivative is then the orthogonal onto the T_{F(p)}M of the ambient : \nabla_X Y = \proj_{T_{F(p)}M} \left( dY^\text{ext}(F(p)) (X) \right), where dY^\text{ext}(F(p)) denotes the of Y^\text{ext} at F(p), and the projection ensures the result remains to M. This construction is particularly intuitive for visualizing the covariant derivative on surfaces embedded in \mathbb{R}^3. For a surface M \subset \mathbb{R}^3 with unit normal \nu(p) at p \in M, the projection operator is \Pi(p) = I - \nu(p) \nu(p)^T, so \nabla_X Y(p) = \Pi(p) \, dY^\text{ext}(p) X, which isolates the tangential component of the Euclidean derivative of the extended field. For instance, on the unit sphere S^2 \subset \mathbb{R}^3, this yields the tangential part of the ordinary derivative, aligning the change in a vector field with the sphere's intrinsic geometry rather than its extrinsic curvature in \mathbb{R}^3. While effective for submanifolds of , this embedding method does not apply directly to general abstract Riemannian manifolds, as not all such manifolds admit into \mathbb{R}^n for finite n. It underscores the distinction between extrinsic (dependent on the ambient ) and intrinsic (independent of it), where quantities like the covariant derivative can be defined solely on M without reference to \mathbb{R}^n. This perspective traces back to Carl Friedrich Gauss's foundational work on surfaces, where in his 1827 paper Disquisitiones generales circa superficies curvas, he established the , proving that the of a surface is an intrinsic invariant that can be computed from the metric without relying on its embedding in .

Vector Bundle Perspective

The TM of a smooth manifold M is a over M, where each fiber T_p M consists of tangent vectors at point p \in M, and smooth sections of TM correspond precisely to vector fields on M. This perspective frames the covariant derivative as a tool for differentiating these sections in a manner compatible with the bundle's , allowing comparison of vectors across different fibers without a global trivialization. A connection on the vector bundle E \to M (such as TM) provides a covariant differentiation operator by splitting the tangent space TE at each point into a direct sum of horizontal and vertical subbundles, TE = HE \oplus VE, where the vertical subbundle VE is tangent to the fibers and the horizontal subbundle HE defines directions of "parallel" transport. For a curve \gamma(t) in M, the horizontal lift to a path \tilde{\gamma}(t) in E ensures that a section along \gamma remains parallel if its derivative lies entirely in the horizontal direction, enabling the covariant derivative \nabla_X s of a section s along a vector X to extract the vertical component of the pushforward Ts(X). In the case of a trivial bundle, such as the over \mathbb{R}^n, the connection reduces to the ordinary , as fibers can be globally identified via a constant frame. However, for nontrivial bundles like the of the , the topology introduces twisting, where around a non-contractible loop fails to return to the starting vector, manifesting as a holonomy that the connection must account for. This setup aligns with the notion of an Ehresmann , which specifies the subspaces as a smooth complementary to the vertical one, allowing of sections while respecting the bundle's structure. The of the then arises as the obstruction to flatness, measuring how the fails to integrate to a and capturing the intrinsic twisting in the bundle's through the non-commutativity of iterated covariant derivatives.

Formal Definition

Affine Connections on Manifolds

In , the study of affine connections begins with the foundational structures of smooth manifolds. A smooth manifold M is a that locally resembles and is equipped with a , allowing for the definition of differentiable functions and maps. The TM \to M is the of all spaces T_p M at points p \in M, where each T_p M is a isomorphic to \mathbb{R}^n for \dim M = n. Smooth vector fields on M are sections of the , denoted \Gamma(TM), which can be viewed as derivations on the ring of smooth functions C^\infty(M) satisfying the Leibniz rule for products. An on a smooth manifold M is a \nabla: \Gamma(TM) \times \Gamma(TM) \to \Gamma(TM) that satisfies linearity in both arguments over \mathbb{R} and the Leibniz rule for scalar multiplication by smooth functions f \in C^\infty(M). Specifically, for vector fields X, Y \in \Gamma(TM), \nabla_{fX} Y = f \nabla_X Y, \quad \nabla_X (f Y) = (X f) Y + f \nabla_X Y. This structure provides a way to differentiate vector fields along directions specified by other vector fields, generalizing the notion of directional derivatives in . The operation of the connection is commonly denoted by \nabla_X Y or sometimes X \cdot Y, emphasizing its role in directional differentiation. Affine connections axiomatically capture the idea of along curves on the manifold, where a is parallel if its vanishes along the curve's direction. In the general setting, an extends to any smooth E \to M as a map \nabla: \Gamma(TM) \times \Gamma(E) \to \Gamma(E) satisfying the same linearity and Leibniz properties for sections \sigma \in \Gamma(E): \nabla_{fX} \sigma = f \nabla_X \sigma and \nabla_X (f \sigma) = (X f) \sigma + f \nabla_X \sigma. On a (M, g), where g is a , there exists a unique \nabla that is both torsion-free and metric-compatible, meaning \nabla g = 0. This unique connection is known as the , which uniquely determines the geometry compatible with the metric.

Covariant Derivative of Vector Fields

The covariant derivative provides a means to differentiate vector fields on a manifold, accounting for the geometry via an affine connection. For vector fields X and Y on a smooth manifold M, the covariant derivative \nabla_X Y represents the directional derivative of Y in the direction of X, producing another vector field tangent to M. This operation satisfies the axioms of an affine connection, such as \mathbb{R}-linearity in both arguments and the Leibniz rule \nabla_X (f Y) = (X f) Y + f \nabla_X Y for smooth functions f: M \to \mathbb{R}. In abstract terms, the local expression of \nabla_X Y involves differentiating the components of Y with respect to a varying local basis of the tangent spaces, adjusted by the connection coefficients to ensure the result lies in the appropriate tangent space. This adjustment is crucial on curved manifolds, where tangent spaces at different points are not canonically identified, unlike in flat space. For instance, on \mathbb{R}^n equipped with the flat connection (where connection coefficients vanish), \nabla_X Y reduces precisely to the directional derivative D_X Y = \sum_i (X Y^i) \partial_i, matching the calculus of vector fields in . A key application arises in the study of curves on the manifold: a curve \gamma: I \to M is a geodesic if its velocity vector field satisfies the geodesic equation \nabla_{\dot{\gamma}} \dot{\gamma} = 0, defining "straight lines" that locally minimize distance and generalize lines to curved spaces. This equation encodes the intrinsic , with solutions depending on the . The covariant derivative interacts with the , which measures their non-commutativity via [X, Y] f = X(Y f) - Y(X f) for functions f. In general, \nabla_X Y - \nabla_Y X \neq [X, Y]; the difference is captured by the of the , which vanishes for torsion-free connections like the Levi-Civita connection on Riemannian manifolds.

Extension to Other Fields

Covariant Derivative of Covector Fields

The covariant derivative extends naturally to covector fields, or 1-forms, on a manifold equipped with an \nabla. For a 1-form \omega and vector fields X, Y on the manifold, the covariant derivative \nabla_X \omega is defined to be the 1-form satisfying (\nabla_X \omega)(Y) = X(\omega(Y)) - \omega(\nabla_X Y) for all vector fields Y. This definition ensures that the covariant derivative respects the duality between vector fields and covector fields, allowing the differentiation of linear functionals on the tangent spaces in a manner consistent with the connection's . This extension preserves the Leibniz rule for tensor products of covector fields. Specifically, for 1-forms \omega and \eta, the covariant derivative satisfies \nabla_X (\omega \otimes \eta) = (\nabla_X \omega) \otimes \eta + \omega \otimes (\nabla_X \eta). This maintains the bilinear structure of tensor operations under differentiation, enabling the consistent application of the covariant derivative to more general tensor fields derived from covectors. In local coordinates (x^i) on the manifold, where \omega = \omega_j \, dx^j, the components of the covariant derivative \nabla_i \omega are given by (\nabla_i \omega)_j = \partial_i \omega_j - \Gamma^k_{ij} \omega_k, with \Gamma^k_{ij} denoting the of the . This expression accounts for the transformation of covector components under changes in coordinate basis, subtracting the connection terms to yield a tensorial object. In the context of a Riemannian manifold with metric tensor g, metric compatibility of the connection (\nabla g = 0) implies that the covariant derivative of g vanishes, leading to (\nabla_X g)(Y, Z) = g(\nabla_X Y, Z) + g(Y, \nabla_X Z) for all vector fields X, Y, Z. Since g is a (0,2)-tensor built from covectors, this relation highlights how the covariant derivative of covector fields underlies the preservation of the metric's inner product structure along directional derivatives. For torsion-free connections, such as the , the d\omega of a 1-form \omega relates directly to the covariant derivative via the alternation operator: d\omega = \mathrm{Alt}(\nabla \omega). This connection demonstrates how the antisymmetric part of \nabla \omega captures the intrinsic differential structure of the 1-form, independent of the connection's symmetric contributions.

Generalization to Tensor Fields

A tensor field of type (k, l) on a smooth manifold is a smooth section of the (k, l)-tensor bundle, assigning to each point p in the manifold a multilinear map T_p: (T_p^* M)^{\otimes k} \times (T_p M)^{\otimes l} \to \mathbb{R}, where T_p^* M denotes the at p. This generalizes vector fields as (1, 0)-tensors and covector fields as (0, 1)-tensors, allowing the covariant derivative to extend naturally to arbitrary s while preserving their multilinear structure under . The extension of the covariant derivative to a type (k, l) tensor field T along a vector field X follows the pattern established for vectors and covectors: for each contravariant (upper) index, a positive connection term is added, and for each covariant (lower) index, a negative term is subtracted, ensuring compatibility with the Leibniz rule and tensor transformation laws. In local coordinates, where \Gamma^i_{jk} are the Christoffel symbols of the connection, the components of the covariant derivative (\nabla_X T)^{i_1 \dots i_k}_{j_1 \dots j_l} are given by (\nabla_X T)^{i_1 \dots i_k}_{j_1 \dots j_l} = X(T^{i_1 \dots i_k}_{j_1 \dots j_l}) + \sum_{m=1}^k \Gamma^{i_m}_{p \, X} T^{\dots p \dots}_{j_1 \dots j_l} - \sum_{n=1}^l \Gamma^q_{j_n \, X} T^{i_1 \dots i_k}_{\dots q \dots}, with the sums replacing the m-th upper index and n-th lower index, respectively. This formula guarantees that \nabla_X T is a tensor field of type (k, l+1), transforming correctly under coordinate changes, unlike the partial derivative which fails to do so on curved manifolds. For scalar fields, which are type (0, 0) tensors, the covariant derivative reduces to the : \nabla_X f = X f. The covector case, as a special instance of (0, 1)-tensors, applies the rule with a single negative Christoffel term. A key application is the of a scalar f, defined as the (0, 2)-tensor \nabla^2 f(Y, Z) = (\nabla_Y \nabla f)(Z) = Y(Z f) - (\nabla_Y Z) f, where Y and Z are vector fields. For the , the is symmetric, satisfying \nabla^2 f(Y, Z) = \nabla^2 f(Z, Y), reflecting the torsion-free property of the connection.

Coordinate Description

Expression in Local Coordinates

In a smooth manifold equipped with an affine connection, the expression for the covariant derivative in local coordinates (x^1, \dots, x^n) takes a concrete form using the \Gamma^i_{jk}, which are the connection coefficients with respect to the coordinate basis \{\partial_i\}. For a Y = Y^i \partial_i, the covariant derivative along a basis vector field is \nabla_{\partial_j} Y = \left( \partial_j Y^i + \Gamma^i_{jk} Y^k \right) \partial_i, where summation over repeated indices is implied, and \partial_j = \partial / \partial x^j. This formula corrects the ordinary partial derivative for the variation in the basis vectors induced by the connection. The general case for an arbitrary vector field X = X^j \partial_j follows by linearity: \nabla_X Y = \left( X^j \partial_j Y^i + \Gamma^i_{jk} X^j Y^k \right) \partial_i. This expression, originally developed in the context of absolute differential calculus, allows computation of how vector fields change along arbitrary directions while respecting the manifold's geometry. The covariant derivative extends naturally to covector fields and higher-rank tensors via the Leibniz rule and linearity. For a covector field \omega = \omega_i \, dx^i, it is \nabla_{\partial_j} \omega = \left( \partial_j \omega_i - \Gamma^k_{ji} \omega_k \right) dx^i. Note the sign change in the connection term compared to the vector case, reflecting the dual nature of covectors. For a general tensor field of type (r,s), the formula involves the partial derivative plus +\Gamma terms for each of the r contravariant indices and -\Gamma terms for each of the s covariant indices. As an illustrative example, for a (1,1)-tensor T = T^i_j \, \partial_i \otimes dx^j, \nabla_{\partial_k} T = \left( \partial_k T^i_j + \Gamma^i_{kl} T^l_j - \Gamma^l_{kj} T^i_l \right) \partial_i \otimes dx^j. These expressions ensure the covariant derivative behaves tensorially under contractions and tensor products. Under a change of local coordinates from x to x', the Christoffel symbols do not transform as a tensor, but their specific transformation law guarantees the tensoriality of the overall covariant derivative. The formula is \Gamma'^i_{jk} = \frac{\partial x^p}{\partial x'^i} \frac{\partial x'^j}{\partial x^q} \frac{\partial x'^k}{\partial x^r} \Gamma^p_{qr} + \frac{\partial x^p}{\partial x'^i} \frac{\partial^2 x'^j}{\partial x^q \partial x^r} \frac{\partial x^q}{\partial x'^k}, where primes denote quantities in the new coordinates. The first term resembles a tensor transformation, while the second accounts for the second derivatives arising from the coordinate change. This non-tensorial behavior of \Gamma is essential for defining a consistent derivative operator across charts.

Role of Christoffel Symbols

The Christoffel symbols of the second kind, denoted \Gamma^i_{jk}, represent the connection coefficients of an \nabla with respect to a local coordinate basis \{\partial/\partial x^j\} on a smooth manifold. Specifically, they are defined by the relation \nabla_{\partial/\partial x^j} \partial/\partial x^k = \Gamma^i_{jk} \partial/\partial x^i, which encodes how the differentiates basis vectors. Introduced by in his 1869 paper on the transformation of quadratic differential forms, these symbols provide a coordinate-based description of the but are not tensorial objects; under a coordinate transformation x \to x', they transform as \Gamma'^i_{jk} = \frac{\partial x'^i}{\partial x^l} \frac{\partial x^m}{\partial x'^j} \frac{\partial x^n}{\partial x'^k} \Gamma^l_{mn} + \frac{\partial x'^i}{\partial x^l} \frac{\partial^2 x^l}{\partial x'^j \partial x'^k}, incorporating second derivatives that prevent tensor transformation properties. On a (M, g) equipped with a g_{ij}, the is the unique that is both torsion-free and compatible with the , meaning \nabla g = 0. The for this connection are computed explicitly as \Gamma^i_{jk} = \frac{1}{2} g^{il} \left( \partial_j g_{kl} + \partial_k g_{jl} - \partial_l g_{jk} \right), where g^{il} is the inverse and \partial_j denotes partial with respect to x^j. This formula arises from solving the conditions of metric compatibility, which implies \partial_j g_{ik} = \Gamma^l_{ji} g_{lk} + \Gamma^l_{jk} g_{il}, and vanishing torsion, combined via cyclic summation and . The resulting symbols satisfy \Gamma^i_{jk} = \Gamma^i_{kj}, reflecting the torsion-free property. For the , the vanish in Cartesian coordinates on the \mathbb{R}^n with the flat metric g_{ij} = \delta_{ij}, since the metric components are constant and all partial derivatives \partial_j g_{kl} = 0. In curved spaces, however, they capture the geometry; for instance, on the unit 2-sphere S^2 with spherical coordinates (\theta, \phi) and metric ds^2 = d\theta^2 + \sin^2 \theta \, d\phi^2, the non-vanishing symbols are \Gamma^\theta_{\phi\phi} = -\sin\theta \cos\theta and \Gamma^\phi_{\theta\phi} = \Gamma^\phi_{\phi\theta} = \cot\theta, computed directly from the using g_{\theta\theta} = 1, g_{\phi\phi} = \sin^2\theta. For general affine connections, the Christoffel symbols \Gamma^i_{jk} need not be symmetric in the lower indices. The torsion tensor T, measuring the failure of the connection to be torsion-free, is defined by T(X,Y) = \nabla_X Y - \nabla_Y X - [X,Y] for vector fields X, Y; in coordinates, where the Lie bracket [\partial_j, \partial_k] = 0, its components are T^i_{jk} = \Gamma^i_{jk} - \Gamma^i_{kj}. Thus, arbitrary choices of \Gamma^i_{jk} allow for non-zero torsion, generalizing beyond the Levi-Civita case while still serving as the local representation of the connection.

Properties and Algebraic Structure

Basic Properties (Linearity, Leibniz Rule)

The covariant derivative exhibits bilinearity over the real numbers in both of its arguments. For real scalars a, b \in \mathbb{R}, vector fields X, Y, and a Z, it satisfies \nabla_{aX + bY} Z = a \nabla_X Z + b \nabla_Y Z and \nabla_X (aZ + bW) = a \nabla_X Z + b \nabla_X W for another W. This linearity ensures that the covariant derivative behaves as a between appropriate tensor bundles, preserving the structure over \mathbb{R}. A fundamental property is the Leibniz rule, which extends the product rule from ordinary differentiation to tensor fields. For tensor fields S and T of arbitrary type, and a vector field X, \nabla_X (S \otimes T) = (\nabla_X S) \otimes T + S \otimes (\nabla_X T). In the special case of a smooth function f (a scalar field) and a vector field Y, \nabla_X (f Y) = (X f) Y + f \nabla_X Y, where X f is the directional derivative of f along X. This rule underscores the covariant derivative's role as a derivation on the algebra of tensor fields. The covariant derivative is compatible with tensor contractions, meaning it commutes with the operation of contracting indices. For a T with components T^\alpha{}_\alpha (sum over repeated index), \nabla_\beta (T^\alpha{}_\alpha) = (\nabla_\beta T)^\alpha{}_\alpha, ensuring that contractions of the differentiated tensor yield the same result as differentiating after contraction. This property preserves the and other operations under differentiation. In flat Euclidean space \mathbb{R}^n equipped with the standard flat connection (where Christoffel symbols vanish in Cartesian coordinates), the covariant derivative reduces to the ordinary partial derivative operator on tensor components. For covector fields (differential forms), this specializes further, aligning with the exterior derivative d in the sense that the antisymmetrized covariant derivative recovers d on forms. On a (M, g), there exists a unique \nabla that is both torsion-free and compatible with the metric g (i.e., \nabla g = 0). This unique connection is the , which satisfies the metric compatibility condition \nabla_\alpha g_{\beta\gamma} = 0 and ensures preserves lengths and angles. The uniqueness follows from solving the system of partial differential equations imposed by these conditions on the connection coefficients.

Torsion and Curvature Tensors

The torsion tensor associated to an \nabla on a manifold is defined by T(X, Y) = \nabla_X Y - \nabla_Y X - [X, Y] for vector fields X and Y, yielding a of type (1,2). This tensor quantifies the antisymmetric part of the connection and vanishes identically for the induced by a pseudo-Riemannian , ensuring compatibility with the and absence of "twist" in . The curvature tensor of the connection is defined by R(X, Y) Z = \nabla_X \nabla_Y Z - \nabla_Y \nabla_X Z - \nabla_{[X, Y]} Z for vector fields X, Y, and Z, resulting in a of type (1,3). In local coordinates \{x^i\}, the components of the tensor, expressed in terms of the \Gamma^i_{jk}, take the form 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}. These components capture the nonlinear obstruction to the being flat, extending the linear structure of the itself. The curvature tensor satisfies two fundamental Bianchi identities. The first Bianchi identity is the algebraic relation obtained by cyclic summation over its lower indices: R(X, Y) Z + R(Y, Z) X + R(Z, X) Y = 0, which holds when the connection is torsion-free and reflects the cyclic symmetry inherent in the geometry. The second Bianchi identity is a differential constraint: \nabla_X R(Y, Z) W + \nabla_Y R(Z, X) W + \nabla_Z R(X, Y) W = 0, whose appropriate contraction yields \nabla^k \mathrm{Ric}_{k j} = \frac{1}{2} \nabla_j [R](/page/R), where [R](/page/R) is the Ricci scalar; this implies the vanishing divergence of the Einstein tensor \nabla^\mu [G_{\mu\nu}](/page/Einstein_tensor) = 0, crucial for conservation laws in gravitational theories. Geometrically, the torsion tensor measures the path-dependence in the construction of infinitesimal parallelograms via : for vectors X and Y, T(X, Y) gives the closure failure when transporting Y along X and , indicating a "screw-like" deviation from metric compatibility./05%3A_Curvature/5.09%3A_Torsion) The curvature tensor, in the context of an Ehresmann connection on a , assesses the integrability of the horizontal distribution: vanishing implies the horizontal subbundle is integrable (by Frobenius theorem), allowing local trivializations without obstructions.

Specialized Applications

Derivative Along a Curve

In differential geometry, the covariant derivative can be specialized to differentiation along a smooth \gamma: I \to M on a manifold M equipped with a \nabla. For a V defined along \gamma, the covariant derivative along the curve is given by \frac{\nabla V}{dt} = \nabla_{\dot{\gamma}(t)} V, where \dot{\gamma}(t) = \frac{d\gamma}{dt} is the to the curve; this is often termed the total covariant derivative and measures the rate of change of V relative to the connection as one moves along \gamma. A V along \gamma is said to be parallel if \frac{\nabla V}{dt} = 0 at every point, which corresponds to of vectors along the curve. In local coordinates, this condition yields a of linear equations (ODEs): \frac{dV^i}{dt} + \Gamma^i_{jk}(\gamma(t)) V^j \dot{\gamma}^k = 0, where \Gamma^i_{jk} are the of the ; the unique starting from an initial V(0) at t=0 is given by the parallel transport map, which in general can be expressed via the path-ordered exponential of the pulled-back along the curve. This transport preserves the inner product induced by the metric when \nabla is metric-compatible, enabling consistent comparison of vectors at different points along \gamma. In the context of , for non-geodesic timelike curves (such as worldlines of accelerated observers), the standard is modified to Fermi-Walker transport to account for physical non-rotation and preserve of spatial frames relative to the velocity vector. The Fermi-Walker derivative along a curve with u^\mu and a^\mu is \frac{D_F V^\mu}{d\tau} = \frac{\nabla V^\mu}{d\tau} + (V \cdot a) u^\mu - (V \cdot u) a^\mu, where \tau is ; vectors satisfying \frac{D_F V^\mu}{d\tau} = 0 define non-rotating frames, as realized by idealized gyroscopes. This variant ensures that scalar products between transported vectors remain constant, crucial for maintaining physical orientations in curved . A key example arises in the study of geodesics, the "straightest" curves where the tangent vector \dot{\gamma} is parallel along itself, satisfying \frac{\nabla \dot{\gamma}}{dt} = 0. In coordinates, this geodesic equation takes the form \ddot{\gamma}^i + \Gamma^i_{jk}(\gamma) \dot{\gamma}^j \dot{\gamma}^k = 0, describing the acceleration-free motion under the connection. Applications of covariant derivatives along curves include solving for Jacobi fields, which are vector fields J along a geodesic \gamma satisfying the Jacobi equation \frac{\nabla^2 J}{dt^2} = R(\dot{\gamma}, J) \dot{\gamma}, where R is the Riemann curvature tensor; this reduces to \frac{\nabla^2 J}{dt^2} = 0 in flat space and generalizes the notion via the connection and curvature; these fields quantify the infinitesimal separation between nearby geodesics in a variation, providing a measure of how curves diverge or converge along \gamma.

Relation to Lie Derivative

The Lie derivative and the covariant derivative are both differential operators on tensor fields over a manifold, but they differ fundamentally in their construction and interpretation. The Lie derivative \mathcal{L}_X T of a tensor field T along a vector field X measures the rate of change of T under the infinitesimal flow generated by X, without requiring any additional geometric structure like a connection. In contrast, the covariant derivative \nabla_X T relies on a linear connection to define parallel transport, ensuring that the result remains a tensor of the same type, independent of coordinate choices. This distinction is particularly evident for vector fields, where the Lie derivative \mathcal{L}_X Y = [X, Y] for vector fields X and Y captures the commutator, while the covariant derivative \nabla_X Y incorporates the connection's Christoffel symbols to account for the manifold's geometry. A key relation between the two operators arises when expressing the of a in terms of the . In general, for any linear with T, the formula is \mathcal{L}_X Y = \nabla_X Y - \nabla_Y X - T(X, Y), where the torsion term T(X, Y) accounts for the antisymmetric part of the . This identity links the intrinsic [X, Y] to the symmetric aspects of the , highlighting how torsion measures the of the covariant derivative to commute in a way that matches the Lie bracket exactly. In the common case of a torsion-free , such as the on a , the torsion vanishes (T = 0), simplifying the relation to \mathcal{L}_X Y = \nabla_X Y - \nabla_Y X. Here, the emerges as an antisymmetrized combination of covariant derivatives, emphasizing its role in capturing directional changes without net torsion effects. For general tensor fields, the Lie derivative acts as a derivation, satisfying the Leibniz rule \mathcal{L}_X (T \otimes S) = (\mathcal{L}_X T) \otimes S + T \otimes (\mathcal{L}_X S), but it does not inherently preserve tensor type without a ; its coordinate expression involves partial derivatives and adjustments for the . The covariant derivative, however, is defined to map tensors to tensors of one higher , preserving the multilinear intrinsically through the connection's on each . This difference is illustrated in flat , where the connection vanishes, and both operators reduce to ordinary partial derivatives: (\nabla_X Y)^j = X^i \partial_i Y^j and \mathcal{L}_X Y = X^i \partial_i Y - Y^i \partial_i X, coinciding with the Lie bracket in Cartesian coordinates. On a curved manifold, the Lie derivative measures the deformation of tensor fields under the of X, reflecting how the field is "dragged" along integral curves, whereas the covariant derivative quantifies intrinsic changes relative to , independent of the specific . When a g is present, the Koszul formula provides a specific link via the connection's . For a torsion-free that is not necessarily metric-compatible, the of the along X satisfies (\mathcal{L}_X g)(Y, Z) = (\nabla_X g)(Y, Z) + g(\nabla_Y X, Z) + g(Y, \nabla_Z X), where the first term captures deviations from . In the standard case of a metric-compatible (e.g., Levi-Civita, where \nabla g = 0), this reduces to \mathcal{L}_X g = 2\, g(\nabla_{(X} \cdot, \cdot)), or in components, \mathcal{L}_X g_{\mu\nu} = \nabla_\mu X_\nu + \nabla_\nu X_\mu, symmetrizing the lowered covariant derivative of X. This relation underscores how the detects symmetries preserving the (Killing vector fields satisfy \mathcal{L}_X g = 0), while the covariant derivative enforces in geometric equations. In physical applications, such as , the is employed to identify via Killing vectors, where \mathcal{L}_X g = 0 implies isometries, whereas the covariant derivative formulates coordinate-independent s, like the \nabla_{\dot{\gamma}} \dot{\gamma} = 0. This complementary use highlights the 's role in global flow analysis and the covariant derivative's focus on local, connection-based differentiation.

References

  1. [1]
    [PDF] 1. Covariant derivative and parallel transport
    V (t) = ∇˙c(t)X (notice that this is a local statement, and holds whenever ˙c(t) 6= 0). We call this operation D dt the covariant derivative. Proof: Choose a ...
  2. [2]
    [PDF] On the history of Levi-Civita's parallel transport - arXiv
    Aug 6, 2016 · [35] T. Levi-Civita, ”Nozione di parallelismo in una varietà qualunque e conseguente specificazione geometrica della curvatura riemanniana”, ...
  3. [3]
    [PDF] Differential geometry Lecture 16: Parallel transport and the Levi ...
    Jun 26, 2020 · Let (M, g) be a pseudo-Riemannian manifold. A connection ∇ in TM → M is called Levi-Civita connection if it is metric and torsion-free.
  4. [4]
    [PDF] TENSOR CALCULUS 7 - 7. Covariant Derivative - OSU Math
    Covariant Derivative of a Tensor. The covariant derivative can be extended from vectors and covectors ("I-forms") to tensors of arbitrary rank, the key idea ...
  5. [5]
    1854: Riemann's classic lecture on curved space
    Jun 1, 2013 · First, the question of how we might define an n-dimensional space resulted in the definition of Riemann space, including the Riemann tensor.
  6. [6]
    [PDF] On the Hypotheses which lie at the Bases of Geometry. Bernhard ...
    It is known that geometry assumes, as things given, both the notion of space and the fi rst princip les of constructions in space . S he gives de fi nitions of ...
  7. [7]
    Elwin Christoffel (1829 - 1900) - Biography - MacTutor
    Elwin Christoffel was noted for his work in mathematical analysis, in which he was a follower of Dirichlet and Riemann.
  8. [8]
    [PDF] On the genesis of the concept of covariant differentiation - Numdam
    context, the algorithm of covariant differentiation was used by Christoffel as a well-defined technique in a particular field of research, that of differen ...
  9. [9]
    Some remarks on the history of Ricci's absolute differential calculus
    Oct 9, 2024 · The article offers a general account of the genesis of the absolute differential calculus (ADC), paying special attention to its links with ...Missing: Gregorio 1887-1900
  10. [10]
    Some remarks on the history of Ricci's absolute differential calculus
    Oct 9, 2024 · The article offers a general account of the genesis of the absolute differential calculus (ADC), paying special attention to its links with the history of ...
  11. [11]
    2. Manifolds - Lecture Notes on General Relativity - S. Carroll
    There are three important exceptions: partial derivatives, the metric, and the Levi-Civita tensor. Let's look at the partial derivative first. The unfortunate ...
  12. [12]
    [PDF] Introduction to Tensor Calculus for General Relativity - MIT
    This confirms that the Christoffel symbols are not tensor components ... partial derivatives ∂µ. Sup- pose, however, that we differentiate a vector ...
  13. [13]
    [PDF] Some remarks on the history of Ricci's absolute differential calculus
    Indeed, it suffices to consider for example the introduction to [Ricci and Levi-Civita] where the authors explicitly emphasized the intimate connection between ...
  14. [14]
    [PDF] The Meaning of Einstein's Equation - Stanford University
    Mar 10, 2001 · If we parallel transport a tangent vector from the north pole to the equator by going straight down a meridian, we get a different result than ...
  15. [15]
    [PDF] Math 396. Covariant derivative, parallel transport, and General ...
    A smooth section of a vector bundle along a path is an element of the pullback bundle. Parallel transport is a linear isomorphism between fibers of the bundle.
  16. [16]
    [PDF] INTRODUCTION TO DIFFERENTIAL GEOMETRY - ETH Zürich
    Chapter 3 introduces the Levi-Civita connection as covariant derivatives of vector fields along curves.4 It continues with parallel transport, introduces.
  17. [17]
    [PDF] INTRODUCTION TO DIFFERENTIAL GEOMETRY - UC Homepages
    Jan 11, 2011 · covariant derivative of a vector field along a curve. With this ... pdf/Cartan.pdf. [4] Simon K. Donaldson, Lie algebra theory without ...
  18. [18]
    [PDF] General Investigations of Curved Surfaces - Project Gutenberg
    In 1827 Gauss presented to the Royal Society of Göttingen his important paper on the theory of surfaces, which seventy-three years afterward the eminent ...
  19. [19]
    [PDF] Chapter 3 Connections
    so the covariant derivative is literally the “vertical part” of the tangent map. For a section s(t) ∈ E along a path γ(t) ∈ M, we have the analogous.<|control11|><|separator|>
  20. [20]
    [PDF] Vector Bundles and Connections
    The exterior covariant derivative is connected to the curvature tensor which we discuss further below, see Definition 3.0.1. 2.2.4. Exercises. 1) Compute the ...
  21. [21]
    [PDF] Connections - IME-USP
    Let M be a smooth manifold. A (Koszul) connection in M is a bilinear map ∇ : Γ(TM)×Γ(TM) → Γ(TM), where we write ∇XY instead of ∇(X, Y ), such that a. ∇fX Y = ...
  22. [22]
    [PDF] Kudinoor Fundamental Theorem of Riemannian Geometry
    Definition (Affine Connection). Let X(M) be the set of all C∞ vector fields on M. An affine connection on a manifold M is an R-bilinear map. V : X(M) × X(M) ...
  23. [23]
    [PDF] Affine Connections - UCSD Math
    Outline: We are starting a new chapter on affine and Riemannian connections. These concepts, which enable us to differentiate vector fields, will bring us ...
  24. [24]
    affine connection in nLab
    Apr 4, 2021 · An affine connection ∇ \nabla on a smooth manifold M M is a connection on the frame bundle F M F M of M M , i.e., the principal bundle of frames ...
  25. [25]
    Existence and uniqueness of the Levi-Civita connection on ...
    1. Introduction. The fundamental theorem of Riemannian geometry asserts that there exists a unique metric compatible and torsion-free connection on the bimodule ...
  26. [26]
    book:gdf:levicivita - Geometry of Differential Forms
    May 15, 2013 · A connection satisfying (\ref{tfree}) is called torsion free, and a connection which is both torsion free and metric compatible is called a Levi ...
  27. [27]
    [PDF] Introduction to Geometry and geometric analysis
    3.7 Covariant derivative . ... 1-forms = covector fields) by. ∇X t(Y ) = Xt(Y ) − t(∇X Y ) . Having defined the connection on 1-forms, we can define ∇X ...
  28. [28]
    [PDF] Differential Geometry in Physics - UNCW
    3.3 Covariant Derivative . ... A connection is called torsion-free if T(X, Y ) = 0. In this case,. ∇XY − ∇Y X = [X, Y ]. We will elaborate later on the ...
  29. [29]
    [PDF] Locally conformally Kдhler manifolds - Misha Verbitsky
    Also, dω = 0, because ∇ is torsion-free, and dω = Alt(∇ω). The implication (i) ⇒ (ii) is proven by the same argument as used to construct the Levi-Civita ...
  30. [30]
    [PDF] General Relativity - Kevin Zhou
    with the covariant derivative. • We define the covariant derivative with the following postulates. – ∇ is a map from (k, l) tensor fields to (k, l + 1) ...
  31. [31]
    [PDF] Second-Order Geometry - Optimization Algorithms on Matrix Manifolds
    The result states that the horizontal lift of the covariant derivative of ξ with respect to η is given by the horizontal projection of the covariant derivative ...
  32. [32]
    None
    Below is a merged summary of the covariant derivative and related concepts from "Introduction to Riemannian Manifolds" by John M. Lee (2018), consolidating all information from the provided segments into a comprehensive response. To retain maximum detail, I will use a structured format with tables where appropriate, followed by a narrative summary for additional context. The response avoids any "thinking" tokens and focuses solely on presenting the information.
  33. [33]
    Ricci and Levi-Civita's tensor analysis paper - Academia.edu
    Ricci and Levi-Civita's paper is foundational in modern tensor calculus and differential geometry. The original paper was published in 1901 in Mathematische ...
  34. [34]
    [PDF] Introduction to Tensor Calculus
    The covariant derivative of a covector can also be defined with this symbol,. ∇µwα = ∂µwα − rν. µαwν . (7.5). The covariant derivative of a tensor tαβ is then.
  35. [35]
    [PDF] Riemannian Manifolds: An Introduction to Curvature
    the covariant derivative with respect to the Riemannian connection of g. Therefore, we can interpret the second fundamental form as a measure of the ...
  36. [36]
    [PDF] General Relativity Fall 2018 Lecture 6: covariant derivatives
    Sep 21, 2018 · The covariant derivative provides a geometric (i.e. coordinate independent) way to define the gradient of tensors. A covariant derivative ∇ is a ...
  37. [37]
    [PDF] Lecture Notes, September 30, 2014
    Sep 30, 2014 · We define the covariant derivative as. ∇µvν = ∂vν. ∂xµ. + Γν. µσvσ ... Leibniz rule. ∇(S ⊗ T) = ∇S ⊗ T + S ⊗ ∇T . Connections exist ...
  38. [38]
    [PDF] 6.1 More on the covariant derivative - MIT
    covariant derivative obey the Leibniz rule for the derivative of the product of quantities: ∇β (pαAα) = pα∇βAα + Aα∇βpα . (6.6). Comparing Eqs. (6.5) and ...
  39. [39]
    Lecture Notes on General Relativity - S. Carroll
    A covariant derivative is an operator that reduces to the partial derivative in flat space, but transforms as a tensor on a manifold.
  40. [40]
    [PDF] 1 The Levi-Civita Connection and its curva- ture - MIT Mathematics
    Let's prove, from this point of view, the basic uniqueness theorem the Levi-. Civita connection. Lemma 1.3. There is a unique torsion free metric compatible ...
  41. [41]
    [PDF] Riemann-Christoffel curvature tensor—23 Mar 2010
    Mar 23, 2010 · There is one Christoffel symbol for each upper index. The covariant derivative of a covariant vector is. Aa;b = Aa,b - Gg ab Ag.
  42. [42]
    [PDF] Bianchi identities
    Jan 22, 2025 · The curvature tensor field R satisfies the following properties. Lemma 4. For any Riemannian manifold (M,ϕ) with a Riemannian connection, the ...
  43. [43]
    [PDF] Ehresmann, Kozul, and Cartan connections
    Let H ⊂ TP be a G-invariant horizontal distribution. Then let HΦ ≡ Φ∗H be the gauge-transformed distribution. Lemma 4.14. HΦ ⊂ TP is an Ehresmann connection.
  44. [44]
    [PDF] Chapter 12 Connections on Manifolds - UPenn CIS
    However, if M is curved, for example, a sphere, then it is not obvious how to “parallel transport” a tangent vector at c(0) along a curve c.
  45. [45]
    [PDF] On the construction of Fermi-Walker transported frames - arXiv
    Apr 15, 2008 · Fermi-Walker frames are created by transforming tetrad fields using a local Lorentz transformation, which ensures the timelike component ...
  46. [46]
    [PDF] Fermi–Walker transport and Thomas precession
    Apr 23, 2018 · The Fermi-Walker transport equation is used to derive the Thomas precession formula, which is not intuitive and was surprising even to Einstein.
  47. [47]
    [PDF] JACOBI FIELDS As we have seen, in the second variational formula ...
    Theorem 1.5. A vector field X along a geodesic γ is a Jacobi field if and only if. X is the variation field of some geodesic variation of γ.
  48. [48]
    [PDF] the theory of lie derivatives and its applications
    ... Lie derivative of a linear connexion. 15. CHAPTER II. LIE DERIVATIVES OF ... covariant derivative of O respectively. Cf. SCHOUTEN [8], p. 124. 3 The ...
  49. [49]
    [PDF] General Relativity - DAMTP
    This course covers manifolds, tensors, metric tensors, covariant derivative, curvature, physical laws in curved spacetime, and Einstein's equation.<|control11|><|separator|>
  50. [50]
    [PDF] GR lecture 5 Covariant derivatives, Christoffel connection ...
    One can use them to derive a much simpler formula for the determinant's ... The source-dependent part is derived, as before, by varying the action (31) ...Missing: history | Show results with:history
  51. [51]
    [PDF] Covariant and Lie Derivatives - GR at NC State
    Lie derivatives do not rely on the presence of a metric for their definitions. When a metric is present, they can be written in terms of covariant derivatives ...