Fact-checked by Grok 2 weeks ago

Vector field

A vector field is a mathematical function that assigns a unique vector to every point in a subset of , such as \mathbb{R}^2 or \mathbb{R}^3. This assignment allows the vector field to describe directional quantities that vary continuously across the domain, with the vector at each point indicating both magnitude and direction. In , vector fields form the foundation for key operations including the (which produces a vector field from a ), (measuring the field's source or sink behavior at a point), and (capturing the rotational aspect of the field). These operators enable the analysis of how the field behaves under , essential for theorems like Stokes' and the that relate surface and volume integrals. Vector fields have broad applications in physics, where they model phenomena such as the velocity of , the direction and strength of wind patterns, gravitational forces, and electromagnetic fields around charges. For instance, the arising from static charges in the plane is a classic example of a , derivable from a . In more advanced contexts, such as , vector fields extend to manifolds and underpin the study of flows and symmetries in dynamical systems.

Definitions

In Euclidean space

In \mathbb{R}^n, a vector field on an open subset U \subset \mathbb{R}^n is defined as a V: U \to \mathbb{R}^n that assigns point x \in U a V(x) to \mathbb{R}^n at x. This mapping can be viewed as a collection of arrows, one at each point in U, indicating direction and magnitude. In a coordinate system (x^1, \dots, x^n), the vector field V is represented as V(x) = \sum_{i=1}^n V^i(x) \frac{\partial}{\partial x^i}, where V^i: U \to \mathbb{R} are the component functions, and \frac{\partial}{\partial x^i} form the of tangent vectors. These components transform contravariantly under a change of coordinates x' = \phi(x), given by V'^j(x') = \sum_{i=1}^n \frac{\partial x'^j}{\partial x^i}(x) V^i(x), ensuring the vector field remains well-defined independently of the coordinate choice. For applications in , such as computing derivatives, vector fields are typically required to be , meaning their component functions V^i are infinitely differentiable (C^\infty), or at least continuously differentiable (C^1). A simple example is the constant vector field on \mathbb{R}^2 defined by V(x,y) = (1, 0), which assigns the horizontal to every point.

On manifolds

A vector field on a smooth manifold M is defined as a smooth section of the TM, that is, a smooth map V: M \to TM such that the canonical \pi: TM \to M satisfies \pi \circ V = \mathrm{id}_M. This ensures that to each point p \in M, the vector field assigns a unique V(p) \in T_p M, preserving the bundle structure without introducing extraneous fibers. This definition generalizes the case, where the tangent bundle is trivialized as M \times \mathbb{R}^n, but applies intrinsically to curved spaces without a global . Equivalently, in a coordinate-free manner, a vector field can be characterized as a on the of functions C^\infty(M). Specifically, V is a \mathbb{R}- V: C^\infty(M) \to C^\infty(M) satisfying the Leibniz rule: for all f, g \in C^\infty(M), V(fg) = f \, V(g) + g \, V(f). This perspective emphasizes the operational role of vector fields as generators of directional , independent of any bundle formalism, and aligns with their action on functions via pointwise evaluation. The two definitions are mutually equivalent on smooth manifolds, with derivations corresponding bijectively to sections of TM. In local coordinates given by a (U, (x^i)) on M, a vector field V admits the expression V = \sum_i V^i \frac{\partial}{\partial x^i}, where the component functions V^i: U \to \mathbb{R} are . Under a change of coordinates to another (V, (x'^j)) with transition map x'^j = x'^j(x^k), the components transform contravariantly according to the law V'^j = \sum_k \frac{\partial x'^j}{\partial x^k} V^k. This tensorial transformation ensures that the vector field is well-defined globally on M, as the local expressions glue consistently across overlapping charts. The notion of vector fields on manifolds is independent of the dimension of M and does not presuppose , allowing application to manifolds of arbitrary finite , orientable or otherwise. However, if a manifold admits a global frame of n linearly independent vector fields (where n = \dim M), it is parallelizable and hence orientable, as such a frame induces a consistent via the of the . This connection highlights intrinsic topological constraints but does not affect the local definition or existence of individual vector fields. The formalization of vector fields on manifolds, including their bundle-theoretic and derivation-based characterizations, was advanced by in the early through his development of exterior differential systems and moving frames, which provided tools for analyzing local geometry without coordinates.

Examples

Gradient fields

A gradient vector field is one derived from a scalar potential function, defined as \mathbf{V} = \nabla f, where f: U \to \mathbb{R} is a smooth scalar function on an open subset U of Euclidean space. This construction assigns to each point in U a vector pointing in the direction of the steepest ascent of f, with magnitude equal to the rate of that ascent. In \mathbb{R}^n, the components of the are given explicitly by \nabla f = \left( \frac{\partial f}{\partial x^1}, \frac{\partial f}{\partial x^2}, \dots, \frac{\partial f}{\partial x^n} \right). fields exhibit key properties: they are conservative, so line integrals of \mathbf{V} between two points depend only on the endpoints, not the path taken; and they are irrotational, satisfying \curl \nabla f = \mathbf{0} in three dimensions. A representative example is the gravitational field near , approximated as the of the \phi = -\frac{GM}{r} in radial coordinates, where G is the , M is Earth's mass, and r is the distance from Earth's center; this yields \mathbf{g} = -\nabla \phi = -\frac{GM}{r^2} \hat{r}. For visualization, the direction of a vector field is always perpendicular to the level sets of f, meaning field lines are normal to the surfaces or curves where f is constant, illustrating how the field "flows" from low to high potential values.

Central fields

A central vector field in Euclidean space \mathbb{R}^n is defined by \mathbf{V}(\mathbf{x}) = g(|\mathbf{x}|) \frac{\mathbf{x}}{|\mathbf{x}|} for \mathbf{x} \neq \mathbf{0}, where g is a scalar determining the based on the distance from the . This form ensures that the direction of each points radially toward or away from the fixed point at the , with the field's strength varying solely with radial distance. In three dimensions, central vector fields are commonly expressed in spherical coordinates as \mathbf{V}(\mathbf{r}) = F(r) \hat{\mathbf{r}}, where r = |\mathbf{r}| is the radial distance and \hat{\mathbf{r}} is the unit vector in the radial direction. A prominent example is the field F(r) = -\frac{k}{r^2}, which models attractive forces decreasing with the square of the distance. Such fields are rotationally symmetric and irrotational, meaning their vanishes: \nabla \times \mathbf{V} = \mathbf{0} in spherical coordinates when the angular components are zero. The divergence of a central vector field in three dimensions, computed in spherical coordinates, simplifies to \nabla \cdot \mathbf{V} = \frac{1}{r^2} \frac{d}{dr} \left( r^2 F(r) \right), revealing how the field acts as a or at the depending on the behavior of F(r). For instance, the inverse-square yields a divergence of zero everywhere except at the origin, where it behaves like a . Geometrically, the field lines of a central vector field consist of straight rays radiating from or converging to the origin, illustrating the pure radial flow without tangential components. This structure underscores the field's symmetry around the central point, making it ideal for modeling phenomena with spherical invariance. Historically, central vector fields gained prominence through Isaac Newton's law of universal gravitation, which posits a central force field proportional to -\frac{GM m}{r^2} \hat{\mathbf{r}}, directing motion along radial lines in the absence of other influences. Newton's formulation in the Philosophiæ Naturalis Principia Mathematica (1687) established this as the foundational model for gravitational interactions, influencing classical mechanics profoundly.

Other common examples

A constant vector field on \mathbb{R}^n assigns the same vector \mathbf{c} \in \mathbb{R}^n to every point \mathbf{x}, so \mathbf{V}(\mathbf{x}) = \mathbf{c}. This models uniform translation or steady without variation in direction or speed across . A satisfies \nabla \cdot \mathbf{V} = 0 at every point in its , implying no net through closed surfaces. In \mathbb{R}^2, the \mathbf{V}(x,y) = (-y, x) provides a basic example, as its is zero, and it represents incompressible with circulatory patterns. Such fields can always be expressed locally as the of a . In , a field \mathbf{V}(\mathbf{x}, t) describes the motion of particles by assigning a to each spatial point \mathbf{x} at time t, allowing for time dependence that captures evolving flows. For instance, \mathbf{V}(x,y) = (-y, x) models rigid-body rotation about the with constant speed. On a two-dimensional , such as \mathbb{R}^2 equipped with the standard symplectic form \omega = dx \wedge dy, the generated by a smooth H(x,y) is \mathbf{V}(x,y) = \left( \frac{\partial H}{\partial y}, -\frac{\partial H}{\partial x} \right). This construction satisfies \iota_{\mathbf{V}} \omega = -dH, ensuring the flow preserves the structure and level sets of H. Vector fields are typically visualized through arrow representations, with arrows placed at discrete points to convey and proportional for . plots extend this convention computationally, rendering a of scaled arrows on a for clear depiction of field patterns.

Vector calculus operations

Line integrals

A line integral of a vector field \mathbf{V} defined on an U \subseteq \mathbb{R}^n along a smooth C parametrized by \gamma: [a, b] \to U with \gamma(a) and \gamma(b) as the endpoints is given by \int_C \mathbf{V} \cdot d\mathbf{r} = \int_a^b \mathbf{V}(\gamma(t)) \cdot \gamma'(t) \, dt. This formulation arises from approximating the integral as a of products between the vector field values and displacements along the , capturing the component of \mathbf{V} to C. The integral depends on the of C, reversing sign if the parametrization is traversed backward. Physically, when \mathbf{V} represents a force field, the line integral computes the work done by \mathbf{V} on a particle traversing the path C, as it sums the tangential force contributions over the displacement. In general, this work is path-dependent, meaning different curves connecting the same endpoints can yield different values, reflecting how the field's direction and magnitude vary along the route. However, if \mathbf{V} is conservative, meaning \mathbf{V} = \nabla f for some f, the becomes path-independent and equals the difference in potential at the endpoints: \int_C \mathbf{V} \cdot d\mathbf{r} = f(\gamma(b)) - f(\gamma(a)). This result is encapsulated in the fundamental theorem for s, which states that for a \mathbf{V} = \nabla f on a simply connected , the along any piecewise smooth curve C from \mathbf{a} to \mathbf{b} is f(\mathbf{b}) - f(\mathbf{a}), independent of the specific path taken. The theorem generalizes the one-dimensional to higher dimensions, highlighting the role of gradients in reversible work scenarios. To illustrate path dependence for non-conservative fields, consider \mathbf{V}(x,y) = \left( -\frac{y}{x^2 + y^2}, \frac{x}{x^2 + y^2} \right) in \mathbb{R}^2 \setminus \{\mathbf{0}\}, a field tangent to circles centered at the origin. Parametrizing the unit circle as \gamma(t) = (\cos t, \sin t) for t \in [0, 2\pi], the line integral is \int_C \mathbf{V} \cdot d\mathbf{r} = \int_0^{2\pi} \left( -\frac{\sin t}{\cos^2 t + \sin^2 t} \cdot (-\sin t) + \frac{\cos t}{\cos^2 t + \sin^2 t} \cdot \cos t \right) dt = \int_0^{2\pi} 1 \, dt = 2\pi. This nonzero value for a closed curve underscores the field's rotational nature, confirming it is not conservative.

Divergence

The divergence of a vector field provides a scalar measure of how much the field acts as a or at a given point, quantifying the local expansion or contraction of the field lines. In \mathbb{R}^n, for a vector field \mathbf{V} = (V^1, \dots, V^n), the divergence is defined as \div \mathbf{V} = \sum_{i=1}^n \frac{\partial V^i}{\partial x^i}, where the partial derivatives are taken with respect to the standard coordinates x^1, \dots, x^n. This operator arises naturally in the context of calculations, where positive indicates a net outflow of the field from an , akin to a , while negative suggests an inflow, like a . Physically, the represents the of through an enclosing a point; by the , the integral of \div \mathbf{V} over a equals the total through its surface, providing a bridge between local and global properties of the field./16:_Vector_Calculus/16.05:_Divergence_and_Curl) On a (M, g), the extends to a coordinate-free using differential forms and the structure. Specifically, for a vector field V on an oriented of dimension n, the is given by \div V = * \, d \, (* V^\flat), where V^\flat denotes the musical lowering the index via the g to obtain the associated 1-form, d is the , and * is the mapping k-forms to (n-k)-forms. This formulation ensures compatibility with the manifold's geometry and is essential for applications in and , where it measures the trace of the of V. A notable example is the divergence of a central (radial) vector field in \mathbb{R}^n, \mathbf{V}(\mathbf{x}) = F(r) \hat{\mathbf{r}}, where r = \|\mathbf{x}\| and \hat{\mathbf{r}} = \mathbf{x}/r. The computation yields \div \mathbf{V} = \frac{1}{r^{n-1}} \frac{d}{dr} \left( r^{n-1} F(r) \right), which simplifies to (n-1) F(r)/r if F(r) is homogeneous of degree zero (constant magnitude), highlighting how such fields exhibit divergence away from the , while fields with F(r) \propto 1/r^{n-1} (as in ) have zero divergence away from the unless sources are present. This formula underscores the role of dimensionality in flux behavior, as seen in where F(r) = 1/r^{n-1} yields constant divergence proportional to sources. The divergence operator possesses key algebraic properties that facilitate computations and proofs. It is linear: for scalar constants a, b and vector fields \mathbf{V}, \mathbf{W}, \div (a \mathbf{V} + b \mathbf{W}) = a \, \div \mathbf{V} + b \, \div \mathbf{W}. Additionally, it satisfies a : for a scalar f and vector field \mathbf{V}, \div (f \mathbf{V}) = f \, \div \mathbf{V} + \nabla f \cdot \mathbf{V}, allowing the decomposition of scaled fields into intrinsic and contributions. These properties mirror those of the standard and ensure the divergence behaves consistently under affine transformations and scaling.

Curl

The curl of a vector field quantifies the local or of the field at a given point, indicating the tendency of the field lines to circulate around that point. In three-dimensional , it is a applied to a vector field \mathbf{V} = V^x \mathbf{i} + V^y \mathbf{j} + V^z \mathbf{k}, producing another vector field that points in the direction of the axis of with equal to the rotational speed. The explicit formula for the curl in \mathbb{R}^3 is given by \nabla \times \mathbf{V} = \begin{vmatrix} \mathbf{i} & \mathbf{j} & \mathbf{k} \\ \frac{\partial}{\partial x} & \frac{\partial}{\partial y} & \frac{\partial}{\partial z} \\ V^x & V^y & V^z \end{vmatrix} = \left( \frac{\partial V^z}{\partial y} - \frac{\partial V^y}{\partial z} \right) \mathbf{i} + \left( \frac{\partial V^x}{\partial z} - \frac{\partial V^z}{\partial x} \right) \mathbf{j} + \left( \frac{\partial V^y}{\partial x} - \frac{\partial V^x}{\partial y} \right) \mathbf{k}, or in component form, (\nabla \times \mathbf{V})_x = \frac{\partial V^z}{\partial y} - \frac{\partial V^y}{\partial z}, \quad (\nabla \times \mathbf{V})_y = \frac{\partial V^x}{\partial z} - \frac{\partial V^z}{\partial x}, \quad (\nabla \times \mathbf{V})_z = \frac{\partial V^y}{\partial x} - \frac{\partial V^x}{\partial y}. Two important properties of the curl operator are that the curl of the gradient of any scalar potential function f vanishes, \nabla \times (\nabla f) = \mathbf{0}, and the divergence of any curl is zero, \nabla \cdot (\nabla \times \mathbf{V}) = 0. In two dimensions, the curl analogue for a vector field \mathbf{V} = V^x \mathbf{i} + V^y \mathbf{j} is a scalar defined as \operatorname{curl}_2 \mathbf{V} = \frac{\partial V^y}{\partial x} - \frac{\partial V^x}{\partial y}, which measures the counterclockwise in the xy-plane. On a three-dimensional oriented , the curl generalizes to \operatorname{curl} \mathbf{V} = \star (d \mathbf{V}^\flat), where \star denotes the and \mathbf{V}^\flat is the 1-form obtained by lowering the index of \mathbf{V} using the . A physical example arises in magnetostatics, where Ampère's law relates the curl of the \mathbf{B} to the \mathbf{J} via \nabla \times \mathbf{B} = \mu_0 \mathbf{J}, with \mu_0 the permeability of free space./07:Magnetostatics/7.09:Ampere's_Law(Magnetostatics)-_Differential_Form)

Physical interpretations

In mechanics

In , vector fields frequently model fields, assigning a vector to each point in space that dictates the acceleration of a particle at that location via Newton's second law, \mathbf{F} = m \mathbf{a}. A prominent example is the gravitational field exerted by a M on a test m, given by \mathbf{F} = -\nabla \left( -\frac{GMm}{r} \right), where G is the and r is the distance from M, resulting in an attractive directed toward the source. This field governs planetary motion and other gravitational interactions, with the work done by the field along a particle's path computed via line integrals./16:_Vector_Calculus/16.01:_Vector_Fields) Vector fields also describe velocity in kinematics, where for a particle at position \mathbf{q} in configuration space, the velocity field is \mathbf{V}(\mathbf{q}) = \frac{d\mathbf{q}}{dt}, tangent to the particle's path and specifying its instantaneous direction and speed./28:_Fluid_Dynamics/28.02:_Velocity_Vector_Field) This representation captures the geometric aspects of motion without reference to underlying forces, focusing on trajectories in space. In the Hamiltonian formulation of mechanics, a vector field \mathbf{V}_H arises as the symplectic gradient of the Hamiltonian function H, generating the flow that evolves the system in phase space while preserving the symplectic structure. This field encodes the dynamics of conservative systems, such as those with time-independent potentials, ensuring energy conservation along trajectories. For central force motion, where the force depends only on radial distance from a fixed center, particle orbits trace integral curves of the resulting velocity field, yielding conic sections like ellipses for inverse-square laws./06:_General_Planar_Motion/6.03:_Motion_Under_the_Action_of_a_Central_Force) These paths, such as Keplerian orbits, exhibit conserved angular momentum due to rotational symmetry. From the Lagrangian perspective, the velocity field \mathbf{V} emerges from solutions to the Euler-Lagrange equations derived from the L = T - V, defining a Lagrangian vector field on the whose integral curves satisfy the . This approach unifies and for systems with , applicable to constrained motions like pendulums./09:_Hamilton's_Action_Principle/9.03:_Lagrangian)

In electromagnetism

In , the \mathbf{E} and \mathbf{B} are vector fields that govern the forces on charged particles and the propagation of electromagnetic waves. These fields arise from charges, currents, and their time variations, as encapsulated in . The \mathbf{E} is expressed in terms of a \phi and \mathbf{A} as \mathbf{E} = -\nabla\phi - \frac{\partial\mathbf{A}}{\partial t}, highlighting its connection to potential functions and time-dependent effects. This form underscores that \mathbf{E} is irrotational in the static but acquires from changing magnetic fields. In electrostatics, where time derivatives vanish, \mathbf{E} = -\nabla\phi, rendering it a conservative vector field whose line integral around closed paths is zero. A prototypical example is the Coulomb field due to a point charge q at the origin, given by \mathbf{E}(\mathbf{r}) = \frac{1}{4\pi\epsilon_0} \frac{q}{r^2} \hat{r}, which is the negative gradient of the potential \phi(\mathbf{r}) = \frac{1}{4\pi\epsilon_0} \frac{q}{r}. Gauss's law relates the divergence of \mathbf{E} to charge density via \nabla \cdot \mathbf{E} = \frac{\rho}{\epsilon_0}, indicating that electric field lines originate from positive charges and terminate at negative ones, with the divergence measuring local sources. The magnetic field \mathbf{B} satisfies \nabla \cdot \mathbf{B} = 0, implying it is sourceless and has no divergence, consistent with the nonexistence of magnetic monopoles in classical theory. This solenoidal nature means magnetic field lines form closed loops, as seen in configurations like the uniform field inside an ideal solenoid. The Ampère-Maxwell law governs the curl of \mathbf{B}: \nabla \times \mathbf{B} = \mu_0 \mathbf{J} + \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}, where steady currents \mathbf{J} and changing electric fields produce magnetic circulation. The Lorentz force law combines these vector fields to describe the force on a charge q moving with velocity \mathbf{v}: \mathbf{F} = q(\mathbf{E} + \mathbf{v} \times \mathbf{B}), with the electric term acting directly and the magnetic term perpendicular to \mathbf{v}. This force drives charged particle motion in electromagnetic environments, such as cyclotrons or dynamics. Maxwell's equations employ divergence and curl of \mathbf{E} and \mathbf{B} to interlink sources and field evolutions.

Flows and dynamics

Integral curves

An integral curve of a vector field V defined on a smooth manifold M is a map \gamma: I \to M, where I \subseteq \mathbb{R} is an open interval, such that \gamma'(t) = V(\gamma(t)) for all t \in I. Typically, one specifies an \gamma(0) = p for some point p \in M, ensuring the curve passes through p at time zero. This equation means that at every point along the curve, the to \gamma coincides exactly with the value of the vector field V at that point, providing a geometric realization of the field's directions. The existence and uniqueness of such integral curves are addressed by the , which applies when V is locally continuous with respect to the manifold's coordinates. Specifically, for any p \in M, there exists a unique maximal open I_p \subseteq \mathbb{R} containing 0 and a unique \gamma_p: I_p \to M satisfying \gamma_p(0) = p and \gamma_p'(t) = V(\gamma_p(t)) for all t \in I_p. This guarantees a local solution near t = 0, and the curve can be extended along the maximal until it either reaches the of the domain of V or ceases to remain in compact subsets of M. Local curves thus capture short-term behavior, while global curves exist only if the maximal is all of \mathbb{R}. Geometrically, integral curves trace paths that are instantaneously aligned with V, effectively "following the arrows" of the field without deviation. For instance, consider a constant vector field V(x) = v on \mathbb{R}^n, where v \in \mathbb{R}^n is fixed; the integral curves are then straight lines \gamma(t) = p + t v emanating from any initial point p, parametrized at constant speed \|v\|. In , these curves represent the trajectories of particles governed by a velocity field derived from deterministic forces.

Complete vector fields

A vector field V on a smooth manifold M is said to be complete if every of V can be extended to the entire real line, meaning that for every point p \in M, there exists an \gamma: \mathbb{R} \to M satisfying \gamma(0) = p and \gamma'(t) = V(\gamma(t)) for all t \in \mathbb{R}. This property ensures that the local behavior of the vector field globalizes without singularities or escapes to in finite time. Equivalently, V is complete if it generates a global \phi_t: M \to M defined for all t \in \mathbb{R}, where \phi_0 is the map and the flow covers the entire manifold M. curves correspond to the orbits under this flow, specifically as the slices \{ \phi_t(p) \mid t \in \mathbb{R} \} for fixed p \in M. is guaranteed under certain conditions on V. For instance, if V has compact —meaning V vanishes outside a compact subset of M—then V is complete, as the bounded support prevents curves from escaping the manifold in finite time. Similarly, if the zero section M \hookrightarrow TM composed with V: M \to TM yields a proper (i.e., preimages of compact sets are compact), then V is complete. Moreover, every vector field on a compact manifold is complete, since compactness bounds the trajectories. Examples illustrate these concepts clearly. Constant vector fields on \mathbb{R}^n, such as V = \sum a_i \frac{\partial}{\partial x_i} with constant coefficients a_i, are complete, as their flows are global translations \phi_t(x) = x + t a, defined for all t \in \mathbb{R}. In contrast, certain polynomial vector fields are incomplete; for example, on \mathbb{R}, the field V = x^2 \frac{\partial}{\partial x} has integral curves solving \frac{dx}{dt} = x^2, which yield \gamma(t) = \frac{\gamma(0)}{1 - t \gamma(0)} for \gamma(0) > 0, blowing up at finite time t = 1/\gamma(0). Complete vector fields play a key role in the structure of , where left-invariant vector fields generate one-parameter subgroups. Specifically, for a G viewed as a manifold, a left-invariant vector field X (determined by its value at the identity) is complete, and its flow \phi_t realizes the one-parameter subgroup \{ \exp(t X) \mid t \in \mathbb{R} \}, providing a homomorphism from (\mathbb{R}, +) to G.

Lie bracket

The Lie bracket of two smooth vector fields V and W on a smooth manifold M is the vector field [V, W] defined by its action on smooth functions f \in C^\infty(M) as [V, W] f = V(W f) - W(V f). This operation arises naturally as the of the derivations associated to V and W, and it endows the space of smooth vector fields with a structure. In local coordinates (x^1, \dots, x^n) on M, where V = V^i \frac{\partial}{\partial x^i} and W = W^j \frac{\partial}{\partial x^j}, the components of the Lie bracket are [V, W]^k = V^i \frac{\partial W^k}{\partial x^i} - W^j \frac{\partial V^k}{\partial x^j}. This coordinate expression confirms that [V, W] is independent of the choice of coordinates and transforms as a vector field. The Lie bracket satisfies several key algebraic properties: it is bilinear over \mathbb{R}, skew-symmetric ([V, W] = -[W, V]), and obeys the [V, fW] = f [V, W] + (V f) W for any function f \in C^\infty(M). These properties, along with the , make the space of vector fields into an infinite-dimensional . Geometrically, the Lie bracket captures the extent to which the flows generated by V and W fail to commute, with [V, W] to the orbits of these flows; it is central to the integrability of , as a smooth distribution is integrable (i.e., to a ) if and only if it is closed under the Lie bracket. A concrete example in \mathbb{R}^2 with standard coordinates illustrates non-vanishing brackets: if V = \frac{\partial}{\partial x} and W = x \frac{\partial}{\partial y}, then [V, W] = \frac{\partial}{\partial y}.

Advanced concepts

In , given manifolds M and N and a f: M \to N, two vector fields V on M and W on N are said to be f-related if, for every point p \in M, the df_p: T_p M \to T_{f(p)} N satisfies df_p(V_p) = W_{f(p)}. This relation captures how vector fields transform under the mapping f, ensuring compatibility between their actions at corresponding points. When f is a , for any smooth field V on M, there exists a unique smooth field W on N that is f-related to V; this W is called the pushforward of V by f, denoted f_* V, and is explicitly defined by (f_* V)_{f(p)} = df_p(V_p) for all p \in M. The operation thus provides a way to transport fields between diffeomorphic manifolds while preserving their . Moreover, the respects the : for smooth fields V and W on M, [f_* V, f_* W] = f_* [V, W] on N. This property establishes that the induces a between the spaces of smooth fields on M and N, maintaining the algebraic relations intrinsic to these spaces. A example arises in coordinate transformations on a manifold, where a f between coordinate charts induces related fields via the ; in local coordinates, this corresponds to the standard transformation rule for components of contravariant fields under , ensuring invariance of geometric quantities like line elements. In the context of symmetries, f-related fields are central to equivariant structures under actions. Specifically, if a G acts ly on a manifold M via diffeomorphisms \phi_g: M \to M for g \in G, a field V on M is G-equivariant if V is \phi_g-related to itself for every g \in G, meaning d\phi_{g,p}(V_p) = V_{\phi_g(p)} for all p \in M. Such equivariant fields commute with the , facilitating the study of invariant dynamics, symmetry reduction, and conserved quantities in geometric .

Topological index

The topological index of a vector field V at an isolated zero p is defined as the degree of the , which sends a point on a small S surrounding p to the normalized vector V(q)/\|V(q)\| on the unit S^{n-1}, where n = \dim M. This integer-valued invariant captures the local winding behavior of the vector field around the singularity and is independent of the choice of surrounding , provided V has no other zeros nearby. The states that if M is a compact, oriented manifold without boundary and V is a smooth vector field on M with finitely many isolated zeros, then the sum of the indices of V at these zeros equals the \chi(M). This result, originally due to in two dimensions and extended by to higher dimensions, bridges local analytic properties of vector fields with global topological invariants of the manifold. In two dimensions, the index can be computed explicitly from the local phase portrait: a source or sink (radial node) has index +1, while a saddle has index -1. For instance, on the 2-sphere S^2 with \chi(S^2) = 2, any vector field with isolated zeros must have a total index of +2, such as two sources. A representative example is the gradient vector field of the height function on the standard embedded torus T^2 \subset \mathbb{R}^3, which has \chi(T^2) = 0. This field has four isolated zeros: one maximum and one minimum, each with index +1, and two saddles, each with index -1, yielding a sum of zero as required by the theorem. The has key applications in , where the indices of critical points of a function (whose negative is a vector field) sum to \chi(M), providing a topological obstruction to the of certain functions and linking critical point counts to Betti numbers via Morse inequalities. It also underpins s; for example, on even-dimensional spheres S^{2k} where \chi(S^{2k})=2 \neq 0, the theorem implies the non-existence of a nowhere-zero vector field, relating via topological arguments to the Brouwer fixed-point theorem for the (2k+1)-.

Generalizations

To other geometric structures

Vector fields generalize naturally to the setting of fiber bundles, where they are interpreted as smooth sections of associated . In this framework, given a P \to [M](/page/M) with structure group G and a \rho: [G](/page/G) \to GL(V) of a V, the associated vector bundle E = P \times_\rho V \to [M](/page/M) consists of fibers that are copies of V, and a vector field on [M](/page/M) corresponds to a global section \sigma: [M](/page/M) \to E that assigns to each point p \in [M](/page/M) an element \sigma(p) \in E_p \cong V, transforming appropriately under the . This construction allows vector fields to encode additional structure, such as gauge fields in physics, where sections of associated bundles describe configurations invariant under local symmetries. In pseudo-Riemannian manifolds, vector fields are tangent vectors equipped with an indefinite , enabling classifications into timelike, spacelike, or null vectors based on the sign of the inner product g(X, X), where g has (p, q) with p + q = \dim M. For instance, in four-dimensional with Lorentzian metric ( (1,3) or (3,1)), such vector fields model worldlines of particles in , where timelike fields correspond to causal trajectories. The geometry preserves the structure but alters lengths and angles, impacting notions like and completeness of geodesics generated by these fields. On complex manifolds, vector fields extend to holomorphic sections of the holomorphic tangent bundle T^{1,0}M, which is a where local coordinates allow expression as \sum \xi_j \frac{\partial}{\partial z_j} with holomorphic coefficients \xi_j. These fields preserve the complex structure J, satisfying L_X \omega = 0 for the Kähler form \omega in compatible cases, and generate holomorphic flows that respect the manifold's analytic properties. Such sections are crucial in , as their governs infinitesimal deformations of the manifold. Spinor fields represent a further of vector fields in quantum contexts, arising as sections of associated to the double cover of the , capturing spin representations that vectors cannot. In , Dirac spinors, which are sections of the spinor bundle S = S^+ \oplus S^- over , describe fermionic particles like electrons, transforming under the Spin(1,3) rather than the full . This structure resolves paradoxes in rotating frames, where spinors acquire phase factors under 360-degree rotations, unlike vector fields. A key example is the on spinor bundles, which acts on sections \psi \in \Gamma(S) as D\psi = \sum e_i \cdot \nabla_{e_i} \psi, where \{e_i\} is an orthonormal and \cdot denotes Clifford , effectively deriving vector-like behavior through its , the Clifford mimicking tangential directions. This , central to the iD\psi = m\psi, links dynamics to vector field flows via index theory and spectral properties on curved manifolds.

Infinite-dimensional vector fields

Infinite-dimensional vector fields arise in the study of manifolds modeled on infinite-dimensional topological vector spaces, such as Banach or Fréchet spaces, where the local model spaces replace finite-dimensional spaces. A vector field on such a manifold M is defined as a smooth section of the TM, meaning a smooth map V: M \to TM with \pi \circ V = \mathrm{id}_M, where \pi: TM \to M is the , ensuring V(p) \in T_p M for each p \in M. This generalizes the finite-dimensional notion, but the smoothness is defined using charts to the model spaces, often requiring careful handling of topologies like those induced by countable families of seminorms in Fréchet spaces. For rigor, Banach manifolds use complete normed spaces as models, while Fréchet manifolds employ complete metrizable locally convex spaces, allowing for more flexible structures like spaces of smooth functions. A concrete example is the space of smooth functions C^\infty(\mathbb{R}), which carries a natural Fréchet manifold structure via the seminorms \|f\|_{k} = \sup_{x \in \mathbb{R}} \sum_{j=0}^k |f^{(j)}(x)|. On this space, the translation vector field is given by V: C^\infty(\mathbb{R}) \to C^\infty(\mathbb{R}), where V(f) = f' and f' denotes the of f, acting as an element of the at f, which is identified with C^\infty(\mathbb{R}) itself due to the linear . This field generates translations along derivatives, illustrating how operators on function spaces serve as vector fields in infinite dimensions. Applications of infinite-dimensional vector fields are prominent in partial differential equations (PDEs), particularly in . The Navier-Stokes equations for incompressible fluids can be reformulated as the evolution equation \frac{du}{dt} = A(u), where u belongs to a suitable Sobolev or Besov (e.g., H^1(\Omega)^d for domain \Omega \subset \mathbb{R}^d), and A is a nonlinear vector field capturing the onto divergence-free fields, viscous terms, and . This viewpoint enables the analysis of global attractors and long-time behavior using infinite-dimensional . Key challenges in infinite-dimensional settings include the potential non-uniqueness of flows generated by vector fields, even for smooth fields, due to the lack of finite-dimensional compactness properties; for instance, divergence-free vector fields in \mathbb{R}^2 may admit multiple integral curves starting from the same point. Additionally, differentiability concepts diverge from the finite-dimensional case: Gâteaux differentiability, defined via directional derivatives along curves, is weaker and does not imply in infinite dimensions, whereas Fréchet differentiability requires uniform bounds over neighborhoods and is rarer for nonlinear operators on Banach spaces. These issues necessitate specialized techniques, such as regularization or Fréchet topologies, to ensure well-posedness.

References

  1. [1]
    Calculus III - Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · A vector field on two (or three) dimensional space is a function →F F → that assigns to each point (x,y) ( x , y ) (or (x,y,z) ( x , y , z ) ) a ...
  2. [2]
    [PDF] 1 Vector Fields - Michigan State University
    Definition(s) 1.1. 1. Let D be a set in R2 (plane region). A vector field on R2 is a function F that assigns to each point (x, y) in D a two-dimensional vector ...
  3. [3]
    [PDF] Lecture 5 Vector Operators: Grad, Div and Curl
    The three vector operators are the gradient of a scalar field, the divergence of a vector field, and the curl of a vector field.
  4. [4]
    16.5 Divergence and Curl - Vector Calculus
    Divergence and curl are two measurements of vector fields that are very useful in a variety of applications. Both are most easily understood by thinking of ...<|control11|><|separator|>
  5. [5]
    2 Differential Calculus of Vector Fields - Feynman Lectures - Caltech
    There are also vector fields. The idea is very simple. A vector is given for each point in space. The vector varies from point to point. As an example, consider ...
  6. [6]
    [PDF] V1. Plane Vector Fields
    From physics, we have the two-dimensional electrostatic force fields arising from a distri- bution of static (i.e., not moving) charges in the plane. At each ...
  7. [7]
    [PDF] 5 Vector fields
    We can use this as an alternative definition: Definition 5.2 (Vector fields – second definition). A vector field on M is a linear map. X : C•. (M) ! C. •. (M).
  8. [8]
    [PDF] Vector Calculus - People
    Definition 41.1. (Vector Field). Let E be a subset of a Euclidean space. A vector field F on E is a rule that assigns to each point P of E a unique vector F(P) ...
  9. [9]
    [PDF] 1.3 Vector Fields and Flows.
    Mar 1, 2012 · This section introduces vector fields on Euclidean space and the flows they determine. This topic puts together and globalizes two basic ...
  10. [10]
    [PDF] Vector fields and differential forms - Arizona Math
    Apr 28, 2006 · ... vector field is expressed in new co- ordinates, then these new coordinates are related to the old coordinates by a linear transformation.
  11. [11]
    [PDF] INTRODUCTION TO DIFFERENTIAL GEOMETRY - ETH Zürich
    These are notes for the lecture course “Differential Geometry I” given by the second author at ETH Zürich in the fall semester 2017. They are based on.
  12. [12]
    [PDF] 2.2 The derivative 2.3 Vector fields
    A derivation is a linear map D from smooth functions to R satisfying the Leibniz rule. D(fg) = fDg + gDf. The tangent bundle allows us to make sense of the ...
  13. [13]
    [PDF] Differential Geometry
    Aug 14, 2020 · We have defined vector fields as sections in the tangent bundle of a smooth manifold. ... Nomizu, Foundations of Differential Geometry Vol. I, ...
  14. [14]
    [PDF] Lecture notes on Differential Geometry
    Mar 2, 2023 · The goal of the course is to introduce the language of basic differential geometry. For this reason, a major part consists in building correct ...<|separator|>
  15. [15]
    None
    ### Summary of Élie Cartan's Contributions to Vector Fields and Differential Geometry on Manifolds
  16. [16]
    [PDF] Conservative Fields. A vector field is called gradient if ... - UCSD Math
    Theorem. A vector field F defined and continuously differentiable throughout a simply connected domain D is conservative if and only if it is irrotational in D.Missing: properties | Show results with:properties
  17. [17]
    [PDF] Gradient
    Every curve r(t) on the level curve or level surface satisfies d dt f(r(t)) = 0. By the chain rule, Vf(r(t)) is perpendicular to the tangent vector r′(t).
  18. [18]
    [PDF] Gradient: proof that it is perpendicular to level curves and surfaces
    We will show that at any point P = (x0,y0,z0) on the level surface f(x, y, z) = c (so f(x0,y0,z0) = c) the gradient v f| P is perpendicular to the surface. By ...
  19. [19]
    [PDF] Lecture 22: Path independence
    A vector field is called irrotational if curl(F) = 0 everywhere in the plane. Gradient fields are irrotational. Proof: If <P, Q> = <fx,fy>, then Qx - Py = fyx ...
  20. [20]
    [PDF] Chapter 5
    Sep 25, 2019 · The gravitational potential is defined as the scalar function Φ whose gradient is equal to the opposite of the gravitational field: One way ...
  21. [21]
    [PDF] Physics 162: Assignment 4
    Feb 2, 2015 · is a conservative force field, that is ∇ × f = 0. Eq. (21) implies the useful result that any central vector field of the form g(r)ˆr, that ...
  22. [22]
    [PDF] 16.1 Vector Fields - Montana State University
    A radial vector field is a vector field where all the vectors point straight towards (f (r) < 0) or away (f (r) > 0) from the origin, and which is rotationally ...
  23. [23]
    [PDF] Chapter 6 Vector Calculus - Computational Mechanics Group
    2. Note: Divergence of any central vector field is a central scalar field. Example 2: Gradient and divergence of a central field in spherical coordinates. If. , ...
  24. [24]
    7 The Theory of Gravitation - Feynman Lectures - Caltech
    What is this law of gravitation? It is that every object in the universe attracts every other object with a force which for any two bodies is proportional to ...
  25. [25]
    [PDF] Gradient, Divergence, Curl and Related Formulae - UT Physics
    To integrate this divergence over the volume of the ball, we use spherical coordinates in which the volume element is d. 3. Vol = r. 2 dr d. 2. Ω. −→. 4πr. 2 dr ...
  26. [26]
    Vector fields - Ximera - The Ohio State University
    A vector field in is a function where for every point in the domain, we assign a vector to the range.
  27. [27]
    [PDF] Chapter 6 Gravitation and Central-force motion - Physics
    And Newton's laws of motion with central gravitational forces are still very much in use today, such as in designing spacecraft trajectories to other planets.
  28. [28]
    6.1 Vector Fields - Calculus Volume 3 | OpenStax
    Mar 30, 2016 · In this section, we examine the basic definitions and graphs of vector fields so we can study them in more detail in the rest of this chapter.Missing: mathematics | Show results with:mathematics<|separator|>
  29. [29]
  30. [30]
    Divergence - Richard Fitzpatrick
    Such a field is called a solenoidal vector field. The simplest example of a solenoidal vector field is one in which the lines of force all form closed loops.
  31. [31]
  32. [32]
    [PDF] Chapter 3 Cartesian Vectors and Tensors: Their Calculus
    A solenoidal vector field is defined as one in which the divergence vanishes. This implies that the flux across a closed surface must also vanish. A vector ...
  33. [33]
    [PDF] Lectures on Symplectic Geometry - UC Berkeley math
    Jan 9, 2023 · The lectures cover topics such as quadratic Hamiltonians, symplectic manifolds, contact manifolds, and Arnold's conjecture, among others.
  34. [34]
    Vector field overview - Math Insight
    Vector fields are visualized by plotting vectors at points, in 2D and 3D. In 2D, vectors are plotted at (x,y); in 3D, at (x,y,z).Missing: conventions quiver
  35. [35]
    Quiver or vector plot - MATLAB - MathWorks
    quiver(X,Y,U,V) plots arrows with directional components U and V at the Cartesian coordinates specified by X and Y.
  36. [36]
    Calculus III - Line Integrals of Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will define the third type of line integrals we'll be looking at : line integrals of vector fields.
  37. [37]
    15.3 Line Integrals over Vector Fields
    Line integrals over vector fields share the same properties as line integrals over scalar fields, with one important distinction. The orientation of the curve C ...
  38. [38]
    Introduction to a line integral of a vector field - Math Insight
    One interpretation of the line integral of a vector field is the amount of work that a force field does on a particle as it moves along a curve.Missing: central | Show results with:central
  39. [39]
    16.3 The Fundamental Theorem of Line Integrals
    If a vector field F is the gradient of a function, F=∇f, we say that F is a conservative vector field. If F is a conservative force field, then the integral ...
  40. [40]
    [PDF] Lecture 20: Theorem of lineintegrals
    For a gradient field, the line-integral along any closed curve is zero. When is a vector field a gradient field? F(x, y) = Vf(x, y) implies Py(x, y) ...
  41. [41]
    [PDF] Discussion 16 Worksheet Answers - Conservative vector fields and ...
    C −y x2+y2 dx + x x2+y2 dy. P dx + Qdy = 2π where γ is the unit circle, oriented counterclockwise. the desired integral is 2π. ~F ◦dr = 0.
  42. [42]
    [PDF] class notes on hodge theory - John Etnyre
    Notice that we can use the above diagram to generalize the classical notions of divergence, gradient and curl to a general Riemannian 3-manifold (of course the ...
  43. [43]
    Calculating the divergence of a central vector field.
    Aug 24, 2014 · I am trying to calculate the divergence of a central electric field, namely the electric field due to a point charge and my book begins like this.Calculating the divergence of the Gravitational field ∇⋅→FDivergence of Curl for a vector space of dimension $nMore results from math.stackexchange.com
  44. [44]
    The idea of the curl of a vector field - Math Insight
    The curl of a vector field captures the idea of how a fluid may rotate. Imagine that the below vector field F represents fluid flow.
  45. [45]
    Calculus III - Curl and Divergence - Pauls Online Math Notes
    Nov 16, 2022 · There is also a definition of the divergence in terms of the ∇ ∇ operator. The divergence can be defined in terms of the following dot product.
  46. [46]
    Formal definition of curl in three dimensions (article) - Khan Academy
    F ‍ is a three-dimensional vector field. ( x , y , z ) ‍ is some specific point in 3d space. curl F ( x , y , z ) ‍ returns a three-dimensional vector.
  47. [47]
    2d curl formula (video) - Khan Academy
    Sep 26, 2016 · We can say as a formula, that the 2d curl, 2d curl, of our vector field v, as a function of x and y, is equal to the partial derivative of q with respect to x.
  48. [48]
    curl in nLab
    Jul 31, 2020 · In 3D Riemannian geometry, the curl of a vector field is defined as g − 1 ( ⋆ g d dR g ( v ) ) or via the limit integral formula.Definitions · In terms of differential forms · Via cross products · Examples
  49. [49]
    5.3 Newton's Second Law - University Physics Volume 1 | OpenStax
    Sep 19, 2016 · Newton's second law is closely related to his first law. It mathematically gives the cause-and-effect relationship between force and changes ...
  50. [50]
    13.1 Newton's Law of Universal Gravitation - OpenStax
    Sep 19, 2016 · Gravity is the force that forms the Universe. Problem-Solving Strategy. Newton's Law of Gravitation. To determine the motion caused by the ...
  51. [51]
    [PDF] the very, very basics of hamiltonian actions on symplectic manifolds
    The Hamiltonian vector field of f, (sometimes called the symplectic gradient of f), is the smooth vector field Xf corresponding to the differential 1-form ...
  52. [52]
    [PDF] Kepler's Laws - Central Force Motion - MIT OpenCourseWare
    When the only force acting on a particle is always directed to wards a fixed point, the motion is called central force motion.
  53. [53]
    [PDF] arXiv:1907.02947v3 [math-ph] 13 Jun 2020
    Jun 13, 2020 · This sode XL ≡ ΓL is called the Euler–Lagrange vector field associated with the La- grangian function L. Proof. It follows from the coordinate ...
  54. [54]
    20 Solutions of Maxwell's Equations in Free Space
    From a mathematical view, there is an electric field vector and a magnetic field vector at every point in space; that is, there are six numbers associated with ...
  55. [55]
    [PDF] 20.1 Revisiting Maxwell's equations - MIT
    Apr 28, 2005 · The source free Maxwell's equations show us that ~E and ~B are coupled: variations in ~E act as a source for ~B, which in turn acts as a ...
  56. [56]
    5.14: Electric Field as the Gradient of Potential
    Sep 12, 2022 · The electric field intensity at a point is the gradient of the electric potential at that point after a change of sign.Missing: static | Show results with:static
  57. [57]
    5 Application of Gauss' Law - Feynman Lectures - Caltech
    Since in a conducting material the electric field is everywhere zero, the divergence of E is zero, and by Gauss' law the charge density in the interior of the ...
  58. [58]
    Maxwell's Equations - HyperPhysics
    The divergence of a vector field is proportional to the point source density, so the form of Gauss' law for magnetic fields is then a statement that there are ...
  59. [59]
    [PDF] 3. Magnetostatics - DAMTP
    solenoid, the constant magnetic field is given by. B = µ0IN z. (3.7). Note that, since K = IN, this is consistent with our general formula for the disconti-.
  60. [60]
    18 The Maxwell Equations - Feynman Lectures - Caltech
    In particular, the equation for the magnetic field of steady currents was known only as ∇×B=jϵ0c2. Maxwell began by considering these known laws and expressing ...
  61. [61]
    The Feynman Lectures on Physics Vol. II Ch. 13: Magnetostatics
    The total electromagnetic force on a charge can, then, be written as F=q(E+v×B). This is called the Lorentz force. Fig. 13–1.
  62. [62]
    [PDF] Existence and uniqueness for integral curves - Academic Web
    Clearly smooth vector fields are locally Lipshitz, and we have the following: Theorem B (Picard–Lindelöf theorem). Let ~v: U → Rn be a locally Lipshitz vector ...
  63. [63]
    [PDF] lecture 14: integral curves of smooth vector fields
    In calculus and in ODE, we learned the conception of integral curves of such a vector field: an integral curve is a parametric curve that represents a specific ...
  64. [64]
    [PDF] Introduction - UC Davis Math
    The following result, due to Picard and Lindelöf, is the fundamental local ex- istence and uniqueness theorem for IVPs for ODEs. It is a local existence theorem.
  65. [65]
    [PDF] 2.4.2 The Flow of a Vector Field - UCI Mathematics
    A vector field X Vect(M) is called complete if, for each po M, there is an integral curvey: R→ M of X with y(0) = Po⋅. Page 3. 42. CHAPTER 2. FOUNDATIONS. Lemma ...
  66. [66]
    [PDF] DYNAMICS OF VECTOR FIELDS 1. Integral Curves Suppose M is a ...
    Any smooth vector field on a compact manifold is complete. Proof. The set Supp(X), as a closed set in the compact manifold, is compact. D. 2.<|control11|><|separator|>
  67. [67]
    [PDF] 5 Vector fields
    If two vector fields X,Y 2 X(M) are tangent to a submanifold S ✓. M, then their Lie bracket is again tangent to S. Proposition 5.2 can be proved by using the ...Missing: law | Show results with:law
  68. [68]
    [PDF] Complete polynomial vector fields on C , Part I
    Nonetheless we immediately check that the resulting vector field X is not complete in this case. Finally let us assume that i = 3. It is again easy to see ...
  69. [69]
    Incomplete vector field - differential geometry - Math Stack Exchange
    May 3, 2013 · For instance on R, X=(x2+1)∂∂x, is incomplete as it shoots off to infinity in finite time as evidenced by its flow F=tan(t−C). Should I be able ...Prove that a specific vector field is not completeHow to show that a given vector field is not complete in $\mathbb{R}^2More results from math.stackexchange.com
  70. [70]
    [PDF] Teleparallelism as anholonomic geometry - arXiv
    ... Lie bracket – that is useful in differential geometry: [v, w] f = v (wf) – w (vf) = (v j ∂j w i. – w j ∂j v i. ) ∂i f . (4.3). That is, the components of the ...
  71. [71]
    [PDF] TOPOLOGY FROM THE DIFFERENTIABLE VIEWPOINT
    (A 2-dimensional version of this theorem was proved by Poincare in 1885. The full theorem was proved by Hopf [14] in 1926 after earlier partial results by ...
  72. [72]
    [PDF] the euler characteristic, poincare-hopf theorem, and applications
    In section 5, we discuss Morse theory and indicate how it can be used to identify a smooth vector field with the Euler characteristic. Section 6 quickly proves ...
  73. [73]
    [PDF] differential topology: morse theory and the euler characteristic
    The height function on T2 is a Morse function with four critical points: the topmost and the bottom-most points of the torus have indices 2 and 0, respectively, ...
  74. [74]
    [PDF] poincaré-hopf theorem and morse theory
    Jun 9, 2020 · The general goal of this thesis is to compare two different ways of proving the Poincaré-Hopf theorem. One of them includes Morse theory and ...
  75. [75]
    [PDF] Principal Bundles and Associated Vector Bundles - Clear Physics
    In a vector bundle, the fiber is a vector space V . For some types of classical fields, a configuration of the field may be described as a smooth section of a ...
  76. [76]
    [PDF] Connections on fibre bundles - Spin Geometry
    A connection on a principal bundle allows us to define a covariant derivative (a.k.a. a Koszul con- nection) on sections of any associated vector bundle. If E → ...
  77. [77]
  78. [78]
    [PDF] SEMI-RIEMANNIAN GEOMETRY
    This book is an exposition of semi-Riemannian geometry (also called pseudo-Riemannian geometrytthe study of a smooth manifold fur- nished with a metric ...
  79. [79]
    [PDF] 1. Complex Vector bundles and complex manifolds
    Holomorphic maps between complex manifolds are defined so that their restriction to each chart is holomorphic. Note that a complex manifold is an almost complex ...
  80. [80]
    [PDF] Math 222B, Complex Variables and Geometry - UCI Mathematics
    Definition 3.4. A holomorphic vector field on a complex manifold (M,J) is vector field Z ∈ Γ(T1,0) which satisfies Zf is ...
  81. [81]
    [PDF] Two-component spinor techniques and Feynman rules for quantum ...
    Jun 4, 2022 · Abstract. Two-component spinors are the basic ingredients for describing fermions in quantum field theory in 3 + 1 spacetime dimensions.
  82. [82]
    Emergent Spinor Fields from Exotic Spin Structures - Oxford Academic
    May 7, 2024 · Spinor fields, according to the classical definition, carry irreducible representations of the Spin group, constructed upon the twisted Clifford ...<|separator|>
  83. [83]
    Connections and the Dirac operator on spinor bundles - ScienceDirect
    Spinor fields are then defined as sections of a vector bundle Σ → M carrying a representation of the Clifford bundle. Karrer proved the existence of a covariant ...Missing: like | Show results with:like
  84. [84]
    [PDF] Lectures on Dirac Operators and Index Theory - UCSB Math
    Jan 7, 2015 · Definition: [C-spinor bundles] The complex spinor bundle S → M is the associated vector bundle. S = Pspin ×ρn ∆n. Remark If PG → M is a ...
  85. [85]
    infinite-dimensional manifold in nLab
    ### Summary of Infinite-Dimensional Manifold from nLab
  86. [86]
    Chapter 1 Frechet manifolds - Project Euclid
    A Hausdorff topological vector space F with the topology, defined by a countable system of seminorms, is called a Frechet space iff it is complete. Example 1.
  87. [87]
    [PDF] Infinite dimensional manifolds
    In these lectures I will try to give an overview on infinite dimensional manifolds with special emphasis on: Lie groups of diffeomorphisms, manifolds of ...<|control11|><|separator|>
  88. [88]
    978-1-4612-0645-3.pdf
    Infinite-dimensional dynamical systems in mechanics and physics/. Roger Temam. ... Ghidaglia and R. Temam [6]). Also the theory of attractors for stochastic.
  89. [89]
    [PDF] Infinite Dimensional Dynamical Systems and the Navier-Stokes ...
    Aug 2, 2007 · I will discuss the existence and properties of invariant manifolds for dynamical systems defined on Banach spaces and review the theory of ...
  90. [90]
  91. [91]
    [PDF] infinite-dimensional carnot groups and gآteaux differentiability - ARPI
    Fréchet differentiability clearly implies Gâteaux differentiability, but the opposite does not hold in general in the infinite dimensional setting.
  92. [92]
    [PDF] arXiv:2411.03265v1 [math.DG] 5 Nov 2024
    Nov 5, 2024 · infinite dimensional tame Fréchet manifold. Proof. There are different ways of exhibiting a differentiable manifold struc- ture of this space.