Fact-checked by Grok 2 weeks ago

Four-gradient

In , the four-gradient (or 4-gradient) is the covariant four-vector analogue of the ordinary three-dimensional operator from , defined in four-dimensional as the with respect to the coordinates \partial_\mu = \frac{\partial}{\partial x^\mu}, where the coordinates are typically x^\mu = (ct, x, y, z) with the \eta_{\mu\nu} = \operatorname{diag}(-1, +1, +1, +1). When applied to a \phi, it produces a \partial_\mu \phi, whose components are \left( \frac{1}{c} \frac{\partial \phi}{\partial t}, \frac{\partial \phi}{\partial x}, \frac{\partial \phi}{\partial y}, \frac{\partial \phi}{\partial z} \right) in the mostly-plus convention, ensuring under transformations between inertial frames. The four-gradient transforms as a covariant , meaning its components adjust according to the rules to maintain the invariance of physical laws across reference frames. In the mostly-minus metric convention (common in ), its explicit form is \partial^\mu = \left( \frac{\partial}{\partial t}, -\nabla \right), where \nabla is the spatial , highlighting the sign flip in the contravariant spatial components to align with the metric's structure. This is fundamental in formulating relativistic field theories, such as electrodynamics, where it appears in in covariant form: for instance, the of the four-current j^\mu satisfies \partial_\mu j^\mu = 0, expressing local invariantly. Key applications include deriving the wave equation via the d'Alembertian \square = \partial^\mu \partial_\mu = \frac{1}{c^2} \frac{\partial^2}{\partial t^2} - \nabla^2, which is a governing propagation in relativistic fields. In , the four-gradient extends to curved through the , but in flat , it remains the , underscoring its role as a building block for tensorial descriptions of physical quantities. Its contravariant counterpart \partial^\mu = \eta^{\mu\nu} \partial_\nu facilitates contractions that yield invariants, essential for ensuring the consistency of equations across boosted frames.

Fundamentals

Notation

The four-gradient in relativistic notation is represented using partial derivative operators with indices to denote its four-vector nature. The contravariant form is denoted as \partial^\mu, where the raised index \mu indicates the contravariant components, while the covariant form is \partial_\mu with the lowered index. Greek indices such as \mu run from 0 to 3, corresponding to the time component and the three spatial components, with the coordinate x^0 = ct where c is the and t is time. The explicit components depend on the convention. In the mostly plus signature (- ,+,+,+), the contravariant four-gradient has components \partial^\mu = \left( -\frac{1}{c} \frac{\partial}{\partial t}, \nabla \right), where \nabla is the three-dimensional . In the mostly minus signature (+,-,-,-), the components are \partial^\mu = \left( \frac{1}{c} \frac{\partial}{\partial t}, -\nabla \right). These conventions affect the relative signs, particularly for the spatial components in one case and the time component in the other. The indices are raised and lowered using the Minkowski \eta_{\mu\nu}, which is diagonal with entries determined by the chosen signature. A brief historical note: the Einstein , which implies automatic summation over repeated indices (e.g., A^\mu B_\mu), was introduced by in his 1916 paper on to simplify tensor expressions.

Definition

In four-dimensional Minkowski spacetime, the four-gradient generalizes the three-dimensional gradient operator from to incorporate the time dimension, assuming familiarity with partial derivatives of scalar s. Commonly, the four-gradient refers to the covariant form. For a field \phi (a invariant under Lorentz transformations), the covariant four-gradient is defined as the \partial_\mu \phi, with components given by the partial derivatives with respect to the spacetime coordinates x^\mu = (ct, \mathbf{x}), where c is the and \mathbf{x} = (x, y, z). To align with the mostly plus metric convention \eta_{\mu\nu} = \operatorname{diag}(-1, +1, +1, +1) used in the article introduction, the components are \partial_\mu \phi = \left( \frac{1}{c} \frac{\partial \phi}{\partial t}, \nabla \phi \right), where \nabla \phi = \left( \frac{\partial \phi}{\partial x}, \frac{\partial \phi}{\partial y}, \frac{\partial \phi}{\partial z} \right) is the ordinary three-dimensional gradient. This definition ensures dimensional consistency, as all components have units of inverse length, and aligns with the structure of special relativity by treating time and space on equal footing via the factor of $1/c. The contravariant four-gradient \partial^\mu \phi is obtained by raising the index using the : \partial^\mu \phi = \eta^{\mu\nu} \partial_\nu \phi = \left( -\frac{1}{c} \frac{\partial \phi}{\partial t}, \nabla \phi \right). The negative sign in the time component arises from the , which reflects the opposite signs in the interval ds^2 = -c^2 dt^2 + d\mathbf{x}^2. While both forms are used depending on context, the covariant version \partial_\mu \phi is often denoted simply as the four-gradient in applications. Under Lorentz boosts, the components of \partial_\mu \phi transform as a covariant according to the inverse : \partial'_\mu \phi = (\Lambda^{-1})^\nu{}_\mu \partial_\nu \phi. This property holds because \phi is a scalar (unchanged under s), so the transformation of its derivatives mirrors the inverse transformation of the coordinates, preserving the vector nature across inertial frames. For a boost along the x-direction with v = \beta c, the matrix elements involve \gamma = 1/\sqrt{1 - \beta^2}, ensuring relativistic invariance of physical laws involving the four-gradient.

Derivation

Generalization from 3D Gradient

The generalization of the three-dimensional gradient to the four-gradient emerged within the framework of , particularly through Hermann Minkowski's 1908 formulation of as a unified four-dimensional manifold that treats space and time on equal footing. This approach resolved inconsistencies in by introducing a geometry where physical laws remain invariant under Lorentz transformations, extending Euclidean concepts to a pseudo-Riemannian structure. In three-dimensional Euclidean space, the gradient of a scalar field \phi is defined as the vector \nabla \phi = \left( \frac{\partial \phi}{\partial x}, \frac{\partial \phi}{\partial y}, \frac{\partial \phi}{\partial z} \right), capturing the spatial rate of change and direction of steepest increase. To extend this to four-dimensional spacetime while preserving invariance under Lorentz transformations, the time dimension is incorporated by treating time as a coordinate scaled by the speed of light c. The process begins by considering the partial derivative with respect to the time-like coordinate ct, yielding the components \frac{\partial \phi}{\partial (ct)} alongside the spatial derivatives. This step ensures that the resulting object transforms as a four-vector, maintaining the structure of physical equations across inertial frames. Dimensional consistency is achieved through the Minkowski metric, where the spacetime interval is ds^2 = -c^2 dt^2 + dx^2 + dy^2 + dz^2, homogenizing the units of time and space into length squared. The factor of c in ct aligns the time derivative \frac{\partial}{\partial (ct)} = \frac{1}{c} \frac{\partial}{\partial t} with the spatial ones, all having dimensions of inverse length. The full four-gradient operator is then expressed as \partial_\mu \phi = \frac{\partial \phi}{\partial x^\mu}, where the coordinates are x^\mu = (ct, x, y, z). This construction is limited to flat Minkowski spacetime, where the metric is constant, and does not extend directly to curved spacetimes requiring the in .

Expression in Minkowski Spacetime

In Minkowski spacetime with (-1, +1, +1, +1), the four-gradient is defined in covariant form as the set of partial derivatives with respect to the spacetime coordinates x^\mu = (ct, \mathbf{x}), where c is the and \mathbf{x} = (x,y,z). Specifically, \partial_\mu = \left( \frac{1}{c} \frac{\partial}{\partial t}, \nabla \right), with the temporal component \partial_0 = \frac{1}{c} \frac{\partial}{\partial t} and the spatial components \partial_i = \frac{\partial}{\partial x^i} for i=1,2,3, where \nabla = \left( \frac{\partial}{\partial x}, \frac{\partial}{\partial y}, \frac{\partial}{\partial z} \right) is the three-dimensional . The contravariant four-gradient is obtained by raising the index using the inverse Minkowski metric \eta^{\mu\nu} = \operatorname{diag}(-1, +1, +1, +1), via \partial^\mu = \eta^{\mu\nu} \partial_\nu. This yields \partial^0 = -\frac{1}{c} \frac{\partial}{\partial t} and \partial^i = \frac{\partial}{\partial x^i} for i=1,2,3, so \partial^\mu = \left( -\frac{1}{c} \frac{\partial}{\partial t}, \nabla \right). This operator satisfies the invariance property under coordinate transformations in , expressed as \partial_\mu x^\nu = \delta^\nu_\mu, where \delta^\nu_\mu is the , reflecting its role as the basis for in the . For example, consider the simple \phi = ct = x^0. The covariant four-gradient is then \partial_\mu \phi = (1, 0, 0, 0), a constant independent of position, illustrating how the operator extracts coordinate differentials. The four-gradient ties directly to the ds^2 = \eta_{\mu\nu} dx^\mu dx^\nu, as the infinitesimal change in a is d\phi = \partial_\mu \phi \, dx^\mu, connecting local variations to the .

Mathematical Roles

As a 4-Divergence Operator

The four-divergence of a contravariant four-vector field A^\mu is defined as the scalar quantity obtained by contracting it with the operator \partial_\mu, yielding \partial_\mu A^\mu. This operation generalizes the three-dimensional to four-dimensional Minkowski , where the \partial_\mu acts on the components of A^\mu using the Minkowski to raise or lower indices appropriately. Physically, the four-divergence measures the net of the field through the of a four-dimensional in , analogous to how the ordinary quantifies source strength or net outflow in . In integral form, by the four-dimensional generalization of , the volume integral of \partial_\mu A^\mu equals the surface integral of A^\mu over the enclosing , capturing the relativistic interplay between temporal and spatial flows. For the charge four-current J^\mu = (c\rho, \mathbf{J}), where \rho is the and \mathbf{J} is the three-current density, the four-divergence is \partial_\mu J^\mu = \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{J}, linking the rate of change of to the of the current in a manifestly covariant manner. This form preserves units and ensures the expression transforms correctly under Lorentz boosts. The four-divergence satisfies the Leibniz product rule for a scalar field \phi and a four-vector A^\mu, given by \partial_\mu (\phi A^\mu) = (\partial_\mu \phi) A^\mu + \phi \partial_\mu A^\mu, allowing differentiation of products in a way that maintains covariance. Regarding antisymmetry, the contraction \partial_\mu A^\mu inherits properties from the Minkowski metric's signature, where spatial components contribute positively to the sum while the time component does so with the opposite sign due to index raising/lowering, ensuring the overall scalar is invariant. This specific contraction \partial_\mu A^\mu is unique as the Lorentz-invariant definition of divergence for a four-vector, as any other index pairing would not yield a scalar under Lorentz transformations, preserving the relativistic structure of physical laws.

As the d'Alembertian Operator

The d'Alembertian operator, denoted as \square, arises as the second-order contraction of the with itself in . For a \phi, this is expressed as \square \phi = \partial^\mu \partial_\mu \phi, where \partial^\mu is the contravariant four-gradient and the Einstein summation convention is used over the spacetime indices \mu = 0, 1, 2, 3.<grok:render type="render_inline_citation"> 30 </grok:render> In where c = 1, and consistent with the mostly plus (-+++), the explicit form is \square = \partial^\mu \partial_\mu = -\frac{\partial^2}{\partial t^2} + \nabla^2, with the time derivative contributing negatively and the spatial Laplacian positively.<grok:render type="render_inline_citation"> 30 </grok:render> Restoring the c, the operator becomes \square = -\frac{1}{c^2} \frac{\partial^2}{\partial t^2} + \nabla^2.<grok:render type="render_inline_citation"> 31 </grok:render> The sign structure of the d'Alembertian depends on the choice of in Minkowski , conventionally either (+,-,-,-) or (- ,+ ,+ ,+). In the mostly-minus signature (+,-,-,-), the time component yields a positive contribution \left(\frac{1}{c} \frac{\partial}{\partial t}\right)^2, while the spatial components yield negative contributions -\left(\frac{\partial}{\partial x^i}\right)^2 for i=1,2,3, resulting in the wave-like form \square = \frac{1}{c^2} \frac{\partial^2}{\partial t^2} - \nabla^2.<grok:render type="render_inline_citation"> 30 </grok:render> The opposite mostly-plus signature (- ,+ ,+ ,+) reverses these signs, producing \square = -\frac{1}{c^2} \frac{\partial^2}{\partial t^2} + \nabla^2, though the physical content remains under consistent usage.<grok:render type="render_inline_citation"> 42 </grok:render> This contraction ensures that \square is a , under relativistic transformations.<grok:render type="render_inline_citation"> 30 </grok:render> Unlike the three-dimensional Laplacian \nabla^2, which is an associated with steady-state problems, the d'Alembertian is due to its mixed signs, leading to wave propagation solutions with finite speed rather than instantaneous .<grok:render type="render_inline_citation"> 32 </grok:render> This character is fundamental to relativistic theories, phenomena like light cones and from non-relativistic elliptic equations. A key application of the d'Alembertian appears in the Klein-Gordon equation for a free of mass m: \square \psi + \frac{m^2 c^2}{\hbar^2} \psi = 0, where \psi is the field, \hbar is the reduced Planck constant, and the d'Alembertian acts consistent with the mostly-plus .<grok:render type="render_inline_citation"> 31 </grok:render> This second-order relativistic wave equation generalizes the non-relativistic to incorporate , yielding solutions that describe massive particles with dispersion relations E^2 = p^2 c^2 + m^2 c^4.<grok:render type="render_inline_citation"> 31 </grok:render>

As a Jacobian for Metric Transformations

In special relativity, coordinate transformations in Minkowski spacetime that preserve the form of the are described by the matrix whose elements are the components of the four-gradient applied to the transformed coordinates. Specifically, for a transformation from coordinates x^\nu to new coordinates x'^\mu, the is \Lambda^\mu{}_\nu = \frac{\partial x'^\mu}{\partial x^\nu}, where the partial derivatives constitute the four-gradient operator \partial_\nu. This matrix encapsulates how four-vectors, including those derived from the four-gradient, transform under such changes. The role of this Jacobian in metric preservation is central to the structure of Lorentz transformations. The Minkowski metric \eta_{\mu\nu} remains invariant, satisfying \eta'_{\alpha\beta} = (\Lambda^{-1})^\mu{}_\alpha (\Lambda^{-1})^\nu{}_\beta \eta_{\mu\nu} = \eta_{\alpha\beta}, or equivalently in matrix form, \Lambda^T \eta \Lambda = \eta, ensuring that the spacetime interval ds^2 = \eta_{\mu\nu} dx^\mu dx^\nu is unchanged across inertial frames. For an explicit example, consider a Lorentz boost along the x-direction (1-direction) with relative velocity v. The relevant non-zero components of the are \frac{\partial x'^0}{\partial x^0} = \gamma, \frac{\partial x'^0}{\partial x^1} = -\gamma \beta, \frac{\partial x'^1}{\partial x^0} = -\gamma \beta, and \frac{\partial x'^1}{\partial x^1} = \gamma, where \beta = v/c and \gamma = (1 - \beta^2)^{-1/2}, with c the (often set to 1 in ). These components arise directly from differentiating the boost transformation equations ct' = \gamma (ct - \beta x) and x' = \gamma (x - \beta ct). The determinant of the Jacobian matrix for transformations in the proper is |\Lambda| = 1, which maintains the orientation of and ensures volume preservation under these metric-preserving changes. In the context of infinitesimal transformations, the four-gradient connects to the generators of the through the variation \delta x^\mu = \varepsilon^\nu \partial_\nu x^\mu, where \varepsilon^\nu parameterizes the small displacement; in , this yields \delta x^\mu = \omega^\mu{}_\nu x^\nu for the linear Lorentz generators \omega^\mu{}_\nu, highlighting the four-gradient's role in deriving the of symmetries.

In 4D Stokes' Theorem

The four-gradient, acting as the four-divergence operator on a contravariant A^\mu, is central to the in four-dimensional Minkowski . This equates the volume of the to the surface over the : \int_V \partial_\mu A^\mu \, d^4 x = \int_S A^\mu \, d\Sigma_\mu, where V denotes a compact four-volume in , S = \partial V is its oriented three-dimensional , and the is taken with respect to the flat Minkowski measure d^4 x = c \, dt \, dx \, dy \, dz. This relation generalizes the classical three-dimensional to four dimensions, expressing the net "source" of the field A^\mu within V—as measured by its four-divergence—in terms of the flux through the enclosing S. It underpins integral formulations of conservation principles over volumes, where the local vanishing of the four-divergence implies global balance via boundary fluxes. The oriented surface element d\Sigma_\mu is a covariant density representing the infinitesimal area on S projected along the \mu-direction, with orientation induced by the relative to the volume V (typically outward). In coordinates, for a parametrized by three variables, d\Sigma_\mu = n_\mu \, dS, where n_\mu is the unit covector normal to S and dS is the scalar area element; the Minkowski metric raises or lowers indices as needed to ensure . Mathematically, the theorem relies on the equivalence between and differential forms in flat : the four-divergence \partial_\mu A^\mu corresponds to the d applied to the associated (n-1)-form dual to A^\mu in n=4 dimensions, yielding the general \int_V d\omega = \int_S \omega for a 3-form \omega. An illustrative application involves integrating over a worldtube—a volume consisting of a worldline segment extended by a spatial cross-section—for . If the four-current j^\mu obeys \partial_\mu j^\mu = 0, the theorem yields \int_V \partial_\mu j^\mu \, d^4 x = 0 = \int_S j^\mu \, d\Sigma_\mu, equating the enclosed charge (from the time-like faces) to the lateral current flux, thus verifying conservation along the tube without internal sources.

Transformations in Special Relativity

Relation to Lorentz Transformations

The four-gradient operator \partial_\mu transforms as a covariant under , ensuring the relativistic covariance of physical laws involving derivatives. Specifically, if the coordinates transform as x'^\rho = \Lambda^\rho{}_\sigma x^\sigma, where \Lambda^\rho{}_\sigma is the matrix, the chain rule yields the transformation law \partial'_\mu = \frac{\partial x^\nu}{\partial x'^\mu} \partial_\nu = (\Lambda^{-1})^\nu{}_\mu \partial_\nu. In mixed , this is often expressed as \partial'_\mu = \Lambda_\mu{}^\nu \partial_\nu, where \Lambda_\mu{}^\nu denotes the appropriate components of the transformation adjusted for the . This transformation mixes the components of the time and space derivatives. For a boost along the x-direction with velocity v (setting c=1), the Lorentz matrix has \gamma = 1/\sqrt{1 - \beta^2} and \beta = v, leading to \partial'_0 = \gamma (\partial_0 + \beta \partial_1) and \partial'_1 = \gamma (\beta \partial_0 + \partial_1), while \partial'_2 = \partial_2 and \partial'_3 = \partial_3. The transverse components remain unchanged, but the longitudinal ones couple the temporal derivative \partial_0 = \partial / \partial t with the spatial derivative \partial_1 = \partial / \partial x, reflecting the . The four-gradient of a Lorentz scalar \phi(x) such that \phi'(x') = \phi(x) transforms as a covariant four-vector: \partial'_\mu \phi' = \Lambda_\mu{}^\nu \partial_\nu \phi. This ensures that contractions involving the four-gradient, such as \partial^\mu A_\mu for a contravariant four-vector A^\mu, are Lorentz invariants, as the raised four-gradient \partial^\mu = \eta^{\mu\nu} \partial_\nu and A^\mu transform oppositely, maintaining the Minkowski inner product. In the context of plane waves, the phase factor e^{i k_\mu x^\mu} is a Lorentz scalar, requiring the four-momentum covector k_\mu to transform covariantly as k'_\mu = \Lambda_\mu{}^\nu k_\nu to preserve invariance under the coordinate change x'^\mu = \Lambda^\mu{}_\rho x^\rho. This ensures that wave equations involving the four-gradient, such as the Klein-Gordon equation (\partial^\mu \partial_\mu + m^2) \psi = 0, remain form-invariant. The proper orthochronous , consisting of transformations with determinant +1 and preserving the time orientation, maintains the structure of the four-gradient as a covector , as these transformations preserve the Minkowski \eta_{\mu\nu}. Transformations outside this , such as or time reversal, alter the sign conventions but still respect the overall tensorial form when appropriately defined.

As Part of Proper Time Derivative

In relativistic kinematics, the four-gradient operator plays a central role in defining the proper time derivative along a particle's worldline. The proper time \tau is the invariant interval measured by a clock moving with the particle, given by d\tau = dt \sqrt{1 - v^2/c^2}, where t is coordinate time and v is the particle's speed. The four-velocity u^\mu = dx^\mu / d\tau is the tangent vector to the worldline, with components u^\mu = \gamma (c, \mathbf{v}), where \gamma = 1/\sqrt{1 - v^2/c^2}. This four-velocity satisfies the normalization condition u^\mu u_\mu = -c^2 in the Minkowski metric with signature (- , + , + , +), ensuring its magnitude is fixed and independent of the reference frame. The derivative of a \phi along the worldline is expressed as \frac{D\phi}{D\tau} = u^\mu \partial_\mu \phi, where \partial_\mu denotes the four-gradient components \partial_\mu = (\partial / \partial (ct), \nabla). This operator d/d\tau = u^\mu \partial_\mu generalizes the classical to four dimensions, capturing the total rate of change of \phi as observed in the particle's instantaneous . Physically, it represents the invariant rate of variation experienced by the particle, unaffected by Lorentz boosts, since both proper time and the four-gradient transform covariantly. For tensor fields, such as a V^\nu, the extension of the derivative takes the form of the along the : \mathcal{L}_u V^\nu = u^\mu \partial_\mu V^\nu - V^\mu \partial_\mu u^\nu. In flat Minkowski , this reduces to the convective transport of the along the , preserving the tensorial structure under coordinate changes. This formulation ensures that the evolution of vector quantities remains covariant, aligning with the principles of .

Definition of 4-Wavevector

The four-wavevector, denoted k^\mu, is a four-vector in Minkowski spacetime that characterizes the propagation of waves, particularly in the context of . It is defined as the four-gradient of the phase function \phi of a wave, specifically k^\mu = -i \partial^\mu \phi for complex phase representations, ensuring consistency with the eikonal approximation in wave optics and . For a plane wave described by the form \exp[i (\omega t - \mathbf{k} \cdot \mathbf{x})], where \omega is the angular frequency and \mathbf{k} is the three-dimensional wavevector, the components are k^\mu = (\omega/c, \mathbf{k}), with the spatial part \mathbf{k} having magnitude |\mathbf{k}| = 2\pi / \lambda for wavelength \lambda. In the plane wave case, the \phi = k_\mu x^\mu satisfies \partial_\mu \phi = k_\mu, where the applied to the phase directly yields the covector form, and raising the index with the gives the contravariant four-wavevector. The time component \omega / c has units of inverse (wave number), matching the units of the spatial components to ensure dimensional homogeneity in relativistic formulations. For massless such as , the k_\mu k^\mu = 0 holds, implying \omega = c |\mathbf{k}| and classifying the four-wavevector as null. Under Lorentz transformations, the four-wavevector transforms as k'^\mu = \Lambda^\mu{}_\nu k^\nu, where \Lambda^\mu{}_\nu is the matrix, preserving the k'_\mu k'^\mu = 0 due to the invariance of the Minkowski metric. This transformation property ensures that the k_\mu x^\mu remains invariant across inertial frames, maintaining the physical coherence of wave fronts.

Applications in

Construction of Faraday Tensor

The electromagnetic Faraday tensor F^{\mu\nu}, also known as the field strength tensor, is constructed antisymmetrically from the four-potential A^\mu using components of the four-gradient operator \partial^\mu. The four-potential is defined as A^\mu = \left( \frac{[\phi](/page/Phi)}{[c](/page/Speed_of_light)}, \mathbf{A} \right), where [\phi](/page/Phi) is the electric and \mathbf{A} is the , with [c](/page/Speed_of_light) denoting the . Explicitly, the tensor is given by F^{\mu\nu} = \partial^\mu A^\nu - \partial^\nu A^\mu, where the partial derivatives \partial^\mu = \frac{\partial}{\partial x_\mu} act in the contravariant sense, incorporating the Minkowski metric to raise or lower indices as needed. This construction encodes the electric and magnetic fields as the independent components of F^{\mu\nu}. In a specific inertial frame, the non-zero components are F^{0i} = -\frac{E_i}{c} for the electric field contributions (with i = 1,2,3) and F^{ij} = -\epsilon^{ijk} B_k for the magnetic field, where \epsilon^{ijk} is the Levi-Civita symbol and summation over repeated indices k is implied. The antisymmetric nature of the tensor follows directly from its definition, yielding F_{\mu\nu} = -F_{\nu\mu} upon lowering indices with the . Consequently, F^{\mu\nu} possesses exactly six independent components, corresponding to the three components each of the \mathbf{E} and \mathbf{B}. This structure unifies the two fields into a single relativistic entity. The Faraday tensor exhibits gauge invariance under transformations of the four-potential of the form A^\mu \to A^\mu + \partial^\mu \Lambda, where \Lambda is an arbitrary scalar function. The added term vanishes in the antisymmetric difference, as \partial^\mu \partial^\nu \Lambda - \partial^\nu \partial^\mu \Lambda = 0 due to the equality of mixed partial derivatives, ensuring that physical observables derived from F^{\mu\nu} remain unchanged. This tensorial formulation originated with Hermann Minkowski's 1908 work on spacetime, where he first demonstrated the unification of electric and magnetic fields as components of a single antisymmetric second-rank tensor in four-dimensional Minkowski space.

Derivation of Maxwell's Equations

In the relativistic formulation of electromagnetism, Maxwell's equations emerge naturally from the action of the four-gradient operator on the Faraday tensor F^{\mu\nu}, which encodes the electric and magnetic fields in a Lorentz-covariant manner. The homogeneous set of Maxwell's equations, corresponding to Faraday's law and the absence of magnetic monopoles, follows from the Bianchi identity for the antisymmetric tensor F_{\mu\nu}: \partial_\lambda F_{\mu\nu} + \partial_\mu F_{\nu\lambda} + \partial_\nu F_{\lambda\mu} = 0. This cyclic sum arises because F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu, where A^\mu is the four-potential, making the exterior derivative dF = 0 identically, which in component form yields the above relation with four independent equations. The inhomogeneous equations, incorporating electric charges and currents via the four-current J^\nu, are obtained by contracting the four-gradient with the contravariant Faraday tensor: \partial_\mu F^{\mu\nu} = \mu_0 J^\nu, where [\mu_0](/page/Vacuum_permeability) is the and J^\nu = (c\rho, \mathbf{J}) in units. This single tensor equation encapsulates Ampère's law with Maxwell's correction and for electricity. These relativistic forms recover the familiar three-dimensional in the of an observer. For instance, setting \nu = 0 in the inhomogeneous equation gives \partial_i F^{i0} = \mu_0 J^0, which, with the standard components F^{i0} = \frac{E^i}{c} and J^0 = c\rho, simplifies to \nabla \cdot \mathbf{E} = \rho / \epsilon_0 after accounting for c^2 = 1/(\epsilon_0 \mu_0). Similarly, the spatial components \nu = k yield \nabla \times \mathbf{B} - \frac{1}{c^2} \frac{\partial \mathbf{E}}{\partial t} = \mu_0 \mathbf{J}. For the homogeneous set, the Bianchi identity implies \nabla \cdot \mathbf{B} = 0 and \nabla \times \mathbf{E} + \frac{\partial \mathbf{B}}{\partial t} = 0 (in appropriate units). The tensorial structure ensures Lorentz invariance, as both F^{\mu\nu} and J^\nu transform covariantly under Lorentz transformations, unifying all four equations into forms that hold identically in any inertial frame without frame-dependent adjustments. In the presence of hypothetical magnetic sources ( currents K^\nu), the homogeneous equations generalize using the Hodge tensor {}^*F^{\mu\nu} = \frac{1}{2} \epsilon^{\mu\nu\rho\sigma} F_{\rho\sigma}, yielding \partial_\mu {}^*F^{\mu\nu} = \mu_0 K^\nu; in standard without monopoles, this reduces to \partial_\mu {}^*F^{\mu\nu} = 0.

Applications in Relativistic Mechanics

Role in Hamilton–Jacobi Equation

In , the governs the motion of a through the principal function S, incorporating the four-gradient as the operator that relates the action to coordinates. The equation takes the covariant form g^{\mu\nu} \partial_\mu S \, \partial_\nu S + m^2 c^2 = 0, where g^{\mu\nu} is the , \partial_\mu denotes the components of the four-gradient, m is the particle's rest mass, and c is the . This formulation arises from the relativistic action principle and ensures invariance under Lorentz transformations, with the four-gradient enabling the expression of the equation in four-dimensional . The four-momentum p_\mu of the particle is directly given by the four-gradient of as p_\mu = \partial_\mu S. Substituting this into the yields g^{\mu\nu} p_\mu p_\nu + m^2 c^2 = 0, which is the on-shell condition for the four-momentum, equivalent to p^\mu p_\mu = -m^2 c^2 in the . In curved , solutions to this describe geodesics, the shortest paths for particles, where the contravariant four-gradient \partial^\mu S is proportional to the tangent vector along the worldline, generating the particle's . For stationary problems, such as motion in a time-independent potential, the principal function separates into temporal and spatial parts: S = -E t + S_\text{space}(\mathbf{x}), where E is the conserved . Substituting this into the reduces it to a three-dimensional spatial equation, facilitating the solution for orbital parameters like in central force problems, such as the relativistic field. In the high-energy or massless limit, the connects to the eikonal approximation in , where the mass term vanishes, yielding g^{\mu\nu} \partial_\mu \psi \, \partial_\nu \psi = 0 for the eikonal function \psi. Here, the four-gradient of \psi defines the wave , analogous to the , describing ray propagation along null geodesics in refractive media.

Source of Conservation Laws

In relativistic field theories, conservation laws arise from symmetries of the action via , which associates each transformation with a conserved current whose four-divergence vanishes on solutions to the . For a \phi transforming as \delta \phi, the Noether current is given by J^\mu = \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \delta \phi, where \mathcal{L} is the , and its follows from the invariance of the action, yielding \partial_\mu J^\mu = 0 when the equations (Euler-Lagrange equations) are satisfied, known as the on-shell condition. A key application occurs for spacetime translation invariance, which corresponds to conservation of the energy-momentum four-vector. The associated Noether current is the stress-energy tensor T^{\mu\nu}, symmetric in flat spacetime for many theories. For a real scalar field with Lagrangian \mathcal{L} = \frac{1}{2} \partial^\mu \phi \partial_\mu \phi - V(\phi), the canonical stress-energy tensor takes the form T^{\mu\nu} = \partial^\mu \phi \partial^\nu \phi - g^{\mu\nu} \mathcal{L}, where g^{\mu\nu} is the metric tensor. The four-divergence of this tensor satisfies the continuity equation \partial_\mu T^{\mu\nu} = 0, expressing local conservation of energy and momentum on-shell. This local conservation implies a form via the four-dimensional generalization of the , where the flux of T^{\mu\nu} through the of a worldtube (a region bounded by two spacelike hypersurfaces connected by timelike surfaces) vanishes for an . Specifically, for a closed worldtube V with \partial V, \int_{\partial V} T^{\mu\nu} d\Sigma_\nu = 0, ensuring that the total entering equals that leaving the region. This integral formulation underscores the role of the four-gradient in encoding relativistic laws without reference to .

Applications in Quantum Mechanics

Schrödinger Relations

In relativistic quantum mechanics, the four-gradient \partial_\mu plays a central role in the Schrödinger relations, which extend the foundational operator-wave function relations of non-relativistic quantum mechanics to a Lorentz-covariant framework. These relations define the four-momentum operator as p_\mu = -i \hbar \partial_\mu, yielding the equation p_\mu \psi = -i \hbar \partial_\mu \psi, where \psi is the scalar wave function and the index \mu runs over spacetime coordinates. This covariant form unifies the treatment of energy and momentum, generalizing the non-relativistic Schrödinger equation i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where the Hamiltonian \hat{H} governs time evolution, and the three-momentum operator \hat{\mathbf{p}} = -i \hbar \nabla acts on spatial derivatives. The time component \mu = 0 corresponds to energy-time evolution, while the spatial components \mu = 1,2,3 encode three-momentum, ensuring consistency with special relativity's spacetime symmetry. For a free massive scalar particle, the Schrödinger relations lead to a second-order equation by contracting the four-momentum with itself, invoking the relativistic mass-shell condition p_\mu p^\mu = m^2 c^2. This produces the Klein-Gordon equation in operator form: (-i \hbar \partial_\mu)(-i \hbar \partial^\mu) \psi = m^2 c^2 \psi, or equivalently, \left( \square + \frac{m^2 c^2}{\hbar^2} \right) \psi = 0, where \square = \partial_\mu \partial^\mu is the d'Alembertian operator. Independently derived by Walter Gordon in 1926, this equation emerged from applying the relativistic energy-momentum relation E^2 = p^2 c^2 + m^2 c^4 to the wave function via the four-gradient operators. Erwin Schrödinger's 1926 efforts to construct a relativistic were pivotal in this development, as he sought to reconcile de Broglie's matter waves with Einstein's relativistic formula during the formulation of wave mechanics. His initial relativistic proposal, explored in unpublished notes from late 1925 and early 1926, anticipated the Klein-Gordon form but was abandoned due to difficulties in reproducing the atom's ; instead, he published the non-relativistic version that revolutionized . later refined the equation in 1927, confirming its relativistic invariance. Despite its formal success, the Klein-Gordon equation derived from the Schrödinger relations reveals significant limitations in quantum interpretation. The conserved j^\mu = \frac{\hbar}{2 m i} (\psi^* \partial^\mu \psi - \psi \partial^\mu \psi^*) can yield negative densities for certain solutions, arising from the equation's second-order time dependence and negative-energy branches (E = -\sqrt{p^2 c^2 + m^2 c^4}). These issues, including non-positive-definite probabilities, undermined the equation's viability as a fundamental quantum description and prompted to develop a in 1928 that avoids such pathologies while incorporating .

Covariant Quantum Commutation Relations

In relativistic quantum mechanics, the canonical commutation relations are extended to a covariant form to incorporate spacetime structure, treating position and momentum as components of four-vectors. The fundamental relation is [x^\mu, p^\nu] = i \hbar \eta^{\mu\nu}, where \eta^{\mu\nu} is the Minkowski metric tensor (\eta^{\mu\nu} = \operatorname{diag}(-1, +1, +1, +1)), x^\mu = (ct, \mathbf{x}) is the four-position, and p^\mu = (E/c, \mathbf{p}) is the four-momentum operator. In the position representation, the momentum operator is defined using the four-gradient as p^\mu = -i \hbar \partial^\mu, ensuring the relations hold covariantly across spacetime. This covariant structure generalizes the non-relativistic position-momentum commutation relations [x^i, p^j] = i \hbar \delta^{ij} (for spatial indices i, j = 1, 2, 3) to the full spacetime, preserving the algebraic foundation of quantum mechanics while respecting special relativity. The spatial components retain the familiar form, but the inclusion of temporal components introduces new dynamics. Seminal formulations, such as the Stueckelberg-Horwitz-Piron (SHP) approach, impose these relations on off-shell trajectories parametrized by proper time \tau, enabling a manifestly Lorentz-invariant evolution equation i \hbar \frac{\partial \psi}{\partial \tau} = K \psi, where K = \frac{p^\mu p_\mu}{2m} + V is the invariant Hamiltonian. The time component of the commutation relation, [x^0, p^0] = i \hbar \eta^{00}, implies [t, H] = -i \hbar (with H = E the operator and appropriate unit conventions), treating time as a Hermitian operator conjugate to the . However, this raises fundamental issues: in standard , the single-particle (e.g., \sqrt{\mathbf{p}^2 c^2 + m^2 c^4}) is bounded from below, violating the requirement for a self-adjoint time operator under Pauli's theorem, which prohibits unbounded spectra for conjugate pairs in . This leads to non-unitary evolution and ill-defined time observables, complicating interpretations like energy-time uncertainty \Delta E \Delta t \geq \hbar / 2. The SHP formalism circumvents this by allowing off-shell momenta (p^\mu p_\mu \neq m^2 c^2), rendering K unbounded and permitting a well-defined time operator. To avoid operator-ordering ambiguities in defining covariant observables (e.g., in powers of x^\mu and p^\nu), Weyl quantization maps classical phase-space functions to symmetric products via the Weyl correspondence, ensuring the commutation relations are preserved without ad hoc choices. Alternatively, path-integral formulations, as in Feynman's relativistic extensions, bypass explicit time operators altogether by integrating over worldlines parametrized by \tau, directly yielding transition amplitudes consistent with the covariant algebra. The commutation tensor i \hbar \eta^{\mu\nu} transforms covariantly under Lorentz transformations, as both x^\mu and p^\nu are four-vectors, ensuring the algebra is invariant and consistent with . This covariance underpins applications like deriving Lorentz-invariant conservation laws from in the SHP framework.

Relativistic Wave Equations and Probability Currents

In , the four-gradient operator plays a central role in defining conserved probability currents for wave equations that respect Lorentz invariance, such as the Klein-Gordon and Dirac equations. These currents arise as four-vectors whose vanishes, ensuring the conservation of probability in a covariant manner. The four-gradient terms in the wave equations lead to bilinear forms that interpret the time component as a probability density and the spatial components as a probability flux, though challenges like non-positive definiteness persist in some cases. For the Klein-Gordon equation, which describes spin-0 particles, the probability four-current is given by j^\mu = \frac{i \hbar}{2m} \left[ \psi^* \partial^\mu \psi - (\partial^\mu \psi^*) \psi \right], where \psi is the complex scalar wave function, m is the particle mass, and \hbar is the . This current satisfies the \partial_\mu j^\mu = 0, which follows directly from the Klein-Gordon equation and its , ensuring that the total probability is preserved under Lorentz transformations. The current emerges from varying associated with the Klein-Gordon \mathcal{L} = \partial_\mu \psi^* \partial^\mu \psi - m^2 \psi^* \psi, corresponding to the Noether current for global U(1) phase symmetry, where the four-gradient terms generate the bilinear structure under infinitesimal variations \delta \psi = i \epsilon \psi. However, the time component j^0 of the Klein-Gordon current, proportional to i (\psi^* \partial_t \psi - \psi \partial_t \psi^*), is not positive definite, leading to interpretation issues such as negative probabilities for certain superpositions of positive-frequency solutions. This non-positive definiteness arises because the flow lines defined by \partial_\mu S (from \psi = e^{iS/\hbar}) can become space-like, allowing superluminal velocities and closed loops in the probability interpretation. Consequently, the Klein-Gordon equation is typically treated as a field theory for particles and antiparticles rather than a single-particle , with j^\mu reinterpreted as a charge current. In contrast, the Dirac equation for spin-1/2 particles yields a positive-definite probability current j^\mu = \bar{\psi} \gamma^\mu \psi, where \bar{\psi} = \psi^\dagger \gamma^0 is the Dirac adjoint, and \gamma^\mu are the gamma matrices satisfying \{ \gamma^\mu, \gamma^\nu \} = 2 \eta^{\mu\nu}. The conservation \partial_\mu j^\mu = 0 is derived by multiplying the Dirac equation (i \gamma^\mu \partial_\mu - m) \psi = 0 by \bar{\psi} from the left and its adjoint by \psi from the right, then adding to eliminate mass terms, resulting in a four-divergence form involving the four-gradient. This current provides a satisfactory single-particle probability interpretation, with j^0 = \psi^\dagger \psi > 0 normalizing to unity upon spatial integration. To address relativistic effects in probability densities, particularly for boosted frames, a four-velocity weighting is introduced, where the density \rho_0 = \sqrt{|j_\alpha j^\alpha|} is associated with the u_\alpha = j_\alpha / \rho_0, ensuring Lorentz-invariant normalization and avoiding configuration-space issues in multi-particle systems. This approach ties the probability flow to the particle's parametrization, with the four-current j_\alpha constructed from initial and final wave functions via real parts of bilinear forms.

Derivation of Quantum Equations from Special Relativity

In , the energy-momentum relation for a of rest mass m is given by E^2 = p^2 c^2 + m^2 c^4, where E is the total , \mathbf{p} is the three-momentum, and c is the . This relation, originally derived from the principles of Lorentz invariance and laws, sets the foundation for relativistic . To bridge this classical to , the de Broglie hypothesis posits that particles exhibit wave-like properties, with the p^\mu = (E/c, \mathbf{p}) related to the four-wavevector k^\mu = (\omega/c, \mathbf{k}) via p^\mu = \hbar k^\mu, where \hbar = h / 2\pi is the reduced Planck's constant and \omega = 2\pi f is the . Here, the four-wavevector arises from the four-gradient operator applied to the phase of a \psi = e^{i S / \hbar}, yielding k_\mu = \partial_\mu S / \hbar, with S the classical action. The quantization procedure promotes the classical four-momentum components to differential operators: the energy E to i \hbar \partial_t and the momentum \mathbf{p} to -i \hbar \nabla, ensuring compatibility with the non-relativistic in the low-energy limit. Substituting these into the relativistic relation and assuming a wave function \psi satisfying the correspondence principle leads to the Klein-Gordon equation. In covariant form, the four-momentum operator is p^\mu = -i \hbar \partial^\mu, where \partial^\mu is the four-gradient (\partial_t / c, \nabla). The on-shell condition p_\mu p^\mu = m^2 c^2 then yields the Klein-Gordon equation (\square + m^2 c^2 / \hbar^2) \psi = 0, with \square = \partial_\mu \partial^\mu the d'Alembertian operator. This second-order relativistic wave equation was first obtained independently by and Walter Gordon in 1926, motivated by quantizing the Compton effect within . However, the Klein-Gordon equation suffers from issues such as negative-probability densities in its single-particle interpretation, stemming from the squared structure of the operator. To address this and obtain a first-order equation, Paul Dirac proposed a linearization of the relativistic relation in 1928. He sought matrices \gamma^\mu satisfying the \{\gamma^\mu, \gamma^\nu\} = 2 \eta^{\mu\nu}, allowing the (i \hbar \gamma^\mu \partial_\mu - m c) \psi = 0 or equivalently \gamma^\mu p_\mu \psi = m c \psi, where \psi is a four-component . Squaring this yields the Klein-Gordon equation, but the linear form naturally incorporates spin-1/2 degrees of freedom and resolves the negative-energy problem through the interpretation. This derivation from via the four-gradient thus unifies relativistic invariance with quantum de Broglie relations, paving the way for .

Covariant Derivative in Relativistic Quantum Mechanics

In relativistic quantum mechanics, the four-gradient operator \partial_\mu is extended to the covariant derivative to account for local gauge symmetries associated with internal degrees of freedom, such as electric charge in quantum electrodynamics (QED) or color charge in quantum chromodynamics (QCD). This replacement ensures that the theory remains invariant under local transformations of the fields, incorporating interactions with gauge bosons in a manifestly covariant manner. The general form of the covariant derivative in non-Abelian gauge theories is D_\mu = \partial_\mu - i g A_\mu^a T^a, where g is the coupling constant, A_\mu^a are the components of the gauge field, and T^a are the generators of the gauge group in the appropriate representation. In the specific case of QED, which describes the interaction of electrons with the under the U(1) gauge group, the covariant derivative simplifies to D_\mu = \partial_\mu + i e A_\mu, where e > 0 is the magnitude and A_\mu is the . This form arises from , where the ordinary derivative in the free is replaced by the to introduce the interaction without altering the structure of the relativistic wave equation. The resulting becomes i \gamma^\mu D_\mu \psi = m \psi, where \gamma^\mu are the Dirac matrices, \psi is the spinor field, and m is the mass; this equation governs the dynamics of fermions in an electromagnetic field while preserving Lorentz invariance and gauge symmetry. A key consequence of this gauge-invariant formulation is the of currents. In the free theory, the four-divergence of the satisfies \partial^\mu j_\mu = 0. In the interacting theory, this generalizes to covariant D^\mu j_\mu = 0 (or D^\mu j_\mu^a = 0 for non-Abelian cases), ensuring that the Noether currents associated with global symmetries are protected under local gauge transformations. This property is essential for the consistency of the theory and leads to Ward-Takahashi identities that constrain scattering amplitudes. The uniqueness of this approach lies in the principle of , which systematically replaces partial derivatives with to couple fields to fields, extending the flat-space four-gradient to include dynamics in internal symmetry spaces while remaining within . Although this construction can be further generalized to curved using the full \nabla_\mu in , the here is on special relativistic contexts with non-trivial internal structures. This gauge-covariant extension parallels the role of the Faraday tensor in describing electromagnetic fields but applies to arbitrary groups.

References

  1. [1]
    [PDF] Special Relativity: Tensor Calculus and Four-Vectors
    In four dimensions and flat space, special relativity tells us that the invariant interval is. Page 4. defined as.
  2. [2]
    Electrodynamics in Relativistic Notation - Feynman Lectures - Caltech
    The principle of relativity is a fact of nature that in the notation of four-dimensional vectors the equations of the world will look simple.
  3. [3]
    [PDF] Introduction to Tensor Calculus for General Relativity - MIT
    The gradient of a scalar field f is a (0, 1) tensor with components (∂µf). 4.2 Gradient of a vector: covariant derivative. The reason that we have introduced a ...
  4. [4]
    [PDF] The Lorentz transformation - Physics Department, Oxford University
    It is more natural to take the 4-gradient as (−∂/∂ct, ∂/∂x, ∂/∂y, ∂/∂z) since then it more closely resembles the familiar 3-gradient. The 4-gradient ...
  5. [5]
    [PDF] Classical Action Principles - OU Physics & Astronomy
    The four-gradient is. ∂µ = 1 c. ∂. ∂t. , V ,. (1.40) ... ∂µ∂νφ − (∂µδxν)∂νφ. (1.77). We will take this ... How this is modified when currents are present is ...
  6. [6]
    [PDF] Minkowski space
    Mar 21, 2013 · signature. The bilinear form η has signature (−,+,+,+) or (+,−,−,−). Which signature is used is a matter of convention. Both are fairly common.Missing: sources | Show results with:sources
  7. [7]
    [PDF] Die Grundlage der allgemeinen Relativitätstheorie
    1916. ANNALEN DER PHYSIK. VIERTE FOLGE. BAND 49. 1. Die Grundlage der allgemeinen Relativitätstheorie; von A. Einstein. № 7. +. Die im nachfolgenden ...
  8. [8]
    (Special) Relativity - The CERN Accelerator School
    Why Special Relativity ? Most beams are ... We can define a four-gradient now: ~. ∇. ∂ = (. 1 ... T = m0(γ − 1) · c. 2. T = γm0 − m0. Page 103 ...
  9. [9]
    [PDF] Space and Time - UCSD Math
    It was Hermann Minkowski (Einstein's mathematics professor) who announced the new four- dimensional (spacetime) view of the world in 1908, which he deduced from ...
  10. [10]
    None
    Summary of each segment:
  11. [11]
    [PDF] Electromagnetism - DAMTP
    From the 4-derivative ∂µ = (∂/∂(ct),∇) and the 4-vector Aµ = (ϕ/c,−A), we can form the anti-symmetric tensor. Fµν = ∂µAν − ∂νAµ. This is constructed ...
  12. [12]
    [PDF] Lecture 13 Notes, Electromagnetic Theory II
    - In regular three-dimensional Galilean relativity, the dot product of two position vectors is invariant under transformations. Define the 4-vector (covariant) ...
  13. [13]
  14. [14]
    [PDF] The Special Theory of Relativity
    Nov 9, 2001 · ... four-vector, it is a contravariant four-vector! Either of these four-vectors is called the four-divergence. Let's introduce some new ...
  15. [15]
    [PDF] Key formula summary
    Nov 3, 2005 · These can be combined into the current 4-vector (or “4-current”) ... J. c. We have solved Maxwell's equations. This equation shows how.
  16. [16]
    [PDF] Lorentz Invariance and the 4-vector Dot Product
    To re-iterate, any dot product of two 4-vectors is Lorentz invariant. There are a few particularly useful invariant quantities for solving problems. The first ...Missing: uniqueness | Show results with:uniqueness
  17. [17]
    [PDF] Lorentz Transformations in Special Relativity
    (8) shows that a Lorentz transformation can be regarded as a matrix that is orthogonal with respect to the Minkowski metric gµν. A matrix Λ satisfying Eq.
  18. [18]
    [PDF] 7. Special Relativity - DAMTP
    We'll find it very useful to introduce a simple spacetime diagram to illustrate the physics of relativity. In a fixed inertial frame, S, we draw one direction ...
  19. [19]
    [PDF] Calculus in four dimensions
    In general, the exterior derivative maps k- forms to (k + 1)-forms. In four dimensions, there are 4 derivatives, the gradient, the curl, the hypercurl and the ...
  20. [20]
    1. Special Relativity and Flat Spacetime
    This is why it makes sense to think of SR as a theory of 4-dimensional spacetime, known as Minkowski space. (This is a special case of a 4-dimensional manifold, ...
  21. [21]
    [PDF] ELECTROMAGNETIC FIELD THEORY
    Sep 13, 2011 · ... transforms as ?G itself only if is time-independent, and that r ?G ... four-gradient of a four-scalar is a covariant vector, formed as ...
  22. [22]
    [PDF] The velocity and momentum four-vectors
    The two four-vectors Uµ and U′ µ are related by a Lorentz boost,. U′ µ = Λµ. νUν. (5) with the boost matrix Λµ ν given by1. Λ = γw. −γw. ~β. −γw. ~β δij + ...
  23. [23]
    The electromagnetic field tensor - Richard Fitzpatrick
    The electromagnetic field tensor ; $\displaystyle F_{i4}$ · $\displaystyle -F_{4i} = -E_i,$, (1479) ; $\displaystyle F^{i4}$ · $\displaystyle -F^{4i} = +E^i,$, ( ...Missing: 0i} = | Show results with:0i} =<|separator|>
  24. [24]
    [PDF] Notes on gauge theory (S. Naculich, July 2024) 1 Electromagnetism
    Jun 2, 2025 · The four-current. jµ = (ρ,j). (1.5) obeys. ∂µjµ = 0. (1.6). The Faraday tensor is a second-rank antisymmetric tensor defined by. Fµν ...
  25. [25]
  26. [26]
    [PDF] Lecture Notes on Electromagnetism and Gauge Invariance
    Sep 3, 2020 · 2.2 Invariance of electromagnetic fields under a gauge transformation . ... But the Faraday tensor Fµν is gauge invariant since F0. µν = A0. [µ,ν].
  27. [27]
    [PDF] Unit 7-2: Maxwell's Equations in Relativistic Form
    Feb 7, 2023 · According to special relativity, there is no absolute “rest frame.” Or alternatively, there is no way to determine a unique value of the ...Missing: four- | Show results with:four-<|control11|><|separator|>
  28. [28]
    [PDF] LECTURE NOTES 19
    constant scalar quantity. ∴ Jμ is a proper four vector, i.e. Jμ = proper current density 4-vector. Thus: J J. J J μ μ.
  29. [29]
    [PDF] Today in Physics 218: relativistic electrodynamics in tensor form
    Apr 23, 2004 · Aside: the four-gradient. The four-gradient with respect to contravariant coordinates, turns out to be a covariant four-vector operator, as.<|control11|><|separator|>
  30. [30]
    [PDF] The Classical Theory of Fields
    for their assistance in checking formulas. Moscow, September 1961. L. D. Landau, E. M. Lifshitz. Page 11. PREFACE. TO THE THIRD ENGLISH EDITION ... formulas of ...
  31. [31]
    The Hamilton-Jacobi Equation for a Relativistic Particle. - OSU Math
    The Hamilton-Jacobi Equation for a Relativistic Particle. Being defined in terms of the action integral, the dynamical phase satisfies a differential equation ...
  32. [32]
    [PDF] Quantum Field Theory I, Chapter 4
    Noether's Theorem. Every continuous global symmetry of the action leads to a conserved current and thus a conserved charge for solutions of the equations of.
  33. [33]
    [PDF] 8.323 Relativistic Quantum Field Theory I (Spring 2023), Problem ...
    Feb 15, 2023 · Consider the field theory of a complex valued scalar field ϕ(x) with action ... By Noether's theorem, the conserved current is given by. jµ = ∂L.
  34. [34]
    [PDF] Deriving the energy momentum tensor for a scalar field
    The energy momentum tensor is defined as the variation of the action with respect to the metric gµν. For inflation we are interested in the action of a ...
  35. [35]
    [PDF] Special Relativity - Caltech PMA
    2.12 Volumes, Integration, and Conservation Laws . ... Perhaps the most central of special relativistic laws is the one stating that the speed.
  36. [36]
  37. [37]
    [PDF] Quantum Mechanics of Klein-Gordon Fields I - arXiv
    Feb 19, 2006 · construct the observables of the theory,. • formulate it in terms of the position wave functions,. • identify the probability current density ...
  38. [38]
    [PDF] Copyright c 2019 by Robert G. Littlejohn
    This is one covariant version of the Klein-Gordon equation. The energy-momentum 4-vector in special relativity is. pµ = E/c p. ,. pµ = E/c. −p. ,. (19) in both ...
  39. [39]
    [PDF] arXiv:quant-ph/0109059v1 13 Sep 2001
    A general proof of the existence of a non positive-definite time-component of the current ... tories for the Klein-Gordon equation which yields time-like ...
  40. [40]
    None
    ### Summary of Dirac Probability Current from the Document
  41. [41]
    [PDF] Lagrangian Description for Particle Interpretations of Quantum ...
    are the rest density and 4-velocity of the probability flow, respectively. Note that these quantities are quite different from the particle rest density 0 σ ...
  42. [42]
    [PDF] recherches sur la théorie des quanta
    RECHERCHES SUR LA THÉORIE. DES QUANTA. Par M. Louis de BROGLIE. Annales de Physique. 10° Série Tome III Janvier-Février 1925. MASSON & C", ÉDITEURS. 120 ...
  43. [43]
    [PDF] The Quantum Theory of the Electron - UCSD Math
    Oct 26, 2013 · By P. A. M. DIRAC, St. John's College, Cambridge. (Communicated by R. H. Fowler, F.R.S.-Received January 2, 1928.) ... Quantum Theory of Electron.
  44. [44]
    [PDF] Oskar Klein (1926): Quantentheorie und fünfdimensionale ...
    Oskar Klein (1926). Quantentheorie und fünfdimensionale Relativitätstheorie. Zeitschrift für Physik 37: 895–906. Page 3. 66 ...
  45. [45]
    [PDF] The Compton effect according to Schrödinger's theory
    “Der Comptoneffekt nach der Schrödingerschen Theorie,” Zeit. f. Phys. 40 (1926), 117-133. The Compton effect according to Schrödinger's theory. By W. GORDON ...Missing: Walter | Show results with:Walter