Fact-checked by Grok 2 weeks ago

Continuous symmetry

Continuous symmetry refers to a property of mathematical objects or physical systems that remain invariant under a continuous family of transformations, such as rotations, translations, or scalings by arbitrary real parameters, rather than discrete steps. In mathematics, these symmetries are precisely captured by Lie groups, which are smooth manifolds equipped with group operations that model infinitesimal changes and their compositions, providing a framework for analyzing geometric and algebraic structures like rotations in Euclidean space or Lorentz transformations in special relativity. In physics, continuous symmetries play a foundational role through , which establishes a one-to-one correspondence between such symmetries of the laws of nature and conserved quantities; for instance, time-translation invariance implies , while spatial-translation invariance yields conservation. This connection, first proven by in 1918, underpins much of modern , from to , where symmetries dictate the form of Lagrangians and Hamiltonians. Examples include the of isotropic media, leading to conservation, and the gauge symmetries in , which enforce . The study of continuous symmetries extends beyond invariance to include symmetry breaking, where spontaneous or explicit mechanisms disrupt perfect symmetry, resulting in phenomena like the in or phase transitions in condensed matter. Lie algebras, derived from the tangent spaces of Lie groups, offer tools to classify these symmetries via infinitesimal generators, enabling computations in diverse fields from to .

Definitions and Concepts

Definition

In and physics, refers to a property of a that remains under certain transformations, meaning the system's essential features are unchanged after applying the transformation. Continuous specifically arises when these transformations are parameterized by continuous variables, such as real numbers, allowing for an infinite variety of intermediate states between any two transformations. For instance, rotations around an exemplify this, where the angle of rotation can take any value in the real numbers \mathbb{R}, smoothly varying the without discrete jumps. A classic example is the of a , which remains indistinguishable under by any \theta \in \mathbb{R} about its , illustrating how continuous parameterization enables invariance across a of transformations rather than isolated steps. This contrasts with symmetries but highlights the seamless nature of continuous ones, often modeled using Lie groups for their smooth structure. Continuous symmetries are fundamentally tied to the concept of infinitesimal transformations, which are limiting cases of small changes in the transformation parameter, providing the building blocks for the entire group of symmetries. These infinitesimal generators capture the local behavior and facilitate the analysis of how finite transformations emerge from successive small variations. Mathematically, a G acts continuously on a X if the action map G \times X \to X, defined by (g, x) \mapsto g \cdot x, is a , ensuring the transformations vary differentiably across the group.

Discrete vs. Continuous Symmetries

symmetries refer to transformations that form finite or countable groups, such as reflections or rotations by discrete angles in the \mathbb{Z}_n. These groups consist of a limited number of distinct elements, where the symmetry operations cannot be parameterized continuously but instead occur in isolated steps. In contrast, continuous symmetries involve transformations that form groups, allowing for smooth, parameterized variations, such as arbitrary rotations or translations. A fundamental structural difference lies in their : discrete symmetries yield finite-dimensional representations and rely on categorical invariants like character tables for analysis, whereas continuous symmetries permit infinite-dimensional representations and necessitate differential methods, such as , to study generators. This continuity enables the decomposition of transformations into arbitrarily small changes, facilitating tools like the from the to the group. An illustrative example in physics is , a discrete symmetry that inverts spatial coordinates (\mathbf{x} \to -\mathbf{x}), forming a like \mathbb{Z}_2, compared to time , a continuous symmetry that shifts time by any real amount (t \to t + \epsilon), parameterized by a continuous . While both preserve the form of physical laws in systems, the discrete nature of parity leads to binary outcomes like even or odd states, without implying ongoing . The consequences of this distinction are profound in physics: continuous symmetries generally lead to conserved quantities through , associating each independent symmetry parameter with a , whereas discrete symmetries do not typically produce such continuous conservation laws. For instance, time translation yields , but parity invariance does not enforce a similar ongoing quantity. This disparity underscores why continuous symmetries underpin much of classical and , enabling predictions of long-term dynamical behavior.

Mathematical Foundations

Lie Groups and Lie Algebras

A is defined as a group that is also a manifold, such that the group operations of multiplication and inversion are maps compatible with the manifold structure. This compatibility ensures that the group structure respects the differential geometry of the manifold, allowing for the study of continuous transformations through . Lie groups provide the foundational for modeling continuous symmetries, as their elements represent , parameterized families of transformations. The associated with a captures the structure of the group and is identified with the at the . This encodes the generators of the group, which are the first-order approximations to group elements near the identity, obtained via curves passing through it. The structure arises from the Lie bracket operation on left-invariant vector fields, which measures the non-commutativity of these transformations. The Lie bracket [X, Y] for two vector fields X and Y on the manifold is defined by [X, Y] f = X(Y f) - Y(X f) for any smooth function f, making the space of left-invariant vector fields into a Lie algebra. A prominent example is the special orthogonal group SO(3), which consists of all 3×3 orthogonal matrices with determinant 1 and represents the continuous symmetries of rotations in three-dimensional Euclidean space. Its Lie algebra \mathfrak{so}(3) is the three-dimensional vector space of skew-symmetric 3×3 matrices, with the Lie bracket corresponding to the cross product; in physics, these basis elements generate the angular momentum operators L_x, L_y, L_z, satisfying [L_i, L_j] = i \hbar \epsilon_{ijk} L_k. Continuous symmetries in physical or geometric contexts correspond to actions of groups on the underlying spaces, where the group elements induce diffeomorphisms that preserve the system's structure. The associated then describes the local, infinitesimal behavior of these symmetry transformations through its generators.

One-Parameter Subgroups

In the theory of groups, a one-parameter of a G is defined as a group homomorphism \phi: \mathbb{R} \to G, satisfying \phi(s + t) = \phi(s) \phi(t) for all s, t \in \mathbb{R}, with \phi(0) = e the and the map being differentiable. This structure captures continuous families of transformations parameterized by a , forming a one-dimensional isomorphic to (\mathbb{R}, +). One-parameter subgroups are intimately connected to the \mathfrak{g} of G through the \exp: \mathfrak{g} \to G. Specifically, for each X \in \mathfrak{g}, the curve \phi(t) = \exp(tX) defines a one-parameter subgroup, where the associates to each X a of group elements. This correspondence is bijective: every one-parameter subgroup arises uniquely from some X \in \mathfrak{g} via this exponentiation, with the at the identity \frac{d}{dt} \phi(t) \big|_{t=0} = X. In matrix Lie groups, this manifests explicitly as \phi(t) = e^{tX}, where the matrix exponential provides the concrete realization. The evolution of a one-parameter is governed by the flow equation \frac{d}{dt} \phi(t) = X(\phi(t)), where X denotes the left-invariant on G generated by the element X \in \mathfrak{g}. This describes how the infinitesimal action at the propagates along the curve, ensuring that \phi(t) remains a . Solving this yields the starting at the , uniquely determining the for each X. In the context of continuous symmetries, one-parameter subgroups illustrate how infinitesimal symmetries—represented by elements of the as vector fields on the manifold—generate finite, global transformations through integration of their flows. This mechanism bridges local actions to full group elements, enabling the realization of continuous symmetry operations as parameterized paths in G. A fundamental result states that in a compact connected , every element lies in some one-parameter subgroup. This underscores the generative role of these subgroups, as the is surjective in this setting, ensuring the entire group is covered by flows from the identity.

Physical Applications

Noether's Theorem

Noether's theorem establishes a profound connection between continuous symmetries of physical systems and conservation laws, originating from Emmy Noether's 1918 paper "Invariante Variationsprobleme," which addressed questions raised by and regarding the and momentum in the context of general relativity's . Noether, working in , developed the theorem amid discussions on the invariance properties of variational principles in Einstein's theory, proving that symmetries imply conserved quantities even when the underlying is curved. The states that if the action S = \int L \, dt of a system, derived from a L(q, \dot{q}, t), is under a continuous symmetry generated by an variation \delta q, then there exists a conserved quantity associated with that symmetry. In field theory formulations, for a Lagrangian density under a symmetry, a conserved Noether current j^\mu arises such that \partial_\mu j^\mu = 0, where the index \mu runs over spacetime coordinates. This applies to both spacetime symmetries (like translations or rotations) and internal symmetries (like phase transformations), provided the symmetry leaves the action unchanged up to a total divergence. To derive the theorem, consider the variation of the under the transformation. The total variation \delta L decomposes into an explicit change \frac{\partial L}{\partial q} \delta q + \frac{\partial L}{\partial \dot{q}} \delta \dot{q} plus a term from the coordinate transformation itself. For an on-shell variation (satisfying the ), the Euler-Lagrange equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) - \frac{\partial L}{\partial q} = 0 imply that the symmetry condition \delta L = \frac{d}{dt} F (for some F, often zero for strict invariance) leads to \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \delta q - F \right) = 0. Thus, for time-independent symmetries, the quantity \frac{\partial L}{\partial \dot{q}} \delta q is conserved along the system's trajectory. In field-theoretic settings, this generalizes to the j^\mu = \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \delta \phi - K^\mu, where K^\mu accounts for the term, satisfying the . The conserved charge is given by Q = \int j^0 \, d^3x, which remains constant under along the symmetry flow, reflecting the invariance. This charge generates the symmetry transformations via Poisson brackets in . The theorem assumes the symmetry is a quasi-symmetry of , meaning the transforms as \delta L = \partial_\mu K^\mu (a total ), ensuring the action's variation vanishes upon . It accommodates both finite-dimensional groups (yielding finitely many conserved currents) and infinite-dimensional groups (yielding differential identities), handling and internal symmetries without requiring the to be explicitly time-independent.

Conservation Laws from Symmetries

In classical mechanics, Noether's theorem establishes a direct correspondence between continuous symmetries of the Lagrangian and conserved quantities, providing a systematic derivation of fundamental conservation laws. One of the most prominent examples is time-translation symmetry, which arises when the laws of physics are invariant under shifts in time, implying that the Lagrangian L has no explicit time dependence. In this case, the theorem yields the conservation of energy, where the Hamiltonian H = \sum_i p_i \dot{q}_i - L remains constant along the system's dynamical trajectory. This conserved quantity represents the total energy of the system and holds for any time-independent potential, underscoring the uniformity of temporal evolution in isolated systems. Spatial translation symmetry, reflecting the homogeneity of space, similarly leads to momentum conservation via . When the is independent of the coordinates \mathbf{q}_i, the total linear \mathbf{p} = \sum_i \frac{\partial L}{\partial \dot{\mathbf{q}}_i} is conserved. This result applies to systems in uniform gravitational or electromagnetic fields where forces derive from potentials that do not explicitly vary with position, ensuring that the center-of-mass motion proceeds at constant velocity. Rotational symmetry, characteristic of isotropic environments, further implies the conservation of . For a invariant under rotations, the total \mathbf{L} = \sum_i \mathbf{r}_i \times \mathbf{p}_i is preserved, as demonstrated by the theorem's application to angular transformations. These laws collectively govern the dynamics of particle systems, from planetary orbits to molecular vibrations, without relying on ad hoc assumptions. The framework extends naturally to relativistic field theories, where continuous symmetries under the —encompassing translations and Lorentz transformations—generate conservation laws through the stress-energy tensor T^{\mu\nu}. dictates that this tensor is conserved, satisfying \partial_\mu T^{\mu\nu} = 0 on-shell, which encodes the local (\nu=0) and momentum (\nu=j) in . The of T^{\mu\nu} arises from the field's variation under infinitesimal coordinate shifts, with improvements like the Belinfante symmetrization ensuring gauge invariance in theories such as or scalar fields. This tensor plays a central role in , sourcing the via Einstein's equations. In , the implications of these symmetries manifest through unitary representations on the , where continuous transformations are implemented by operators U satisfying U^\dagger U = I. The infinitesimal generators of these symmetries are Hermitian observables that commute with the , [G, H] = 0, ensuring under via the i\hbar \frac{d}{dt} |\psi\rangle = H |\psi\rangle. For , the itself serves as the generator, dictating the phase evolution e^{-iHt/\hbar} of energy eigenstates. Analogously, the generates spatial translations and generates rotations, with these observables' expectation values remaining constant, thereby bridging classical laws to quantum probabilities.

Examples and Illustrations

Rotational and Translational Symmetries

refers to the invariance of a under arbitrary displacements in , mathematically expressed as the \mathbf{x} \to \mathbf{x} + \mathbf{a} for any constant \mathbf{a}. In free space, this is continuous, implying that the laws of physics are homogeneous and independent of position, which underpins the uniformity of physical properties across . For example, the motion of a follows the same equations regardless of its starting location, reflecting this spatial homogeneity. In contrast, crystalline solids exhibit only discrete translational , where invariance holds for specific translations rather than arbitrary ones, representing a spontaneous breaking of the continuous present in the underlying atomic interactions. Rotational symmetry involves invariance under arbitrary rotations in , governed by the special orthogonal group SO(3), which parameterizes all proper rotations. A prominent example is spherical symmetry in central force problems, such as gravitational attraction, where the force depends solely on the distance between bodies and points radially, leaving the system's dynamics unchanged under any rotation about the center. In the , describing the motion of planets orbiting under inverse-square , this rotational invariance leads to the conservation of orbital , ensuring that the plane of the orbit remains fixed while the area swept by the radius vector is constant over time. These conserved quantities arise from the underlying symmetries, as formalized in conservation laws. To visualize these symmetries, consider the infinitesimal generators represented as vector fields: for translations, the field is constant and uniform, resulting in parallel flow lines with zero curl, indicating no local rotation. In contrast, infinitesimal rotations produce a vector field proportional to \boldsymbol{\omega} \times \mathbf{r}, where \boldsymbol{\omega} is the rotation axis vector and \mathbf{r} the position; the flow lines form closed loops around the axis, with a nonzero curl that quantifies the local rotational tendency. In real materials, these continuous symmetries are often broken. Translational symmetry becomes discrete in crystals due to the periodic lattice structure, limiting invariance to specific shifts matching the unit cell dimensions. Similarly, is broken in anisotropic materials, such as certain or composites, where properties like electrical or elasticity vary with direction, lacking full SO(3) invariance—for instance, in iron-based superconductors where orbital emerges above the transition temperature.

Gauge Symmetries

Gauge symmetries represent a of continuous internal symmetries that are local, meaning the transformation parameters can vary independently at different points in . In and , a prototypical example is the U(1) gauge symmetry under which a charged field \psi transforms as \psi \to e^{i\alpha(x)} \psi, where \alpha(x) is a spacetime-dependent . This locality distinguishes gauge symmetries from global symmetries, as the transformation must leave the action invariant only when accompanied by appropriate adjustments to other fields. A concrete illustration arises in , where the gauge symmetry acts on the A_\mu via the A_\mu \to A_\mu + \partial_\mu \Lambda, with \Lambda(x) an arbitrary scalar function. To maintain invariance under this local U(1) transformation in the presence of fields, the \partial_\mu is replaced by the D_\mu = \partial_\mu - i e A_\mu, where e is the . This substitution ensures that the kinetic term for the , \bar{\psi} i \gamma^\mu D_\mu \psi, remains unchanged under the combined field transformations. The requirement of gauge invariance necessitates the introduction of gauge fields, such as the in , which mediate interactions and propagate the associated with the . In non-Abelian gauge theories, this extends to more complex structures; for instance, (QCD) is based on the SU(3) , where gluons serve as the gauge bosons that bind quarks through the strong force, enabling the unification of fundamental interactions within the . These gauge fields acquire self-interactions in the non-Abelian case, leading to phenomena like in QCD. Applying to local gauge symmetries yields identities known as Ward identities, rather than conserved currents in the usual sense, due to the spacetime dependence of the transformations. These identities constrain scattering amplitudes and correlation functions, ensuring consistency with the underlying gauge invariance, as derived from for systems with redundant . The concept of gauge symmetry originated with Hermann Weyl's 1918 attempt to unify and through a local scaling invariance in a generalized , though this initial formulation faced challenges with observational predictions. It was later refined in the 1954 work of Ning Yang and Robert Mills, who developed the general framework of non-Abelian theories invariant under local isotopic spin rotations, laying the groundwork for modern .

References

  1. [1]
    DOE Explains...Symmetry in Physics - Department of Energy
    In physics, symmetry refers to how particles behave when space, time, or quantum numbers are reversed. We're used to seeing simple types of symmetry in ...
  2. [2]
    [PDF] Lie groups and Lie algebras (Winter 2024)
    Lie groups, by contrast, are about so-called continuous symmetries (more aptly called smooth sym- metries), such as translational or rotational symmetries of a ...
  3. [3]
    [PDF] Emmy Noether and Symmetry
    Noether's (First) Theorem states that to each continuous symmetry group of the action functional there is a corresponding conservation law of the physical ...
  4. [4]
    [PDF] 3 Classical Symmetries and Conservation Laws
    Noether's theorem: For every continuous global symmetry there exists a global conservation law. Before we prove this statement, let us discuss the connection ...
  5. [5]
    Symmetry and Symmetry Breaking
    Jul 24, 2003 · The definition of symmetry as “invariance under a specified group of transformations” allowed the concept to be applied much more widely, not ...
  6. [6]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    classes of Lie groups (namely compact Lie groups and semisimple Lie groups, to be defined later) ... Bump, Lie Groups, Graduate Texts in Mathematics, vol. 225, ...
  7. [7]
  8. [8]
    [PDF] Lie Group Actions
    A manifold M endowed with a continuous G-action is called a (left or right). G-space. • If M is a smooth manifold and the action is smooth, M is called a smooth.
  9. [9]
    [PDF] Lecture Notes on Group Theory in Physics (A Work in Progress)
    ... Lie groups . . . . . . . . . . . . . . . . . . . . . . . 29. 1.5.3 ... continuous symmetries. Group theory – big subject! Our concern here lies in its ...
  10. [10]
    [PDF] and its applications in physics - Group Theory
    Infinite dimensional symmetries. String theory. 85. 2015-01-24 iii ... So far we have considered discrete symmetry groups with at most countable number of.
  11. [11]
    Lie Group -- from Wolfram MathWorld
    A Lie group is a smooth manifold obeying the group properties and that satisfies the additional condition that the group operations are differentiable.
  12. [12]
    [PDF] lie groups and lie algebras womp 2007
    Definition. A Lie group is a group G that is also a smooth manifold, such that the multipli- cation G × G → G, (g ...
  13. [13]
    What is a Lie group?
    A Lie group is a group of symmetries where the symmetries are continuous. A circle has a continuous group of symmetries.
  14. [14]
    [PDF] About Lie Groups
    Oct 6, 2005 · The tangent space TeG at the identity is a real vector space. Using the three classes of maps inherent in the Lie group structure, we can equip ...
  15. [15]
    [PDF] Differential Geometry and Lie Groups A Computational Perspective
    This book is written for a wide audience ranging from upper undergraduate to advanced graduate students in mathematics, physics, and more broadly ...
  16. [16]
    Introduction to Smooth Manifolds, Second Edition
    This book is an introductory graduate-level textbook on the theory of smooth manifolds. Its goal is to familiarize students with the tools they will needMissing: definition | Show results with:definition
  17. [17]
    [PDF] Theory of Angular Momentum and Spin
    Show that in this case holds ρ0(AB) = ρ0(B)ρ0(A). Generators. One can assume then that R(~ϑ) should also form a Lie group, in fact, one isomorphic to SO(3).
  18. [18]
    [PDF] Chapter 7 Continuous Groups, Lie Groups, and Lie Algebras
    This Lie group is called the general linear group in two dimensions and is denoted by GL(2,R), where the 'R' signifies that the entries are real; the ...
  19. [19]
    [PDF] The SO(3) and SE(3) Lie Algebras of Rigid Body Rotations ... - arXiv
    This document discusses how to modify these methods so they can be applied to non-Euclidean state vectors, such as those containing rotations and full motions ...
  20. [20]
    [PDF] One parameter subgroups
    In a compact connected Lie group G, every element lies on some. 1-p subgroup. This is not true in a non-compact G, i.e. there are elements in G which do not ...
  21. [21]
    [PDF] Fall, 2022 Lecture III The Exponential Map, Local Lie Groups, and ...
    Sep 26, 2022 · Definition 1.4. We define the exponential map, expG : g → G by sending. A ∈ g to γA(1) where γA is the one-parameter subgroup whose tangent.
  22. [22]
    [PDF] Colloquium: A Century of Noether's Theorem - arXiv
    Jul 11, 2019 · In the summer of 1918, Emmy Noether published the theorem that now bears her name, establishing a profound two-way connection.
  23. [23]
    Emmy Noether's Wonderful Theorem (rev ed.). - AIP Publishing
    Dec 1, 2018 · In 1918, the mathematician Emmy Noether published two wonderful theorems that had a tremendous impact in physics, mathematics, and beyond.<|control11|><|separator|>
  24. [24]
    [PDF] Noether's theorem - Physics Department, Oxford University
    Feb 27, 2025 · In order to understand the Euler-Lagrange equations correctly, it is important to note what is held constant in each partial derivative, and ...
  25. [25]
    [physics/0503066] Invariant Variation Problems - arXiv
    Mar 8, 2005 · Authors:Emmy Noether, M. A. Tavel. View a PDF of the paper titled Invariant Variation Problems, by Emmy Noether and M. A. Tavel. View PDF.
  26. [26]
    [gr-qc/0608096] Noether's theorem, the stress-energy tensor ... - arXiv
    Noether's theorem is reviewed with a particular focus on an intermediate step between global and local gauge and coordinate transformations, namely linear ...Missing: field | Show results with:field
  27. [27]
    [PDF] Symmetries and conservation laws in quantum me- chanics
    In quantum mechanics, a conserved quantity's probability is time-independent, and a symmetry is a unitary transformation that maps states to equivalent states.
  28. [28]
    [PDF] Transformations and Symmetries - Rutgers Physics
    Finite Transformations. • As noted above, a finite continuous transformation can be built up out of an infinite number of infinitesimal transformations. ˆ. Ux ...
  29. [29]
    [PDF] Time Crystals - Stony Brook University
    A standard example of symmetry breaking is the existence of crystals: the symmetry breaking of the continuous spatial translation symmetry to the discrete ...
  30. [30]
    [PDF] Units, limits, and symmetries When solving physics problems it's ...
    ... central forces (like gravity) have spherical symmetry because the force depends only on the distance from the origin. In this case, spherical symmetry means ...<|separator|>
  31. [31]
    [PDF] Central Forces - LIGO-Labcit Home
    Thus rotational symmetry about the z-axis implies the conservation of the z-component of the angular momentum. Full rotational symmetry, about any axis, implies ...
  32. [32]
    Kepler Orbits - Galileo and Einstein
    The conservation of angular momentum comes from the spherical symmetry of the system: the attraction depends only on distance, not angle. In quantum mechanics, ...
  33. [33]
    [PDF] so(4) Symmetry of the Kepler Problem - Columbia Math Department
    Then there exists a corresponding conserved quantity. In this problem, we saw conservation of angular momentum, and conservation of energy is also present ...
  34. [34]
    [PDF] Curl, Divergence and Laplacian - Purdue Math
    The curl operator encodes information about infinitesimal rotations of a vector field. Think of a small plastic ball inside the stream of a river. If the ball ...
  35. [35]
    UM Ma215 Examples: 16.5 Curl
    Curl of a Vector Field​​ Intuitively, the curl measures the infinitesimal rotation around a point. This is difficult to visualize in three dimensions, but we ...
  36. [36]
    Symmetry-breaking orbital anisotropy observed for detwinned Ba ...
    Apr 11, 2011 · Nematicity, defined as broken rotational symmetry, has recently been observed in competing phases proximate to the superconducting phase in ...
  37. [37]
    [PDF] Gauge Theory - DAMTP
    ... gauge symmetry which underlies the Maxwell equations, the Standard Model and, in the guise of diffeomorphism invariance, general relativity. Gauge symmetry ...
  38. [38]
    [PDF] GAUGE THEORIES - UT Physics
    also called a gauge symmetry — the field transformations at. different points x have independent parameters. For example, a local ...
  39. [39]
    Conservation of Isotopic Spin and Isotopic Gauge Invariance
    The possibility is explored of having invariance under local isotopic spin rotations. This leads to formulating a principle of isotopic gauge invariance.
  40. [40]
    [PDF] P528 Notes #3: Non-Abelian Gauge Theories
    Jan 31, 2017 · From this point of view, it is completely natural to try to construct field theories with a gauge invariance under non-Abelian transformation ...
  41. [41]
    Noether's Second Theorem and Ward Identities for Gauge Symmetries
    Oct 23, 2015 · We reintroduce Noether's second theorem and discuss how to work with the physical remnant of gauge symmetry in gauge fixed systems.
  42. [42]
    Noether's second theorem and Ward identities for gauge symmetries
    Feb 4, 2016 · We reintroduce Noether's second theorem and discuss how to work with the physical remnant of gauge symmetry in gauge fixed systems.
  43. [43]
    Gauge theory: Historical origins and some modern developments
    Jan 1, 2000 · In this article the authors review the early history of gauge theory, from Einstein's theory of gravitation to the appearance of non-Abelian gauge theories in ...