Fact-checked by Grok 2 weeks ago

Noether's second theorem

Noether's second theorem, formulated by in her 1918 paper "Invariante Variationsprobleme," is a cornerstone of the and that establishes a profound link between symmetries of action functionals and the structure of the resulting . Specifically, it states that if a variational problem is invariant under an infinite-dimensional of transformations—depending on arbitrary functions rather than finite parameters—then the Euler-Lagrange equations derived from are not independent but satisfy a set of differential identities among themselves, rather than producing conserved quantities as in Noether's first theorem. This theorem applies to systems where the number of equations equals the number of dependent variables, leading to under-determined systems whose solutions are constrained by these identities on the solution manifold (often called "on-shell"). The theorem emerged in the context of early 20th-century efforts to understand , where had noted apparent paradoxes in due to coordinate invariance; Noether's result resolved this by showing that such "trivial" conservation laws arise from infinite-dimensional , yielding identities like the Bianchi identities in instead of integrable conserved currents. Mathematically, for a Lagrangian density L invariant under transformations generated by a symmetry characteristic Q = D^* F, where D is a linear and F is an arbitrary smooth function, the theorem implies an identity of the form F \cdot D E(L) = D^*(F) \cdot E(L) + \operatorname{Div} P[F, E(L)], where E(L) denotes the Euler-Lagrange expressions; this holds identically when the equations E(L) = 0 are satisfied, confirming the dependencies. In modern physics, Noether's second theorem underpins the analysis of gauge theories, such as electromagnetism and Yang-Mills theories, where local gauge symmetries (infinite-dimensional) lead to Ward-Takahashi identities that ensure consistency of quantum field theories and constrain scattering amplitudes without implying new conservation laws. It also extends to broader contexts, including anomalous diffusion models and plasma physics algorithms, where it reveals hidden structural relations in under-determined systems. Unlike the first theorem, which connects finite-dimensional symmetries to Noether currents and charges, the second theorem highlights how infinite symmetries enforce relational constraints, making it indispensable for understanding the foundational symmetries of fundamental interactions.

Historical Context

Emmy Noether's Contributions

Amalie Emmy Noether was born on March 23, 1882, in , , into a Jewish family of mathematicians; her father, , was a prominent algebraist at the University of Erlangen. She began her studies at the University of Erlangen in 1900, auditing classes despite formal restrictions on women, and earned her PhD in 1907 with a dissertation on algebraic invariants under Paul Gordan. From 1907 to 1915, Noether lectured unpaid at Erlangen, often substituting for her father. In 1915, she moved to the at the invitation of and to study advanced topics, though she had no formal position there until her in 1919. In 1915, Noether returned to Göttingen at the invitation of and to assist with their research on differential invariants and the . Due to gender restrictions, from 1916 she lectured under Hilbert's name without pay. Under Hilbert and Klein's influence, she developed innovative approaches to ideal theory and ring structures, emphasizing axiomatic methods that revolutionized modern algebra, including her ideal theorem on chain conditions. This period solidified her reputation as a leader in abstract algebraic structures, influencing fields from to non-commutative geometry. During , amid wartime disruptions in , Noether collaborated closely with Hilbert on , where symmetry and invariance principles were central to foundational challenges. Motivated by these discussions, particularly Hilbert's efforts to establish in curved , Noether published her seminal 1918 paper "Invariante Variationsprobleme" in the Nachrichten der Königlichen Gesellschaft der Wissenschaften zu ; due to restrictions on women presenting, the paper was read by Klein on July 26, 1918. In this work, she introduced both of her invariance theorems, with the first linking continuous symmetries to laws as a precursor, while rigorously addressing gaps in Hilbert's informal assertions about energy theorems in by clarifying conditions for variational symmetries. Her analysis provided the mathematical framework to resolve ambiguities in applying principles to generally covariant theories, bridging with physical invariance.

Development in the 1910s

In the early , physicists and mathematicians increasingly turned to variational principles to formulate field theories, building on the principle of least action that had long unified and . By this period, the least action principle was routinely applied to electromagnetic fields and charged particles, as demonstrated in Joseph Larmor's derivation of field equations from a combined for matter and , which highlighted symmetries in the action . This approach gained renewed prominence in the context of emerging relativistic theories, where variational methods provided a framework for deriving while preserving invariance under transformations, setting the stage for more complex gravitational theories. David Hilbert's investigations from 1915 to 1917 exemplified these efforts, as he sought to develop a unified theory of gravitation and electromagnetism using variational calculus based on a general action principle. In his November 1915 communications to the Göttingen Academy, Hilbert derived field equations from a variational principle incorporating the metric tensor and electromagnetic potentials, aiming to establish a rigorous conservation law for energy-momentum. However, his attempt to prove global energy conservation faltered due to the general covariance of the theory, which rendered energy expressions dependent on the choice of coordinates and precluded a unique, coordinate-independent conserved quantity. This failure underscored the challenges of applying classical conservation ideas to curved spacetime, prompting deeper scrutiny of symmetries in variational problems. These issues fueled intense debates on and invariance in among , Hilbert, and during 1915–1918. Einstein, having finalized the field equations in November 1915, emphasized as a core requirement for the theory's , arguing it ensured physical laws' independence from coordinate choices. Hilbert and Klein, however, questioned the physical implications of full , with Klein critiquing it in letters to Einstein as potentially undermining meaningful laws by allowing arbitrary diffeomorphisms that alter definitions. Their exchanges, particularly Klein's 1917–1918 with Hilbert, highlighted tensions between geometric invariance and empirical , influencing the of symmetries in relativistic theories. Klein's earlier (1872), which classified geometries by their invariance groups, provided a foundational influence here, framing as a geometry invariant under the in special cases and broader transformation groups in , thereby bridging with physical symmetries. Amid these discussions, Klein and Hilbert invited to in 1915 to assist with applications to , recognizing her expertise from as vital for resolving open questions in variational symmetries. Noether arrived that year and participated actively, contributing to the local seminar on . This collaboration extended into the 1916–1918 correspondence between Klein and Hilbert, where they grappled with in covariant theories and acknowledged Noether's insights on differential invariants, which helped clarify the role of infinite-dimensional symmetry groups in field equations. Her algebraic background in invariants, honed through studies under Paul Gordan, enabled these contributions, though formal recognition at the university was delayed until 1919.

Background Concepts

Variational Principles

Variational principles form the cornerstone of by providing a to derive through the optimization of a functional known as the action. In , the action S is defined as the time integral of the L, typically expressed as S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt, where q represents and \dot{q} their time derivatives. Hamilton's principle states that the physical path of a system makes the action stationary, meaning its first variation vanishes: \delta S = 0. This condition leads to the Euler-Lagrange equations, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) - \frac{\partial L}{\partial q} = 0, which govern the dynamics of the system. The development of these ideas traces back to the 18th and 19th centuries, with Leonhard Euler formalizing the in the 1740s, introducing the formulation in his 1788 work Mécanique Analytique, and establishing the principle of least action in 1834. In field theories, the action extends to an integral over , S = \int \mathcal{L}(x, \phi, \partial \phi) \, d^4 x, where \mathcal{L} is the density depending on field variables \phi(x) and their derivatives. Applying the yields the field Euler-Lagrange equations, \frac{\partial \mathcal{L}}{\partial \phi} - \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) = 0, which describe the evolution of fields such as electromagnetic potentials or scalar fields. These variational methods unify diverse areas of physics, including through particle trajectories, via derived from the electromagnetic action, and through the Einstein-Hilbert action that encodes .

Symmetries in Lagrangian Mechanics

In , a symmetry is defined as a transformation of the q(t) and possibly time t, denoted by infinitesimal variations \delta q(t) and \delta t, such that the L(q, \dot{q}, t) remains up to a total time : \delta L = \frac{dF}{dt} for some function F(q, t). This condition ensures that the equations of motion, derived from the principle of least action, are unchanged under the transformation. Continuous symmetries arise from Lie groups, which are smooth groups of transformations parameterized by a continuous \epsilon, where the group operation combines parameters additively. For small \epsilon, the transformation is approximated by its infinitesimal generator, a that dictates the direction of the change, such as \delta q = \epsilon X(q) where X is the generator. These symmetries contrast with discrete symmetries, which involve finite, non-parameterized operations like reflections or inversions. Symmetries are further classified as global or local. Global symmetries, also known as rigid transformations, feature parameters that are constant across spacetime, applying uniformly to the entire system. Local symmetries, or flexible transformations, allow parameters to vary with position and time, such as \epsilon(x, t), often leading to gauge structures in more advanced theories. Noether's first theorem provides a brief recap: for every continuous global symmetry of the Lagrangian, there exists a conserved quantity, such as linear momentum arising from invariance under spatial translations. In general, such symmetries reduce the effective degrees of freedom in the system by imposing constraints and systematically explain the origin of conservation laws in physical theories.

Core Mathematical Formulation

First Variation Formula

The first variation of the action plays a central role in the calculus of variations for field theories, providing the mathematical foundation for deriving equations of motion and analyzing symmetries. In Lagrangian field theory, the action is defined as S[\phi] = \int \mathcal{L}(\phi, \partial_\mu \phi) \, d^4 x, where \phi represents the fields, \mathcal{L} is the Lagrangian density, and the integral is over a spacetime region with Minkowski metric (or more generally, a curved metric with \sqrt{-g}). Under an infinitesimal transformation, the variation \delta S of the action is computed to identify conditions for extrema, such as \delta S = 0 for solutions satisfying the equations of motion. For general infinitesimal transformations, including both and coordinate changes, the total variation of a \phi is given by \delta \phi = \phi'(x') - \phi(x) = \epsilon X[\phi] + O(\epsilon^2), where \phi'(x') is the transformed at the transformed point x' = x + \epsilon \delta x, \epsilon is an parameter, and X[\phi] is the generator of the acting on \phi. This total variation accounts for both the explicit change in the value and the implicit shift due to coordinate , decomposed as \delta \phi = \delta_0 \phi + \delta x^\mu \partial_\mu \phi, with \delta_0 \phi being the intrinsic variation at fixed coordinates. The corresponding variation of the action takes the form \delta S = \int \left[ \delta \mathcal{L} + \partial_\mu (\mathcal{L} \delta x^\mu) \right] d^4 x, where \delta \mathcal{L} arises from the changes in \phi and \partial_\mu \phi, while the divergence term originates from the Jacobian of the coordinate and the change in the integration measure. To connect this to the equations of motion, integration by parts is applied to \delta S, assuming the variations have compact support so that boundary terms vanish. This yields \delta S = \int \left[ E(\phi) \delta \phi + \partial_\mu K^\mu \right] d^4 x, where E(\phi) = \frac{\partial \mathcal{L}}{\partial \phi} - \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) is the Euler-Lagrange operator, representing the field equations, and K^\mu collects the remaining terms forming a total divergence (often referred to as a current in symmetry contexts). For extrema of the action, \delta S = 0 implies E(\phi) = 0 off-shell only if \delta \phi is arbitrary; otherwise, the on-shell condition E(\phi) = 0 ensures the variation vanishes for admissible \delta \phi. The assumption of smooth transformations with compact support ensures the integral is well-defined and boundary contributions are absent, applicable to fields in a finite spacetime region. In the context of symmetries, this framework identifies cases where \delta S = 0 holds on-shell for specific transformation-generated \delta \phi.

Variational Symmetries

In the context of Noether's second theorem, a variational symmetry is defined as a transformation of the fields that leaves the density invariant up to a total derivative, specifically satisfying the condition \delta \mathcal{L} = \partial_\mu F^\mu for some vector field F^\mu that depends on the coordinates, fields, and transformation parameters. This condition ensures that the variation of the action \delta S = \int \delta \mathcal{L} \, d^4x reduces to a boundary term, which vanishes for appropriate boundary conditions, thereby preserving the without altering the on-shell. These symmetries are particularly characterized by field-dependent transformations of the form \delta \phi^i = \varepsilon^\alpha(x) X_\alpha^i[\phi], where \varepsilon^\alpha(x) are arbitrary functions representing local parameters, and X_\alpha^i[\phi] explicitly depends on the fields \phi and possibly their derivatives; more generally, the variation may involve derivatives of \varepsilon^\alpha via linear differential operators. Substituting such transformations into the first variation formula yields an off-shell identity of the form E_i(\phi) X_\alpha^i[\phi] + \partial_\mu K^\mu_\alpha = 0, where E_i(\phi) denotes the Euler-Lagrange operator components, and K^\mu_\alpha is a suitable current-like term; this identity holds identically, independent of whether the Euler-Lagrange equations E_i(\phi) = 0 are satisfied. For the general case with derivatives on parameters, integration by parts isolates differential relations among the E_i. In contrast to Noether's first theorem, which applies to field-independent transformations (global symmetries) and produces conserved currents corresponding to fixed constants of motion, the explicit field dependence here results in identities among the Euler-Lagrange equations rather than non-trivial conservation laws. A canonical example arises in gauge theories, such as , where the transformation \delta A_\mu = \partial_\mu \varepsilon(x) acts on the gauge field A_\mu with a local \varepsilon(x), rendering the invariant up to a total and leading to the off-shell identity \partial^\nu (\partial^\mu F_{\mu\nu}) = 0, which holds identically due to the antisymmetry of F_{\mu\nu} and reflects the Bianchi identity structure. These are often termed "open" or gauge symmetries due to their infinite-dimensional nature, parametrized by arbitrary functions, and they do not fixed constants of motion but instead enforce differential relations among the field equations.

Statement of Noether's Second Theorem

Noether's second theorem concerns variational symmetries of the action functional that depend on arbitrary functions, leading to differential identities among the equations of motion rather than conserved currents. Specifically, consider a Lagrangian density L(\phi, \partial \phi) for fields \phi^i in a spacetime manifold, with the action S[\phi] = \int L \, d^4x. If the action is invariant under field transformations of the form \delta \phi^i = \sum_{|I|=0}^s R_a^{i,I} (\phi, \partial \phi, \dots) \lambda^a_I (x), where \lambda^a(x) are arbitrary smooth functions (a=1,\dots,q), \lambda^a_I their derivatives up to order s, and R the corresponding characteristics, then the Euler-Lagrange equations E_i(\phi) = 0, defined by E_i(\phi) = \frac{\partial L}{\partial \phi^i} - \partial_\mu \left( \frac{\partial L}{\partial (\partial_\mu \phi^i)} \right), satisfy the identities Q_a = \sum_i \sum_{|I|=0}^s (-1)^{|I|} d_I \left( E_i R_a^{i,I} \right) = 0 holding identically (off-shell) for each a, where d_I denotes multi-index total derivatives. The derivation follows from the variational symmetry condition. Since the transformation is a symmetry, the variation of the action is a total divergence: \delta S = \int \partial_\mu K^\mu \, d^4x for some K^\mu. On the other hand, the general first variation of the action is \delta S = \int \left( \sum_i E_i \, \delta \phi^i + \partial_\mu T^\mu \right) d^4x, where T^\mu is the standard Noether current term. Substituting the general \delta \phi^i and equating the two expressions yields, after to account for derivatives on the parameters, \sum_a \int \lambda^a_I \left( \sum_i (-1)^{|I|} d_I (E_i R_a^{i,I}) \right) d^4x + \text{total divergence terms} = 0. Because the \lambda^a_I are arbitrary, their coefficients must vanish identically, giving the identities Q_a = 0. Unlike Noether's first theorem, no nontrivial conserved current arises here; the arbitrariness of the parameters forces the relations directly among the Euler-Lagrange equations themselves. For symmetries parameterized by multiple arbitrary functions \epsilon^a (a = 1, \dots, r), the theorem generalizes to r independent identities, which are linear combinations of the Euler-Lagrange equations. These identities reflect dependencies among the , reducing the number of independent conditions. In gauge theories, such as , the gauge symmetry \delta A_\nu = \partial_\nu \epsilon (with \epsilon arbitrary) yields the identity \partial_\mu (\partial^\nu F^{\mu\nu}) = 0 identically, akin to the Bianchi identity, ensuring consistency of the equations without introducing conserved quantities beyond the trivial ones.

Converse Theorem

The converse of Noether's second theorem establishes that certain linear identities among the Euler-Lagrange equations imply the existence of variational symmetries. Specifically, if the Euler-Lagrange equations E_a(\phi) = 0 for a Lagrangian \mathcal{L}(\phi, \partial \phi) satisfy non-trivial linear identities of the form \sum_a c^a E_a(\phi) \equiv 0, where the coefficients c^a are arbitrary functions (or more generally, span an infinite-dimensional space parameterized by arbitrary functions), then there exists a variational symmetry transformation \delta \phi^i = \sum_a c^a D_a^i[\phi], where the D_a^i are differential operators depending on \phi and its derivatives, such that the variation of the Lagrangian satisfies \delta \mathcal{L} = \partial_\mu F^\mu for some F^\mu. A proof proceeds constructively by treating the arbitrary functions parameterizing the identities as additional dependent variables in an extended variational problem. Applying the Euler-Lagrange to this augmented system yields the required characteristics Q^i = \sum_a c^a (D_a^i)^\dagger, where (D_a^i)^\dagger denotes the formal , ensuring the preserves the action up to a total . This inversion leverages the structure of the identities to generate the generators directly. For the general case, the extends to include of the parameters. Key assumptions include that the identities are non-trivial (i.e., not identically zero without restricting the solution space) and that the arbitrary functions span the of the Euler-Lagrange on-shell, ensuring the symmetries are variational and correspond to the full dependency structure. These conditions guarantee that the derived transformations act non-trivially on the solutions. In theories, such as Yang-Mills theory, the converse theorem demonstrates the completeness of the Noether identities, where the linear dependencies among the precisely match the gauge freedom parameterized by arbitrary Lie-algebra-valued functions, confirming that all on-shell redundancies arise from variational symmetries. Although the converse was explicitly stated and proven in Noether's original paper, its implications and modern formalizations gained prominence in the physics literature during the , particularly in the context of and theories.

Applications in Physics

Gauge Theories

Gauge symmetries constitute a class of local, spacetime-dependent transformations under which the action of a theory remains , introducing redundancies in the field configurations that describe the same physical state. These symmetries form infinite-dimensional groups parameterized by arbitrary functions of the coordinates, distinguishing them from global symmetries. In Abelian gauge theories like , the U(1) gauge transformation acts on the as \delta A_\mu = \partial_\mu \Lambda(x), where \Lambda(x) is an arbitrary smooth function, leaving the field strength F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu invariant. Noether's second theorem elucidates the consequences of such local symmetries by deriving identities that relate the equations of motion, rather than conserved currents. For Maxwell's theory with Lagrangian density \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu}, the Euler-Lagrange equations yield the inhomogeneous Maxwell equations \partial_\mu F^{\mu\nu} = J^\nu, which are dynamical. Noether's second theorem provides the identity \partial_\nu (\partial_\mu F^{\mu\nu}) = 0, holding identically due to the antisymmetry of F^{\mu\nu}, which ensures that the continuity equation \partial_\nu J^\nu = 0 follows on-shell. This identity ensures that the four equations are not independent, as the divergence of the inhomogeneous equation vanishes identically due to the antisymmetry of F^{\mu\nu}. In non-Abelian gauge theories, such as Yang-Mills theory, the structure generalizes with the gauge group G acting via . The is defined as D_\mu = \partial_\mu - i g [A_\mu, \cdot], where A_\mu = A_\mu^a T^a is the gauge field in the , and the field strength is F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu - i g [A_\mu, A_\nu]. The local gauge transformation is \delta A_\mu = D_\mu \lambda, with \lambda an algebra-valued function, leading via Noether's second theorem to identities that imply the D_\nu J^{\nu a} = 0 on-shell from the Yang-Mills equations \partial_\mu F^{\mu\nu a} + g f^{abc} A_\mu^b F^{\mu\nu c} = J^{\nu a}, manifesting constraints among the equations rather than laws. The Bianchi identities arise from the structure of the field strength. This identity underscores the non-independence of the Yang-Mills equations D_\mu F^{\mu\nu a} = J^{\nu a}, where —such as the Lorenz or gauge—is subsequently imposed to resolve the redundancy and define a transverse of physical configurations. The theorem thus provides the foundational explanation for why the full set of equations overdetermines the system, with the identities ensuring consistency. A profound extension arises in the quantization of theories through the BRST formalism, where Noether's second theorem is adapted to BRST symmetries—nilpotent, anticommuting transformations that incorporate fields to preserve invariance at the quantum level. These s, introduced as Grassmann-odd parameters replacing the original functions, generate transformations \delta_{\text{BRST}} \phi = s \phi \cdot c, where c denotes s and s is the BRST differential, leading to Noether identities in the extended that underpin the measure and eliminate unphysical . This framework, essential for perturbative , relies on the second theorem to derive the on-shell vanishing of BRST variations, ensuring the consistency of -antighost interactions in higher-stage reducible s.

Examples in Field Theories

In field theories beyond the standard gauge paradigms, Noether's second theorem manifests through identities arising from symmetries with arbitrary parameters, illustrating its broad utility in deriving on-shell relations without invoking conserved currents. Consider the massless Dirac field, described by the Lagrangian \mathcal{L} = \bar{\psi} i \gamma^\mu \partial_\mu \psi. The theory possesses a chiral symmetry under the transformation \delta \psi = i \epsilon \gamma^5 \psi, where \epsilon is a constant parameter, but extensions to local variations \epsilon(x) highlight the theorem's role in infinite-dimensional symmetries. Applying Noether's second theorem yields an identity relating the left- and right-handed components of the , expressed through the axial current J^\mu_5 = \bar{\psi} \gamma^\mu \gamma^5 \psi, where the charge Q_{5\epsilon} = \int d^3x \, J^0_{5\epsilon} connects the difference in left- and right-handed densities as Q_{5\epsilon} = -e \int d^3x \, \epsilon (\eta^\dagger \eta - \xi^\dagger \xi), enforcing consistency between the projected equations i \gamma^\mu \partial_\mu P_L \psi = 0 and i \gamma^\mu \partial_\mu P_R \psi = 0. In nonlinear sigma models, such as those modeling fields in low-energy , the approximate SU(2)_L × SU(2)_R chiral symmetry acts non-linearly on the fields \phi^a via \delta \phi^a = \epsilon^{ab} \phi^b / f_\pi + \cdots, where higher-order terms reflect the non-linear nature. Noether's second , applied to these infinite-parameter transformations, generates Ward identities that constrain scattering amplitudes and correlation functions, ensuring the symmetry's implications hold on-shell; for instance, the double-soft follows directly from these identities, dictating the behavior of amplitudes under soft limits without additional dynamical input. General relativity provides a gravitational example, where the Einstein-Hilbert action S = \frac{1}{16\pi G} \int d^4x \sqrt{-g} R is invariant under diffeomorphisms \delta x^\mu = \xi^\mu(x), an infinite-dimensional . Noether's second theorem implies that this invariance leads to differential identities among the , specifically the \nabla_\mu G^{\mu\nu} = 0, where G^{\mu\nu} is the ; these hold identically due to the antisymmetry of the Riemann tensor and ensure compatibility of the field equations without relying on their dynamical satisfaction. In (QED), while gauge theories represent a primary application, the theorem elucidates the status of : the \partial_\mu j^\mu = 0, with j^\mu = i q (\bar{\psi} \gamma^\mu \psi), emerges as an identity from the local U(1) symmetry rather than a dynamical consequence of the matter equations alone, as the second theorem demonstrates that \partial_\mu \partial_\nu F^{\mu\nu} \equiv 0 enforces current conservation off-shell via the antisymmetry of F^{\mu\nu}. A modern application appears in string theory, where the worldsheet action exhibits conformal invariance under holomorphic transformations z \to z + \epsilon(z). Noether's second theorem associates this local symmetry with the holomorphic stress-energy tensor T(z), yielding conserved currents J_z = T(z) \epsilon(z) whose conservation \bar{\partial} J_z = 0 implies the Virasoro constraints; these underpin the Virasoro algebra [L_m, L_n] = (m - n) L_{m+n} + \frac{c}{12} m (m^2 - 1) \delta_{m+n,0}, essential for anomaly cancellation and spectrum consistency in critical dimensions post-1980s developments.

Implications and Extensions

On-Shell Identities

In the context of Noether's second theorem, on-shell identities refer to the differential relations among the Euler-Lagrange that emerge from symmetries parameterized by infinitely many arbitrary functions, holding true precisely when the fields satisfy the equations of motion, denoted as E(\phi) = 0. These identities constrain the solutions by revealing dependencies among the equations, thereby reducing the number of independent conditions; for instance, in , the four Maxwell equations are subject to two independent Noether identities, leaving only two physically distinct constraints. The physical implications of these identities are profound, as they describe gauge orbits in the configuration space, where distinct field configurations related by transformations represent the same physical state, modulo redundancies that eliminate unphysical . This structure ensures that quantities remain under local , with the identities enforcing consistency by linking the to the symmetry generators. In theories with infinite-dimensional symmetry groups, such as those involving local parameters, these on-shell identities guarantee the well-posedness of the , preventing inconsistencies that could arise from overcounting dynamical variables. In , the on-shell identities from Noether's second theorem manifest as Ward-Takahashi identities, derived from the invariance of the under gauge transformations, which impose relations on correlation functions and ensure the unitarity and renormalizability of gauge-invariant theories. These quantum counterparts highlight how classical redundancies translate to constraints on scattering amplitudes and . An important extension involves off-shell formulations, where auxiliary fields or are introduced to reformulate the identities such that they hold without requiring the , facilitating the treatment of gauge symmetries in both classical and quantum settings. This approach is particularly useful in constrained Hamiltonian systems and covariant quantization procedures.

Modern Generalizations

One prominent modern generalization of Noether's second theorem arises in the covariant , developed by Wald and collaborators in the , which provides a framework for deriving conserved charges and identities in diffeomorphism-invariant theories like . In this approach, the theorem's on-shell identities are extended to yield variational principles for the presymplectic structure, enabling the identification of entropy as an integrable Noether charge associated with horizon symmetries. This has profound implications for , where emerges as a consequence of the on the covariant . In two-dimensional conformal field theories (CFTs), Noether's second theorem accommodates infinite-dimensional Lie algebras, such as affine Kac-Moody symmetries arising from global current algebras. These symmetries, parameterized by arbitrary functions on the worldsheet, lead to operator product expansions that realize the Kac-Moody algebra, with the theorem ensuring the consistency of conserved currents under quantization. This structure underpins the exact solvability of many CFTs and their role in dualities. For supersymmetric theories, the theorem generalizes to Noether identities that intertwine bosonic and fermionic transformations, treating parameters as gauge-like variables dependent on arbitrary Grassmann functions. In , these identities constrain the supercurrents and ensure the closure of the supersymmetry algebra on-shell, facilitating the construction of consistent supersymmetric actions even under spontaneous breaking. In higher-spin theories, which involve fields of spin greater than two and often require higher-order derivatives, Noether's second theorem has been reformulated using geometry to handle reducible symmetries and their dependencies on higher derivatives. This extension yields generalized identities that maintain consistency in the infinite-dimensional gauge structure typical of such theories. Recent applications in , particularly within the /CFT since the , leverage the theorem to derive holographic identities that relate redundancies to conformal anomalies. These identities provide constraints on symmetries across the duality, refining our understanding of quantum corrections to gravitational charges and entanglement in holographic systems.

References

  1. [1]
    [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.
  2. [2]
    English trans. of E. Noether Paper - UCLA
    Welcome to CWP at physics.UCLA.edu. E. Noether, "Invariante Variationsprobleme," Nachr. v. d. Ges. d. Wiss. zu Göttingen 1918, pp235-257.
  3. [3]
    [PDF] Noether's Two Theorems
    Second Theorem. An infinite-dimensional variational symmetry group depending upon an arbitrary function corresponds to a nontrivial differential relation among ...
  4. [4]
    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.Missing: primary | Show results with:primary
  5. [5]
    Anomalous diffusion and Noether's second theorem | Phys. Rev. E
    Mar 12, 2021 · The source of this multiplicity is the high degeneracy of the eigenvalues of Eq. (9) which generate a hierarchy of conserved charges. At ...Missing: primary | Show results with:primary
  6. [6]
    Gauge Structure in Algorithms for Plasma Physics - DataSpace
    Fourth, a formulation of Noether's second theorem is developed and applied toward a variational PIC algorithm using the formalism of discrete exterior calculus.<|control11|><|separator|>
  7. [7]
    Emmy Noether (1882 - 1935) - Biography - University of St Andrews
    Emmy Noether is best known for her contributions to abstract algebra, in particular, her study of chain conditions on ideals of rings. Thumbnail of Emmy Noether
  8. [8]
    Emmy Noether's Paradise - Ideas | Institute for Advanced Study
    Nov 23, 2016 · Amalie Emmy Noether was born in 1882 into an affluent family from the Bavarian town of Erlangen. She followed her father's footsteps to study ...
  9. [9]
    Emmy Noether, David Hilbert and Felix Klein | La Matematica
    Aug 12, 2025 · While Emmy Noether became both the first woman to be granted the title Privatdozentin (private lecturer) at the University of Göttingen in 1919 ...<|separator|>
  10. [10]
    [PDF] Emmy Noether on Energy Conservation in General Relativity - arXiv
    Dec 4, 2019 · In. 1918, Emmy Noether was working closely with Felix Klein, who was deter- mined to decipher the mathematical meaning of Hilbert's invariant ...<|separator|>
  11. [11]
    [PDF] The Noether Theorems in Context Introduction - PhilSci-Archive
    applies equally well to Emmy Noether's “Invariante Variationsprobleme”, pu- blished in the Nachrichten von der Königlichen Gesellschaft der Wissenschaften.<|control11|><|separator|>
  12. [12]
    [2103.17160] Noether's Theorems and Energy in General Relativity
    Mar 31, 2021 · This paper has three main aims: first, to give a pedagogical introduction to Noether's two theorems and their implications for energy conservation in general ...
  13. [13]
    [PDF] Noether's Theorems and Energy in General Relativity - PhilSci-Archive
    Mar 30, 2021 · In short, her second theorem could account for the four identities that Hilbert had pointed out, so that the conservation laws were 'improper', ...
  14. [14]
    [DOC] The Principle of Least Action as a Philosophical Shibboleth
    In the two decades around 1900, the PLA enjoyed a remarkable renaissance as a formal unification of mechanics, electrodynamics, thermodynamics, and relativity ...
  15. [15]
    Hilbert's 'Foundations of Physics': Gravitation and electromagnetism ...
    The usual implication is that Hilbert's principal intent in November 1915 was to arrive at a theory of gravitation based on the principle of general covariance ...Missing: debates | Show results with:debates
  16. [16]
    A Note on General Relativity, Energy Conservation, and Noether's ...
    In GR the energy-momentum conservation equation is implied by Noether's theorem and follows from the invariance of the Einstein-Hilbert action under ...
  17. [17]
    Substantive general covariance and the Einstein-Klein dispute - arXiv
    Sep 16, 2021 · Famously, Klein and Einstein were embroiled in an epistolary dispute over whether General Relativity has any physically meaningful conserved ...
  18. [18]
    [PDF] 5 Geometries in collision: Einstein,
    Special relativity and Klein's Erlangen Program. The primary burden ofEinstein's 1905 special theory ofrelativity was to deny physical significance tothe ...
  19. [19]
    [PDF] The Philosophy and Physics of Noether's Theorems
    In 1915, the great mathematicians Felix Klein (1849–1925) and David Hilbert (1862–1943) invited Noether to Göttingen in the hope that her expertise in invariant ...
  20. [20]
  21. [21]
    [PDF] Classical Action Principles
    The action principle states that under infinitesimal variations, the change in the ... CLASSICAL FIELD THEORY. 7 Version of January 12, 2015 is a four-vector. The ...
  22. [22]
    [PDF] cornelius lanczos - Variational Principles of Mechanics
    There is a tremendous treasure of philosophical meaning behind the great theories of Euler and Lagrange, and of Hamilton and. Jacobi, which is completely ...
  23. [23]
    [PDF] VARIATIONAL PRINCIPLES in CLASSICAL MECHANICS Douglas ...
    Aug 9, 2017 · ... variational principles that underlie the Lagrangian and Hamiltonian analytical formulations of classical mechanics. These variational ...
  24. [24]
    [PDF] chapter 2. lagrangian quantum field theory §2.1 general formalism
    Jan 2, 2010 · principle was used to derive the Euler-Lagrange field equations which describe the dynamical space-time evolution of the fields we must ...
  25. [25]
    [PDF] CLASSICAL FIELD THEORY - UT Physics
    which obey the. Euler–Lagrange equations of motion — yield the lowest action among all ...
  26. [26]
  27. [27]
    [PDF] Noether's theorem - Physics Department, Oxford University
    Feb 27, 2025 · Noether's theorem asserts that whenever a physical system has a symmetry (in a sense to be described) then there is a conserved quantity (i.e. ...
  28. [28]
    [PDF] Lie, Noether, and Lagrange
    Apr 15, 2017 · Lie groups are infinite and there is a group element for each value of a continuous parameter [3]. Lie groups correspond to the symmetries of ...
  29. [29]
    [PDF] 3 Classical Symmetries and Conservation Laws
    Some symmetries involve discrete operations, hence called discrete symme- tries, while others are continuous symmetries.
  30. [30]
    [PDF] 13.1 Field theories from Lagrangians - MIT
    Requiring δS = 0, this analysis yields Euler-Lagrange equations for the field: ... We expand that first variation using δg. /. -g = -. 1. 2. /. -ggαβδgαβ ...
  31. [31]
    [PDF] The Noether theorem
    is required when considering the variation of φ, since the φ depend on the xµ. The total variation of φ(xµ) is given by δφ(xµ) = φ. 0. (xµ0) − φ(xµ). (3.12).<|control11|><|separator|>
  32. [32]
    [PDF] Symmetry Transformations, the Einstein-Hilbert Action, and Gauge ...
    Symmetry transformations are changes in the coordinates or variables that leave the action invariant. It is well known that continuous symmetries generate ...
  33. [33]
    [PDF] Noether's Theorems and Gauge Symmetries - arXiv
    We present, in a general form, all the main results relating to the Noether variational problem for gauge theories, and we show the rela- tionships between them ...
  34. [34]
    [PDF] Emmy Noether and Symmetry
    Second Theorem. An infinite-dimensional variational symmetry group depending upon an arbitrary function corresponds to a nontrivial differential relation among ...
  35. [35]
    [PDF] The Noether Theorems
    Aug 14, 2009 · Euler–Lagrange derivative or Euler–Lagrange differential . Denote it ... arbitrary functions and their derivatives up to order σ, then there.
  36. [36]
    None
    Summary of each segment:
  37. [37]
  38. [38]
    [PDF] Extensions of Noether's Second Theorem - University of Kent
    A simple local proof of Noether's Second Theorem is given. This proof immedi- ately leads to a generalization of the theorem, yielding conservation laws ...Missing: source | Show results with:source<|control11|><|separator|>
  39. [39]
    A new approach to the converse of Noether's theorem - IOPscience
    This condition does not mention any second-order differential equation field but is expressed in terms of the geometry of the second-order tangent bundle.
  40. [40]
    [hep-th/0009058] Noether's Theorems and Gauge Symmetries - arXiv
    Sep 8, 2000 · We present, in a general form, all the main results relating to the Noether variational problem for gauge theories, and we show the relationships between them.
  41. [41]
    [PDF] arXiv:2503.07525v1 [hep-th] 10 Mar 2025
    Mar 10, 2025 · ψ is the Dirac field, which is massless in this study. 3 Noether's second theorem. In this study, we address the charges in the case that ...
  42. [42]
    None
    Nothing is retrieved...<|separator|>
  43. [43]
    [PDF] Noether's Theorems and Energy in General Relativity - arXiv
    Mar 31, 2021 · Noether's second theorem, applied to the matter part of the action, gives the covariant analogue of the energy and momentum conservation ...<|control11|><|separator|>
  44. [44]
    [PDF] 4. Introducing Conformal Field Theory
    The spirit of Noether's theorem in quantum field theories is captured by operator equations known as Ward Identities. Here we derive the Ward identities ...
  45. [45]
    Noether's second theorem for BRST symmetries - AIP Publishing
    May 6, 2005 · ... Noether identities. We present Noether's second theorem in the case of BRST transformations depending on derivatives (jets) of dynamic ...
  46. [46]
    [gr-qc/9403028] Some Properties of Noether Charge and a Proposal ...
    Mar 15, 1994 · We propose a local, geometrical prescription for the entropy, S_{dyn}, of a dynamical black hole. This prescription agrees with the Noether charge formula.
  47. [47]
    Noether's second theorem for BRST symmetries - math-ph - arXiv
    Dec 10, 2004 · We present Noether's second theorem for graded Lagrangian systems of even and odd variables on an arbitrary body manifold X in a general case of ...Missing: derivation | Show results with:derivation