Noether's second theorem

Summary

In mathematics and theoretical physics, Noether's second theorem relates symmetries of an action functional with a system of differential equations.[1] The action S of a physical system is an integral of a so-called Lagrangian function L, from which the system's behavior can be determined by the principle of least action.

Specifically, the theorem says that if the action has an infinite-dimensional Lie algebra of infinitesimal symmetries parameterized linearly by k arbitrary functions and their derivatives up to order m, then the functional derivatives of L satisfy a system of k differential equations.

Noether's second theorem is sometimes used in gauge theory. Gauge theories are the basic elements of all modern field theories of physics, such as the prevailing Standard Model.

The theorem is named after its discoverer, Emmy Noether.

Mathematical formulation edit

First variation formula edit

Suppose that we have a dynamical system specified in terms of   independent variables  ,   dependent variables  , and a Lagrangian function   of some finite order  . Here   is the collection of all  th order partial derivatives of the dependent variables. As a general rule, latin indices   from the middle of the alphabet take the values  , greek indices take the values  , and the summation convention apply to them. Multiindex notation for the latin indices is also introduced as follows. A multiindex   of length   is an ordered list   of   ordinary indices. The length is denoted as  . The summation convention does not directly apply to multiindices since the summation over lengths needs to be displayed explicitly, e.g.

 
The variation of the Lagrangian with respect to an arbitrary variation   of the independent variables is
 
and applying the inverse product rule of differentiation we get
 
where
 
are the Euler-Lagrange expressions of the Lagrangian, and the coefficients   (Lagrangian momenta) are given by
 

Variational symmetries edit

A variation   is an infinitesimal symmetry of the Lagrangian   if   under this variation. It is an infinitesimal quasi-symmetry if there is a current   such that  .

It should be remarked that it is possible to extend infinitesimal (quasi-)symmetries by including variations with   as well, i.e. the independent variables are also varied. However such symmetries can always be rewritten so that they act only on the dependent variables. Therefore, in the sequel we restrict to so-called vertical variations where  .

For Noether's second theorem, we consider those variational symmetries (called gauge symmetries) which are parametrized linearly by a set of arbitrary functions and their derivatives. These variations have the generic form

 
where the coefficients   can depend on the independent and dependent variables as well as the derivatives of the latter up to some finite order, the   are arbitrarily specifiable functions of the independent variables, and the latin indices   take the values  , where   is some positive integer.

For these variations to be (exact, i.e. not quasi-) gauge symmetries of the Lagrangian, it is necessary that   for all possible choices of the functions  . If the variations are quasi-symmetries, it is then necessary that the current also depends linearly and differentially on the arbitrary functions, i.e. then  , where

 
For simplicity, we will assume that all gauge symmetries are exact symmetries, but the general case is handled similarly.

Noether's second theorem edit

The statement of Noether's second theorem is that whenever given a Lagrangian   as above, which admits gauge symmetries   parametrized linearly by   arbitrary functions and their derivatives, then there exist   linear differential relations between the Euler-Lagrange equations of  .

Combining the first variation formula together with the fact that the variations   are symmetries, we get

 
where on the first term proportional to the Euler-Lagrange expressions, further integrations by parts can be performed as
 
where
 
in particular for  ,
 
Hence, we have an off-shell relation
 
where   with  . This relation is valid for any choice of the gauge parameters  . Choosing them to be compactly supported, and integrating the relation over the manifold of independent variables, the integral total divergence terms vanishes due to Stokes' theorem. Then from the fundamental lemma of the calculus of variations, we obtain that   identically as off-shell relations (in fact, since the   are linear in the Euler-Lagrange expressions, they necessarily vanish on-shell). Inserting this back into the initial equation, we also obtain the off-shell conservation law  .

The expressions   are differential in the Euler-Lagrange expressions, specifically we have

 
where
 
Hence, the equations
 
are   differential relations to which the Euler-Lagrange expressions are subject to, and therefore the Euler-Lagrange equations of the system are not independent.

Converse result edit

A converse of the second Noether them can also be established. Specifically, suppose that the Euler-Lagrange expressions   of the system are subject to   differential relations

 
Letting   be an arbitrary  -tuple of functions, the formal adjoint of the operator   acts on these functions through the formula
 
which defines the adjoint operator   uniquely. The coefficients of the adjoint operator are obtained through integration by parts as before, specifically
 
where
 
Then the definition of the adjoint operator together with the relations   state that for each  -tuple of functions  , the value of the adjoint on the functions when contracted with the Euler-Lagrange expressions is a total divergence, viz.   therefore if we define the variations
 
the variation
 
of the Lagrangian is a total divergence, hence the variations   are quasi-symmetries for every value of the functions  .

See also edit

Notes edit

  1. ^ Noether, Emmy (1918), "Invariante Variationsprobleme", Nachr. D. König. Gesellsch. D. Wiss. Zu Göttingen, Math-phys. Klasse, 1918: 235–257
    Translated in Noether, Emmy (1971). "Invariant variation problems". Transport Theory and Statistical Physics. 1 (3): 186–207. arXiv:physics/0503066. Bibcode:1971TTSP....1..186N. doi:10.1080/00411457108231446. S2CID 119019843.

References edit

Further reading edit

  • Noether, Emmy (1971). "Invariant Variation Problems". Transport Theory and Statistical Physics. 1 (3): 186–207. arXiv:physics/0503066. Bibcode:1971TTSP....1..186N. doi:10.1080/00411457108231446. S2CID 119019843.
  • Fulp, Ron; Lada, Tom; Stasheff, Jim (2002). "Noether's variational theorem II and the BV formalism". arXiv:math/0204079.
  • Bashkirov, D.; Giachetta, G.; Mangiarotti, L.; Sardanashvily, G (2008). "The KT-BRST Complex of a Degenerate Lagrangian System". Letters in Mathematical Physics. 83 (3): 237–252. arXiv:math-ph/0702097. Bibcode:2008LMaPh..83..237B. doi:10.1007/s11005-008-0226-y. S2CID 119716996.
  • Montesinos, Merced; Gonzalez, Diego; Celada, Mariano; Diaz, Bogar (2017). "Reformulation of the symmetries of first-order general relativity". Classical and Quantum Gravity. 34 (20): 205002. arXiv:1704.04248. Bibcode:2017CQGra..34t5002M. doi:10.1088/1361-6382/aa89f3. S2CID 119268222.
  • Montesinos, Merced; Gonzalez, Diego; Celada, Mariano (2018). "The gauge symmetries of first-order general relativity with matter fields". Classical and Quantum Gravity. 35 (20): 205005. arXiv:1809.10729. Bibcode:2018CQGra..35t5005M. doi:10.1088/1361-6382/aae10d. S2CID 53531742.