You are on page 1of 17

ANNALSOFPHYSICS127,257-273(1980)

Symmetries and Conservation Laws in Gauge Theories*

R. JACKIW AND N. S. MANTON

Center for Theoretical Physics, Laboratory for Nuclear Science and Department of Physics,
Massachusetts Insfitute of Technology, Cambridge, Massachusetts 02139

Received December 19, 1979

The relationship between conservation laws and symmetries of space-time is familiar.


Here it is shown that in a symmetric background gauge field these conservation laws
persist, but in modified form. A further contribution to the conserved quantity occurs. It
is determined by the gauge transformation which, when acting together with some co-
ordinate transformation, leaves the symmetric background gauge-potential invariant.
The addition to the constant of motion can also be interpreted as arising from the dynamical
interaction of the gauge field with the system. A classical example is the angular momentum
conservation law for a charged particle moving in the field of a magnetic monopole. Generali-
zations of this are here derived.

1. INTRODUCTION

Attention has recently been drawn to the problem of defining the response of gauge
potentials to coordinate transformations and characterizing those configurations
that are invariant under specific coordinate changes. The reason that this classical
problem retains interest is the presence of gauge-invariance [I]. If one were discussing
a vector field A,(x) which did not undergo gauge transformations, then its response to
a coordinate transformation, whose infinitesimal form is

x” + X” = x’” + 8X”,
(1.1)
6.P = f”(X)

would be a familiar formula involving the Lie derivative,

* This work is supported in part through funds provided by the U.S. Department of Energy
(DOE) under contract EY-76-C-02-3069.
257
0003-4916/80/080257-17$05.00/O
Copyright 0 1980 by Academic Press, Inc.
All rights of reproduction in any form reserved.
258 JACKIW AND MANTON

[Recall that the Lie derivative Lf of a general tensor field II”;:::: produces a tensor of
the same type.

The Lie derivative of products of tensors follows Leibnitz’ rule and contractions are
respected, so that the order of taking the Lie derivative and of contracting indices is
immaterial.] An invariant field is one that does not respond to the transformation;
i.e., its Lie derivative vanishes. However, gauge freedom modifies the above. The
concept of coordinate invariance must be extended to allow for non-invariance that
can be compensated by a gauge transformation. Thus an invariant gauge-potential
need not have vanishing Lie derivative; rather it must satisfy the following condition:

L,A, = zZuW, . (1.3)

Here ZSUis the gauge-covariant derivative

caII= a,- [A,, (1.4)

and W, , as well as A, , take values in the space of anti-Hermitian matrices which


span the Lie algebra of the gauge group.
Even the formula for the response of A, to a coordinate transformation can be
changed by adjoining a gauge transformation. The observation that the Lie derivative
may also be written in terms of the field strength,

LA = f% + ~,(f*A,), (1.5)

suggests defining a gauge-covariant coordinate transformation, which differs from


the usual (1.2) by a gauge transformation [2].

&A, = f”Fau . (14

The invariance condition requires the above to be an infinitesimal gauge transforma;


tion.
pFarr = -.Q,,yf. (1.7)

This also shows that certain projections of an invariant field strength are [gauge-
covariant] derivatives of a scalar field [3]. Comparison of (1.3), (1.5), and (1.7) gives
the relationship between Y and W.

W, = f”A, - ‘yf. (l-8)


A change of gauge
A, --+ UA,V + (a,,U) U-l (1.9)
CONSERVATION LAWS IN GAUGE THEORIES 259

induces a gauge-covariant change in ?P.

?Pf - uYfu-l. (1.10)

However, as is seen from (1.8), W, undergoes an inhomogeneous transformation,


which makes reference to the form of the coordinate transformation, J

w, -+ UWJJ-I $ (L&J) U-l. (1.11)

Two infinitesimal coordinate changes f and g, performed in the two opposite


orders, result in a third one, h.

ha=f@,ga -g@,f" = L,g” = -L$“. (1.12)

The composition law for the scalar fields W and Y follows directly by iterating (1.2)
and using (1.12). W satisfies [4]

L, w, - L, w, - [ w, , W,] - w, = 0, (1.13)

while from (1.8) we obtain the corresponding formula for !P, which also reflects the
composition law for gauge-covariant coordinate transformations (1.6) [2].

f"gBFas=f"gaul, -gagYf + Wf, Y,l- ph. (1.14a)

Moreover, (1.7) permits rewriting the first two terms on the right-hand side of (1.14a)
in terms of the same projection of F,, occurring on the left-hand side. This leaves

f%%o = -P-‘f , Y,l + UT,> (1.14b)

a formula which supplements (1.7) to give an expression for further projections of an


invariant F,, in terms of the scalar fields Y.
The above summarizes known results [1 J. In this paper we continue to explore the
properties of the scalar fields Y, and their close relatives W. We show that they have
the following physical significance. When we consider a free or self-interacting
“matter” system in the absence of gauge fields, the dynamics will be characterized by
various constants of motion which are present as consequence of invariance under
[infinitesimal] coordinate transformations J [Examples are the energy as a
consequence of time-translational invariance; the angular momentum arising from
rotational invariance etc.] Tf the matter interacts with a dynamical gauge field in a
fashion consistent with invariance under the same coordinate transformations, the
constants of motion will continue to be present and will have the form

Cf = Cfnatter + Ci&namlcaI gauge field . (I. 15)


The first part represents the contribution of the matter and of the matter-field inter-
action terms; the second is the contribution of the gauge fields themselves.
260 JACKIW AND MANTON

On the other hand, if the matter system interacts with an external, prescribed
gauge field, the invariance will be absent in the general case, since a generic external
field removes all symmetries. Consequently, there will no longer be any constants of
motion. However, in the special case when the external gauge fields themselves are
invariant under the infinitesimal coordinate transformation f, then the matter
dynamics will continue to be characterized by constants whose form will now be

Cf = Glatter+ Gern&lgaugefleld
. (1.16)

The second term on the right-hand side allows for a contribution to the total con-
served quantity from the external gauge field. We show here that such contributions
are indeed present; they are essentially determined by Y. In other words, for invariant
fields, the scalar field defined in (1.7) has the physical significance of carrying the
external gauge field contribution to the total constant of motion.
We shall make frequent use of Noether’s method for finding constants of motion.
Let us record how it works. Consider an action I depending on dynamical variables
x through a Lagrangian 2.
I = j d4x g11z9. (1.17)

[It is assumed that no derivatives of x higher than the first occur in A?.] Dynamical
equations of motion for x are determined by requiring that I be invariant against
arbitrary variations of x. Consider next a specific infinitesimal transformation x -+
x + 6x. If under this transformation one can demonstrate, without using equations
of motion, that the action changes into

(1.18)

i.e., apart from surface terms it is invariant, then the transformation is a symmetry
operation and the current P, given by

J* = w’2g) 6x - K”,
Wccx)
is conserved,
&Ja = 0. (1.19b)

If the coordinate x0 is time, then there is a constant of motion

C = 1 dr Jo. (1.20)

There is a general procedure which we shall use to find transformations where


Noether’s theorem can be applied. Consider a theory of gauge fields and some other
matter fields in curved space. Regarding both the fields and the metric as dynamical
variables, the action possesses several large invariance groups. There will be in-
CONSERVATION LAWS IN GAUGE THEORIES 261

variance under general coordinate transformations and gauge invariance. Further,


if the matter includes spinor fields, it will be necessary to introduce a vierbein, and
then the action will be invariant under local Lorentz transformations. Finally, in
some cases, for example, if gauge fields are coupled to massless spinors, there will be
Weyl invariance. All these invariances are local, so that the parameters of the trans-
formations are arbitrary functions. Correspondingly the invariance groups are
infinite dimensional.
Moreover, we know quite generally the effect of these transformations on the
action. Under gauge, local Lorentz, and Weyl transformations, the action is strictly
invariant. Under an infinitesimal coordinate transformation, the action is not strictly
invariant. The change is

61 = f d4x i3,(g’ppTq (1.21)

which could be expressed as a surface integral. Notice that (1.21) is true for any
Lagrangian with no explicit coordinate dependence. This result follows from the
properties of Lie derivatives quoted earlier. We just suppose that all tensorial quanti-
ties transform according to the Lie derivative, and that the Lagrangian 9 is a scalar
constructed from contractions of products of tensors.
Suppose now that the metric and associated geometric quantities are not dynamical.
This is the usual situation of a field theory in a curved space which is not a dynamical
theory of gravity. Furthermore, suppose that the gauge potential is an external
background field, so that only the other matter fields are dynamical. If there is some
combination of transformations which leave both the geometrical quantities and the
gauge-potential invariant, then we can obtain a conservation law for the matter
fields. The argument is simple. Applying this combination of transformation to the
matter, gauge, and geometrical fields, we know that the change in the action is (1.21).
But the gauge and geometrical fields did not change, so we could have applied the
transformations to the matter fields alone, still obtaining (1.21). In this way we find
a transformation of the matter fields which changes the action by a total divergence,
and Noether’s theorem guarantees a conservation law.
The above argument is a generalization, to theories with background gauge fields,
of the usual one relating conservation laws to invariances of the background geometry.
The group of invariances of both the geometry and the background gauge potential
is finite dimensional. In flat space, for example, a gauge-potential could be static and
rotationally invariant, so there would be a conserved energy and angular momentum,
but translation invariance would in general be broken and the conservation of momen-
tum would disappear.
The remainder of this paper is devoted to demonstraring and examplifying the
above. Jn Section II we discuss matter fields, coupled to dynamical gauge fields,
whereas in Section III the gauge fields will be taken to be external and fixed. In
Section IV the matter is taken to be a point particle. Here we encounter a classical
instance of our general theory: the non-kinematical addition to the angular momen-
tum of a charged particle moving in the field of an Abelian magnetic monopole.
262 JACKIW AND MANTON

II. INVARIANCES OF DYNAMICAL FIELDS

In this section we review invariance properties of a theory of dynamical gauge


fields coupled to matter. These are of course familiar; they are here presented for the
sake of completeness to establish notation and to set the stage for our new results.
For the moment we ignore the matter.
The action for the gauge fields

Zgsuge= 1 d4x gl’z_EPgauge,


(2.1)
9 gauge = 3 tr F,,F,,g”“g”

is invariant under coordinate transformations and Weyl transformations, provided


the background metric is also transformed. The rules for coordinate transformations
are
&guy = Lfguv, (2.2a)
&A, = LA, (2.2b)

while for the Weyl transformation we have

hvg P” = -2AgwJ, (2.3a)


&A, = 0, (2.3b)

where f and A are arbitrary functions.


These rules imply that
6c gll2 = a,( g’/2f”), (2.4)
8wgl/2 = 4Ag112. (2.5)

To check for invariance of the dynamics in a given, background metric we must


not transform the metric tensor, which is not a dynamical variable. Thus the dynamics
will be invariant only under those combinations of coordinate and Weyl trans-
formations which leave the metric tensor invariant. We therefore require

8fP” = Lfg w - 2AgL(v = 0. (2.6a)

In this case A is determined by f

A = -*p:, (2.6b)

and f must satisfy the conformal Killing equation

AL:”
+fy:u= w:&,Y~
[The semicolon denotes coordinate covariant differentiation.]
CONSERVATION LAWS IN GAUGE THEORIES 263

Using an fsatisfying (2.6c), the relevant transformation of A, is

&A, = L,A, - SBu(f”A,) = forF,* (2.7)

as in (1.6), and the change of the action is

Qggauge
=sd’x
?4g”2fa=%auge) (2.8)

as demonstrated quite generally in the last section.


The current, obtained from Noether’s theorem, is

J” = g1j2 2 tr(F”‘F,,, - $F,,F”%,“) f’. (2.9)

Note that the symmetric, gauge-invariant, and traceless stress-energy tensor has
been constructed, because of our use of (2.7).

eglge = 2 tr(Fb18F$ - $FwyF”V,ay). (2.10)

In the absence of other interactions the current and the stress-energy tensor are
covariantly conserved. The gauge-field contribution to the constant of motion (1.15) is

= 2 tr dr g1’2(P0Fy,fy - $Fu,F”“fo). (2.11)


s

Now we couple matter to the gauge field. The only matter field theory that we shall
consider in detail is a fermionic, spin-4 Dirac model. The formalism in curved space
is complicated by the need to introduce a vierbein ez [5]. The Greek indices are
raised and lowered by the metric tensor g,,,; the Latin ones by the flat space metric
?lab . The two are related through the vierbein.
ab
guv = %beMev . (2.12)

The Dirac field # transforms according to a definite representation of the gauge


group, and we shall continue to denote the gauge-potential by A,, even in this represen-
tation. The Dirac adjoint 4 is obtained from # by Hermitian conjugation and multi-
plication by r”, the flat-space Dirac matrix. The curved space gamma matrices are
constructed from the flat-space ones with the help of the uierbein.

P = ezy”. (2.13)

Before writing down an action we must still introduce the spin connection, CO:‘.
This is defined by
GM - wzbeb,) - (p t+ v) = 0. (2.14)
264 JACKIW AND MANTON

An explicit expression for WE’ in terms of the vierbein can be found, but we shah not
need this [5]. The action is now

I matter = d4x @“ymatter , (2.15a)


s

$P,atter = (i/2) $r”(a,$ - ~~4 - A,#) + h.c. (2.15b)


Here w II = 10~~~0
2 )1 ab , where o,, is the Dirac spin matrix

uab = f[y”, yb] (2.16)

and h.c. indicates that the Hermitian conjugate is to be added.


As is well-known Imatter is invariant, apart from the surface term (1.21), under
infinitesimal coordinate transformations

(2.17a)

provided the vierbein is transformed like a covariant vector.

&e,” = L,eE = f”aaeye,”+ (a,f*) e,” . (2.17b)

The action is also invariant under local Lorentz transformations which transform
only the spinors and the vierbein.

SL# = -J-&k (2.18a)

sLe,” = -!ii@beE , (2.18b)

fPb(X) = -Qba(X), 52 = &Qabff&, . (2.19)

The action is, of course, gauge invariant, and finally, owing to the masslessness of the
Dirac fields, the action possesses Weyl symmetry, but to avoid further inessential
complications we shall henceforth ignore this.
The transformation laws of the vierbein, together with the detinition (2.14), imply
that the spin connection transforms like a gauge-potential for the local Lorentz
group, namely,
&%L = Lr% 9 (2.20)
&o, = --a,!2 + [w, ) sz]. (2.21)

Just as in the pure gauge case, when the curved space background is not dynamical,
the Fermi field dynamics will be invariant only under that combination of coordinate
and Lorentz transformations which leave the vierbein [and the metric] invariant. For
there to be such a combination, we need only assume that f is a Killing vector, so
that
L,g,, = 0 (2.22a)
CONSERVATION LAWS IN GAUGE THEORIES 265

or equivalently
.LY +.L = 0. (2.22b)

Then the vierbein will be invariant, since the equation

S,e,” = L,e: - ,R”,eE = 0 (2.23)

uniquely determines the matrix I&, , and this matrix does represent an infinitesimal
Lorentz transformation because of (2.12) and (2.22b). We denote this local Lorentz
transformation associated with f by s2,. Clearly, a combination of transformations
which leaves the vierbein invariant also leaves the spin connection invariant.
Having found a combination of transformations leaving the geometrical quantities
fixed, we can now construct a conservation law for the system. The transformation of
the Dirac field is
84 = faDa+ - &+h (2.24)

where D, is the gauge-covariant derivative. Just as in (1.6) and (2.7) we have supple-
mented the coordinate transformation with a gauge transformation. In this way we
obtain a gauge-invariant current

J hatter = (i/2) g1’2$l?$# + h.c. (2.25)

for the fermion. Note that the expected term involving Smatter vanishes when the
field equations are satisfied. The matter contribution to the constant of motion (1.15)
takes the form
CLatter = i s dr g1’2t,6T08,# + h.c. (2.26)

Obviously, when both the gauge field and the matter field undergo the coordinate
transformation, the combined action is invariant, and the combined current

J; = Jkatter + J&nammgaugemci . (2.27)


is conserved, leading to the constant of motion (1.15).

III. MATTER COUPLED TO AN EXTERNAL GAUGE FIELD

When the gauge field is nondynamical, i.e., taken to be an external, prescribed


function of space and time, the relevant action is just the matter part, and the gauge-
potentials are not varied. In that circumstance, the symmetries will in general be
broken by the external field and no constants of motion will be present. However,
some constants of motion will persist if the gauge-potential itself possesses some
invariances.
266 JACKIW AND MANTON

We take f and Ln, to be the same quantities as before, that is, a coordinate and
local Lorentz transformation which together leave the metric and vierbein invariant.
Now we also assume that the gauge-potential has the invariance property

(3.1)
Recall that this equation has the interpretation that the (gauge-covariant) coordinate
transformation generated by f, followed by the further infinitesimal gauge trans-
formation generated by Y, , leaves A, invariant.
We know, from the general argument in Section I, how to find a conservation law
for the matter. We consider the transformation of the Dirac field

A,$ = fa&$ - Q,lcI+ Yd. (3.2)


This is just as in (2.24) but with the gauge transformation generated by Yf included.
As usual, the change in the action is

A&atter = s d4x Q?‘zfa%natter). (3.3)

This is because the coordinate, local Lorentz, and gauge transformation acting on
the geometrical quantities and all the fields produces the change in the action (1.21),
but this particular combination of transformations leaves the geometry and gauge-
potential invariant, so the same change 81 results from the transformation of the
matter field alone. It follows that the conserved current is

J; = (i/2) g?,PAf# + h.c., (3.4)


where the term proportional to 9 matter again vanishes by virtue of the matter-field
equations. Part of the current (3.4) is just the conventional matter current of (2.25);
the extra piece comes from the external gauge field and can also be written in terms of
the matter source currentj”.
J,” = J&tter + ju”‘G, (3.5)
ji = ig1’2$lYa*. (3.6)
Here t, are the representation matrices of the gauge algebra. This source current
appears in the field equation for the dynamical gauge field coupled to the fermionic
matter.
(9Gg”2F”“), = ji . (3.7)
The constant of motion is as in (1.16) with

where the (non-Abelian) charge density p is the time component of,jU.


CONSERVATION LAWS IN GAUGE THEORIES 267

In the transformation law (3.2), each term is separately gauge invariant. There is an
equivalent approach which is perhaps more canonical. In view of (1.8), we can
rewrite (3.2) as
44 = folad - Qf* - Wf# (3.9)

which shows that the constant of motion is

Cf = 1 dr gl” 1; $r”(f”a,# - .n,# - W,t,,h) + h.c. . (3.10)


I

This is still gauge invariant as a whole, because of (1.11). Moreover (3.10) makes no
reference to the gauge-potential, but only to W, . This is interesting since the fields W,
which are determined by (1.13), depend only on the geometry of the symmetry group.
For given W’s, there is a large class of invariant gauge-potentials satisfying (1.3) [l],
but the constant of motion (3.10) is the same for each member of the class.
Now it might be argued that there is no such thing as a background gauge-poten-
tial, and that the conserved quantity we have obtained should be the same as that
obtained when regarding the gauge field as dynamical. In some sense this is the case,
but the demonstration requires some further assumptions, and a little care.
We assume first that the symmetric background gauge-potential satisfies the
vacuum-field equations. Then matter is introduced, and the matter source current is
assumed to be small, so that the matter behaves like a test charge. This source induces
a perturbation of the gauge field, and the dynamical gauge-potential can be written

A, = A: + a,, (3.11)

where A: is the symmetric background potential. a, is small and has no particular


symmetry properties. Both the external gauge-field picture and the dynamical picture
are now correct, for the gauge-potential is dynamical, but the matter interacts with
just the background gauge field if we ignore the back reaction of a, . To show that the
constant of motion arising from dynamical gauge fields reduces in this case to that for
external gauge fields, we make one further assumption: that the coordinate trans-
formation is purely spatial, i.e., f” = 0.
The proof begins by rewriting (2.11) as

C&namicalgauge field = 2 tr dr g1”Fo4(f ‘FiB + BD?Py)


s

- 2 tr dr g1’2FoEBb!Pf . (3.12)
I

Integrating by parts, ignoring the surface term, and using the non-Abelian Gauss’ law
for the dynamical gauge field.
9ug1/2Fwo = p, (3.13)
268 JACKIW AND MANTON

show that the second term of (3.12) gives just the integral (3.8), so that if the first
term of (3.12) vanishes, then we shall have demonstrated the equivalence of the
dynamical gauge field and the external gauge-field contributions to the constant of
motion. By substituting (3.11) into the first term of (3.12) and expanding to first order
in a, , we obtain

2 tr dr g”“FBoB{fi(~~ao - 9faJ - [as , ul,]}. (3.14)


s

The zeroth-order terms have disappeared because of the symmetry of the background
field (3.1). An integration by parts and use of -FL = [9:, @] casts the above into
the form

2 tr dr {-L3i(g”2FoBfi) + 9,( gl”FcyfB) - g”“g”“gs”[5B~ . B,,] u/,} a0 , (3.15)


s

where all quantities, except a, , now refer to the background gauge-potential. Next
we use the symmetry equation for the background field to rewrite the covariant
derivative of !Pf in terms of components of the field tensor, giving

2 tr dr (-9i(g1’2Foafi) + -9,(g”2FoyfB)
s

+ g”2go*gsv9~(fiFi,) - g”2go~gs”~~(fiF~~)} a, . (3.16)

Finally, the Killing equation satisfied by f, the Bianchi identity satisfied by Fm, , and
the field equation for the background field yield the desired result, that the above
vanishes.
We conclude that the contribution (3.8) to the constant of motion can be inter-
preted as being carried by the gauge field itself, and arises from the interaction of the
small induced uL1with the background potential A: , in the dynamical picture.
A useful generalization of this result is possible. Suppose the background gauge-
potential satisfies the field equations with a fixed background source, which itself is
symmetric. This will allow a much larger class of symmetric background gauge-
potentials. Again the constant of motion obtained by perturbing the gauge-potential
dynamically by the matter will be the same as that obtained by regarding only the
matter as dynamical. Both of these constants will however be different from before.

IV. POINT PARTICLES COUPLED TO GAUGE FIELDS

Another application of our general theory is to a point particle moving in curved


space-time and interacting with a gauge field. The equations governing this dynamics
were derived by Wong [6]. We wish to formulate them in terms of an action principle,
CONSERVATION LAWS IN GAUGE THEORIES 269

hence we take for the total action a sum of the gauge-field action [for the moment the
gauge field is assumed to be dynamical] and the particle action.

Iparticle = -c j ds + 2 tr 1 dzp A,(z) T. (4.1)

In the first term, describing free motion, ds is the proper interval and c a normaliza-
tion constant.
ds = (g,“(z) dz@dz”)1/2 = (guy(z) i’ti”)1/2 dh. (4.2a)

The particle coordinate is zfi. In the first equality a parametrization-invariant expres-


sion is given; in the second, we have explicitly parametrized in terms of an arbitrary
quantity X, with the dot signifying differentiation with respect to h. The parametriza-
tion allows rewriting the last term in (4.1), the gauge-field-particle interaction term,
as
dz“A,(z) T = dhi”AA,(z) T. (4.2b)

T is the total isospin [or generalizations thereof] of the particle. It is taken to be an


external, nondynamical quantity which is not varied in the derivation of the equations
of motion. Nevertheless T depends on h and satisfies an equation given below.
Variation of A, produces the field equation.

9J gl/2Ffqx)) = j”(x),
(4.3)
j”(x) = 1 dh svT a4(x - z).

An integrability condition can be derived from this by contracting another covariant


derivative in the free index. As a consequence jU must be covariantly conserved, a
requirement which is met provided

rP - [.?A,, T] = 0. (4.4)

This also guarantees the gauge- and parametrization-invariance of the interaction


term in (4.1). We shall henceforth take (4.4) to be a subsidiary condition satisfied by
T. Finally, varying with respect to zLL,the coordinate of the particle, gives with the
help of (4.4),

(4.5)

Equations (4.3)-(4.5) are the Wong equations; they may alternatively be derived by
considering action principles in which T is a dynamical quantity. These more compli-
cated approaches are not needed by us.
270 JACKIW AND MANTON

Coordinate transformations of zu with Killing vectors leave the kinetic term in-
variant. when the particle is massless, iup, = 0, conformal Killing vectors also lead
to invariance.] In order to study invariances of the interaction term in (4.1), we vary
separately zUaccording to 6zu = - f” [note the opposite sign from (1. l)], A,, according
to (1.6), but not T.

i$ (2 tr s dzfi A,T) = -2 tr j dx (*‘(a,f”) A, + k:“f” %A,> T + 2 tr 1 dzuf”F,,T

= -2 tr
I dX ad (f”A,T). (4.6)

The second equality follows from the first when (4.4) is used. The particle contribution
to the constant of motion follows from Noether’s theorem

C&W = (--A + 2 tr &N--f 9 + 2 tr(f u-4T)


= PLLf” (4.7)

and the total constant of motion [in the parametrization h = x”] is

Cf = j dr g1’2~&ueefu + p‘X, . (4.8)

We have here treated the gauge field as a dynamical variable; when it is an external
quantity whose prescribed form undergoes no variation, the change of the total
action, which is only the particle action, follows from (4.6):

~tJparme= -2 tr I dh -$ (f uA,T) - 2 tr I dz”f*F,,T.

For invariant gauge fields, we may use (1.7) and then (4.4) shows that (4.9a) is equal to

%particle= -2 tr
s
dX -$ (f”A, - Y,) T.

The constant of motion becomes

Cf = p,f u - 2tr y/,T. (4.10)

Just as in Section III, we can recognize, for spatial coordinate transformations, the
external gauge-field contribution to (4.10) as the “frozen” remnant of the dynamical
expression in (4.8). This follows from (3.8) when (4.3) is used for an evaluation of
gl/2p, in the parametrization X = x0.

gll”p = TS3(r - z). (4.11)


CONSERVATION LAWS IN GAUGE THEORIES 271

The two terms in (4.10) are separately gauge invariant. Moreover, the first involves
the velocity of the particle, and so can be regarded as the mechanical contribution. It
is therefore natural in the matter-field case, (3.9, to regard as mechanical the gauge-
invariant matter contribution to the total constant of motion [even though, unlike the
corresponding term in (4. lo), the field analog contains the gauge-potential].
Another form for the constant of motion may be given. When it is noted from (4.1)
that the canonical momentum is

I&, = -pu -I- 2 tr A,T, (4.12)

we see that (1.8) allows (4.10) to be presented as

Cf = --17,f@ + 2 tr W,T. (4.13)

In view of the geometrical and universal nature of W, the above non-gauge-invariant


decomposition is also useful. We call it the canonical decomposition with the first
term involving the canonical particle momentum. Analogously for the matter field
expression (3.10), we identify the first two terms in the parentheses of that formula
with the canonical matter current. [It must be remarked that in spite of the natural-
ness of the above two decompositions, there is an unavoidable ambiguity in spliting a
total conserved quantity into matter and field parts.]

V. DISCUSSION

We have shown that the quantities W and ?P, which initially arise in a formal,
mathematical context as the infinitesimal gauge transformations which must supple-
ment coordinate changes in order to exhibit the invariance of symmetric gauge fields,
also carry the direct physical significance of contributing to conserved quantities
associated with the symmetry. This provides another example of the familiar identity
between constants of motion and generators of transformations, and in special cases
our results have been encountered previously. We now discuss the known examples in
the context of our general theory.
It can happen that the W’s satisfying the consistency condition (1.13) are position
independent, and the symmetry compensating gauge transformation is global. In
that case (3.10) shows that the total constant of motion can be written as

(5.la)
where
T = s dr g1j2p (5.lb)

is the total isospin [or its generalizations]. This situation obtains for spatial rotations,
where a possible form for W, is just nata. Here fi = efgkn*rk, nG$ = 1, so II is just the

595h7b2
272 JACKIW AND MANTON

axis of rotation, and ta {a = 1,2,3} are generators of the gauge group forming an
SU(2) subalgebra. In that case (5.1) is the familiar formula that total angular momen-
tum is composed of orbital, spin, and isospin parts. The Wu-Yang monopole [7],
the ‘t Hooft-Polyakov [8] and Julia-Zee [9] solutions, and the Witten ansatz [lo] are
all examples of spherically symmetric configurations with constant W. The one-
instanton Euclidean solution [l l]

-2ia,,x~
A,, = (5.2a)
x2 + x2 ’
cPy = (1/4i)(&Y - I?&), (52b)
(p = t-h 0, 2 = (iq I) (5.2~)

provides a similar example. It is O(4) invariant with constant W [12].

(5.3a)
(5.3b)

However, constant W is not the generic case. There exist field configurations which
are spherically symmetric yet for no gauge choice are the w’s constant. Examples
have been constructed which are essentially non-Abelian as well as Abelian. The
former are not widely known [3]. The latter have been appreciated in the context of
Dirac monopoles for a long time. The magnetic monopole with magnetic field

B = g(r/rs) (5.4)

satisfies the symmetry criterion (1.7) for rotational invariance. That equation here
reduces to
(n~r)hB=-VY (5.5a)
and from (5.4) we find
Y =gIl*P. (5.5b)

Also it is easy to check that W is not constant. When this is considered in the context
of charged particle mechanics, discussed in Section IV, we recognize the Poincare [13]
result that the angular momentum for the system has a nonkinematical contribution

J=rhp-gi+. (5.6)

Notice also that the field contribution, the second term in (5.6), is due to the inter-
action of the electric field of the electrically charged particle with the magnetic field
of the monopole; a result first discussed by Saha and Wilson [14]. This illustrates the
need to include, as we did in (3.1 l), in addition to the symmetric background gauge-
potential, the perturbation a, due to the back reaction of the charged matter. Indeed
CONSERVATION LAWS IN GAUGE THEORIES 213

we see that the discussion of Eq. (3.12) et seq. is just a generalization to arbitrary
transformations of the Saha-Wilson observation.
For a final example of local rotations, we recall that the instanton (5.2) is invariant
under O(5) rotations, but this symmetry is realized with a position-dependent W.

f” = 2xux . c - c”(x2 + X2),


(5.7)
W, = 2in!a*c”Y .

This has been show in Ref. [12].

REFERENCES

1. Mathematicians studied symmetry properties of gauge fields over a quarter century ago; for
a discussion see S. KOBAYASHI AND K. NOMIZIJ, “Foundations of Differential Geometry,”
Interscience, New York, 1963. More recently physicists have looked at pieces of the problem.
See P. FORGACS AND N. MANTON (Comm. Math. Phys., in press) for a discussion and for refer-
ences to the relevant physics literature.
2. R. JACKIW, Phys. Rev. Lett. 41 (1979), 1635.
3. This observation is the starting point for various derivations of Higgs fields from Yang-Mills
fields; see N. MANTON, Nucl. Phys. B 158 (1979), 141.
4. In fact one really derives only the covariant gradient of (1.13), and the condition given in the
text is sufficient but not necessary. For an example that makes use of this distinction see L. BROWN
AND W. WEISBERGER, Nucl. Phys. B 157 (1979), 285.
5. An excellent summary is given by B. ZUMINO, in “Lectures on Elementary Particles and Quantum
Field Theory” (S. Deser, M. Grisaru, and H. Pendelton, Eds.), MIT Press, Cambridge, 1970.
6. S. WONG, Nuovo Cimento A 65 (1970), 689. This derivation is in flat space, but the generalization
to curved space is obvious.
7. T. T. Wu AND C. N. YANG, in “Properties of Matter under Unusual Conditions” (H. Mark and
S. Fernbach. Eds.), Interscience, New York, 1969.
8. G. ‘T HOOFT, Nucl. Phys. B 79 (1974), 276; A. POLYAKOV, Zh. Eksp. Teor. Fiz. Pis’ma Redaktsiyn
20 (1974), 430 [JETP Lett. 20 (1974), 1941.
9. B. JULIA AND A, ZEE, Phys. Rev. D 11 (1975), 2227.
10. E. WITTEN, Phys. Rev. Lett. 38 (1977), 121.
11. A. BELAVIN, A. POLYAKOV, A. SCHWARTZ, AND Y. TYUPKIN, Phys. Lett. B 59 (1975), 85.
12. R. JACKIW AND C. REBBI, Phys. Rev. D 14 (1976), 517.
13. H. POINCAR~, Compr. Rend. 123 (1896), 530.
14. M. SAHA, Indian J. Phys. 10 (1936), 145; H. WILSON, Phys. Rev. 75 (1949), 309.

You might also like