Flato 1970

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

ANNALS OF PHYSICS: 61, 78-97 (1970)

Conformal Covariance of Field Equations

MOSHE FLATO, JACQUESSIMON, AND DANIEL STERNHEIMER

Physique MathPmatique, Cokge de France, Paris, France

ReceivedApril 21, 1970

A rigorousdefinitionof conformalcovarianceof field equationsin first and second


quantizedfield theoriesis given here.The generalimplicationsof thesedefinitionsas
well as the necessary and sticient conditionsfor Poincan5plusdilatationcovariance
to imply conformalcovariance are then discussed. Familiesof existingfield equations
are then divided into two different seriesaccordingto the relationsbetweenthe con-
formal degreeof the field and its spin.In addition,remarksare madewhich connect
conformalhehaviourof field equationswith theoriesof stronginteractions.

I. INTRODUCTION

In recent years the conformal group has gained much interest and extensive
study in the literature. Many interesting ideas in quantum field theory as well as
new concepts in strong interaction physics were formulated and examined with
the help of the conformal group.
Why is this group so attractive to the theoretical physicist? Considered as
a space-time symmetry group, and in the most orthodox way, this group has a
striking advantage: It is the most general group which (beyond singularities)
leaves invariant the light cone. This last fact led physicists to try using this group as
a possible generalization of the PoincarC group-the symmetry group of the special
theory of relativity.
An immediate difficulty arisesif one takes this last point of view [I]: It can be
proved that the group of all one-to-one applications of the four dimensional
Minkowski space into itself which conserve the causal order of pairs of vectors
in the space, is the semidirect product of the positive dilatations by the orthochron-
ous Poincare group, often called the Weyl group.
However [2], one can utilize the conformal group (which contains the Weyl
group) without violating “physical causality”: It is simple to realize that to any
bounded region M containing the origin in Minkowski space there corresponds
78
CONFORMAL COVARIANCE 79

a neighborhood of identity V, in the conformal group such that V,,., acts on M


causally. In other words, one can say that the Lie algebra of the conformal group
is causal. Globally one can only state (a fact which is also quite sufficient for physical
applications) that beyond singularities the conformal group (which can then change
the time ordering of events) transforms any two physically connectable events in
space-time (viz., two events having time-like separation) into two other physically
connectable events.
In particle physics this group has gained much interest, not only because of the
simple fact which was mentioned above, but because of many other valuable and
fundamental features of this group, some of which we mention here:
(a) It has long been known that the de-Sitter SO(4, 1) group, which contains
the proper symmetry group of the hydrogen atom, can be utilized as a spectrum-
generating group for the hydrogen atom. The search for a slightly bigger group
playing a similar role for the hydrogen atom (in a “better” way) led [3] naturally
to the study of special unitary representations of SO(4,2), which is locally
isomorphic to the conformal group.
(b) In hadron physics this group was also considered [4] as a group acting on
infinite supermultiplets of hadrons, or equivalently, acting on the indices of infinite-
component fields (viz., fields transforming under infinite-dimensional representa-
tions of the homogeneous Lorentz group). In other words, some of the self-adjoint
generators of the unitary irreducible representations of this group were utilized in
constructing wave equations for infinite-component fields, generalizing the Dirac
equation.
Evidently this last approach can be easily written in current-algebra terms.
The only reasons for utilizing this group in this last context are those of simplicity
(this group is a low-dimensional simple group which is a candidate for such
purposes) and of a vague analogy with the hydrogen atom case.
(c) Structurally speaking, this group contains the SU(2, 1) subgroup which
a priori is a noncompact alternative to SU(3) [5]. Moreover, the maximal compact
subgroup of SU(2, 1) is U(2), which is the invariance group of the strong interac-
tions (isotopic spin plus hypercharge). The Lie algebra of SU(2, 2) is a unification
of the Lie algebras of Poincare and of SU(2,l) [6].
The suggested interpretation was, therefore, the following: The Poincare sub-
group of the conformal group, with its usual unitary irreducible representations,
describes the relativistic kinematics of free hadrons. The SU(2, 1) subgroup, with
its representations, describes the internal degrees of freedom and the multiplet
structure of these particles. The spectral unification SU(2,2) itself has only one
important aim: To describe the mass levels of hadron spectroscopy.
Indeed, in Ref. [7] a mass formula in a local (infinite dimensional) representation
80 FLATO, SIMON, AND STERNHEIMER

was found. However this representation was not integrable on the Poincare
subgroup and therefore the spectrum of the energy-momentum vector was
discrete.
This difficulty was overcome in Ref. [8], in which a local representation is con-
structed of a contraction algebra of SU(2, 2), giving rise to a discrete mass spectrum,
and integrable to the Poincare subgroup.
But let us come back to field theory and to our first point of view on the conformal
group.
The fact that this group can in a sense be considered as a “good” space-time
group brings us to examine rigorously the covariance of various wave equations,
Lagrangians, and action integrals under this group. To do so, one should of course
first define rigorously what is meant by covariance under this group in field theory.
Once this is done, one can classify various field equations with respect to their
covariance type and then one can try to deduce some consequences.
It has always been mentioned that, from the above point of view, only equations
describing the motion of zero-mass particles are distinguished, as they are the
only ones which are covariant under the conformal group. Indeed, as the conformal
Lie algebra contains among its generators, in addition to the Poincare generators,
that of homogeneous dilatations and four generators of local (point-dependent)
dilatations, only zero mass will be invariant under such dilatations (the mass
having the dimension of [cm]-‘).
However, it is trivial to remark that, if one interpretes the action of the five extra
generators as just the possibility of changing the unit scale in a space-time dependent
manner (a kind of passive transformation), no difficulty is caused for massive
particle equations as far as conformal (‘extended covariance” is considered. This
very evident point of view in classical field theory, which requires only redefining
the mass in the transformed equation, finds a quantum field counterpart in
the present paper.
If one sticks to the orthodox interpretation, and wants to put the accent on
zero-mass particles in conformal theories, one can of course argue that in reactions
with very high momentum transfer one can expect approximate conformal co-
variance to prevail, even for massive particles, since in such situations the mass of
the hadrons is presumably negligible.
If one adopts this generally accepted point of view, then one can look upon the
usual relativistic quantum field theory as a very badly broken conformal covariant
(or, for simplicity, Weyl covariant) field theory. Such an approach can give, among
other things, a semitheoretical connection between the amount of symmetry-
breaking on one hand and the observed hadronic mass spectrum on the other hand.
Irrelevant to which approach one is trying to push, we shall study in what follows
some general aspects of the conformal group considered as a space-time covariance
group in classical and quantum field theories.
CONFORMAL. COVARIANCE 81

2. CONFORMAL COVARIANCE OF FIELD EQUATIONS

A. Definition and General Properties


(Finite) covariance is usually defined in the following way: If A(+,@) = 0 is a
field equation, we say it is Poincare covariant if when x E M (Minkowski space)
transforms into x’ = P. x (P E B, Poincare group), then Z/J(X) cotransforms
according to a finite-dimensional representation S of SL(2, C), #‘(x’) = S(A) #(x)
[where we denote P = (Ll, b), n E SO,(3, l), b E M, and &(1” -+ (1 in the homo-
morphism (l/2, l/2): X(2, C) --f SO,,(3, I)], so that the equation keeps the same
form (A(Y) #(x’) = 0) in the new coordinate system.
This notion can be (and has been) extended in three possible directions:

(a) Consider more general representations, e.g., an infinite-dimensional uni-


tary representation S of SL(2, C). This gives the so-called infinite-component field
theories, which are known to have some unpleasant features.
(b) Introduce more (or the most) general transformations on fields which are
compatible with Poincare covariance on M. This new approach [9] seems promising.
(c) Consider a more general group acting on M. We shall study this aspect
of the problem here, dealing with the identity component %?of the conformal
group.

DEFINITION. If x -+ x’ = t . x is a conformal transformation, we say that the


equation A(L) #(x) = 0 is conformally covariant if it goes into A@‘) +‘(x’) = 0
when z,Gcotransforms according to $J’(x’) = S(t, x) #(x), where S(t, x) is some
matrix dependent on the transformation t and on the point in space-time. In order
that this definition be consistent with the group structure of % we must have,
whenever both sides are defined (nonsingular),

S(tt’, x) = S(t, t’x) S(t’, x). (2.1)

Compatibility with Poincare covariance requires that whenever we restrict ourselves


to Poincare transformations x ---f x’ = P . x = (A, b) . x, we have

S(P, x) = S(A). (2.2)

[This already implies that S must in fact be related to the conformal group via some
covering group, viz., SU(2,2), and that S must depend on the point x, since it is
trivial on the translation subgroup and conformal group is simple.]
We shall see that these properties determine S(t, x) almost completely. More
exactly, if we write (as in Ref. [lo]) Q(a, x) = X-l(1 + 2ax + a2x2), QkJ =

595/61/1-6
82 FLATO, SIMON, AND STERNHEIMER

exp(kJlog Q), where .I is the matrix (‘I!1 -I),,I of the order of S and I1 is the identity
matrix, and recall that a real irreducible representation is said to be absolutely
irreducible if its complex extension is still irreducible, we have the following
entirely geometrical result:

THEOREM. Let S be anyfinite-dimensional real or complex irreducible representa-


tion of SL(2, C) (depending on which linear combinations, real or complex, of the
field componentswe allow). Denote by Ct the (algebraic) set of singularpoints of the
conformal transformation t. Then any real or complex (accordingly) matrix function
S(t, x) satisfying (2.1) and (2.2) and analytic in t in someneighbourhoodof B in V
-and thus analytic in x in the complex Minkowski space M, outside of singularities
-is given, for special conformal transformations

t = (A, a) : x --t x’ = sZ;l(a, x)[x + ax2],

S(X, a; x) = Q,(a, x)n+*v &(a, x), (2.3)

which determines its vaiues for any t E V by (2.1) and (2.2). Here n E R, and is
called the conformal degree of S, y = i(i2 = -1) in the complex case, y = J in
the real case,k E R, and k = 0 in the real absolutely irreducible case;&(a, x) is the
uniquely defined analytic (for x in the connectedcomponentof the origin in M - Ct ,
and therefore also outside singularities in M,) unimodular solution of some matrix
d#erential system &!$,(a, x) = NU(a, x) &(a, x) with initial value &(a, 0) = Z,
the identity matrix.
First we note that since S(A) = S(A) S(e, x), where e denotes the identity of V,
S(e, x) = I. More generally, S(T, , x) = I, where T, denotes the translation
x + x + b. Therefore (still denoting by h the multiplication by X E R+) we have
by (2.1) that S(& x + b) = S(X, x) and thus is independent of x, whence
S@, x) S(4 = W) m x >, since LI and X commute. Therefore S(A, x) = f(A)1
in the absolutely irreducible case (with real- or complex-valued f according to the
real or complex character of S) and S(A, x) = f,(h)Z + f2@)J in the remaining case.
If we setf(A) = fi(X) + if2(X) in the latter, we see from (2. 1) that we must have in
all cases the functional equation f(N) = f(A)f(x’), the only analytic nonzero
solution of which is f(A) = A-(n+ik) (n, k E R), whence

S(X, x) = h-(n+kv) * 1 (2.4)

Now if we denote by A, = exp(a@A,) the special conformal transformation


x -+ In,(a, x)--l (x + ax”) and for simplicity S(A, , x) = S(a, x), consider transla-
CONFORMAL COVARIANCE 83

tions Tb = exp(PP,J and Lorentz transformations A, = exp(&UVM,,,), and denote


by D the infinitesimal generator of the dilatation A, we have

A,TbA-, = exp(b*P, + uLLbV[AU, PJ + &zUuvba[A,, , [A,, PJ])


= (1 - 2ubD)(l + b”P,,)(l + 2a”b’MJ(l + &‘A, - 2(ub) u”A,)
+ O(l b I”)
(where 1b I2 = x 1b, I2 while ub = uub,). But we have
S(u, x + b) = S(A,T, , x) = S(A,T,A-, , AJ) S(u, x).
Thus if we write A,, = (&S’(A, , x)/&‘~)~~ and J$ = (&S(u, X)/&P),,, , the
fact that @S(X, x)/~A),=, = -(n + ky) Z implies that
S(u, x + b) = (1 + 2(n + ky) ub + 2~“b”~~~ + (u2by - 2(ub)a”) dv(Aax))
x S(u, 4 + 00 b I”>.
Thus S(u, X) is differentiable with respect to x and we have
(as@, xyaxq f a,qu, x)
= [2(n + ky)uol + 2d‘Jif,, + u2dm(A, . X) - 2u,y~‘~~(A,x)] S(a, x).

Therefore we have also, by derivation with respect to a and taking a = 0, since


W, ,x1 = 1,

Now S(A, , x * x) = S(Aha , x) whence (since S is continuous with respect to x


in 0) S(A, , 0) = Z and thus -91’,(O)= 0. Therefore
J3$(x) = 2.M,,xa + 2(n + ky) x, I.
Remembering that
(Aa * x)” = Li’I(a, x)-l (x’ + @‘x2),
the equation for S then becomes
w(a, x) = C(n+ f%wJwl + mG 41 e, 4
where Sz = D,(u, x) and SN,(u, x) = @‘Au0 + (2a,u”A,, + u2u/ia,)(Aa * x)“. We
now write
S(u, x) = S(u, x)“+h ’ &)(a, x), (2.5)
which, together with (2.4), will give us the relation (2.3) of the theorem. The matrix
function S,, must therefore satisfy the system
%%4 xl = N,(u, 4 &I(4 4 (2.6)
84 FLATO, SIMON, AND STERNHEIMER

with initial value &(a, 0) = I, and an equation similar to (2.1), since we also have
J&z + a’, X) = J&z, A,z) $I@‘, x). One may check that the Frobenius integrabil-
ity condition

is satisfied. Moreover, since the matrices iV, are traceless and we have, from (2.6),

a,(det S,,) = (Tr iV,)(det S,,) = 0,

it follows that (at least in M,, , the connected component of 0 in A4 - C,) S, is


unimodular (and that det S, is constant in every other connected component).
Further, the solution of (2.6) with the given initial value is unique (at least in M,,)
since if S, , S,’ are two solutions, then writing S,’ = S&?S,,’ we get a,(S;?S,‘) = 0.
The existence of such a solution can be proved, and an expression for it given,
with the help of (a four dimensional extension of) the notion of the multiplicative
integral [ 1I]

So@,4 = 1%(I+ N,(u, x) dX@),


0

the integral being taken along any path from 0 to x (in MO), e.g., a path composed
of segments parallel to the coordinate axes (which enables us to define So as a
product of ordinary matrizers of the kind c:, (I + A$ dxl) with x,’ < x1 < x1”),
the Frobenius condition enabling us to pass from one path to another.
Moreover, we see that 2(u2xU + au) a,S, = (&.Q)(a,So) = 0, which means that
So is constant along lines “orthogonal” (for the hyperbolic metric) to the hyper-
surfaces Q = const.
The singular points for the transformation A, (and for the calculus developed
here) are on the surface J&z, x) = 0, i.e., the cone (x + u/$)~ = 0 if a2 f 0, and
the plane ax = - Q if u2 = 0. The cone splits A4 into three connected components
(including MO), while the plane gives only two. But if we consider the algebraic
surface B = 0 in the complex space M, (eight-dimensional on R), we see that it is
of real dimension six and its complement is here connected. This enables us to
define arbitrary powers of LI (and log a), as well as So , on all M - Ct from their
definition on MO (where Q > 0). We can do this by analytic continuation to a
connected (and simply connected) Riemann surface defined, e.g., via a cut in i&f,
which rests on the surface Ct : Q = 0, the choice of the cut fixing uniquely the
definition of these functions.
The solution So thus obtained will also depend analytically on the four-parameter
a, and a second-order calculation gives
S,(u, x) = I + 2d‘~‘~~~ + 4~“a”(.N~,~,~ - gm.Mve) x”xp + O(l a j3). (2.7)
CONFORMAL. COVARIANCE 85

The extension of S, , Q, and S to general t E V is straightforward, using a factoriza-


tion of t and formulas (2.1) and (2.2).
Remark. In the real absolutely irreducible case, i.e., S = (j, j), the reality
condition of S imposes k = 0, i.e., global dilatations will cotransform the fields
only through multiplication by A”, II E R (and similarly for local dilatations). Since
the power of X will be related to the dimensionality of the fields 4 considered, it
seems natural to eliminate the phase Aik also in the general complex case S = (j, j’),
i.e., to impose k = 0.
Now it is evident from (2.9, (2.6), and the expression for N, that the real matrices
S(t, x) corresponding to the scalar restriction (real irreducible) (j, j’) @ (j’, j)
of the representation (j, j’) (with j + j’) are just the scalar restrictions of the
matrices S’(t, X) corresponding to (j, j’), the scalar restriction of the complex
matrix A + iB being defined as the real matrix (<-,“). The k’s and the IZ’Sappearing
in the two cases are then the same; this will force us (though L?fJ is unimodular) to
impose k = 0 also in the case (j, j’) @ (j’, j). The only physical covariance
matrices S are therefore given by

S@, a; 4 = %(a, xl” s&h 4, with real n.

THEOREM. For pure spin representations (0, j) @ (j, 0) and (j/2, j/2), the con-
formal degreen is related to the spin j via the relation

(-1)s” = (-I)@

if we impose the mild assumptionthat the extended S defined by (2.1) and (2.2) is
real in M - C, and has an analytic extension to complex Minkowski space.
Indeed (0, j) is the symmetrized tensor product @j(O, l/2) of 2 j times the basic
representation (0, l/2), and similarly the contragredient (j, 0) is @(l/2,0), while
(j, j) = (j, 0) @ (0, j). Equations (2.6) and the expression for N,, show us that a
similar decomposition holds for the general S(t, x), using those related with the
representations (4, 0) and <o, 4). If we then characterize the S = (0, 4) represen-
tation by the relation S*(A) all,!@) = f’l:u”, where u@ are Pauli matrices, then
one can see [12] that in MO one has

&(a, x) = Q(a, x)-li2 (I + ~~~~(3~0~)

and therefore the unique analytic solution of (2.6) is given, for (0, j) @ (j, 0), by
&(a, x) = Q(u, x)-j [G>“j(Z + upx,&7”) 0 @“j(Z + ~~x,3W)*-~]. (2.8)
Whichever cut is chosen in M, , @I2 will be pure imaginary in any connected
component of M - Ct, where Q < 0. Comparing (2.8) with (2.3), and since the
86 FLATO, SIMON, AND STERNHEIMER

quantity between square brackets in (2.8) is real, we see that if we want S(u, X) to
be real in M - C1 we must have (- I)n-j E R. The same result is obtained in the
case (j/2, j/2), whence the theorem.
This implies that n - j is rational, with odd denominator. If we further require
that UN determinations of S (and not only one) be real in M - Ct , then n - j must
be an integer; under this (quite natural) assumption, we have n E 4Z and n is an
integer or half-integer together with the spin.
Remarks. (1) The reality of S for (j, j’) @ (j’, j) (j f j’) and the similar
construction of S, from the basic matrices show us that in this case also n E i&Z,
whence n E 4Z also for (j, j’), since the (real) matrices S of a scalar restriction are
the scalar restrictions of the complex matrices S.
(2) In the case of a tensor field, one form (which is unique up to equivalence)
of S, is determined by the coordinate transformations [lo]

so = .Q(a,x)(*-P)($)” (fg-)”

for a field p times covariant and q times contravariant.


B. Covariance in Second-Quantized Theories
1. All that has been said in (A) is still valid in second-quantized theories,
in which the 4’s are considered (as usual) as operator-valued distributions, i.e.,
weakly continuous mappings from some space of infinitely-differentiable test
functions (Yon M (rapidly decreasing, 9, or with compact support, 9) to a linear
space Op(H) 0 E, where E is some finite-dimensional space and Op(H) is a space
of (in general unbounded) operators having some common invariant dense domain
in the Hilbert space H of states.

2. If U is some representation of V on a Hilbert space H, then for t = A,


the dilatation, and P2 = U(PpPu), we have U(t-l) P”U(t) = h-2P2 = A-lP2X-l.
Now on the Hilbert space H = L2(M) we can define a (spin 0, irreducible) unitary
representation of V on H by (U(t)v)(x) = Q,(t-‘, x)-” y(t-lx). The unitarity follows
from the fact that d4(t, X) = P4(t, X) d4x. This gives for the infinitesimal gener-
ators,
lJ(-Mu,) = X& - xva, , lJ(-P,,) = a,, U(-D) = x8 + 2
(2.9)
U(--A,) = x9, - 2x,(xa + 2).

[Due to the definition of U, the generators give the opposite Lie algebra from that
obtained through the geometrical action of V by the Cartan calculus; the opposite
group representation U’ can be obtained by writing the group action on the right,
CONFORMAL COVARIANCE 87

and thus defining (U’(t) * v)(x) = Q(t, x)-” v&t), where x -+ xt through t E V:
the generators obtained are the opposite of (2.9) and satisfy the usual commutation
relations (cf., e.g., Ref. [12], in which S = --D); the final results of this part would
not be changed should the opposite group be represented.]
A long but straightforward calculation shows that, defining A, = exp(@A,), we
have in this representation

U(A_,) P”U(A,) = SZ(a, x) P2Q(a, x),

so that more generally for any t E V,

U(t-1) P”U(t) = i&t, x) mqt, x). (2.10)

Incidentally, for the given representation U, we obtain from (2.10)

Q(t’, tx) = U(t-1) x2@‘, x) U(t).

3. According to our definition of conformal covariance, we cotransform


the field # into #‘(x’) = S(t, X) #(x) when x -+ x’ = tx. Therefore we have
[SO,xl ~Wl(4 = #’(B1,w here 6~is defined by the equality (of differential forms)
a(x) dx = a(t-lx) d(t-lx) = Q(t-I, x)-” c&-lx) dx. (2.11)

Now given any unitary representation U of %?on the Hilbert space H of states, we
define y’ = U(t-l)q~ and write the covariance condition for states and fields,

(dJ(4 * 99’ = $04 * 9)‘. (2.12)

This gives us (the integral being considered formally, as usual)

U(F) #(a) U(t) = #‘(a) = 1 S(t, t-lx) &t-lx) a(x) dx,

whence

U(t) I&X) U(P) = j S(t-l, x) #(x) a(x) dx. (2.13)

This can be written formally as

U(t) #(x) U(t-1) = s(t, x)-l #(tx), (2.14)

and is, when limited to B, the exact formulation of the second Wightman axiom.
This justifies the definition of v’ and the covariance condition (2.12), which can
be considered as a preformulation of the axiom. [In the case where we write
on the right the group action on M: x + xt = x’, we should of course define
88 FLATO, SIMON, AND STERNHEIMER

q’ = V’(t)q+ where V’ is any given unitary representation of V, and we would then


also obtain V’(t) #(x) V’(t-l) = S(t, x)-l #(a).].

4. Two remarks should be made here. First, one cannot have here,
owing to the restriction to 9, a pure Wightman theory in a straightforward manner,
since the spectrality condition cannot be obtained by the restriction from V to 9
(the spectrum of P2 cannot contain a positive mass without containing a continuum
of all positive masses). A more refined interpretation of global and local dilatations
can, however, be given (using the behaviour of the constant field P”) which gives a
reasonable physical interpretation to this mathematical disease.
Second, we notice that the change of variable in the integral (x to tx) used to
derive (2.13) is legitimate only when t is nonsingular. Moreover, in the right side
of (2.14) we have the factor SZ+; therefore, in order to obtain (2.13) we must check
that L+(t, x) a(x) is still a test function; the factor S,, will also give a similar condi-
tion (due to relations like (2.8)) with, eventually, other powers of L?. Therefore,
01 should vanish for Q(t, x) = 0 and relation (2.14) will only make sense when
applied to such test functions, or rather test forms, a(x) dx. (Since tx = co for
52 = 0, this condition is quite natural.) This may lead to some difficulties, especially
concerning the topology of the spaces of functions or distributions, if a global
formulation is desired.

5. The factor Q-” in (2.14) is related to the fact that our fields have dimen-
sion. To make its interpretation clearer, let us consider the (constant) squared
mass field P”(x) : a(x) -+ P2 s a(x) dx, where P2 = V(P,,Pu), V being some unitary
representation of V. We then have

P”(tx) = S(t, x) v(t) P”(x) V(t-1) = @(t, x) V(t) P”(x) V(t-l),

since here n = 2 (mass is of dimension [cm]-‘).


If t = (A) (or “more generally,” t belongs to the Poincare group plus dilatations),
we know that V(t) P2V(t-l) = h2P2, and Q(t, x) = h-l. Therefore we obtain
P”(tx) = P”(x), which is the quantized expression of the (passive) invariance of the
mass field under (e.g.) dilatations. For general conformal transformations, we
have only the quantized analogue of the classical relation ml2 = Q2m2, viz.,
P”(tx) = Q2(t, x) V(t) P”(x) V(t-‘). However, if V is (e.g.) the representation defined
in 2, denoting by y the variable of the functions in H = L2(M), we have

P”(tx) = SZyt, x) qt-1, y) P”(x) qt-1, y),

which means, acting on f E L2(M),

j- P”(t.4 44 &fW = W-l, Y) (I Q2(t> 4 P”(x) 44 dx) Q(t-‘, y)f(y).


CONFORMAL COVARIANCE 89

This means that if we identify tx with y and neglect the noncommutativity of Q


with P2, we get Pz(tx) = P”(x) in this case also.

3. SCALE AND CONFORMAL COVARIANCE, AND SPECIAL FIELD EQUATIONS

We have just seen that, if PoincarC covariance extends to conformal covariance,


then the fields must necessarily cotransform under the special conformal transfor-
mations (A, a) according to

where S,, is unimodular (and n E +Z, for real Lorentz representations at least).
The transformation law of ds2, given by dsf2 = JZ2 ds2, leads us to identify -n
with the sum of the physical dimensions, in centimetres and seconds, of the field.
Physical quantities should transform according to their physical dimension. We
shall thus postulate, e.g., that e2/fic is invariant under V?,that mass will transform
according to m + m’ = Q(t, x)m when x -+ x’ = tx, etc. The transformation law
of mass (similar to what we have seen in the quantized case) will moreover enable
us to give a sense to conformal “extended covariance” of the equations for massive
particles (and not only zero mass particles, as is usual).
We shall now investigate under which conditions a Lagrangian field equation is
conformally covariant. We shall restrict ourselves, in stating the theorem, to
Lagrangians Z(I,!I, a,#, m) depending on one field #, its first derivatives, and a
mass. The more general situation, in which we have several fields (and eventually
coupling constants), is a straightforward extension of our case (we must not forget
to transform the coupling constants according to their dimensions); this remark
should be kept in mind, especially when the conjugate If of a field occurs together
with the field $ itself in the Lagrangian.

THEOREM. Let 5?(1,4,a,$, m) be a local Lagrangian (i.e., the value of which


depends on/y on the values of the field and its first derivatives at one point) which is
PoincarP invariant, i.e., when x --f x’ = Ax + b, (A, b) E 8, and $(x) + #(x’) =
S(x) $(x), where S is an irreducible representation of SL(2, C), then

=W’, 42, mLs = -W#, 44, m>, . (3.1)


Suppose, moreover, that 64 is scalar covariant, i.e., for all h > 0 there exists an n
such that 2 behaves like an energy density, i.e., defining +‘(x’) = X-n#(x) for
x’ = Ax, and m’ = A-lrn,

LZ’(#‘, a,‘#‘, m’),, = Z(XF%,b, h-(n+l)aulCf, A-lm), = X-45?(~, a,$, m), . (3.2)
90 FLATO, SIMON, AND STERNHEIMER

Then the field equations derived from 8 will be conformally covariant, with

?I44 = Q%, -4 w, 4 $44 f or x’ = tx, t E V, if and only if61R, d4x = 0,

where
R, = ((w,v + ~JW a~/WP). (3.3)

Remarks. (1) For several fields &, we have to introduce their degrees n, and
sum over k to obtain R, ; in particular, 5 is associated with 4.
(2) We have seen that when x + x’ = (A, a)x, then the field must cotrans-
form according to #‘(x’) = Q,P(a, x) &(a, x) 4(x): (3.2) requires n = p, which we
assumed in the theorem.
To prove the theorem, we shall first consider only infinitesimal transformations
A = 1 + E, au intmitesimal, (1 = 1 + 4~’ A&, and thus S(/T) = I + &py-aiBy.
In infinitesimal form (first-order development), (3.1) and (3.2) give

=W - n4(I + W”~J+, (1 - (n + 1)4(Z + B@~&(% - ~“a,)#, (1 - +m),


- (1 - 4~) -WA ad, 4, . (3.4)
Under a conformal transformation (A, a), a first-order development of which can
be written as

x’~ - (1 - a,xp + B)(x@ + (a”xV - aJ”)x”),

the field cotransforms (to first order, and with the help of (2.7)) as

yY(x’) - (1 - nE + 2nad;p)(I + 2a”x”.4!f,,) Jl(x),


and the Lagrangian becomes, writing Z(x) = 9(+, a,#, m), ,

9(x’) - P((l - ne + 2na+P)(I + 2a”x’Af&, (1 - (n + 1)~ + 2(n + 1) a&p)


x (I+ -k~~w+d(a, - ew+ (1 - +& + (@a, + 2~.4LJW
x wc7.m -EpW ad, m>zp
where in this case @’ = @‘(a, x) = 2(a%” - &xu) and where we denote by p
the i-th component of the field #.
Therefore, from (3.4) and the locality of the Lagrangian,

9’(x’) - (1 - 4~ + 8@x,) S(x) + 2auR,,


where R, is given by (3.3), or
Y(x’) - &(a, x)” S(x) + 2a@R, . (3.5)
CONFORMAL COVARIANCE 91

The field equations will thus be conformally covariant, with conformal degree n,
if and only if, given any domain A’ C M (on which the inverse transformation
x’ -+ x is nonsingular), when #’ varies so that S#’ = 0 on the boundary of A’, we
have S J,,, A?‘&‘) d?x’ = 0.
This variation of t,P corresponds to some similar variation of # on some domain
A C M. Now from (3.9, writing for simplicity 9, = Y(x’) for x’ = x, = A,x =
Q(a, x)-l (x + ux2), and denoting by T,(a) the operator

we have Pa,,, = (1 + T,(a) da“) -Lpa, whence we have the system


wD,/aau = T,(U) .Le,,

the (unique, and known a priozi) solution of which is given by the (operational)
multiplicative integral 9, = c (1 + T,(u) duU) 9, , with 2Z0 = P(x), and is
defined (e.g.) as

where TV = (au/j) T,(ku/j), k = l,..., j. Since S acts as a derivation, and remem-


bering that dx’ = Q-* dx, it is easily seen by induction that if 6 J-9 dx = 0 and
6 J R, dx = 0, then 6 Jrri dx = 0 for all i. Therefore (provided that the special
conformal transformation A, is nonsingular in d, and of course that Y depends
smoothly enough on $ in A) we shall also have 6 j,,, 9(x’) dx’ = 0. The extension
to general transformations of V is straightforward.
Conversely, from (3.9, it is also clear that covariance of field equations under V
implies 8 JR, dx = 0, whence the theorem.
Remarks. (a) If R, = 0, the conformal group V is a symmetry group of the
Lagrangian (and therefore of the action integral and the field equations) and con-
versely.
(b) If R, is, a gradient, R, = aUR, Q is a symmetry group of the action
integral J 8 dx (assuming the fields vanish rapidly enough at infinity) and thus
also of the field equations, and conversely.
(c) If we have only 6 JR,, dx = 0, 59 is just a symmetry group of the field
equations.

EXAMPLES. We shall give here some Lagrangian field equations, specifying in


each case the conformal degree n of the field and classifying them according to the
criterion stated in the above remark. We shall omit the (often tedious) calculations
and refer for more details to Refs. [lo] and [13]. It should be mentioned that, in
92 FLATO, SIMON, AND STERNHEIMER

all these cases, n is found to be necessarily real (this difference from the case in which
only global dilatations occur, and the imaginary part of n is not restricted, is essen-
tially due to the fact that, contrary to A”, Sz” does not commute with derivatives).

(a) R, = 0.
(1) The Dirac equation (~“a, + m)# = 0. The Lagrangian is 9 = $(+l, + m)#
and the conformal degree n = 312.
(2) One may check that the continuity equation for currents a,Ju = 0 is con-
formally covariant if and only if n = 3. In particular, R, = 0 implies (under the
hypotheses of the theorem) conformal covariance of the Lagrangian, which in turn
implies n = 3 for the associated conserved currents.
(3) The Maxwell equations written auFuY = 0 and &‘“~auFw = 0 are con-
formally covariant, the field F@” having the conformal degree n = 2. The Lorentz
gauge condition on the vector potential will not be conformally covariant (because
of the last example) but can be replaced by a nonlinear covariant condition such
as B,(A,A”A~) = 0, the A,, having (as expected) the conformal degree n = 1.
(4). The first-order, zero-mass Weinberg equations 114, 151, that write
a~lF>Yl”‘pj”~ = 0 for integral spin j, and aa~@“‘~~j = 0 for half-integral spin j
(Dirac-Fierz equations). In both cases, we find [14] n = j + 1.

(b) R, = a,R f 0.
(5) The real field Klein-Gordon equation (!J + m”)# = 0. We have

L? = g+ya,ga,* - &m2*2, n = 1, and R = $,b2.

(c) 6 JR, dx = 0, but R, is not a gradient.


(6) Second-order spin 1 Weinberg equation [16] y”“a,a,# + m2# = 0.
dip = aUi,@yay~ + m2$,h, n = 1, and R, = -ay$yy,,t,b - t,i&p~.

These examples show that the necessary and sufficient condition obtained is
indeed more general than the sufficient conditions obtained earlier by Gross and
Wess [ 171 (R, = i3,R) or by Mack and Salam [18] (R, = 0).

Remark. For non-Lagrangian field equations, their conformal covariance has


to be checked directly. The main example is that of higher-order (spin j > 1)
Weinberg equations with 2(2j + 1) components [16],
CONFORMAL COVARIANCE 93

We find [13] that they are conformally covariant, but with conformal degree
rr = 2 - j (this formula generalizes the results obtained for j = 4 and j = 1).

4. CONCLUSION

We conclude our article with the following remarks.

1. We have proved, under fairly general conditions, that (- 1)2n = (- 1)2j,


where n stands for the conformal degree of the field and j is its spin. Evidently, to
prove this, we assumed n to be any real number. This fact further implied that n is a
half-integer or integer together with j. In other words, n is connected with the statis-
tics which characterize the field. Therefore (- 1)2” = (- l)B+L, where B and L are,
respectively, the baryon and lepton numbers carried by the field. 12is thus connected
with internal (probably superselection) quantum numbers.
We talk now about the two a priori possible cases: that of relativistic field
equations which are not conformally covariant; and that of most of the existing
relativistic theories, the case of conformal covariance.
In the first case, the noninvariance is generally due to the existence of lower spin
components in the equations, which bring other nonphysical features into the
theory. (In the Rarita-Schwinger case this last remark is evident, since the field
transforms under the (3, 8) @ ((+, 0) @ (0, 4)) rep resentation of the Lorentz group.)
In the second case, the equations can be classified into series according to the
type of dependence of n on j. It seems that for all physically acceptable field
equations the relation n = j + 1 holds (in the case j = 1, y1= 2 will be the con-
formal degree of the tensor field Fgy , the vector potential A, itself having conformal
degree n = 1). The only known case of other series of field equations is the 2j-order
Weinberg equations which are (for j > 1) non-Lagrangian conformally covariant
field equations, of conformal degree n = 2 - j. This “anomalous” behaviour has
its counterpart in the nonphysical properties of these equations. Indeed it is known
that, when coupled to an external (electromagnetic) field, most of the solutions will
be physically nonacceptable. For gauge-type conditions (like generalizations of
the Lorentz condition, etc.) other series are still possible.

2. In the demonstration of the formula (- 1)2” = (- 1)2j we have assumed


a priori that II is real. In the free-field equations that we have treated we have
obtained from covariance principles that IZmust be real. However, one can imagine
situations in which n can be complex with nonrestricted imaginary part:
(a) If one considers only global dilatations, viz., only covariance under the
Weyl group instead of the conformal group;
94 FLATO, SIMON, AND STERNHEIMER

(b) Field equations involving interactions of a certain type.


In such cases, n can be chosen to be complex, and one can try to interpret its
imaginary part as the electric charge, lepton or baryon number, etc., and to
speculate further on this arbitrariness.

3. Special attention must be paid to gauge-invariant theories. To illustrate


our point of view, we discuss briefly the case of the Maxwell equations. In this case,
there exists a small difficulty with the gauge condition. If we maintain the usual
Lorentz gauge a,AU = 0, this condition is not conformally covariant: A necessary
condition for it to be conformally covariant is that A,, will be of conformal degree
three, while the conformal covariance of the Maxwell equations forces A,, to be of
conformal degree R = 1. One can, in the transformed equations, find another
representative in the gauge equivalence class which will still satisfy the Lorentz
condition. However, the dependence of the new vector potential on the conformal
transformation (as well as on the initial vector potential) is very complicated and
by no means manifests the group character of the conformal transformations.

4. Some of the authors mentioned in this article have discussed subjects


similar to the ones discussed here in terms of currents instead of emphasizing the
notion of fields, as we do here.
It should be emphasized that although the current approach has become very
popular in recent years through the striking successesof current algebra in particle
physics, as far as conformal covariance is concerned there is no particular reason to
prefer the current approach to the field approach.
Experimentally speaking, there is a hope of finding conformally invariant
amplitudes. Recently, it was found that for high-energy, large momentum-transfer,
inelastic hadron-lepton scattering, there is some evidence for the existence of a
scale invariant background. Therefore physicists expect [ 171 to study the asymptotic
behaviour of these amplitudes by imposing conformal covariance on them.
It should be mentioned that although the latter point of view seems extremely
promising, it seems to us that in this context the real problem already lies in the
scale invariance of amplitudes, conformal invariance playing no additional striking
role, and that the difficulty actually lies (as will be explained in remark 8) not in
constructing scale-invariant amplitudes but rather in defining operationally a
scale-invariant jield theory.

5. Some particular cases of our general theorem, which specifies


under what necessary and sutTicient conditions the Lagrangian, the action integral,
and the field equations are conformally covariant, were treated previously [ 17, 181.
These authors utilize the fact that in most field theories the divergences of currents
CONFORM& COVARIANCE 95

generated by scale and conformal generators are related (through their relation to
the trace of the symmetric energy-momentum tensor). We did not utilize this fact
in proving our general theorem, since we think that this is a quite general geo-
metrical theorem which is independent of the class of dynamical theories which is
treated.

6. As far as strong interactions are concerned, we have already discussed


in the Introduction the applications of the conformal group to their “internal
behaviour.” If we want to apply to strong interactions the conformal group from
the point of view discussed in this article, we have to remark the following:
Although conformal covariance (and especially scale invariance) seems to play
an important role in high-energy inelastic hadron-lepton scattering and in asymp-
totic quantum electrodynamics, etc., it seems that the usual strong interactions are
highly conformal (and even scale) noninvariant.
Therefore, if one insists upon applying our concepts to strong interactions, one
will be obliged to relate scale (or conformal) symmetry breaking to the internal
symmetry of strong interactions so as to be able to predict physically measurable
quantities.

7. We have already seen that in all cases that were treated by us it was
possible formally to extend the notion of conformal covariance from zero-mass
particles to massive particles, just by substituting (in classical theories) m’ = Szm,
a fact which corresponds to a space-time dependent choice of the scale of length
and time.
In second-quantized field theories the situation is very similar. We know that in
any representation of the conformal Lie algebra, e-aDPPeaD = e-2aP2. Moreover,
we have proved that for the particular representation of the conformal Lie algebra
given by (2.9), we have exp(--(a”& + X0)) P2 exp(u”d, + ho) = QP2Q. This
means that even for second-quantized theories with massive particles, we can in a
way generalize the concept of conformal covariance.
One should first remark that, at least as far as the particular representation
considered above is concerned, we can, without restricting generality, argue only
on the Weyl group instead of V, as the situations are analogous.
Now as for the Weyl group, the unitary irreducible representations are character-
ized by one value of the spin W2/P2 (W, is the usual Pauli-Lubanski vector) and
give a continuum (or zero only) for the spectrum of P2. If we say that each unitary
irreducible representation of the Weyl group characterizes Oozeparticle, the mass of
which is measured in all possible scales, then we can choose the “unit particle”
characterized by 01= 0, and other equivalent representations (transformed from
the former by the inner automorphism defined by e-uD) will describe other particles
with the same spin having mass “measured” by the value of e-OL.
96 FLATO, SIMON, AND STERNHEIMER

However, this interpretation is not very popular in the literature. In this connec-
tion, as well as for the problem of under what condition for the second-quantized
theory Weyl covariance will imply conformal covariance, it will be quite interesting
to find all those unitary irreducible continuous representations of the conformal
group which, when restricted to its Weyl subgroup, remain irreducible. Until now,
we have found,only a partial answer to this question (which includes most, if not all,
physically relevant cases). [Note added in proof: in a recent preprint, J. Mickelsson
and J. Niederle, from Prague, have answered this problem, suggested earlier by
one of us (M. F.)].
For zero-mass particles it was conjectured [19], and proved later [20], that the
most degenerate discrete series of unitary irreducible representations of the
conformal group (constructed, e.g., as ladder representations) remain irreducible
under the Poincare subgroup, thus characterizing a zero-mass particle. This last
fact is quite important for physical applications.

8. We end our article with some suggestions of yet unexplored, interesting


problems concerning the conformal group, which we mention briefly.
(a) To give an operational description of a dilatation covariant field theory.
Though from the geometrical point of view it is very clear what we mean by dilata-
tion and conformal covariance, things are much more obscure as far as dynamics
are concerned. As a matter of fact, if we consider asymptotic quantum electro-
dynamics and suppose the electron has a zero rest mass, we can expect to have a
situation with dilatation invariance. However in this limit we run into the infrared
catastrophy, and to remedy it one has to introduce a cutoff in the theory, which in
turn breaks the scale invariance. It seems that an operational definition of dilatation
(and therefore conformal) invariance is not an easy matter to achieve, and that this
difficulty is closely connected with other typical fundamental difficulties that one
has in quantum field theory.
(b) In connection with the existence of unitary irreducible representations
(the most degenerate ones) of the conformal group which are irreducible under the
Poincare subgroup (therefore describing a zero-mass particle) it will be interesting
to construct in some situations mentioned before a corresponding conformally
covariant S matrix.
(c) To utilize extensively the conformal group for renormalization theory, as
well as for the study of the Goldstone phenomenon.
(d) To study the most general (not necessarilyfree) field equations covariant
under the conformal group, and in casesin which it exists to give a physical meaning
to the unrestricted imaginary part of the complex conformal degree of the various
fields.
CONFORMAL COVARIANCE 97

(e) To utilize conformal symmetry for the problem of localization of zero-


mass particles, as the PoincarC-covariant theory does not seem to give an unam-
biguous answer to this problem.
(f) Finally, as we know that many contractions of conformal group (in the
sense of Segal and Wigner-Inonii) are important for physical applications, it seems
interesting to study these contractions as well as their unitary irreducible represen-
tations and the local (nonintegrable) ones.
We hope to be able to answer, at least partially, some of these problems in
future articles.

REFERENCES

1. E. C. ZEEMAN, J. Marh. Phys. 5 (1964), 490.


2. M. FLATO AND D. STERNHEIMER, C. R. Acad. Sci. 263 (1966), 935.
3. Y. NAMBU, Progr. Theor. Phys. Suppl. 37, 38 (1966), 368; A. 0. BARUT AND H. KLEINERT,
Phys. Rev. 156 (1967), 1541.
4. A. 0. BARUT AND H. KLEINERT, Phys. Rev. 161 (1967), 1464.
5. D. BOHM, M. FLATO, et al., Nuovo Cimento 38 (1965), 1941.
6. M. FLATO AND D. STERNHEIMER, J. Math. Phys. 7 (1966), 1932.
7. M. FLATO AND D. STERNHEIMER, Phys. Rev. Lett. 16 (1966), 1185.
8. M. FLATO AND D. STERNHEIMER, Commun. Math. Phys. 12 (1969), 296.
9. M. FLATO, P. HILLION, AND D. STERNHEIMER, C. R. Acad. Sci. 264 (1967), 82; M. FLATO AND
P. HILLION, Phys. Rev. D 1 (1970), 1667.
10. M. FLATO, P. HILLION, AND J. SIMON, C. R. Acad. Sci. 268 (1969), 1065.
11. F. R. GANTMACHER, “Application of the Theory of Matrices” (Translated from Russian),
Interscience, New York, 1959, and references quoted there in.
12. J. WESS, Nuovo Cimento, 18 (1960), 1086.
13. M. FLATO AND J. SIMON, C. R. Acad. Sci. 269 (1969), 546.
14. F. BAYEN, C. R. Acad. Sci. 269 (1969), 1051; Nuovo Cimento (1970).
15. S. WEINBERG, Phys. Rev. B 134 (1964), 882; 138, (1965), 988.
16. S. WEINBERG, Phys. Rev. B 133 (1964), 1318.
17. D. J. GROSS AND J. WESS, Preprint TH 1076, CERN (Sept. 1969).
18. G. MACK AND A. SALAM, Ann. Phys. (New York) 53 (1969), 174.
19. D. STERNHEIMER, J. Math. Pures Appl. 47 (1968), 289.
20. G. MACK AND I. TODOROV, J. Math. Phys. 10 (1969), 2078.

595/61/1-7

You might also like