1. Introduction
Every consistent theory of interacting higher spin fields necessarily includes an infinite number of such fields. For this reason, it is extremely important to develop a formalism which effectively includes an infinite number of fields into a simpler field-theoretical object. This formalism should yield correct field equations first of all at the free level and then be promoted to an interacting theory. An elegant geometrical approach to higher spin theories of this kind is known as the method of tensorial spaces. This approach was first suggested by Fronsdal [
1]. Its explicit dynamical realization and further extensive developments have been carried out in [
2,
3,
4,
5,
6,
7,
8,
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24,
25,
26,
27,
28].
In a certain sense, the method of tensorial spaces is reminiscent of the Kaluza–Klein theories. In such theories, one usually considers massless field equations in higher dimensions and then, assuming that the extra dimensions are periodic (compact), one obtains a theory in lower dimensions, which contains fields with growing masses. In the method of tensorial (super)spaces, one also considers theories in multi-dimensional space–times, but in this case the extra dimensions are introduced in such a way that they generate the fields with higher spins instead of the fields with increasing masses. A main advantage of the formulation of the higher spin theories on extended tensorial (super) spaces is that one can combine curvatures of an infinite number of bosonic and fermionic higher spin fields into a single “master” (or “hyper”) scalar and spinor field which propagate through the tensorial supesrpaces (also called hyperspaces). The field equations in the tensorial spaces are invariant under the action of group, whereas the dimensions of the corresponding tensorial spaces are equal to . The case of four space–time dimensions is of particular interest since the approach of tensorial (super)spaces comprises all massless higher spin fields from zero to infinity. The free field equations are invariant under the group, which contains a four dimensional conformal group as a subgroup. In fact, the entire structure of the invariant formulation of the higher spin fields is a straightforward generalization of the conformally invariant formulation of the four-dimensional scalar and spinor fields. This allows one to use the experience and intuition gained from the usual conformal field theories for studying the dynamics of higher spin fields on flat and AdS backgrounds, and to construct their correlation functions.
Being intrinsically related to the unfolded formulation [
29,
30,
31,
32,
33] of higher-spin field theory, the hyperspace approach provides an extra and potentially powerful tool for studying higher spin AdS/CFT correspondence (for reviews on higher-spin holography, see, e.g., [
34,
35]). The origin of higher-spin holographic duality can be traced back to the work of Flato and Fronsdal [
36] who showed that the tensor product of single-particle states of a 3D massless conformal scalar and spinor fields (singletons) produces the tower of all single-particle representations of 4D massless fields whose spectrum matches that of 4D higher spin gauge theories. The hyperspace formulation provides an explicit field theoretical realization of the Flato-Fronsdal theorem in which higher spin fields are embedded in a single scalar and spinor fields, though propagating in hyperspace. The relevance of the unfolded and hyperspace formulation to the origin of holography has been pointed out in [
33]. In this interpretation, holographically dual theories share the same unfolded formulation in extended spaces which contains twistor-like variables and each of these theories corresponds to a different reduction, or “visualization”, of the same “master” theory.
In what follows, we will review main features and latest developments of the tensorial space approach, and associated generalized conformal theories. It is mainly based on Papers [
3,
8,
10,
13,
23,
24,
27]. We hope that this will be a useful complement to a number of available reviews on the higher-spin gauge theories which reflect other aspects and different approaches to the subject
Frame-like approach in higher-spin field theory [
37,
38,
39,
40,
41,
42].
Metric-like approach [
43,
44,
45,
46,
47,
48,
49,
50,
51,
52,
53,
54,
55].
Review that address the both approaches [
56].
Reviews which contain both the metric-like approach and the hyperspace approach [
59,
60].
A short review on the hyperspace approach [
61].
A short review that contains frame-like approach, hyperspaces and higher-spin holography [
62].
The review is organized as follows. In
Section 2 we introduce a general concept of flat hyperspaces. To this end we use somewhat heuristic argument, which includes a direct generalization of the famous twistor-like representation of a light-light momentum of a particle to higher dimensional tensorial spaces i.e., to hyperspaces. The basic fields in this set up are one bosonic and one fermionic hyperfield, which contain infinite sets of bosonic and fermionic field strengths of massless fields with spins ranging from zero to infinity. Physically interesting examples are hyperspaces associated with ordinary space–times of dimensions
and 10. In what follows, we will always keep in mind these physical cases, though from the geometric perspective the tensorial spaces of any dimension have the same properties.
We demonstrate in detail that the solutions of wave equations in hyperspace are generating functionals for higher spin fields. These equations are nothing but a set of free conformal higher spin equations in
and 10. The case of
describes only scalar and spinor fields, the case of
comprises the all massless bosonic and fermionic higher spin fields with spins from 0 to ∞ and the cases of
and
describe infinite sets of fields whose field strengths are self-dual multiforms. These fields carry unitary irreducible representations of the higher-dimensional conformal group and are sometimes called “spinning singletons” [
63].
We then describe a generalized conformal group which contains a convention conformal group as its subgroup (for and , respectively) and show how the coordinates in hyperspace and the hyperfields transform under these generalized conformal transformations.
In
Section 3, we consider an example of curved hyperspaces which are
group manifolds. An interesting property of these manifolds is that they are hyperspace generalizations of
spaces. Similarly to the
space which can be regarded as a coset space of the conformal group
, the
group manifold is a coset space of the generalized conformal group
. This results in the fact that the property of the conformal flatness of the
spaces (i.e., the existence of a basis in which the
metric is proportional to a flat metric) is also generalized to the case of hyperspaces. In particular, a metric on the
group manifold is flat up to a rotation of the
group, the property that we call “
–flatness”.
In
Section 4, we briefly discuss how the field equations given in the previous Sections can be obtained as a result of the quantization of (super)particle models on hyperspaces.
In
Section 5, we derive the field equations on
group manifolds. We show that the field equations on flat hyperspaces and
group manifolds can be transformed into each other by performing a generalized conformal rescaling of the hyperfields. We discus plane wave solutions on generalized
spaces and present a generalized conformal (i.e.,
) transformations of the hyperfields on the
group manifolds. In all these considerations, the property of
flatness plays a crucial role.
Section 6 describes a supersymmetric generalization of the construction considered in
Section 2 and
Section 7 deals with the supersymmetric generalization of the field theory on
introduced in
Section 3. The generalization is straightforward but nontrivial. Instead of hyperspace, we consider hyper-superspaces and instead of hyperfields we consider hyper-superfields. The generalized superconformal symmetry is the
supergroup and the generalized super-
spaces are
supergroup manifolds. We show that all the characteristic features of the hyperspaces and hyperfield equations are generalized to the supersymmetric case as well.
The direct analogy with usual
D-dimensional CFTs suggests a possibility of considering generalized conformal field theories in hyperspaces.
Section 8 and
Section 9 deal with such a theory which is based on the invariance of correlation functions under the generalized conformal group
. The technique used in these Sections is borrowed from usual
D-dimensional CFTs and the correlation functions are obtained via solving the generalized Ward identities in (super) hyperspaces.
In
Section 8, we derive
invariant two-, three- and four-point functions for scalar super-hyperfields. The correlation functions for component fields can be obtained by simply expanding the results in series of the powers of Garssmann coordinates. Therefore, we shall not consider the derivation of
invariant correlation functions for the component fields separately.
Finally, in
Section 9, we introduce generalized conserved currents and generalized stress-tensors. Their explicit forms and the transformation rules under
can be readily obtained from the free field equations and the transformation rules of the free hyperfields.
Further, we show how one can compute invariant correlation functions which involve the basic hyperfields together with higher rank tensors such as conserved currents and the generalized stress tensor. We show that the invariance itself does not impose any restriction on the generalized conformal dimensions of the basic hyperfields even if the conformal dimensions of the current and stress tensor remains canonical.
However, the further requirements of the conservation of the generalized current and generalized stress tensor fixes also the conformal dimensions of the basic hyperfields, implying that the generalized conformal theory will not allow for nontrivial interactions.
We briefly discuss possibilities of avoiding these restrictions by considering spontaneously broken symmetry or local invariance, which may lead to an interacting hyperfield theory.
Appendices contain some technical details such as conventions used in the review, a derivation of the field equations on group manifolds and some useful identities.
2. Flat Hyperspace
Let us formulate the basic idea behind the introduction of tensorial space. We shall mainly concentrate on a tensorial extension of four-dimensional Minkowski space–time. A generalization to higher dimensional and spaces will be given later in this Section.
Consider a four dimensional massless scalar field. Its light-like momentum
,
can be expressed via the Cartan–Penrose (twistor) representation as a bilinear combination of a commuting Weyl spinor
and its complex conjugate
Obviously, since the spinors are commuting, one has and therefore , where the spinor indices are raised and lowered with the unit antisymmetric tensors and .
In order to generalize this construction to higher dimensions note that one can equivalently rewrite Equation (
1) in terms of four-dimensional real Majorana spinors
(
)
Due to the Fierz identities
satisfied by the Dirac matrices
one has
. (The four-component spinor indices are raised and lowered by antisymmetric charge conjugation matrices
and
see the
Appendix A.) Let us note that since identities similar to (
3) hold also in
and 10, the Cartan–Penrose relation (
2) is valid in these dimensions as well.
Let us continue with the four-dimensional case. The momentum
is canonically conjugate to coordinates
. One can easily solve the quantum analogue of Equation (
1)
to obtain a plane wave solution for the massless scalar particle
or in terms of the Majorana spinors
with
being an arbitrary spinor function.
Let us now consider the equation
which looks like a straightforward generalization of (
1) and see its implications. A space–time described by the coordinates
(conjugate to
) is now ten-dimensional, since
is a
symmetric matrix. A basis of symmetric matrices is formed by the four Dirac matrices
and their six antisymmetric products
. In this basis,
has the following expansion
The analogue of the wave Equation (
4) is now
whose solution is
At this point, one might ask the question what is the meaning of Equation (
9) and of the extra coordinates
and
? As we shall see, the answer is that Equation (
9) is nothing else but Vasiliev’s unfolded equations for free massless higher-spin fields in four-dimensional Minkowski space–time [
29]. The wave function
depends on the coordinates
,
and
. While
parameterize the conventional four-dimensional Minkowski space–time, the coordinates
(and/or
) are associated with integer and half-integer spin degrees of freedom of four-dimensional fields with spin values ranging from zero to infinity.
2.1. Higher Spin Content of the Tensorial Space Equations
To demonstrate the above statement let us first Fourier transform the wave function (
10) into a conjugate representation with respect to the spinor variable
considered in [
4]
The function
obeys the equation
Let us expand the function
in series of the variables
and insert this expansion into the Equation (
12). Then one finds that all the components of
proportional to the higher powers of
are expressed in terms of two fields the scalar
and the spinor
. As a result of (
13), these fields satisfy the relations [
4]
The basic fields
and
depend on
and
. Let us now expand these fields in series of the tensorial coordinates
Each four-dimensional component field in this expansion is antisymmetric under the permutation of the indices
and
and is symmetric with respect to the permutation of the pairs
with
. In order to answer the question about the physical meaning of these fields, let us first consider the scalar field Equation (
14). Using the expression (
8) for the tensorial coordinates and four-dimensional
-matrix identities, one can decompose (
14) as follows
where
and
. The meaning of Equations (
18) is the following. The first equation is a Klein-Gordon equation. The second equation implies that the trace (with respect to the
Minkowski metric) of the tensor which comes with the
s-th power of
in the expansion (
14) is expressed via the second derivative of the tensor which comes with the
-th power of
. Therefore, traces are not independent degrees of freedom and the independent tensorial fields under consideration are effectively traceless. The third and fourth Equation in (
18) imply that the tensor fields satisfy the four-dimensional Bianchi identities, and the last equation implies that they are co–closed. These are equations for massless higher-spin fields written in terms of their curvatures
. In four dimensions these equations are conformally invariant. Therefore one can conclude that in the expansion (
16) the field
is a conformal scalar,
is the field strength of spin-1 Maxwell field, the field
is a linearized Riemann tensor for spin-2 graviton, etc.
The treatment of Equation (
15) which describes half-integer higher-spin fields in terms of corresponding curvatures is completely analogous to the bosonic one (
14). The independent equations for the conformal half-integer spin fields are
From (
19)–(
20) one can derive the equation
This equation describes the decomposition of the spinor-tensor
into the part which contains the
space–time derivative of
f and the “physical” part which is self-dual and gamma-traceless, i.e.,
Therefore, one can conclude that due to Equations (
19) and (
20) the field
in the expansion (
17) is a spin-
field, the field
corresponds to the field strength of the spin-
Rarita–Schwinger field, while the other fields are the field strengths of the half-integer conformal higher-spin fields in
.
Finally, let us define the hyperspaces associated with
and
space–time. The dynamics of the fields will be again determined by the equation (
7) with the corresponding hyperspaces and the twistor-like variables
defined as follows.
In
the twistor-like variable
is a 16–component Majorana–Weyl spinor. The gamma–matrices
and
form a basis of the symmetric
matrices, so the
tensorial manifold is parameterized by the coordinates
where
are associated with the coordinates of the
space–time, while the anti-self-dual coordinates
describe spin degrees of freedom.
The corresponding field Equations are again (
14) and (
15) and the entire discussion repeats as in the case of
. The crucial difference is that now the expansions (
16) and (
17) is performed in terms of the coordinates
. As a result one obtains a description of conformal fields whose curvatures are self-dual with respect to each set of indexes
. These traceless rank
tensors
are automatically irreducible under
due to the self-duality property, and are thus associated with the rectangular Young diagrams
which are made of five rows of equal length
s (“multi-five-forms”). The field equations, which are ten-dimensional analogues of the four-dimensional Equations (
18), can be found in [
13].
In
the commuting spinor
is a symplectic Majorana–Weyl spinor. The spinor index can be decomposed as follows
(
;
;
). The tensorial space coordinates
are decomposed into
where
, and
(
) provide a basis of
symmetric matrices, They are related to the usual
-group Pauli matrices
. The matrices
(where
) form a complete basis of
antisymmetric matrices with upper (lower) indices transforming under an (anti)chiral fundamental representation of the non-compact group
. For the space of
symmetric matrices with upper (lower) indices, a basis is provided by the set of self-dual and anti-self-dual matrices
and
, respectively,
The coordinates
are associated with
space–time, while the self-dual coordinates
describe spinning degrees of freedom.
The consideration proceeds as in the
and
case. Because of the form of the tensorial coordinates in (
24) the six-dimensional analogue of the expansions (
16) and (
17) contains powers of
. Corresponding field strengths, which again describe conformal fields in six dimensions, are self-dual with respect to each set of the indexes
. In other words, one has an infinite number of conformally invariant (self-dual) “multi-3-form” higher-spin fields in the six-dimensional space–time which form the
-dimensional representations of the group
.
In [
9,
16,
21] Equation (
12) has been generalized to include several commuting spinor variables
where
is a nondegenerate metric. The value of
r is called the “rank”. As we explained above, the free higher-spin fields in
are described by the rank-one equations in the ten-dimensional tensorial space. The higher-spin currents are fields of rank-two
. These currents obey the equations with off-diagonal
[
19]. The currents
are bilinear in the higher-spin gauge fields
and
, which obey the rank-one equation (
27)
.
On the other hand, when considering rank-two equations the corresponding tensorial space can be embedded in the higher-dimensional tensorial space. From the discussion above, it follows that a natural candidate for such higher-dimensional space is the tensorial extension of
space–time. In this way one effectively linearizes the problem since the conformal currents in four dimensions are identified with the fields in
[
21].
2.2. Four Dimensional Unfolded Higher-Spin Field Equations from the Hyperspace Field Equations
Let us rewrite, in the case of the
theory, the hyperspace relations in terms of the Weyl spinors. The momenta (
7) take the form
while Equation (
7) splits into
and
Equations (
29) relate the dependence of
on the coordinates
to its dependence on
. Thus, using this relation, one can regard the wave function
as the fundamental field.
The expansion of
in series of
and
is
where the reality of the wave function implies
, and by construction the spin-tensors are symmetric in the indices
and in
.
The consistency of (
30) implies the integrability conditions
We have thus obtained the equations of the Vasiliev’s unfolded formulation of free higher spin fields in terms of zero–forms. In this formulation the
component (a physical scalar),
and
components of the expansion (
31) correspond to the physical fields, while the other fields are auxiliary. The latter two fields are the self-dual and anti-self-dual components of the spin–
s field strength. The nontrivial equations on the dynamical fields are [
38] the Klein–Gordon equation for the spin zero scalar field
and the massless equations for spin
field strengths
which follow from (
32). All the components of
that depend on
both and
are auxiliary fields expressed by (
30) in terms of space–time derivatives of the dynamical fields contained in the analytic fields
and
and thus one arrives at the unfolded formulation of [
38].
Let us summarize what we have considered by now. To describe the dynamics of higher-spin fields in four dimensions we have introduced extended ten-dimensional tensorial space, hyperspace, parameterized by the coordinates
(
8). The main object is a generating functional for higher-spin fields described by
or by
. The generating functional depends on the tensorial coordinates
and on the commuting spinors
or
. The dynamics is described by the field Equation (
9) or (
12). To obtain from these the higher-spin field equations in the ordinary space–time parameterized by the coordinates
one can use two options. In the first approach one gets rid of the tensorial coordinates
and arrives at Vasiliev’s unfolded formulation in terms of the functional (
31). Alternatively, one can first get rid of the commuting spinor variables and arrive at the equations for the bosonic (
16) and fermionic (
17) hyperfields. Both pictures provide the equations for the field strengths of the higher-spin potentials, the difference being that these field strengths are realized either as tensors or spin-tensors.
2.3. Generalized Conformal Group
Let us consider in more detail the symmetries of Equation (
7) in which now the Greek indices
run from 1 to an arbitrary even integer
. However, as we explained in the previous Section, the physically interesting cases are associated with
and 16, which correspond to the number of space–time dimensions equal to
and 10, respectively.
It turns out that Equation (
7) is invariant under the transformations of the
group [
5,
8]
The constant parameters
and
correspond to the generators of generalized translations
, generalized Lorentz transformations and dilatations
(generated by the
algebra) and generalized conformal boosts
. The differential operator representation of these generators have the form
and
These symmetries are the hyperspace counterparts of the conventional Poincaré translations, Lorentz rotations, dilatations and conformal boosts of Minkowski space–time. The generalized Lorentz rotations are generated by the traceless operators
, forming the
–algebra, whereas dilatations are generated by the trace of
. The generators (
36), (
37) and (
38) form the
algebra which plays the role of a generalized conformal symmetry in the hyperspace
From the structure of this algebra, one can see that the flat hyperspace can be realized as a coset manifold associated with the translations where is the semi–direct product of the Abelian group generated by the generalized conformal boosts and the general linear group.
The generators of the translations, Lorentz rotations and conformal boosts of the conventional conformal group can be obtained from the generators as projections onto the x-space, for example , etc.
Let us note that the
algebra can be conveniently realized with the use of the twistor-like variables
and their conjugate
In the twistor representation the generators of the
group have the following form
Equations (
14) and (
15) are invariant under the
transformations (
35), provided that the fields transform as follows
Note that these variations contain the term
, implying that the fields have the canonical conformal weight
. A natural generalization of these transformations to fields of a generic conformal weight
is [
4]
3. Hyperspace Extension of AdS Spaces
A hyperspace extension of spaces is another coset of the group. Recall that the usual space can be realized as the coset space (Here, K and denote the generalized conformal boosts and dilatation, respectively.) parameterized by the coset element . The generators of the boosts can be singled out from the generators of the four dimensional conformal group by taking a linear combination of the generators of the Poincaré translations and conformal boosts as , where is the inverse of the radius.
Analogously, for the case of the hyperspace extension of the
space let us consider the generators
where
,
stands for the symmetric part of the
transformations
and
is the
-invariant symplectic metric. One can see that the corresponding manifold is an
group manifold [
8] which can be realized as a coset space
with the coset element
. Indeed, let us recall that
group is generated by
symmetric matrices
which form the algebra
As a group manifold,
is the coset
which has the isometry group
, the latter being the subgroup of
generated by
as one may see from the structure of the
algebra (
39). The generators
form the diagonal
subalgebra of
.
Let us note that, for the case of , i.e., for the case of four space–time dimensions, space is a coset subspace of of the maximal dimension. For , an space is also a subspace of manifold but is no longer the maximal coset of this group.
3.1. GL-Flatness of Group Manifolds
Let us describe a property of
-flatness of the
group manifolds which is a generalization of the conformal flatness property of
spaces. By
-flatness we mean that, in a local coordinate basis associated with
, the corresponding
Cartan form
has the form
with the matrix
being
This expression implies that the Cartan form is obtained from the flat differential by a specific rotation of the latter.
This property can be demonstrated by showing that the Cartan forms (
49) satisfy the
-group Maurer–Cartan equations (see [
8,
23], for technical details)
The matrix
inverse to (
50) depends linearly on
and has a very simple form
Note that the possibility of representing the Cartan forms in the form (
49) is a particular feature of the
group manifold since, in general, it is not possible to decompose the components of the Cartan form into a “direct product” of components of some matrix
.
3.2. An Explicit Form of the Metric
Let us now demonstrate that, for the case of
(
), the pure
-dependent part of the matrix
indeed generates the metric on
in a specific parameterization. To this end, we should evaluate the expression
where the dependence of the matrices
on the coordinates
(see Equation (
8)) was discarded, i.e.,
. Denoting
and, using the explicit form (
50) of
, one obtains
In this way, we obtain a four-dimensional space vierbein and spin-connection
The corresponding metric is
It is well-known (see also
Section 5.1) that the metric on
can be represented as an embedding in a flat
-dimensional space
via the embedding constraint
Choosing the embedding coordinates for
to be
one readily recovers the metric (
58), with the parameter
being related to the
radius
r as follows
Finally, computing the Riemann tensor
and the Ricci scalar
one verifies that the metric (
58) indeed corresponds to a space with constant negative curvature, i.e., the
space.
4. Particles in Hyperspaces
In this Section, we would like to explain the physical meaning of the tensorial space coordinates as spin degrees of freedom from the perspective of the dynamics of a particle in hyperspace.
Historically, the first dynamical system in which the Fronsdal hyperspace proposal for higher–spin fields was realized explicitly was the twistor-like superparticle model of Bandos and Lukierski [
2] which, for
, possesses the generalized superconformal symmetry under
. The original motivation behind this model was a geometric interpretation of commuting tensorial charges in an extended supersymmetry algebra. Its higher–spin content was found later in [
3,
64] where the quantum states of the superparticle were shown to form an infinite tower of massless higher–spin fields, and the relation of this model to the unfolded formulation was assumed. This relation was analyzed in detail in [
4,
5,
8,
10,
13]. In addition to the relation to higher spins, the model of Bandos and Lukierski [
2] has revealed other interesting features, such as the invariance under supersymmetry with tensorial charges (which are usually associated with brane solutions of Superstring and M–Theory). Moreover, it has provided the first example of a dynamical BPS system preserving more than half of the bulk supersymmetries. BPS states preserving
supersymmetries (with
for
) were then shown to be building blocks of any BPS states, and this led to a natural conjecture that they can be elementary constituents or “preons” of M–theory [
65].
Let us consider the generic case of a particle moving in an
-invariant hyperspace
described by the action
where
are the hyperspace coordinates of the particle. The auxiliary commuting variables
is a real spinor with respect to
and a vector with respect to
(introduced in
Section 2). Finally
is the pull–back on the particle worldline of the hyperspace vielbein. For flat hyperspace
and for the case of the
group manifold
where
is an
Cartan form. The latter can be taken in the
-flat realization as in (
49). The dynamics of particles on the
supergroup manifolds was considered for
in [
8,
10,
66] and for generic values of
N in [
4,
5], and, as we have already mentioned, the twistor-like superparticle in the
super-hyperspace was considered in [
67] as a point-like model for BPS preons [
65], the hypothetical
-supersymmetric constituents of M-theory.
The action (
65) is manifestly invariant under global
transformations and
implicitly invariant under global
transformations, acting linearly on
and non-linearly on
. Thus, the model possesses the symmetry that Fronsdal proposed as an underlying symmetry of higher–spin field theory in the case
,
[
1]. To make the
invariance manifest, it is convenient to rewrite the action (
65) in a twistor form (for simplicity we consider the flat case (
66))
where
and
form a linear representation of
Hence, the bilinear form
is manifestly
invariant. Note that, as it follows from the action (
68), the variables
and
are canonically conjugate coordinates and momenta of the particle. Upon quantization, they become the operators introduced in
Section 2.3, Equation (
40).
Using the relation (
69) one can easily recover the
transformation (
35) of
.
Applying the Hamiltonian analysis to the particle model described by (
65) and (
66), one finds that the momentum conjugate to
is related to the twistor-like variable
via the constraint
As we have already mentioned, this expression, e.g., in the case
for which
is given in (
8), is the direct analog and the generalization of the Cartan–Penrose (twistor) relation for the particle momentum
. A difference is that in
the Penrose twistor relation is invariant under the phase transformation
or in the two–component Weyl spinor notation
, while Equation (
72) does not possess this symmetry. rather the symmetry of the model is
(
) subgroup of
and as a result in the model under consideration the phase component
of
is a dynamical degree of freedom. It turns out that upon quantization it is associated with the infinite number of massless quantum states (particles) with increasing spin (helicity). This is in contrast to the conventional twistor-like (super) particle models with a finite number of quantum states, considered e.g., in [
68,
69,
70,
71,
72,
73,
74,
75,
76,
77,
78,
79].
To understand the physical meaning of the phase
, let us notice that Equation (
72) is a constraint on possible values of the canonical momenta of the particle in the hyperspace. In the case
the Majorana spinor
has four independent components. One of these components can be associated with the phase
. The momentum
of the particle along the four conventional Minkowski directions
of the hyperspace (
8) is light-like. Therefore,
depends on three components of
. It does not depend on the phase
of
, since it is invariant under the phase transformation (
73). The momentum
of the particle along the six additional tensorial directions
is not invariant under the phase transformations and, hence, depends on the four components of
. However, we have already associated three of them with the light-like momentum
in
. Therefore, the only independent component of the momentum
is associated with the
phase
of
, and as a result the motion of the particle along the six tensorial directions
is highly constrained. This means that, effectively, the particle moves in the four-dimensional Minkowski space and along a single direction in the six additional dimensions whose coordinate is conjugate to the compact momentum–space direction parameterized by the periodic phase
. As shown in [
3,
64], the coordinate conjugate to the compactified momentum
takes, upon quantization, an infinite set of integer and half-integer values associated with the helicities of higher–spin fields. The half-integer and integer–spin states are distinguished by the discrete symmetry
(
).
The resulting infinite tower of discrete higher–spin states can be regarded [
3,
64] as an alternative to the Kaluza–Klein compactification mechanism akin to Fronsdal’s original proposal. In contrast to the conventional Kaluza–Klein theory, in the hyperspace particle model, the compactification occurs in momentum space and not in coordinate space. The phase
in (
73) can be regarded as a compactified component of the momentum (
72), while the corresponding conjugate hyperspace coordinate is quantized and labels the discrete values of spin of fields in the effective conventional space–time.
As we have already seen by virtue of the Fierz identity (
3) the twistor particle momentum is light-like (
) in
and 10. Therefore, in the hyperspaces corresponding to these space–time dimensions the first–quantized particles are massless [
2,
3,
64]. Moreover, since the model is invariant under the generalized conformal group
, the quantum states of this particle in the hyperspaces containing the
and 10 Minkowski spaces as subspaces correspond to the conformal higher–spin fields introduced in
Section 2.
Let us conclude this section with a brief comment on the model describing a particle propagating on the
group manifold. Its action has the form (
65), with the corresponding Cartan form given by (
67). The property of
-flatness greatly simplifies the analysis of this case. Namely, since the Cartan forms of the
group manifold and the flat hyperspace are related as in Equation (
49), one can simply reduce the classical
action to the flat one by redefining the spinor variables as follows
. However, when quantizing this system we should work with variables that appropriately describe the geometry of the
background in which the particle propagates. Thus upon quantization one gets Equation (
92) as explained in detail in [
10].
8. Generalized CFT. Part I. Correlation Functions in -Invariant Models
In the previous sections, we have described the generalized conformal group and generalized conformal supergroup . We introduced the fundamental fields and superfields and showed how they transform under generalized conformal transformations.
In this Section we shall construct two-, three- and four-point correlation functions of these fields, by requiring the
symmetry of the correlators, i.e., by solving the corresponding Ward identities. In other words we will follow the conventional approach adopted in multidimensional CFTs (see e.g., [
113]). In particular, we will consider
invariant correlation functions from which the
invariant correlation functions can be recovered as components of the expansions of the former in series of the Grassman coordinates
.
-invariant correlation functions in the tensorial spaces have been studied in [
11,
23,
24,
27] and in the unfolded formulation in [
114].
8.1. Two-Point Functions
Let us denote the two-point correlation function by
The invariance under supersymmetry transformation generated by the operators
Q, Equation (
120), requires that
which implies
where
is the interval between two points in hyper–superspace which is invariant under the rigid supersymmetry transformations (
117).
We next require the invariance of the correlator under the
S-supersymmetry (
125)
which is solved by
The two-point function (
171) reproduces the correlators of the component bosonic and fermionic hyperfields
and
after the expansion of the former in powers of the Grassmann coordinates
. Since on the mass shell the superfield (
132) has only two non-zero components, all terms in the
-expansion of the two-point function (
171), starting from the ones quadratic in
, should vanish. This is indeed the case, as a consequence of the field equations.
To see this, let us recall that in the separated points the two-point function of the bosonic hyperfield of weight
satisfies the free field equation. Therefore for
one has (when the two points coincide, one can define an analog of the Dirac delta-function in the tensorial spaces, see [
5] for the relevant discussion)
Similarly, for
the fermionic two-point function satisfies the free field equation for the fermionic hyperfield. Written in terms of the superfields, these equations are encoded in the superfield equation (for
)
Expanding the two-point function
in powers of the Grassmann variables
one may see that the terms in the expansion starting from
vanish due to the free field Equation (
172). From Equations (
171) and (
174) and from the explicit form of the superfield (
132), one may immediately reproduce the correlation functions for the component fields [
11]
The two-point functions on the
manifold may now be obtained from (
171) via the rescaling (
157), which relates the superfields in flat superspace and on the
group manifold
Finally, as in the
case, one may derive the superconformally invariant two-point function for superfields carrying an arbitrary generalized conformal weight
, which on flat hyper superspace has the form
8.2. Three-Point Functions
The three-point functions for the superfields with arbitrary generalized conformal dimensions
,
may be computed in a way similar to the two-point functions using the superconformal Ward identities. The invariance under
Q–supersymmetry implies that they depend on the superinvariant intervals
, i.e.,
where
Invariance under
S–supersymmetry then fixes the form of the function
W to be
Let us note that the three-point function is not annihilated by the operator entering the free equations of motion (
131) for generic values of the generalized conformal dimensions, including the case in which the values of all the generalized conformal dimensions are canonical
Again, the three-point functions on the supergroup manifold
can be obtained via the Weyl rescaling (
157), as in the case of the two-point functions (
177)
8.3. Four-Point Functions
Finally, let us consider, first in flat hyper superspace, the correlation function of four real scalar superfields with arbitrary generalized conformal dimensions,
(
)
Invariance under
Q–supersymmetry again implies that the correlation function depends only on the superinvariant intervals
(
181). Following the analogy with conventional conformal field theory we find
with
W being an arbitrary function of the cross-ratios
subject to the crossing symmetry constraints
Furthermore, the
’s are constrained by the invariance of the four-point function under the
S–supersymmetry to satisfy
Similar to the case of two- and three-point functions, the four-point function of the scalar superfields on
can be obtained from (
186) via the Weyl re-scaling (
157).
8.4. An Example. Superconformal Models
As we mentioned earlier, the case of is the simplest example of “hyperspace” which in this case coincides with the three-dimensional space time itself, and the fundamental fields are just the scalar and the two-component spinor . All known results for three-dimensional (super)conformal theories are reproduced from the above generic formulas restricted to the case of and , as we will show on the example of superconformal two– and three-point functions.
The superconformally invariant two- and three-point correlation functions of the
,
scalar supermultiplet model have been constructed in [
115].
Let us use the spinor–tensor representation for the description of the three-dimensional space–time coordinates
where now
are
spinorial indices and
is the vectorial one. Since (
190) provides a representation of the symmetric
matrices
, no extra coordinates, like
, are present and, hence, no higher-spin fields.
The inverse matrix of (
190),
takes the simple form
We may now consider a real scalar superfield in
with
being a physical scalar,
a physical fermion and
an auxiliary field.
If (
193) satisfies the free equation of motion (
131), which in the
case reduces to
This equation implies that on the mass shell the auxiliary field
vanishes, the scalar field
satisfies the massless Klein–Gordon equation and
satisfies the massless Dirac equation. The field Equation (
194) is superconformally invariant if the superfield
has the canonical conformal weight
.
Let us consider a superconformal transformation of (
193). The Poincaré supersymmetry transformations of
are
They encode the supersymmetry transformations of the component fields
where we have made use of the identity
Under conformal supersymmetry,
transforms as follows
where
is the conformal weight of the superfield. The superconformal transformations of the component fields are
The conformal weights of , and F are , and , respectively.
As we have already seen, the two-point function for a superfield of an arbitrary noncannonical dimension has the form (
178). Expanding the expression on the right hand side of (
178) in powers of
, we obtain
Using the identities
and
one may rewrite the expression (
204) as
Thus, from Equation (
204) or (
207), one may immediately read off the expressions for the correlation functions of the component fields of the superfield (
193)
Let us note that when the superfield
has the canonical conformal dimension
, due to the identity
the last term in (
204) is proportional to the
–function if one moves to the Euclidean signature. Then, one has for the two-point function for the auxiliary field
Note that the correlation functions of the auxiliary field F with the physical fields and with itself (for ) vanish.
On the other hand, if the conformal weight of the superfield (
193) is anomalous, i.e.,
, the correlators of the auxiliary field with the physical ones still vanish (in agreement with the fact that their conformal weights are different), but the
correlator is
This situation may correspond to an interacting quantum
superconformal field theory [
116], where the auxiliary field is non-zero, and fields acquire anomalous dimensions due to quantum corrections.
The consideration of three-point functions is analogous. Using the expression for the three-point function (
182) and expanding it in series of the
variables, we get for the component fields whose labels of scaling dimension we skip for simplicity
The remaining three-point functions containing an odd number of fermions, as well as the correlator , vanish. Note that, dimensional arguments would allow for a non-zero correlator, but supersymmetry forces it to vanish. The correlator is zero as well, since it is proportional to
Moreover, from the above expressions we see that superconformal symmetry does not fix the values of the scaling dimensions
. This indicates that quantum operators may acquire anomalous dimensions and the quantum
,
superconformal theory of scalar superfields can be non-trivial, in agreement e.g., with the results of [
116].
If the value of
were restricted by superconformal symmetry to its canonical value and no anomalous dimensions were allowed (for all the operators which are not protected by supersymmetry) one would conclude that the conformal fixed point is that of the free theory. This is the case, for instance, for the
,
Wess-Zumino model in which the chirality of
matter multiplets and their three-point functions restricts the scaling dimensions of the chiral scalar supermultiplets to be canonical. This implies that in the conformal fixed point the coupling constant is zero, i.e., the theory is free [
117,
118].
10. Conclusions
The idea to formulate higher-spin theories in an extended (super) space, where extra coordinates generate higher spins (by analogy with the Kaluza–Klein theories where compact extra dimensions generate “higher masses”) seems to be very attractive, especially taking into account a level of complexity of higher-spin theories formulated in an ordinary space–time.
The underlying symmetry of this formulation is the group, which contains the corresponding D-dimensional conformal group as a subgroup. This allows one to borrow, for the analysis of the -invariant systems, an intuition and techniques from conventional Conformal Field Theories.
To summarize, the reviewed appraoch generalizes familiar concepts to higher-dimensional tensorial spaces and the correspondence looks schematically as follows
Space–time coordinates are extended to tensorial coordinates .
Cartan–Penrose relation gets extended to the hyperspace twistor-like relation which determines free dynamics of fields in the tensorial space with the momentum conjugate to .
space is extended to the group manifold.
Conformal scalar and conformal spinor become the “hyperscalar” and the “hyperspinor” .
D-dimensional conformal group is extended to the group which underlies the Generalized Conformal Field Theory of the fields and .
We have shown that the hyperspace approach describes (in
and 10) free dynamics of an infinite set of massless conformal higher-spin fields in an elegant compact form. An important and non-trivial problem is to find a non-linear generalization of this formulation which would correspond to an interacting higher-spin theory. This problem has been addressed by several authors. As we have seen, it is related to the necessity to break the
symmetry in an appropriate way. Attempts to construct such a generalization in the framework of hyperspace supergravity and a non-linear realization of the
supergroup were undertaken, respectively, in [
12,
14]. Obstacles encountered in these papers may be related to the fact that their constructions utilized only higher-spin field strengths but did not include couplings to higher-spin gauge potentials, while the consistent formulation of nonlinear equations of massless higher-spin fields contains both [
37,
38,
39]. Therefore, to successfully address the problem of interactions it is important to incorporate higher-spin potentials in the hyperspace approach, e.g., by further elaborating on the construction of [
16].
Another issue, which can be related to the previous one, is a question of consistent breaking
symmetry. The manifestation of this breaking was observed e.g. in higher-spin current interactions [
26]. As we have seen in
Section 9, when considering generalized CFT based on global
invariance (see [
27]), the requirement of generalized current conservation turns out to be too strong to allow for the basic hyperfields to have anomalous conformal dimensions and again points at the necessity to (spontaneously) break
invariance.
Theories with spontaneously broken symmetry might be also useful for studying massive higher-spin fields in hyperspaces. A consideration of theories with local invariance i.e., some sort of generalized gravity is yet another interesting and widely unexplored area.
Finally, let us mention that field Equations (
14) and (
15) for the fields in hyperspaces remind (a part of) weak section conditions of exceptional field theories (see [
132] for a review and references). This similarity can be relevant for higher-spin extensions of these theories, provided the section conditions can be properly relaxed (see e.g., [
133,
134] for a discussion of this point). It would be interesting to further elaborate on this issue, as a connection to the
framework [
18].