We now aim at developing a tensor calculus which bears to Sgravity and its diffeomorphism invariance an analogous relation as Einstein gravity in form has to its non-manifest diffeomorphism invariance. For example, the standard tensor calculus is designed such that the Einstein tensor is a symmetric type tensor, although less manifestly so in form. The counterpart of the Einstein tensor in Sgravity does not transform like a conventional tensor. Yet, on account of the full diffeomorphism invariance of the underlying action some adapted notion of an ‘Stensor’ ought to exist.
4.1. Limit Construction of Stensors
We return to the intertwining relations (A48) of the metric frame. As exemplified in (
A52) and (
A53) one can use them to determine the nonlinear transformation law of tensor components with respect to the
,
orthonormal frame. We now consider the
scaling limit of this construction.
Stensors of weight zero. On a foliated pseudo-Riemannian manifold consider the
tuple
, obtained by taking the metric frame components of a conventional type
tensor
U and assume the
to be invariant under
from (
13). The limit (
4) then always exists with zero weights,
. The
tuple with the transformation law
,
, imprinted by the limit is called an Stensor of type
and
weight zero.
In pseudo-Riemannian geometry conventional tensors that are independent of the spacetime metric give rise to Stensors of zero weight. However the metric-independence is not a necessary condition, only the
-invariance of the
with respect to the fiducial foliation matters. Under these conditions the existence of a limit with
in (
4) can easily be seen from the intertwining relations (A48) of the metric frame. As long as the metric frame components
of the conventional tensor considered are
invariant, the only nontrivial response in
to the scaling (
13) is the replacement of
with
in the intertwining relations (A48). In the limit
the
-dependent terms in (A48) drop out, resulting in the
intertwining relations of the SframeIn the limit the metric shift
should be replaced with
and the components in (86) can be viewed as referring to the Sframe expanded in a generic coordinate basis:
,
,
,
. The relations (86) can then also be computed directly from (
17) without referring to (A48) and its scaling limit. The invariance of the orthonormality relations (
A49) carries over to the Sframe. The relations (86) witness again two remarkable properties: the
and
components do not mix with the other frame components under the
action, and the transformation pattern is independent of the spatial metric
and its inverse. The intertwining relations (86) can be used to compute the transformation law for any weight zero Stensor, without reference to the limit construction.
For the sake of illustration we consider vectors and covectors here; the extension to tensors of arbitrary rank and type is straightforward. Consider first the
-invariant Sframe components
of a vector field
. Then
Similarly, the
-invariant Sframe components
of a covector
transform according to
The same transformations arise by taking the scaling limit of (
A52) and (
A53). Expressed in terms of the Sframe components the vectors fields and one-forms read
This highlights again the intertwining character of the construction: the left hand sides are invariant under the standard tensorial action of diffeomorphisms while the right hand sides are invariant under the action.
A type
Stensor of weight zero is called an Sscalar, and examples have already been encountered in the verification of
’s
invariance. A simple example of an Scovector arises from the gradient of a metric-independent Sscalar like
. In adapted coordinates the components are
,
; hence
,
are its Sframe components. By direct computation one finds this pair to transform according to (
88). Hence,
is an Scovector of weight
.
Other Stensors of zero weight can be formed by taking tensor products. In addition, Stensors of zero weight admit a trace operation inherited from conventional tensors that commutes with taking tensor products. Both aspects will be detailed below.
Stensors of integer weight. Often the existence of a nontrivial limit in (
4) requires nonzero weights,
. We begin with a broad but somewhat implicit definition
Definition 3. (Stensors as Limits).On a foliated pseudo-Riemannian manifold, a tuple , , of tensors is called an Stensor of type and weight , if it arises from the metric frame components of a conventional type tensor as the finite nonzero limit . Thendefines its transformation law, which depends only on the type and the weights. In the limits the substitution is understood. Whenever unambiguous the “” left superscript will be omitted. Several features are implicit in this ‘Limit Definition’: first, nonzero weights arise only if the conventional tensor
U is metric-dependent. Second, the mixing pattern of the components
can be computed from the intertwining relations (A48) by keeping only the leading terms in a large
expansion and it is uniquely determined by the type and the weights
. The discussion after Equation (
A53) implies that each
is a
linear function of the
,
, while
may enter nonlinearly. Third, for a fixed type only a finite number of consistent weight assignments exist, see
Section 4.2.
While the Limit Definition successfully defines the realization for generic tensors it does so non-autonomously in terms of . For most purposes the Limit Definition is also too broad. In order to motivate a narrower but more intrinsic notion we begin with the following observation: the metric frame components of a conventional tensor U some of whose indices have been moved with will differ from those of an underlying tensor V with scale invariant metric frame components solely by having some of the indices moved with . This covers most of the situations occurring in gravitational fields theories. The -invariant combinations may include derivatives of like or , etc., but terms with nonzero scaling weight then normally arise from moving indices with the undifferentiated spatial metric. In the following we restrict attention to ‘metric compatible’ Stensors arising as the weighted limits of conventional gravitational tensors of this form. In this situation the weights are always even integers and the Sframe components are related to those of an underlying Stensor of zero weight by moving indices with .
Definition 4. (Metric compatible Stensors).Let M be a manifold foliated by the fields entering the Sframe (36) and equipped with transforming according to (17) and (22). A tuple , of tensors is called a metric compatible Stensor of type and weight , if it arises from an Stensor , of the same rank but zero weight by moving indices with . Metric compatible Stensors arising from variations of the Sgravity action will be denoted by a subscript ‘0’. This is an intrinsic notion in that no reference to the scaling is made. A minor indirect reference to conventional tensors is made through the notion of a tuple. The latter also entails that the occur coordinated in different blocks in a way compatible with moving spacetime indices with . As a consequence the always originate from the scaling limit of at the same position in the tuple which ensures consistency with the Limit Definition. As in the definition of an tuple, the reference to conventional tensors could be avoided by listing the possible patterns that can arise. In any case, the resulting are constructed from , suitable derivatives, and the underlying components. The scaling degree of some such can be read off simply from that of the constituent fields: , , , . Further, the transformation law of the can be computed intrinsically from the realization alone, and thus does not have to be part of the definition.
As an illustration, consider again the case of a vector
and a covector
with
-invariant metric frame components. Then
defines a covector, whose metric frame components
,
are related to those of
by
,
. The components
will transform according to (
A53). However,
now transforms nontrivially
under the scale transformation (
13). As a consequence the
scaling limit of (
A53) is not given by (
88) but rather by (
92) below. Similarly,
defines a vector whose metric frame components
,
, are related to those of
by
,
. Before scaling
transform according to (
A52). However, since
under (
13) this will again affect the
scaling limit. Instead of (
87) the limit of (
A52) is now given by (
93) below.
The interrelations of the Sframe components will be the same, just with
replacing
. By Definition, the tuple
is a metric compatible Scovector of weight
and
is a metric compatible Svector of weight
. The relations to the underlying weight zero Svector
and Scovector
are given by
The transformation laws can be computed using only the
realization, i.e., (
17), (
22), (
87) and (
88), and coincide with the ones obtained via the limit construction:
Scovector of weight (0,2):
Svector of weight (0,−2):
One may verify that the inner product,
, is invariant. Note that the transformation pattern (
92) and (
93) clearly differ from those of the weight
counterparts in (
88) and (
87), respectively.
Tensor product. A basic operation for conventional tensors is the tensor product. In view of (
90) one may ask whether the
scaling limit of a tensor product coincides with the tensor product of the limits and defines again a weighted Stensor. This is indeed the case and can be seen as follows. Let
U be a conventional type
tensor which factorizes into a tensor product
, i.e.,
, in terms of coordinate components, for
,
. Its metric frame components are obtained by contracting with
in all possible ways and therefore factorize as well:
,
, where we now interpret the indices as abstract ones and correspondingly include the tensor product symbol. The weights are additive,
, and we may assume them to be such that the limits
,
, exist separately, i.e., there are no spurious
cancellations in the product. The standard realization
of the diffeomorphism group is such that
, for all
, whence named ‘tensorial realization’. It follows that
The same holds for the linear hull of expressions of the form , which concludes the argument.
The weighted Stensors of a given type
obviously form a linear space. The above result shows that they can also be endowed with a tensor product under which the weight is additive
The resulting tensor algebra
has
as its automorphism group. Symbolically,
in analogy to the standard case (
A15).
We add some comments: (i) Importantly, the weighted Stensors do
not in general admit a well defined trace operation. In particular,
is not equipped with a trace operation. The qualified existence of a trace operation and the circumstances under which it commutes with taking tensor product will be discussed below. (ii) It is not automatic that all weighted Stensors can be obtained by taking the linear hull of tensor products of Svectors and Scovectors. Examples will be discussed later on. However the transformation laws of generic weighted Stensors can be obtained from (
87) and (
88) alone by taking tensor products. (iii) An exception to (i) and (ii) are Stensors of weight zero. These always admit a well-defined trace operation that commutes with taking tensor products. Moreover, for weight zero all higher rank Stensors can be obtained by taking tensor products of rank one Stensors. Both properties follow from the Sframe intertwining relations (86).
Trace operation. The tensor product (
95) applies to all weighted Stensors, in particular to metric compatible ones. One will naturally want a trace operation to be compatible with moving indices; so we restrict attention to metric compatible Stensors. Then only trace operations contracting a covariant with a contravariant index need to be considered. To motivate the definition we return to conventional tensors and their metric frame components. Before taking the metric frame components contraction with
effects the trace over a lower
index and an upper
index. The completeness relation for the metric frame reads
see (
A25) and (
A26). It can be used to expand any type
conventional tensor
U in terms of its metric frame components
, see (
A54). The trace will likewise be governed by (
97) and produce the expansion of a type
tensor. The coefficients
of the latter will be linear combinations of spatial traces of a subset of the original coefficients
. We write
for the coefficients
, assuming that it is clear from the context over which pair of indices the trace is taken. For example, a type
tensor with without symmetries and components
will expand into 8 terms, while the trace
expands into 2 terms with coefficients that involve spatial traces of only 4 of the original 8 coefficients. Explicitly
The Stensor of weight
associated with
U has components
. Whenever well-defined we define its trace by the
same spatially contracted linear combinations
as the original tensor. In the above example, the trace of the type
Stensor
, has the 2 coefficients
,
, where
,
,
,
, are the limiting counterparts of the coefficients in (
98). A necessary condition for this to make sense is that all terms in a linear combination have the same weight. In the example
must have weight 0 since
has, and
and
must have the same weight. Generally, in each linear combination
all terms must have a fixed weight, for which we write
,
. If the original Stensor
has weight zero this is automatically satisfied, otherwise this condition restricts the class of weighted Stensors which allow for a well-defined trace. Whenever this condition is met, one will want the tuple
to transform as an Stensor of type
and weight
. This will normally be the case automatically but is best verified on a case-by-case basis.
In summary, for any Stensor of type and weight one can (for a pair of indices understood) introduce the tracial coefficients , , as a linear combination of spatial contractions of the original coefficients . Assume that the terms in each linear combination carry the same weight and that the set , transforms like an Stensor of type and weight . Then the latter Stensor is called the Strace of , .
A relevant example is the SRiemann tensor (
124) below, which can be modeled as an Stensor of type
and weight
. Its Strace then is a type
Stensor of weight
, which coincides with the independently defined SRicci tensor.
Generally, Stensors of weight zero admit a trace. The condition on the weights is trivially satisfies so that only the transformation law needs to be checked. As far as the transformation law is concerned we can model the Stensors in question as a tensor product of Svectors and Scovectors of zero weight. Since all indices are on the same footing it suffices to verify that the type
Stensor
admits a trace. The trace is given by
and ought to transform as an Sscalar, based on (
87) and (
88). This is indeed the case. It follows that for zero weight the tensor algebra
can be equipped with a trace operation that commutes with taking tensor products, in exact analogy to the standard case.
4.2. Examples of Stensors and Their Gauge Variations
In the second half of this section we present prime examples of Stensors relevant in a gravitational context. Among them are Sgravity counterparts of the metric, the Einstein- and energy–momentum tensor, as well as the Riemann tensor. All of them are metric compatible Stensors in the sense defined and are subject to a specific weight-dependent transformation law. Those related to variations of the Sgravity action will be denoted by a subscript ‘0’ rather than a left superscript ‘’.
For many purposes the linearization of transformation laws under generic diffeomorphisms suffice. For standard tensors on some smooth manifold
M this is implemented by the (generally covariant) Lie derivative. If
M is a foliated pseudo-Riemannian manifold the latter induces
gauge variations on the metric frame components that are discussed in
Appendix A.3. Here we aim at counterparts of these gauge variations for Stensors.
A convenient starting point is the linearized version of the Limit Definition. Consider the tuple , of tensors obtained by taking the metric frame components of a conventional tensor. The tuple transforms under infinitesimal diffeomorphisms according to the gauge variations induced by the spacetime Lie derivative. We write , for these gauge variations and take their scaling limit. By assumption each scales according to and the right hand side of may also explicitly depend on , . As a results there are for each type only a finite number of weight assignments , that lead to a consistent nontrivial limit. These can readily be classified and lead to a list of possible linearized transformation laws for Stensors in the sense of the Limit Definition. We write for these Stensor gauge variations and label them by the weight tuple , . Whenever unambiguous the “” left superscript will be omitted.
Not all of these cases are compatible with moving spacetime indices as required for metric compatible Stensors. In a second step one can screen the previous list for compatibility with these criteria to obtain a subset of possible weights and linearized transformation laws metric compatible Stensors. We illustrate the construction for rank .
For rank 0 there is only one possibility: rank 0 tensors are scalars and their
counterpart has already been introduced. Infinitesimally
characterizes Sscalars.
For rank 1 there are four consistent limits:
Scovector, weight
:
Scovector, weight
:
Svector, weight
:
Svector, weight
:
These are precisely the linearized versions of (
129), (
130), (
92) and (
93). On account of (
91) all four cases are also metric compatible.
The list of transformation laws for rank two Stensors arising as weighted limits can similarly be generated by inspection of (
A62), (
A65) and (
A68). The results are collected in
Table 2.
Most of these are not metric compatible. In the following we focus on three cases directly relevant for Sgravity: covariant symmetric rank two tensors
of weight
; contravariant symmetric rank two tensors
of weight
; and the interpolating case of a mixed rank two tensor
of weight zero. We first note the transformation laws, which can be obtained, for example, from (
A68), (
A65) and (
A62).
Type
, weight
,
:
Type
, weight
,
,
,
:
Type
, weight
,
:
We yet have to show that the two symmetric Stensors are indeed metric compatible in the sense defined. In order to do so, we note that the transformation law (
105) is consistent with the following reduction conditions
The reduced transformation law reads
with
. In terms of the reduced weight zero Stensor
we define
,
,
. Then
transforms according to (
104). By Definition it is a symmetric, metric-compatible Stensor of type
and weight
. For the contravariant case one proceeds similarly. In terms of the reduced weight zero Stensor
one defines
,
,
. Then (
108) implies that
transforms according to (
106) and hence defines a symmetric, metric-compatible Stensor of type
and weight
.
Another aspect of the metric compatibility of
and
is the existence of a well-defined trace operation inherited from the underlying mixed weight zero Stensor. For the latter the trace operation is a special case of the one described after (
97). Explicitly,
, which transforms like an Sscalar. Using the above construction one obtains the induced trace operations
These expressions adhere to what one would expect from the completeness relations (
A61) and (
A67) and are Sscalars on account of (
104) and (
106). These mathematically preferred cases turns out to be also the ones for which Sgravity provides constructible examples.
Sgravity metric as an Stensor. The projected components of the pseudo-Riemannian metric
are
, and the transformation law (
A68) specializes consistently. The Sgravity metric will naturally be assigned projected components
, and as such should be a type
Stensor of weight
. The transformation law (
104) indeed specializes consistently. Moreover, the nontrivial variation,
, coincides with the one in (
19). Note that the nontrivial weight assignment is essential for this to work. The strong coupling contra-metric will similarly be assigned projected components
. As such it should be an Stensor of type
with weight
or
. On account of the preceeding discussion we take the weight to be
. The transformation law (
106) indeed specializes consistently and the nontrivial variation,
, matches the linearized form of (
22). Both
and
have a well defined trace (
109) which equals
.
Energy–momentum Stensor. Next, consider the energy–momentum tensor of the scalar field (
A69). In order to take the
scaling limit we interpret the scalar field as dimensionless and scale invariant as in (
13). For the limit of the energy–momentum tensor’s
components
one then finds
Here ‘subst’ refers to the substitution (
16) and the limits are viewed as functions of
. In parallel to (
112) below the limiting expressions can alternatively be obtained as variations of the matter part of the action
in (
1); keeping to our notational conventions we therefore denote them with a subscript ‘0’. Along different lines this matter action also arises in a first order formulation [
15]. Based on the previously defined transformation laws
one finds by direct computation
This agrees with the transformation law (
104) of a type
Stensor of weight
. Since
the last equation in (
111) is equivalent to
. This pairs with the first equation, as
. It also ensures that the scaling limit
of the trace
in (
A69) transforms like a Sscalar under
.
Field equations Stensor. The gauge variations of the purely gravitational part of the strong coupling field equations must be compatible with (
111). In general relativity this is ensured by the transformation properties of the projected components (
A72) of the Einstein tensor. In Sgravity we
define
with
from (
1). This parallels (
A72), but since
is regarded as a functional of
(not
) the sign in the second term of
is flipped. Taking this into account, the expressions (
112) turn out to be the weighted scaling limits of the Einstein tensor’s components in the metric frame:
As a consequence
should transform as an Stensor of type
with weight
. Explicitly,
It is a nontrivial consistency check that this comes out correctly, based solely on the
variations of the basic Sgravity fields and the definition (
112). Before turning to this check, observe the difference to the transformation law (
A68) satisfied by the projected components of the Einstein tensor: in contrast to (
A68) the mixing
only occurs in the
variation. Importantly, (
114) matches (
111) in the matter sector which renders the Sgravity field equations
foliation-independent.
We now turn to the direct verification of (
114). A bonus feature is that the result turns out to be valid for arbitrary parameter
c in the DeWitt metric, i.e., when
in (
12) is replaced with
. The limit of Einstein gravity corresponds to
. In order to facilitate comparison with (
A72) and [
1,
10] we regard
as a functional of
and the densitized Slapse
instead of
. We write
for the resulting
c-dependent purely gravitational part of the Sgravity action (
12), with ‘
L’ indicating that the Lagrangian not the Hamiltonian version enters. For simplicity we also omit the cosmological constant term, which can be restored from the previously treated matter sector by shifting the potential,
. In this setting the independent variations are:
(the Lagrangian form of the Hamiltonian constraint),
(the Lagrangian form of the Diffeomorphism constraint), and
. We first note the explicit forms of
Since
, all constituent quantities transform according to their Lie derivatives under purely spatial gauge transformations. Hence also
and
, transform according to their spatial tensor type (a
scalar density and a
cotensor density, respectively) under time-dependent spatial gauge transformations. For the temporal part of the gauge transformations we use
as the descriptor, matching
. A lengthy direct computation then shows
Unlike in general relativity, the evolution equations do not mix with the constraints and thus could be imposed separately. The gauge variation of
requires variation of the Christoffel symbols. A slightly more efficient route is to start from the averaged form
for a non-dynamical spatial vector
. In the gauge variation of (
118) one finds that the
terms cancel and that modulo total time derivatives the result can be written as the spatial average of
. Stripping off the
auxiliary vector one obtains
It is instructive to compare this with the Diffeomorphism Ward identity for pure Sgravity
where the · indicates a
integration. Stripping off the gauge descriptors
yields
Specializing (
119) to
thus mirrors the time evolution of
, as expected on general grounds. The second equation in (
121) can also be used to rewrite (
120) in a more suggestive form
Using
,
,
, in (
117) and (
122) one arrives at (
114).
SRiemann tensor. Rank four Stensors warrant a systematic discussion, omitted here. Mostly in order to highlight the difference to other notions of curvature we present the scaling limit of the Riemann tensor. We begin with noting the metric frame components of the type
Riemann tensor
Here the superscripts refer to the scaling weight of the homogeneous parts under the scaling transformation (
13). As before,
is the extrinsic curvature. Since all blocks are built solely from
there is no freedom to choose weight assignments. A nontrivial limit for the blocks arises only for the case
In line with the general conventions we changed the notation in the limiting quantities and regard them as functions of
. Note that the dependence on the Riemann tensor of the spatial metric has dropped out. The Riemann type symmetries are preserved in the limit and imply
Together with the transformation law (
127) below the weights in (
124) identify the quadtuple
as a (specific) Stensor of type
and weight
. We shall refer to it as the
SRiemann tensor.
The
SRicci tensor is defined analogously, starting from the contractions of (
123). Its transformation law is that of a type
Stensor
of weight
, which can be obtained by taking the weighted limit of (
A65). On the other hand, one can take the trace of the limits (
124) and finds
These coincide with the indicated coefficients of the SRicci tensor, rendering the transition from (
126) to (
126) an example of an Strace, as previously discussed.
The transformation law under infinitesimal
transformations for type
Stensors of weight
and with Riemann-like symmetries are as follows
The Riemann-like symmetries (
125) are compatible with the transformation law (
127). Subject to them, one has in particular
. This allows one to express
in terms of
; so one of the middle two relations in (
127) is then redundant.
4.3. Decoupling Maps
So far the covariant Carroll structure of
Section 3.1 did not enter. Making use of it, the transformation law of metric compatible Stensors can be coded in a more concise way.
Weight zero. Stensors of weight zero can be viewed as arising from some conventional tensor by taking components with respect to the Sframe, written as , which turn out to be invariant. This leads to a characteristic mixing pattern for the blocks that can be inferred from the Sframe intertwining relations (86). On the other hand, one can consider the components of the same conventional tensor with respect to the (-rescaled) SDiff frame, , . Concretely, for a type tensor with coordinate components the components arise by contracting in all possible ways with . For clarity’s sake we denote the components with respect to the SDiff frame systematically by a left superscript, . In particular, a left superscript does not necessarily signal a dependence as before.
On account of (
68) the multiplets
do not mix under arbitrary
transformations. The
tuples
and
,
, can be related by expressing one set of frame coordinate components in terms of the other. Using the
rescaled variants and (
39) and (
42) the nontrivial conversion formulas are
For example, in the case of a vector field
with
-invariant Sframe components
one has
For a one-form
with
-invariant Sframe components
one finds similarly
Compared to (
87) and (
88) the mixing terms have disappeared. For later reference we also note the transition formulas for a weight zero Stensor
of type
:
Again, the affine transformation law precisely cancels the mixing terms of the original Stensor blocks. This holds generally:
Consider the tuple , , obtained by taking the SDiff frame components of a conventional type tensor U, and assume them to be -invariant (treating as invariant). Each of these SDiff frame components transforms multilinearly among itself for fixed , with for each lower a index, and with for each upper b index, while contractions do not affect the transformation law. Further, each can be expanded in terms of the Sframe components , for varying , with -dependent coefficients. As a consequence the decoupled transformation laws for the SDiff frame components are in one-to-one correspondence with the mixing transformation law of the Sframe components.
Nonzero weight. Next, we seek to generalize this correspondence to Stensors of nonzero weight. The tuple , , obtained by taking the SDiff frame components of a conventional type tensor U, will then no no longer be -invariant. One may be tempted to simply take the weighted limit , to obtain the nonmixing multiplets. Some experimentation shows that this does not work; the limits in general do mix under transformations.
Instead, we proceed as follows. Restricting attention to metric compatible Stensors the idea is simply to apply the decoupling map of the underlying Stensor of zero weight and rewrite it to obtain a decoupling map for the weighted Stensor. We first illustrate the procedure with several examples.
Consider a weight zero Scovector
and define
. As seen before, then
transforms as an Svector of weight
. Applying the decoupling map (
130) to
, i.e.,
,
, and rewriting it suggests
By direct computation one can verify that (
132) indeed decouples the weight
transformation law (
93). Similarly, starting with a weight zero Svector
one can define
,
. Then
is an Scovector of weight
. Applying the decoupling map (
129), i.e.,
,
, gives
These can be checked to decouple the
transformation law (
92). In fact, the original and the decoupled transformation laws are in one-to-one correspondence.
We remark that both maps (
132), (
133) can alternatively be obtained by moving the spacetime index with the invertible Carroll metric (
74) to get
and
, in a first step. The SDiff frame components of
are given by
,
. Utilizing the weight zero map (
129) and inserting the definitions reproduces (
132). Similarly, the SDiff frame components of
are
,
. Utilizing the weight zero map (
130) and inserting the definitions reproduces (
133).
Next we consider rank 2 metric compatible Stensors and focus again on the two cases from
Section 4.2. We first present the decoupling maps in both cases and then comment on the derivation:
Type
weight
:
Type
weight
:
In
Section 4.2 we noted explicitly the linearized form of the transformation law for each of these weighted Stensors. It may suffice here to verify the decoupling property also at the linearized level. The additional variation needed is the linearized form of (
38) which comes out as
Using (
104)–(
106) and (
136) one can verify that the given combinations indeed decouple the linearized transformation laws, that is, (
138) below holds.
In order to derive (
134) we return to the realization of
in terms of the reduced type
Stensor
of weight
, with
. One has
,
,
. Next, we recall from (
131) the decoupling map for a weight zero type
tensor. Specializing to
, and inserting the defining relations for
one obtains (
134). The derivation of (
135) proceeds similarly. In term of the reduced weight zero Stensor
one realizes
as
,
,
. The decoupling map (
131) specialized to
implies (
135).
The construction principle generalizes and yields the following result and definition:
Definition 5. (Decoupling map and SCarroll tensor).Let , be a metric compatible Stensor of type and weight . Then there exists a homogeneous polynomial of degree in , , and possibly preexisting Stensors such thatdecouples the transformation law. The map is called the decoupling map and the resulting tuple , is called an SCarroll tensor of weight . We add several explanations. (i) The degree is defined by
,
,
,
. (ii) By a decoupled transformation law we mean that each
transforms multilinearly among itself under
generic (foliation-changing as well as non-Carroll) diffeomorphisms in the
realization, where an
occurs for each upper and a
occurs for each lower index. The
component is always invariant. Upon linearization this amounts to
where
is the
d-dimensional Lie derivative acting on
according to its
tensor type. (iii) The term SCarroll tensor is modeled after that of Carroll tensors [
5,
9] which transform covariantly and without mixing under the Carroll subgroup of diffeomorphisms. SCarroll tensors have the same feature but with respect to the
realization of the full diffeomorphism group. In fact, an SCarroll tensor gives rise to a
tuple of Carroll tensors by fixing the
gauge. For SCarroll tensors of weight zero this is clear from the reduction of the corresponding frames, see (
57). In particular, each upper index transforms with
and each lower index with
. The same holds if indices are subsequently moved with
or
as they transform correspondingly under the Carroll subgroup. (iv) The SCarroll tensors are a convenient arena on which an adapted Levi–Civita type connection can be defined, see
Section 5.3.