5.1. The Limit
The strategy that authors of [
96,
97] followed was the following. Knowing that the conformal symmetry was present at the classical level of the theory with
, they also expected it to be present in some unbroken form on the quantum level, here analyzed to the level of one-loop accuracy. This assumption allowed them, as we will explain later, to use the power of conformal symmetry at the end of their computation for clearing the situation with counterterms of the UV-divergent one-loop effective action. As a matter of fact, these counterterms in
are all globally scale-invariant for generic HD gravitational theories. They are, namely, of three possible forms here;
,
, and
, where the last one is a Gauss–Bonnet scalar term.
As analyzed before, the Weyl squared term, given explicitly by
when properly densitized to the form
, is a conformal invariant. That is, we have for infinitesimal conformal transformations (denoted below by
), which only matter here, that
On the other hand, the last invariant is topological and is known as the celebrated Gauss–Bonnet scalar (or as an Euler density
in a densitized form). Its expansion, in other terms quadratic in standard gravitational curvatures, reads,
Since it is a topological term (depending only on the boundary of spacetime), then any of its bulk variation carried out in the volume of spacetime (in particular the conformal infinitesimal one
) vanishes, that is, we have
as an identity of variational four-dimensional calculus.
Finally, the last invariant with respect to diffeomorphisms (scalar) out of the group of three local invariants in curvature in quantum gravity, that is, quadratic in the curvature, is simply the square of the Ricci scalar
. Here, one notices an interesting thing pertaining to conformal transformations of this invariant which is quadratic in gravitational curvatures (similar to Riemann tensor, Ricci tensor, and Ricci scalar). This term is globally conformally invariant in
, i.e., it is scale-invariant because its energy dimension as the operator is
. The coupling coefficient in front of it in the action in (
25), this
, is of course a dimensionless parameter in
spacetime dimensions. This fact is common also for all the other invariants considered above (in (
29) and (
31)), and this happens precisely because all these invariants contain four derivatives on the metric tensor and this number coincides here with the dimension of spacetime. Hence, the
term,
term in (
29), and the
term in (
31) are properly scale-invariant terms when densitized on the level of action functional. As we have seen above for the last two, we have upgrade of these symmetry invariance properties to the full invariance under local conformal symmetry understood in the GR framework.
For the first term
, however, the situation is different and its symmetry invariance obtains some promotion but not to the full conformal invariance. Namely, the
term is invariant in
dimensions under so-called restricted conformal transformations, that is, the transformations from (
9), satisfying the additional condition that
which is a kind of GR-covariant wave equation (GR-covariant d’Alembert equation) for the parameter of the conformal transformations
. This means that this parameter
can be spacetime-dependent here (already this is more than a case of invariance under global transformations when we have that
but this dependence cannot be arbitrary here. It must satisfy an analogue of the wave equation, which in Euclidean signature would require the
field again to be constant. Instead, in the Minkowskian signature, this wave equation for
does have some non-trivial spacetime-dependent solutions. This signifies that the condition of restricted conformal invariance is more general than global scale invariance in GR with constant parameter
, but still it is less than a general invariance under fully local conformal symmetry with completely arbitrary spacetime dependence in
.
One can easily see this fact by recalling the following properties of conformal transformations of the Ricci scalar. For full generality, we write them in arbitrary dimensions
d. We have explicitly, that under local conformal transformations,
and then using the Leibniz property of the box operator:
, and the fact that
, we rewrite above as
Now, exactly in
(and only in
, neglecting the degenerate case of
where
automatically), we see that the last term in the bracket in the above formula disappears, and we are left with the following simple local conformal transformation law in
, that
which indeed shows that for restricted form of conformal transformations satisfying that
, we obtain that the Ricci scalar transforms fully covariantly under such local conformal rescalings. In actuality, one sees that this is the special property of
spacetime dimensions, since in other dimensions even restricted conformal transformations do not make Ricci scalar transform conformally covariantly because of the second term in (
35) multiplied by the factor
and by a mixed term
. From this, one easily concludes the transformation law
which shows that, indeed, for restricted conformal transformations, the densitized term
is (partial) conformal invariant.
5.2. Final Conformal Transformation
Equipped with the above knowledge, we can discuss the trick and the method used by Fradkin and Tseytlin to obtain, a bit heuristically, the UV-divergences of conformal Weyl gravity in dimensions. The computation for the case showed that, generally, the term is present in UV-divergences, and this term is not conformally invariant as a counterterm, when we mean full arbitrary conformal transformations. Therefore, its presence violates the assumed conformal symmetry at the one-loop level. It would be desirable for consistency of the theory to make it vanish. We explain below what was performed to achieve this.
In an HD theory from (
25), the one-loop UV-divergent part of the effective action when the limit
is taken explicitly reads,
where we effectively used the dimensional regularization scheme (DIMREG), and whole divergences are written as singular expressions in the deviation from
spacetime dimensions. As obvious from above, the
term in this naive procedure is generated at the one-loop level; therefore, conformal symmetry seems to be broken by quantum corrections.
Assuming, in contrary and
a posteriori, that the local conformal symmetry was present also on the level of the first quantum loop, the authors decided to, after the computation, take the limit
in a naive way and eventually performed a compensating additional conformal transformation, as in (
9). They simply wanted to exploit the freedom of making arbitrary post-computation conformal transformations on the results of the non-conformally covariant calculation. The parameter of such a conformal transformation
could be chosen at one’s will. The authors chose such
that
in (
36), that is, the result of final conformal transformation completely nullifies the presence of the square of the Ricci scalar
counterterm. Exactly this term before was spoiling the conformal invariance of the full set of counterterms. In this way, and somehow by hand manipulations, the non-fully conformal counterterm
is completely cancelled out, so the problem with the desired and originally-not-present conformal invariance of the UV-divergent action in conformal gravity at one-loop is solved.
For consistency, the following question then arose: Is it possible to find solutions for the spacetime dependence of the
parameter of conformal transformations in (
9), such that indeed we find that in the result
? The solution of the related equation in
,
or to an equivalent equation,
exists and it is given for
by
In the last formula, we performed some formal manipulations with treatment of the operator
as completely algebraic, although of course it is a differential operator, and taking its inverse requires some additional care. One can convince oneself that the solution for
in (
41) satisfies the equation (
40), when one notices that
when acting on 1 gives
as the result of the operatorial linearity, and moreover
is treated here as the operatorial inverse of the operator in the last parentheses. In particular, to define the inverse of such shifted GR-box covariant operator by possibly spacetime-dependent Ricci scalar, understood here as an arbitrary scalar field
on a curved manifold, one must resort to operating with Green functions of the operator since only then can we sensibly speak about the inverse on the general curved backgrounds. Moreover, the last formula in (
41) is valid only when the metric
, whose Ricci scalar
appears there, is asymptotically flat.
Therefore, one concludes that the solution for
in
always exists for any form of the spacetime-dependent Ricci scalar
However, the price is the resulting strong non-locality in the expression for the
in (
41), which is the parameter of the conformal transformation that has to be performed at the end to bring back the conformal invariance to the theory on the level of one loop. This is how the authors of [
96] recover, at the end, conformal invariance in the model under studies on the quantum level. It is well understood that this procedure looks very
ad hoc. This is so because the formalism they used was suitable only for dealing with diffeomorphisms as local symmetries, and the regularization that they used was only preserving this last symmetry. The care was not exerted to properly treat the local conformal symmetry and to gauge-fix it and to add appropriate FP determinant and FP ghosts for conformal symmetry. At the end, the authors performed quite arbitrarily-looking conformal transformations somehow to cure their results obtained in the limit
. For this, they had to know
a priori, but not
a posteriori, that conformal symmetry survives on the quantum level. Somehow, they knew it and understood why the conformal symmetry is still present on the level of one-loop counterterms and they worked out an ingenious solution and procedure to improve on their non-conformally covariantly-looking results. Eventually, they were able to show that, indeed, the conformal invariance was preserved on the level of one-loop divergent effective action.
As emphasized already a few times, it seems that the authors of [
96] assumed what they wanted to prove by explicit computation, that is, that the conformal symmetry is not violated on the first-loop quantum level.
A posteriori, they proved that their initial assumption was indeed correct, but in this way they entered into some kind of logical vicious circle. Somehow, we need an argument that does not rely on the presence of conformal invariance before it is explicitly checked that it is there. This argument is given in the form of analysis of the conformal anomaly at the one-loop level, that the same authors also discussed in later publications. We will discuss it also below. Here, we want to remark that all the considerations of the presence or absence of the conformal symmetry were considered so far only on the level of the divergent effective action, and the issue of conformal anomaly (CA) and its form were completely neglected for these aspects. So far, we have just analyzed whether the counterterms are conformally invariant, while, as emphasized earlier, the mere existence of them breaks full conformal symmetry due to presence of non-vanishing beta functions for couplings, and thus also of non-trivial RG flow and of non-trivial scale dependence of various correlation functions.
The results from the limit
did naively contain the
term, which is dangerous for full local conformal symmetry, but these were only incomplete results since additional care had to be given to them. By performing final conformal transformations, Fradkin and Tseytlin were able to reduce to zero such non-conformal counterterms; however, the parameter of such a transformation was expressed as highly non-local function in (
41). They could still perform the conformal transformations on the form of counterterms as the last operation, because the symmetry they acknowledged earlier had to be present also on the quantum level since it was present also on the classical level before the quantization procedure. Implicitly, they assumed that the quantization act does not destroy conformal symmetry. They just used the freedom to select a conformal gauge, although one has a strong impression that their approach could be more clear since they knew only
a posteriori that local conformal symmetry must have been present on the level of the first quantum loop in this theory. The fact that the solution for
contains this apparent non-locality in local Weyl conformal gravity in four-dimensional QFT framework could be quite disappointing, and one could wonder whether it is possible to perform all such procedures better and in a more clean way. Moreover, the parameter
must be always spacetime-dependent since constant values do not change the form of the Ricci scalar
R according to (
36), because we always have that
for
.
When one is finished with the dangerous conformal symmetry-violating
term in UV-divergences, then this is not the end of the story because, as authors say, the contribution of two scalar degrees of freedom has to be subtracted from the results in (
38) without
term there. These are assumed to be two minimally coupled to gravity scalar modes on a general background. At the end, the form of UV-divergences in Weyl conformal gravity in
spacetime dimensions regularized in the DIMREG scheme reads,
The results in the formula above should be understood as final ones. No additional conformal transformation or subtraction of some contributions should be performed on them.
One also wonders whether the final conformal transformation nullifying the
term can be reversed. It is easily acceptable that one obtains as the result of conformal transformation that
. However, here is visible apparent irreversibility of the conformal transformation because one does not obtain anything from
as the starting point, since zero transforms to zero under homogeneous transformations. Of course, this is only apparent since by inverse conformal transformations which are achieved with
parameter, one obtains from
back to the case
, or one can even perform a new transformation with
, and then from
one will find oneself in the spacetime where
. One, therefore, has a strange impression that the power of conformal transformations in a theory with conformal symmetry on the quantum level was only used instrumentally by authors of [
96] to reach the assumed goals and, for example, after doing this, one is forbidden to use it anymore, or to use it for reversing its effects. Of course, such a situation with understanding of the symmetry in the theory is not satisfactory, and one asks for a natural formalism which makes such transformations obvious, and all of them are allowed to be performed provided the symmetry is realized there on any level of the theory.
In other words, we can say that the term in UV-divergences of the theory is conformal gauge-dependent, that is, by a choice of some conformal gauge one can eliminate it but a priori it is there. We choose a conformal gauge by performing arbitrary conformal transformation in the theory. Is the term there or not?—This statement is not universal and it is actually gauge-dependent in conformal gravity. The value of the beta function is not universal either. Therefore, in the description preserving conformal symmetry we shall not speak at all about this term, and the issue related to its presence or non-existence is not a physical issue, it is just another example of the gauge artifact statement, here due to conformal gauge.
If one looks at the difference between formulas (
42) and (
38) (the last one obtained as the limit in the case
), one sees that the correcting contribution of two scalars is essential and that this together with the resolution of the problem of the
counterterm signifies why the simple naive limit
does not work for the case of quantum conformal gravity. These are all post-computational procedures that one has to perform after the naive limit
is taken. They constitute the theoretical reasons (or the problems within the formalisms of simple QFT of gravitational interactions with only diffeomorphisms symmetries) why the limits in question are discontinuous and why the case of conformal gravity is very special one.
5.3. Subtraction of Two Scalars
There could be various attempts to explain the role of the mysterious two scalar degrees of freedom that have to be subtracted for the last corrections to the theory. The possible questions are why they have to be subtracted, not added, why they are two, and why scalar degrees of freedom, and not, for example, one massless vector (spin-1) or tensor (spin-2), each of the last two coming with two helicity states as well. Finally, one could ask, accepting that they are scalars, why they have to be minimally coupled to a curved gravitational background, and not conformally coupled to gravity. The last option looks to be the most reasonable one, especially in the framework of conformal gravity, where conformal symmetry of the whole physical system (gravity + matter) on the quantum level is the thing we care so much about here.
A possible explanation of the problem with two scalars may come from analyzing the spectrum. Exactly two degrees of freedom are the difference between eight and six perturbative degrees of freedom (as analyzed around flat spacetime background), respectively, for generic HD gravitational theories and for Weyl conformal gravity. Hence, if one wants to obtain six degrees of freedom one would have to subtract two degrees of freedom from all degrees of freedom present in HD gravity with both
and
terms. The arithmetic here agrees. However, there could be some problems with this interpretation. First, as we discussed before, this way of counting degrees of freedom is not completely unambiguous and one may be tempted to assume number of six degrees of freedom also in generic HD gravitational theory without the Einstein–Hilbert term added. Secondly, when one looks at the spectrum of generic HD theory, one does not see that there are two more scalar degrees of freedom compared to Weyl gravity. One sees only one physical scalar degree of freedom in generic HD theory (with standard counting to give
degrees of freedom), while in conformal gravity, there are no scalars at all
10. The actual reshuffling of degrees of freedom (from eight to six) is intricately more complicated: first, one must subtract one scalar d.o.f. and then later divide five d.o.f.s from massive spin-2 field into two helicities of massless spin-2 and another two helicity states of massless spin-1 (fields present in conformal gravity), and finally one has to remove the last remaining degree of freedom. This game with various Lorentz representations of perturbative degrees of freedom would suggest to try something with one scalar, massless vector or massive spin-2 mode. However, it is incorrect to add or subtract here, for example, the contribution to one-loop divergences of traces of a vector (spin-1) and the final results work only when precisely two scalars minimally coupled are subtracted. Hence, this reinterpretation of subtracting contribution of two scalars can be, at most, only viewed as a heuristic tentative explanation.
On the other hand, one could also come up with a different idea that these two scalars are conformal ghosts. This would explain why, at the end, Fradkin and Tseytlin had to take them into account when limiting to Weyl gravity (with enhanced conformal symmetry). These fields obviously were not present in generic HD gravities, but when one moves to conformal theory, one has to gauge-fix this symmetry and also supply proper ghosts (depending on the character of interactions, they have to be either FP ghosts of new local symmetry or third ghosts if HD terms are present, or both). It is true that the authors of [
96] performed an additional simple conformal gauge-fixing, requiring vanishing of the trace
. This performs the first part of the job of treating with due care the eminent conformal symmetry here. However, the next part is to deal with respective ghost fields of conformal symmetry, and, moreover, what is reassuring here is that since they are of Grassmannian nature (as all quantization ghosts are), then to add their contribution means effectively to subtract the contribution of “normal” particles. Hence, this is consistent with the idea that to reach conformal gravity, one has to add contribution of conformal ghosts, or, in a sense, subtract contribution of normal particles.
As it could be explained thoroughly in the next publications on this matter, these conformal ghost fields are here in four-dimensional conformal gravity of the type of third ghosts exclusively, namely, they are not FP conformal ghosts. However, what is more important here is that they are of scalar character, and they do not carry any Lorentz index. Moreover, as will be made clear in the next section, they are just minimally (covariantly, similar to in GR) coupled to the background and not conformally coupled, since such is the feature of the FP quantization procedure. This fits very well with the description of subtraction of two scalars; however, again, there could be some problems with such an interpretation. First, ghosts used during the covariant quantization procedure (that is, FP ghosts or third ghosts) are typically not counted for active perturbative degrees of freedom around a given background, hence they would not entail any change in the number of degrees of freedom between the two theories (limiting HD gravity and conformal gravity). This is in clear contradiction with the counting suggesting that the change is , while it is also in perfect agreement with other counting, where we have advocated before that there is no change and we have . One cannot also say that potential additional two degrees of freedom, as present in the standard counting of degrees of freedom of HD gravity, are these two conformal ghost modes (counted negatively). The ambiguity and the potential problem with this interpretation remain.
However, after all, the role of such ghost contributions (both whether these are FP or third ghosts) is not to change the number of perturbative active (physical to some extent) degrees of freedom, but to provide the correct expression for the partition function. For example, from such an expression, one could, at the one-loop level when expressed via determinants of some differential operators, obtain the one-loop partition function and, finally, read from it the number of degrees of freedom. This number typically agrees with the number obtained via counting of constraints and other similar canonical methods (and, for example, it gives eight degrees of freedom in scale-invariant HD gravity). However, on the level of such an expression written using determinants in various subspaces, it is not evident to understand the origin of various factors and functional traces there. As emphasized and, for example, explained in the case of conformal gravity, one can find there some various auxiliary elements of the theory used for quantization and for writing path integrals, such as contributions from Jacobians of change of field variables under path integrals, from gauge-fixing functionals, and also from FP and third ghost determinants. One does not find there contributions clearly assigned to modes appearing in the perturbative spectrum of the theory around flat spacetime, such as no clear origin with massless spin-2 or spin-1 fields. Eventually, from one-loop partition function, one finds the correct total number of degrees of freedom (including also these spin-2 and spin-1 massless states in conformal gravity). However, eventually, there is no sign that this counting manifestly takes FP ghosts or Jacobians into account. It only has to take them in intermediate levels of computation for the whole consistency of the procedure.
To answer the question of why there are two scalars, one has to recall a few facts. First, it is well known that the third ghost is just one single real field. This would already look like a contradiction to our counting, since we stated that there are two scalar contributions that need to be subtracted, but secondly, one has to pay enough attention to the fact that the mentioned contributions of minimally coupled real scalars are understood as coming from the following two-derivative scalar action coupled to curved spacetime background:
where we do not write any scalar potential term
V (containing mass terms or self-interactions) since it is well known that the terms in
V do not at all contribute to divergence proportional to
and
term. This is undoubtedly true at the one-loop order and for dimension-four counterterms in the effective action, which could receive contributions only from dimensionless couplings, but, for example, here they do not receive any contribution from quartic self-interactions of scalars. Only the form of the kinetic operator (precisely, how many derivatives there are) and the multiplicities of scalars (how many scalars there are) matter for this computation. There is a slight dependence of the contribution to the
counterterm on the non-minimal coupling
with gravity (that is, with a possible non-minimal term
, which is so-called coupling to a scalar curvature of spacetime), but these are all neglected for our purposes. One could say that, at the end, our model is very poor, since it describes here two massless scalar fields with no interactions, no self-interactions, and even without mass parameters, just coupled simplistically and minimally to the background geometry. Even the normalization of the scalar field here is not an issue for UV-divergences, but here we stick to the canonical one.
One has to compare the above with the contribution from the
matrix. To make the story short, the suitable operator in
is of the form
and this acts between two real scalar third ghosts for conformal symmetry
. Hence, due to fractionalization property of functional traces of a logarithm of such an operator as
, one sees that on the level of the trace of the logarithm, its contribution precisely agrees with two contributions from one scalar box
in a simple scalar field representation in the last part of the formula (
43). This explains the presence of two scalars. In more generality, one could say that, in general, even dimensionality of spacetime
d, the number of simple two-derivative scalars, each with action in (
43), the contribution of which have to be subtracted, has to be equal to half of the dimension
. This assertion can place our theoretical explanation in verification, for example, in six-dimensional conformal gravity models.
Within this second interpretation, we have a confirmation for why these two contributions have to be subtracted, but not added, and why they are scalar degrees of freedom, and not massless vectors or symmetric rank-2 tensors. We also see why they are minimally coupled to GR-covariant background and why they are two scalars. However, as remarked before, the whole existence or need for third ghosts (both of diffeomorphism symmetry or of local conformal symmetry) depends on the choice of the background. As we emphasized previously with some clever choice of partially covariant weighting functional , their contribution can be safely eliminated and they can be forgotten on the quantum level. Such is, for example, the situation around flat spacetime, when, in the gauge-fixing functional , one uses only partial derivatives, and from them builds scalar d’Alembertian operator , and on the contrary, one does not use background-covariant derivatives from GR to construct a proper GR-covariant box operator . Then, in such circumstances, there are no conformal third ghosts at all and so their contribution also does not exist and then this interpretation breaks down.
One could try to rescue this idea that the two scalars are third ghosts for conformal symmetry by claiming that for GR-covariant results of one-loop divergences obtained within some covariant framework, such as within the method of covariant Barvinsky–Vilkovisky traces, one ought to use only the form of the weighting functional , which is GR-covariant, so it must be built using the covariant d’Alembertian operator . Otherwise, one could be afraid of losing the general covariance of the results at the end, but they are guaranteed to be provided such that in the whole process of computation of them there are no steps which explicitly and manifestly violate the general covariance. Using only partially covariant functional (or in a sense forgetting about the third ghosts contributions) is not in line with such a formalism, and the results obtained by this oversight are, of course, incorrect. Hence, one can conclude that third ghosts must be remembered in all GR-covariant formalisms of computation, while one can still safely forget about them around flat spacetime and when working with Feynman diagrams.
However, here one must remember that the original computation due to Fradkin and Tseytlin was performed just in manifestly covariant framework and using background field method (BFM), when they sum various contributions and they do not use non-gauge-covariant results from some partial resummation of Feynman diagrams. It is well known that the two methods at the end should agree for final results for UV-divergences, which are proved to be expressed via gauge-invariant and gauge-fixing-independent counterterms. One must remark that in the covariant BFM formalism, as the Barvinsky–Vilkovisky (BV) trace technology is, the partial results, although looking generally covariant (so in this sense they are better looking than results of some subset of Feynman diagram contributions, which are not gauge-invariant), are without any sense if considered separately. Only their total sum has a clear physical meaning. Therefore, one cannot say with a good meaning, for example, what the generally covariantly written and unambiguous contribution of FP ghosts (or third ghosts) is to total divergences. Namely, it is true that they contribute and one usually should not forget about them, but their contribution alone is not physical. One should not be deluded by intermediate results, which look covariant, and would invite an interpretation of such partial results. Therefore, one must remember third ghosts for local conformal symmetry, but, again, their contribution alone is not very meaningful.
It is still a very interesting fact that their contribution (with all the provisos mentioned above) precisely agrees with the discontinuity between conformal gravity and a four-dimensional HD gravity after the limit , thus realizing the subtraction of two scalar degrees of freedom minimally coupled to background geometry. It remains yet to be seen if such a discontinuity has any physical meaning, or it is just an artifact of our clumsy way of approaching conformal gravity from generic HD gravity. Therefore, the interpretation of two scalars as two third ghosts for conformal symmetry in Weyl gravity is quite a plausible one, since it defends itself quite well, and certainly it is worthy studying further. For this, one should also better understand the issues related to conformal third ghosts.
Instead, here we summarize what is known about the contribution of two scalars to UV-divergences. This does not presuppose that these scalar fields are real or physical or whether they really live on a curved spacetime background. Nonetheless, their contribution can be singled out (with all the provisos above). Namely, two minimally coupled scalar fields (each with one degree of freedom and with a two-derivative action from (
43)) on a curved general background contribute twice the following contribution to UV-divergences,
The last term above, proportional to
, was written just for definiteness and we do not use it in the final main computation. The contribution of two scalars in (
42) can also be studied in the
sector of UV-divergences, although the final result is non-universal and it leads to a conformal gauge-dependent coefficient
, which was justifiably neglected when writing the final formula (
42).
5.4. Relation to Conformal Anomaly
When one speaks about UV-divergences at the one-loop level, one should not forget to relate them to the celebrated conformal anomaly. Basically, on the quantum level where there are UV-infinities and also resulting beta functions of couplings, there is an RG flow, and Green functions do not show perfect scale invariance. They depend on energy scale, and this dependence is dictated by RG phenomena. This means that inevitably in such situation, the conformal symmetry of the matter model (global group, as we discussed in
Section 2.1) is broken and we do not have a globally conformally invariant model on flat spacetime. This is also confirmed by analyzing conformal anomaly of global conformal symmetry in the setup of a curved geometric background (as explained in
Section 2.2). Then, the global group of scale transformations from GR in some matter models (which were globally conformally invariant) may be broken on the quantum level when one couples these matter theories to a non-trivial gravitational background. As for global anomaly of conformal symmetry, this is not a big deal in matter theories, where conformality on the classical level might be just a lucky coincidence, and then one does not view the effects of this anomaly as disastrous for the original matter theory. Simply, on the classical (tree-) level before the quantization, the conformal symmetry was present, but just after quantization of matter fields and matter symmetries (which could be also local internal symmetries) due to quantum effects, this symmetry is not there anymore and it does not constrain quantum dynamics anymore. One loses some power and simplicity of the theory but then the quantum physics enters and all correlations have to be tediously computed according to rules of QFT (in distinction from rules of CFT). One cannot use the rigidity of constraints placed on Green functions due to CFT models anymore, since one is out of their realm.
Technically speaking, the conformal anomaly (CA) understood as an anomaly for global conformal symmetry is equivalent to a trace anomaly. The last one signifies that the trace of the energy-momentum tensor of the matter theory does not vanish, although the theory on the classical level possessed scale invariance or even conformal invariance (before this was coupled to gravity). These last two requirements on the classical level imply that the classical EMT
11 is trace-free (
) upon using EOM from the matter sector, and this is completely tantamount to invariance of the theory under infinitesimal conformal transformations understood in the framework of GR (as they were studied in
Section 2.2). Therefore, we end up in the situation that on the quantum level, due to physical quantum effects (but not due to an act of some uncareful quantization), one finds that the trace of EMT on-shell is not zero anymore. For this, one has to use the quantum matter equation of motion obtained from the full quantum effective action restricted to the matter sector. One can specify the situation at the perturbative one-loop and then one correctly expects that this violation of tracelessness condition (thus, violation of scale invariance) is proportional to beta functions of the theory, which in turn are proportional to UV-divergences. Such a simplistic identification is no longer true at higher loop orders, but we can restrict all of the below analysis to this case only.
To summarize, there is an anomaly because some condition known from the classical theory (here, that ) is not present on the quantum level of the theory anymore. If some symmetry is on the classical and was expected to be a symmetry of the theory at all conditions, then its lack on the quantum level is a true anomaly in a physical sense, and this tells us that maybe we had a wrong expectation about the theory, and that theory with this symmetry is simply anomalous and very sensitive to the quantum effects, or that the original matter theory with conformal symmetry was not in a symmetric enough situation to protect the conformal symmetry from violation at the quantum level. Since, in the matter models, this conformal symmetry is not in a gauged form, there actually is not a big problem with such an anomaly. One has to simply accept that symmetries on the quantum level of the model are in the poorer form, and, for example, conformal symmetry is not realized in the quantum dynamics.
Still, in the matter models, but coupled to an external geometry, one expects that the lack of conformal symmetry on the quantum level manifests itself as the presence of non-trivial counterterms. Here, we can concentrate on UV-divergences in the gravitational sector, but one can also consider them in the matter part. For what we are going to say, this last restriction will be completely sufficient. Therefore, we can define that conformal anomaly
is exactly proportional to the
coefficient of UV-divergences of the theory at the one-loop of accuracy. Moreover, one can view it as proper trace of the quantum EMT defined from the effective action, that is, we have
It is important to realize that in the above expression, the divergent part of the effective action is taken in full generality as off-shell, so without using quantum (or even classical) EOM. For matter models, the divergent effective action there
is with all contributions when the matter fields are taken quantum (so they run in the loops of Feynman diagrams), while the gravity is treated as a classical external field to which the quantum matter system is coupled (so graviton lines can be only external lines of the diagrams). Moreover, in the above formula (
45), a regulator which was used to isolate UV-divergences was already extracted, so the expression for
(and also equivalently for
T) is finite and not UV-divergent.
The
coefficient of divergences is, of course, special to
dimensions because then counterterms have the energy dimension equal to the dimensionality of spacetime and couplings in front of them are dimensionless, so naively we already have classical scale invariance. Its relation to the divergent part of the one-loop effective action (part with classical scale invariance of the action) reads,
In the case of only gravitational counterterms that we are interested in, this coefficient is related to the beta functions of couplings in the following way:
One notices a few interesting facts about this expression for anomaly. First, the coefficient is not universal since it is conformal gauge-dependent (for example, in the background conformal gauge, its value is zero). Secondly, the beta function of the total derivative term is known to be ambiguous on the level of the trace of EMT, but not on the level of the coefficient. In the results we can forget about these last two terms and consider only universal contributions to two beta functions: and . Once again, we remind the reader that the presence of CA is not problematic for matter models, where this symmetry does not show up in the gauged form on the quantum level. One really sees, unambiguously, the presence of CA when, off-shell, the expression for it is not proportional to matter EOM.
One can see the consequences of CA for the quantum matter theory not only on the level of ultraviolet divergent expressions in correlation functions, but also in finite parts of these correlators on the quantum level computed to a given accuracy. For example, due to RG flow phenomena and within RG formalism (which basically tells in the simplest way that physical results should be independent on the arbitrary renormalization energy scale), one finds that the finite terms in the semi-classical general off-shell action for gravity (when quantum matter fields are integrated over) take the following form:
where, for simplicity, we wrote only contributions which are universal on the level of the trace of EMT
T related to the
coefficient. The non-local insertion of the logarithm of the GR-covariant box operator is both present between two Weyl tensors and also between Gauss–Bonnet invariant. The last one explicitly rewritten takes the form
which is in the direct analogy with the formula (
31). One sees that these contributions in (
48) explicitly break conformal symmetry, despite that on the level of divergent action, so without non-local logarithms of the box operator, they entail completely conformally invariant counterterms. This shows that in such a setting, the conformal symmetry is explicitly broken and it is no longer there on the quantum level. As one confirms in (
48), the expressions vanish when the corresponding beta functions vanish. Only there we are sure of quantum conformality present in some matter models. In the expressions in (
48), we see highly non-local functions, namely, of the logarithm of the covariant box operator. Their origin is due to RG invariance of the total effective action when the running of couplings is also taken into account and where
is a physical scale of renormalization which cancels with the
-dependence of running coupling parameters for any physical observable computed within the given model.
Thus far, we have only discussed some issues related to conformal anomaly when gravity was the external dynamical classical field. When we use the same procedure towards quantum gravity when it is both dynamical and quantum, and, for example, propagates inside quantum loops (where there is finally graviton’s propagator), and gravitons can be on internal lines of some Feynman diagrams, then in conformal gravity (so, the theory with local conformal symmetry), one has to be very careful since now this symmetry is in the gauged form. If one sees an anomaly, this might be a sign of a just creation of pathological, sick theory. This issue of how CA in conformal gravity may create disastrous effects was discussed at length in various previous works [
36,
37,
38,
39]. We will also discuss these issues and potential resolution of them at full length elsewhere.
Now, if the conformal symmetry is in the gauged (local) form, then one can adapt the discussion which was presented above to the total system composed of matter and gravity when both components are quantum and dynamical. The conformal symmetry was the defining local symmetry on the classical level of gravitational theory, hence the issue with potential conformal anomaly for it is very important and crucial to resolve satisfactorily. Moreover, local conformal symmetry is instrumental in defining the spectrum of the theory, for example, around flat spacetime, so any change with this would ruin the spectrum and classification of irreducible Lorentz representations of modes there. It would also destroy the counting of perturbative degrees of freedom of the theory. All these problems show in other disguise as violation of unitarity, not due to the HD character of gravitational theory, but due to destroying one of the local gauge symmetries, here in this case of local conformal symmetry. We will not comment on these perennial issues here, but we will refer to the vast literature on this topic.
In the case of a total gravitational system where matter is coupled to gravity, one has to analyze the total EMT of the full system. As is well known, it is composed of two pieces—gravitational and matter parts. One also notices that following Weinberg’s definition, the total EMT is zero on-shell as a consequence of invariance of the total action (gravitational part and coupled matter part) under GCT. Thus, on-shell we do not find any problem with CA in the total system. This remark applies both to the classical and quantum level, since in the last case, one can, for example, compute the effective EMT from the full quantum effective action functional, and for this argument to work, the gravitational theory or matter part alone do not have to be even classically conformally symmetric. This is indeed a very robust argument valid in any diffeomorphism invariant theory, which also preserves this symmetry on the quantum level (so, under the condition that there is no quantum anomaly of diffeomorphism symmetry). In any such theories on-shell, we find that the trace anomaly
T vanishes. However, as usual in QFT, the situation off-shell is also very important, and, actually, in these circumstances one can again relate this trace
to the coefficient
of divergences according to the last formulas (
45)–(
47). These formulas, also in the more general case, with quantum gravitational part remain valid. Of course, the respective beta functions now receive contributions from both quantum gravitational interactions and the quantum matter sector.
In the total system interacting gravitationally, the issue of detecting physical consequences from the expressions for is more complicated than just a case of simple matter theory. First, one knows that even on a classical level there exist some matter models for which the trace T does not vanish classically off-shell but only using matter EOM. Now, with gravity, we have this twist where on-shell as a kind of tautology. One must definitely analyze the situation quantum off-shell, but even if one finds there that , this does not necessarily imply that there is a conformal anomaly in the total system. Sometimes one finds for the total T off-shell an expression which does vanish, then one is sure of the absence of CA, or in the gravitational context, if the expression for it is proportional neither to the gravitational EOM nor to the EOM from the matter sector. These EOM should be derived from the respective quantum EOM originated from quantum effective action. This is a conformal Noether identity for the total system (matter + gravity).
One could also in parallel analyze the situation in six dimensions since then the space for all terms in conformal gravity is bigger. It is interesting viewing the situation with divergences in six-dimensional conformal gravity of the type
which still possesses the propagator around flat spacetime. We expect only three conformally invariant counterterms,
,
, and
, out of 10 possible terms in the action with six dimension operators [
98,
99]. There, we also expect that the conformal anomaly
is described in a conformally covariant way such that
and that only three terms,
,
, and
, out of 10 possible terms in the action with six dimension operators appear there. What about the remaining seven terms? If all this formalism is correct they should be made to vanish by just one conformal transformation in six dimensions with just one parameter
. This is quite improbable in a general situation to remove six counterterms (although they are related) by just one conformal transformation, but it should be true in a general case when the reasoning with conformal anomaly is correct. The UV-divergences after conformal transformations should be described by conformally invariant terms only in
, according to the general wisdom that terms in conformal anomalies are written in a way that preserves conformal symmetry (despite that they are in the anomaly heralding the disastrous breaking of this symmetry).